Loading [MathJax]/jax/element/mml/optable/SuppMathOperators.js
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, 2024, Volume 215, Issue 12, Pages 1666–1719
DOI: https://doi.org/10.4213/sm8815e
(Mi sm8815)
 

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

Strong asymptotics of the best rational approximation to the exponential function on a bounded interval

A. P. Magnus, J. Meinguet

Institut de Mathématique Pure et Appliquée, Université Catholique de Louvain, Louvain-la-Neuve, Belgium
References:
Abstract: We apply recent findings of complex approximation theory to best rational approximation of degree n to the function exp((n+ν)x) on a finite interval [0,c]. We show that the error norm behaves like the nth power of the main approximation rate times the νth power of a secondary approximation rate. The computation of the first rate is a consequence of works of Gonchar, Rakhmanov and Stahl done in the 1980s; the complete asymptotic description was achieved by Aptekarev in the first years of the 21st century. The solution is given in terms of elliptic integrals of the third kind.
Bibliography: 92 titles.
Keywords: rational approximation, exponential function, complex potential.
Received: 12.09.2016 and 21.08.2024
Bibliographic databases:
Document Type: Article
Language: English
Original paper language: Russian

En outre, M. Liouville s’était offert à m’exposer un abrégé de la nouvelle théorie des fonctions elliptiques, professée par lui au Collège de France.

(Moreover, M.Liouville agreed to lecture me on an abridged version of the new theory of elliptic functions taught by him in the Collège de France.)

From a report of P. L. Chebyshev on a journey abroad, in his ØE uvres, vol. 2, p. XIV.

§ 1. Introduction

1.1. The subject matter

Our subject is the asymptotic behaviour of best rational approximation to the function exp((n+ν)x) on a finite real interval [0,c].

This subject became heavily investigated when it appeared that the error norm decreases geometrically fast with the degree, even when the approximation interval is unbounded, c= [22], [89]. A full asymptotic description for c= and ν=0 was given by Gonchar and Rakhmanov as an application of their theory [36]. The full asymptotic description for this case was given by Aptekarev in 2002 [6].

We consider here a normally bounded interval [0,c] with c<.

One will establish statements like exp(x)rn(x)sn(x), with a formula for sn(x). The symbol ‘’ means here that AnBn as n if An/Bn1 then. When An and Bn depend on x, it usually means that the limit is reached uniformly in compacts in some set, sometimes uniformly in the whole set. We will sometimes encounter AnBn+Cn, where Bn and Cn oscillate with n. The meaning is then An=Un+Vn with UnBn and VnCn; see some cases in § 5.4.2.

Cleverly designed polynomial and rational approximations to holomorphic functions in some domain are known to involve a potential function related to the approximation region and a valid holomorphy domain, as will be recalled in § 2.

A proper scaling of the variable for the exponential function is necessary, we look at exp(A(n)x) on a fixed interval [0,c]. It appears then that A(n) must behave like a polynomial of first degree; see § 1.2.

We consider the rational approximation qn(z)/pn(z) of degree n to the exponential function exp((n+ν)z) on the bounded interval E=[0,c].

According to the Gonchar–Rakhmanov–Stahl theory [36], the root asymptotics

[pn(z)]1/nexp[Vp(z)],[qn(z)pn(z)exp((n+ν)z)]1/nCexp[2Vz(z)Vp(z)]
hold, where the complex potentials
Vp(z)=Flog(zt)dμp(t)andVz(z)=Elog(zt)dμz(t)
are made with the limit distributions of poles on an arc F joining a to b, where a and b depend on the point c (see §§ 3.2.1 and 3.2.4) and interpolation points on E. The resulting complex potential V(z)=Vz(z)Vp(z) must have a constant real part on E, and the real part of V(z)+z/2 must be (another) constant on F (the ‘external field’ problem). Hence C=exp(2V(a)a), so that the nth root of the error norm converges towards |exp(2V(0)2V(a)a)|.

Formulae for V, a, b and F will be given in § 3. The modulus k=sin(θ/2), where θ=arg(a/(ac)) (Figure 8), is of particular interest (see § 8, and also § 4). It will be found that k increases from 0 when c=0 to k=0.9089 when c=.

The geometry, or topological properties, of the limit locus of poles F is the same for c= as for c<: it is a single arc joining a to b.

Things change when we come to exp(x2), where it seems that F is now made of two arcs (see [52], § 6, [56], § 5, and [80], § 6.3), a case deserving more investigation.

Strong asymptotics are worked according to the theory of Aptekarev [6] (see (5.1) below):

pn(z)[(za)(zb)]1/4exp((n+12)Vz(z))×exp[(n+ν)V(z)+(ν12)V(z)]
and
pn(z)exp((n+ν)z)qn(z)Cn[(za)(zb)]1/4×exp[(n+12)Vz(z)+(n+ν)V(z)(ν12)V(z)],
where
V(z)=Czdtt(tc)(ta)(tb)
is an auxiliary complex potential such that the sum V(z)++V(z) is a constant on E=[0,c] and another constant on F following § 3.1 of [6] (the usual plane condenser problem without an exterior field; see also the end of § 2.1 below). Near the cuts E and F one must add the contributions from the two sides in the formulae above.

It is useful to establish that the various features V, a and so on, when c<, do converge to the better known corresponding items in the c= case [21], [36], [52]–[54], [89]. This is seen in the last rows of Table 1, in an incidental remark at the end of § 4.3.3, § 5.5.2, and in the second paragraph of § 5.5.3.

There are also other results for c=, including on the properties of best approximations to the function ϕ(z)=[ezj=0zj/j!]z+1 on (,0], which are due to Stahl and Schmelzer [81].

We remind in § 6 how the Adamyan–Arov–Krein (AAK) theory, initially a theory of approximation on the unit circle of meromorphic functions [2], has brilliantly been applied to approximation algorithms by Gutknecht and Trefethen [85], [86] in the 1980s.

We conclude the paper by considering some applications, of which some can be considered as open problems: method of electrostatic images, integral Hankel operator, quadratic relations, approximation by B-splines, ‘beyond ’, or exploring the modulus k>k.

1.2. Scaling

Best L rational approximation of degree n to exp(x) on [0,1] yields, for n=1,2,,5, the error norms 1.580103, 1.645106, 7.3451010, 1.8221013, 2.8751017 .

The approximations of degree 1, 3 and 5 have a real negative pole at 1.572, 4.176, 6.813 .

The asymptotic history of this approximation is quite short, as zeros and poles recede to on such a large scale that the interval [0,1] looks like a mere point, so that the approximation comes close to the Padé approximation. This approximant has a real pole which is almost C(n+1/2) for C=1.3255 (Driver and Temme 1999 [24], p. 10).

Also, the asymptotic formula for the error norm (see Meinardus [59], § 9.3, and Braess [16]) adapted1 here to approximation to exp(x) on [0,L] is

exqn(x)pn(x),[0,L]L2n+1(n!)2exp(L/2)24n+1(2n)!(2n+1)!=L2n+1exp(L/2)π(n+1/2)28n+2(Γ(n+3/2))22exp(L2)(Le16n+8)2n+1,
where we have also used Stirling’s formula.

The accuracy of the Meinardus–Braess formula is striking, as we have L=1, n=1: e1/2/384=1.579103; n=2: e1/2/368640=1.6453106.

We avoid the drift of the locus of poles by choosing a moving scale for the x-variable. A simple experiment is to freeze the real pole α of the odd degree approximants by an appropriate scaling βnx. It is found in Figure 1 that βn is strikingly close to a constant times n+1/2. This feature will be discussed at the end of § 5.4.1 and in § 5.5.1.

So we may stick to exp((n+1/2)x) or, more generally, to exp((γn+δ)x). There is no loss of generality in taking γ=1; we will work with exp((n+ν)x) from § 3 onward.

We now consider the approximation of exp((n+1/2)x) on a fixed interval [0,c]. With x changed to (n+1/2)x and c=L/(n+1/2), the error norm behaves like 2(ce/16)2n+1exp((n+1/2)c/2)=2[c2e2c/2/256]n+1/2 for moderately small c=L/(n+1/2)<41.325=5.3.

A more detailed view of the best rational approximation of degree 5 on [0,1] to exp(5.5x) is given in Figure 2.

§ 2. Potential function and approximation in the complex plane

2.1. Potential of a set of charged points

Let us consider a system of charged particles in the complex plane and a function of the form

nk=1wklog|zzk|,z,zkC,
which is the potential function of the system. Remark that a positive charge wk gives + at the particle position, and a negative charge creates a potential well at its position.

This writing is compatible with the notation Clog|zt|dν(t) seen before, by considering dν(t) to be a discrete measure.

As an example of how such potential functions enter approximation, consider polynomial interpolation at z1,,zn of a function defined in the complex plane by an integral f(z)=Cρ(t)dt/(zt), where C may be a contour or a system of arcs, and ρ(t)dt a real or complex measure (Markov- or Stieltjes-type functions); then the Hermite–Walsh formula (see [90], § 3.1) of the interpolation error is

f(z)pn(z)=Cnk=1zzktzkρ(t)dtzt,
so
|f(z)pn(z)|
where W_n(z)=n^{-1}\sum_k \log|z-z_k| is here the potential of a total negative unit charge on the z_ks.

Up to now, the language of potential seems only to be a shorthand for interpolation error function description. But consider now the problem of polynomial best approximation of such a function f on a set E. The interpolation points z_k on E must now be such that the error function has equal local extremal values of E, as in part (a) of Figure 2 (this is exact if f is real on the real set E). Let V_{n,E} be the relevant potential of a total negative unit charge on the z_k. Assuming the behaviour of the error function dominated by the \exp(nV_{n,E}(z)) term, we expect V_{n,E} to be close to the harmonic function V_E taking a constant value on E (as in part (b) of Figure 2). We are now at the centre of potential theory, the determination of a harmonic function in a domain through boundary values! Actually, V_E is the Green function of E with logarithmic singularity at \infty: V_E(z)-\log|z| is bounded as z\to\infty; see [28], Ch. II, § 3, and [90], Ch. 4.

For instance, if E=[a,b], then V_E(z)=\log|\Phi(z)|+\mathrm{const}, where

\begin{equation*} \Phi(z)=\frac{z-(a+b)/2+\sqrt{(z-a)(z-b)}}{|b-a|/2} \end{equation*} \notag
maps conformally the exterior of E to the exterior of the unit disc (and the reverse connection is \Phi+\Phi^{-1}= 2(2z-a-b)/|b-a|, the famous Joukowsky map).

Let p_n be the denominator of a rational approximation of degree n to a function f, with zeros p_1^{(n)},\dots, p_n^{(n)} in a set F, and q_n be the interpolation of p_n f at n+1 points z_1^{(n)},\dots, z_{n+1}^{(n)} in a set E. It is then possible to relate best rational approximations to functions analytic outside F and potential functions taking (different) constant values on the boundaries of E and F [8], [28], [29], [90], also seen as condenser potentials [7], [23], [31]. For instance, when E and F are intervals [a,b] and [c,d], the potential is the real part of

\begin{equation*} C\int_\infty^z \frac{dt}{\sqrt{(t-a)(t-b)(t-c)(t-d)}}, \end{equation*} \notag
which will be encountered in § 5.4.2 (see [28], [51], [57], [73], [74] for more details on the relationship between approximation theory and potential theory).

Remark that Bogatyrev prefers Riemann surfaces (as in the Akhiezer method) to Green functions ([12], p. XX).

2.2. Rational interpolation at 2n+1 points

Rational approximation of degree n defined by n poles and interpolation at n+1 points does not always explain best approximation performance. Instead, we consider now rational approximation of the numerator and denominator of degree n, the n poles being unknowns. These n new degrees of freedom allow normally interpolation at 2n+1 points instead of {n+1}.

Let \{f_n\} be a family of functions analytic in a region containing a contour C. So, for any z inside C,

\begin{equation} f_n(z)= \int_C \frac{\rho_n(t)\,dt}{z-t}, \end{equation} \tag{2.1}
with \rho_n(t)=-f_n(t)/(2\pi i). It may also happen that C shrinks towards an arc or a system of arcs, by deformation (through the point at infinity, if needed); \rho_n(t) is then the difference (f_n^-(t)-f_n^+(t))/(2\pi i) on the two sides of a cut, or even a real interval where the functions \rho_n are positive (then (2.1) is a (true) Markov function); see [32] and [35], and also the history in [71]. An example will be given at (5.7).

The best rational approximation of degree n on E to f_n is expected to interpolate \displaystyle f_n(z)= \int_C \rho_n(t)(z-t)^{-1}\,dt at 2n+1 points instead of n+1 points, say, z_1^{(n)},\dots, z_{2n+1}^{(n)}, so

\begin{equation} p_n(z)f_n(z)-q_n(z)=(z-z_1^{(n)})\dotsb(z-z_{2n+1}^{(n)}) \int_C \frac{p_n(t)\rho_n(t)\,dt}{(z-t)(t-z_1^{(n)})\dotsb(t-z_{2n+1}^{(n)})} \end{equation} \tag{2.2}
by the Hermite–Walsh formula again, and the numerator q_n seems to have degree 2n, with
\begin{equation*} q_n(z)= \int_C \frac{[(t-z_1^{(n)})\dotsb(t-z_{2n+1}^{(n)})p_n(z) -(z-z_1^{(n)})\dotsb(z-z_{2n+1}^{(n)}) p_n(t)]\rho_n(t)\,dt}{(z-t)(t-z_1^{(n)})\dotsb(t-z_{2n+1}^{(n)})}. \end{equation*} \notag
We find indeed2
\begin{equation*} q_n(z)= \int_C \frac{\{[ -z^{2n} +[z_1^{(n)}+\dotsb+z_{2n+1}^{(n)}-t]z^{2n-1}+\dotsb ]p_n(t)+\dotsb\}\rho_n(t)\,dt}{(t-z_1^{(n)})\dotsb(t-z_{2n+1}^{(n)})} \end{equation*} \notag
of degree 2n, unless the integrals of p_n(t), tp_n(t),\dots, t^{n-1}p_n(t) do vanish, that is, when p_n is orthogonal to the polynomials of degree <n on C with respect to the (probably complex) weight function
\begin{equation*} \frac{\rho_n(t)}{(t-z_1^{(n)})\dotsb(t-z_{2n+1}^{(n)})}, \end{equation*} \notag
where we have used the representation
\begin{equation*} (z-t)^{-1}=z^{-1}+tz^{-2}+\dotsb + t^{n-1}z^{-n} + t^nz^{-n}(z-t)^{-1}. \end{equation*} \notag

By replacing (z-t)^{-1} in (2.2) by its interpolation at the zeros of p_n, t=p_1^{(n)},\dots, p_n^{(n)}, the interpolation error is p_n(t)/[(z-t)p_n(z)], so

\begin{equation} \begin{aligned} \, &f_n(z)-\frac{q_n(z)}{p_n(z)} \nonumber \\ &\qquad= (z-z_1^{(n)})\dotsb(z-z_{2n+1}^{(n)}) \int_C \frac{p_n^2(t)\rho_n(t)\,dt}{(z-t)p_n^2(z)(t-z_1^{(n)})\dotsb(t-z_{2n+1}^{(n)})}. \end{aligned} \end{equation} \tag{2.3}
We define now
\begin{equation*} \begin{aligned} \, V_n(z)&= V_{\mathrm z}^{(n)}(z)-V_{\mathrm p}^{(n)}(z) \\ &=\frac{1}{2n+1}\sum_1^{2n+1}\log|z-z_k^{(n)}|-\frac{1}{n}\sum_1^n \log|z-p_k^{(n)}|; \end{aligned} \end{equation*} \notag
then
\begin{equation} \limsup_{n\to\infty} \biggl|f_n(z)-\frac{q_n(z)}{p_n(z)}\biggr|^{1/n} \leqslant \exp\Bigl[ 2V(z) - \min_{t\in C} (2V(t)-\phi(t))\Bigr], \end{equation} \tag{2.4}
where \phi(t)=\lim_{n\to\infty} n^{-1}\log|\rho_n(t)|, should V_n have the limit V as n\to\infty.

2.3. Conjectures and proofs

The conditions of existence and the properties of the limit of V_n as n\to\infty depend on our knowledge of the asymptotic behaviour of orthogonal polynomials with respect to complex weights, which has heavily been investigated [31], [36], [37], [49]–[51], [55], [58], [66], [70]–[72], [76], [77] and [79]. This research was mostly started by Gonchar, as early3 as the 1960s–1970s [30], [31], and culminated in famous conjectures [33], [34], mostly proved in the no less famous works of Stahl [76]–[78].

The final proof by Gonchar and Rakhmanov 1987 [36], [37] establishes that the orthogonal polynomials p_n of above do satisfy |p_n(z)|^{1/n}\to \exp(V_{\mathrm p}(z)) as n\to\infty, if the functions \rho_n of (2.1) are analytic in a domain containing C, and n^{-1}\log\rho_n has a limit \Phi as n\to\infty. Let \phi be the real part of \Phi and the final potential function

\begin{equation*} V(z)=V_{\mathrm z}(z)-V_{\mathrm p}(z) =\int_E \log|z-t|\,d\mu_{\mathrm z}(t)- \int_{F\subseteq C} \log|z-t|\,d\mu_{\mathrm p}(t) \end{equation*} \notag
be a constant, say, \gamma_E/2, on E; then 2V(z)-\phi(z)= is another constant, say, \gamma_F, on F\subseteq C, where \gamma_F>\gamma_E, and 2V-\phi has equal exterior normal derivatives on the two sides of F (the Stahl symmetry property, or S-curve property [58], [70], [71]) in a form adapted to the Cauchy-like formula (2.1) by Aptekarev; see [6], § 1.1.

Then, for the best rational approximations r_n on E, V_n\to V as n\to\infty, and \|f_n-r_n\|_E^{1/n} \to \rho:=\exp(\gamma_E-\gamma_F)<1. The convergence rate is also written as \rho=\exp(-2/\operatorname{cap}(E,F,\phi)), where

\begin{equation*} \operatorname{cap}(E,F,\phi) =\frac2{\gamma_F-\gamma_E}= \frac1{(V-\phi/2)_F -(V)_E} \end{equation*} \notag
is the weighted condenser capacity of the system (E,F,\phi), that is, the ratio of a charge (negative unit charge on E, positive on F) and an augmented potential difference from E to F [74]. Aptekarev considers also more complete, and more symmetric, systems (E,F,\phi,\psi) [6].

There is a wonderful association of the theory of best approximation and the theory of minimal capacity, at least in the weightless case (\phi(t)\equiv 0), but things are less obvious when \phi(t)\not\equiv 0 [19].

The complete theory and history are surveyed by Marcellán and Martínez-Finkelshtein [55], Martínez-Finkelshtein and Rakhmanov [58]; also see Rakhmanov [70], [71].

For strong asymptotics \lim \|f_n-r_n\|_E/\rho^n, see § 5.

2.4. Complex potential

Let -\sum_k w_k \log(z-z_k) be a complex potential, where one must make an appropriate choice of the complex logarithm, such as the usual convention of a real logarithm for a positive argument, with a discontinuity on the negative half line. Then, for real values of (z_k,w_k), the imaginary part of the potential is a staircase function of the real values of z, such as the staircase function shown in part (c) of Figure 2.

A test particle of positive unit charge is submitted to a force which is the complex conjugate of the derivative of the complex potential. Gauss started studies relating complex function theory to two-dimensional mathematical physics [62], [91].

Note however that the actual poles and interpolation points of a particular best rational approximation problem are normally NOT exactly the solutions of a problem of electrostatics. Only the LIMIT distributions \mu_{\mathrm p} and \mu_{\mathrm z} have a physical meaning. As an example, Figure 5 in § 6.2 shows a set of poles not far, but not exactly the same, from a set of equilibrium points found by a discretization of the continuous problem.

The conditions on \mathcal{V} are now (see [6], § 1.2):

From the Sokhotskyi–Plemelj relations (see [39], § 14.1), the limit values of

\begin{equation*} \mathcal{V}'(z)-\frac{\Phi'(z)}2 = \int_E\frac{d\mu_{\mathrm z}(t)}{z-t}- \int_F \frac{d\mu_{\mathrm p}(t)}{z-t}-\frac{\Phi'(z)}2 \end{equation*} \notag
on the two sides of F, namely,
\begin{equation*} \int_E \frac{d\mu_{\mathrm z}(t)}{z-t}- -\kern-10.5pt\int _F \frac{d\mu_{\mathrm p}(t)}{z-t} \pm \pi i \mu'_{\mathrm p}(z)-\frac{\Phi'(z)}2, \end{equation*} \notag
must be opposite, so,
\begin{equation} \mathcal{V}'(z)= \frac{\Phi'(z)}{2}\pm \pi i \mu'_{\mathrm p}(z), \qquad z\in F, \end{equation} \tag{2.5}
where \displaystyle -\kern-10.5pt\int means the principal value. Condition (VI) above means that the first integral of the right-hand side and the principal value add to \Phi'(z)/2 on F.

For \mathcal{V} itself,

\begin{equation*} \mathcal{V}(z)= \mathcal{V}(a)+\frac{\Phi(z)}2 -\frac{\Phi(a)}2 \mp i\pi\mu_{\mathrm p}(z) \end{equation*} \notag
on the right-hand side and left-hand side of F.

The central part of the theory is how to establish the asymptotic behaviour of the denominator p_n as the polynomial orthogonal to all polynomials of degree smaller than n with respect to the complex weight \exp(n(\Phi(t)-2\mathcal{V}_{\mathrm z}(t))) on F, as seen in § 2.2. As \displaystyle \mathcal{V}_{\mathrm p}(z)= \int_F\log(z-t)\,d\mu_{\mathrm p}(t) is the potential of the expected limit distribution of poles, p_n(z)\sim \exp(n\mathcal{V}_{\mathrm p}(z)) is the obvious starting formula. However, \mathcal{V}_{\mathrm p}(z) is singular on F (a discontinuous derivative when crossing F), and where are the zeros of p_n? An exponential function has no zeros.

A more accurate formula for z near F is

\begin{equation*} p_n(z)\sim A(z)\exp(n\mathcal{V}_{{\mathrm p},+}(z))+ B(z)\exp(n\mathcal{V}_{{\mathrm p},-}(z)), \end{equation*} \notag
referring to values on the right-hand side and the left-hand side of F.

Remind that the two parts making the complex potential function are written here \mathcal{V}(z)=\mathcal{V}_{\mathrm z}(z)-\mathcal{V}_{\mathrm p}(z), where

\begin{equation*} \mathcal{V}_{{\mathrm z},{\mathrm p}}(z)= \int_{E,F} \log(z-t)\,d\mu_{{\mathrm z},{\mathrm p}}(t), \end{equation*} \notag
\mu_{\mathrm z} and \mu_{\mathrm p} being positive measures of unit total weight on their supports E and F.

From above, -\mathcal{V}_{\pm}(z)= \mathrm{const}-\Phi(z)/2 \pm i\theta_{\mathrm q}(z), where \theta_{\mathrm q}(z)= \pi(\mu_{\mathrm p}(z)-\mu_{\mathrm p}(a)) increases from 0 to \pi when z runs from a to b on the right-hand side of F (a positive unit charge on F), and from \pi to 2\pi (or from -\pi to 0) when z returns to a on the left-hand side of F.

Remark also that d\theta_{\mathrm q}/dz\,{=}\, i(\mathcal{V}'(z)-\Phi'(z)/2) is equal to \pi\mu'_{\mathrm p}(z) on the right-hand side of F and to -\pi\mu'_{\mathrm p}(z) on the left-hand side.

So, as \mathcal{V}_{{\mathrm p},\pm}(z)= \mathcal{V}_{\mathrm z}(z)- \mathcal{V}_{\pm}(z) = \mathcal{V}_{\mathrm z}(z)- \Phi(z)/2\pm i\theta_{\mathrm q}(z), our estimate of p_n is

\begin{equation*} p_n(z)\sim \exp\biggl(n\mathcal{V}_{\mathrm z}(z)-\frac{n\Phi(z)}2\biggr)\bigl[A(z)\exp(n i\theta_{\mathrm q}(z))+ B(z)\exp(-n i\theta_{\mathrm q}(z))\bigr], \end{equation*} \notag
showing already a satisfactory oscillating behaviour on F! The still unknown functions A and B will follow from a check of orthogonality of p_n and all polynomials of degree <n. We use the test functions p_n(z)/(z-p) for the n zeros p of p_n. Then
\begin{equation*} \begin{aligned} \, &\int_F p_n(t)\frac{p_n(t)}{t-p} \exp(n\Phi(t)-(2n+1)\mathcal{V}_{\mathrm z}(t))\,dt \\ &\qquad \sim \int_F \bigl[A(t)\exp(n i\theta_{\mathrm q}(t))+ B(t)\exp(-n i\theta_{\mathrm q}(t))\bigr]^2 \frac{\exp(-\mathcal{V}_{\mathrm z}(t))\,dt}{t-p}. \end{aligned} \end{equation*} \notag
The integrals of the terms \exp(\pm 2n i\theta_{\mathrm q}(t)) are high index Fourier coefficients and therefore have a vanishing limit as n\to\infty. What remains is the integral of 2A(t)B(t)\exp(-\mathcal{V}_{\mathrm z}(t))/(t-p) and does not depend on n. A more complete discussion of the principal value integral is needed and will be done in § 5.4.2.

§ 3. The complex potential of the present problem

3.1. A theorem

Theorem 3.1. Let q_n/p_n be the best rational approximant of degree n to the exponential function \exp(-(n+\nu)x) on E=[0,c]. The limits

\begin{equation*} (p_n(z))^{1/n} \to \exp[\mathcal{V}_{\mathrm p}(z)] \end{equation*} \notag
and
\begin{equation*} (q_n(z)-p_n(z)\exp(-(n+\nu)z))^{1/n} \to C\exp[2\mathcal{V}_{\mathrm z}(z)-\mathcal{V}_{\mathrm p}(z)], \end{equation*} \notag
exist as n\to\infty, z\notin E\cup F, where F is an arc in the complex plane.

The complex potential \mathcal{V}=\mathcal{V}_{\mathrm z}-\mathcal{V}_{\mathrm p}, where

\begin{equation*} \mathcal{V}_{\mathrm p}(z)=\int_F \log(z-t)\,d\mu_{\mathrm p}(t) \quad\textit{and}\quad \mathcal{V}_{\mathrm z}(z)=\int_E \log(z-t)\,d\mu_{\mathrm z}(t), \end{equation*} \notag
satisfies \mathcal{V}(\infty)=0 and
\begin{equation} \sqrt{\frac{z(z-c)}{(z-a)(z-b)}}\, \mathcal{V}'(z) =\frac{-1}{2\pi i}\int_a^b \sqrt{\frac{t(t-c)}{(t-a)(t-b)}}\, \frac{dt}{z-t}, \end{equation} \tag{3.1}
where the integral is taken on the right-hand side of F=[a,b],4 the square roots being determined as being positive for large positive z and t and continuous outside the cuts E=[0,c] and F.

One also has

\begin{equation} \mathcal{V}''(z)= \frac{\nu_0 z +\nu_1 }{\sqrt{z^3 (z-c)^3 (z-a)(z-b) }}, \end{equation} \tag{3.2}
where
\begin{equation} \nu_0 x+\nu_1= \frac{1}{4\pi i}\int_a^b \sqrt{\frac{t(t-c)}{(t-a)(t-b)}} \biggl[\frac{ab(x-c)}{t}+\frac{(a-c)(b-c)x}{t-c}\biggr]\,dt. \end{equation} \tag{3.3}
The unit charge conditions on E and F lead to
\begin{equation} \int_a^b \frac{(\nu_0 x +\nu_1) dx}{\sqrt{x (x-c)^3 (x-a)(x-b) }}= \pi i . \end{equation} \tag{3.4}

Finally, the rate of error decrease is \rho, where

\begin{equation} \begin{aligned} \, \notag &\biggl\|\exp(-(n+\nu)x)-\frac{q_n(x)}{p_n(x)}\biggr\|_E^{1/n} \to \rho = \exp[ 2V(0)-2V(a)-\operatorname{Re}a] \\ &\qquad= \exp\biggl[2 \operatorname{Re}\int_0^a \frac{(\nu_0 t +\nu_1) \,dt}{\sqrt{t (t-c)^3 (t-a)(t-b) }}\biggr]. \end{aligned} \end{equation} \tag{3.5}

3.2. Proof of Theorem 3.1

In (2.1) we have

\begin{equation*} \exp(-(n+\nu)z)= \int_C \frac{\rho_n(t)\,dt}{z-t} \end{equation*} \notag
on a contour C containing z as an interior point, where
\begin{equation*} \rho_n(t)=- \frac{\exp(-(n+\nu)t)}{2\pi i}, \end{equation*} \notag
so that \Phi(t)=\lim n^{-1}\log\rho_n(t)=-t everywhere in the complex plane.

3.2.1. The first formula

The existence of limits follows from the Gonchar–Rakhmanov–Stahl theory recalled in § 2.3, which we apply to the present problem.

As E is a real set, we enjoy an important simplification by trying a symmetric set F with respect to the real axis, so that the potential V satisfies the Schwarz symmetry, or reflection, property V(\overline{z})=V(z). It also means that the complex potential \mathcal{V} and its derivative \mathcal{V}' at \overline{z} are the complex conjugates of \mathcal{V}(z) and \mathcal{V}'(z) (see [40], Theorem 7.7.2, and [65], § 5.5). On the two sides of E, as the real potential V is constant, its gradient is vertical, so \mathcal{V}' takes opposite purely imaginary values!

The two conditions on \mathcal{V}' on the two sides of E and F will almost immediately give a formula for the complex potential, up to a small number of unknown constants.

We assume F to be an arc joining a to b=\overline{a}, the two main unknown constants to be determined later on.

As \mathcal{V}'(z) must take opposite values on the two sides of [0,c], \sqrt{z(z-c)}\mathcal{V}'(z) is a meromorphic function outside the second cut F (an elementary instance of the homogeneous Privalov problem — see [39], § 14.8, Example 1 with d=1/2), hence the same is true after division by \sqrt{(z-a)(z-b)}, so that

\begin{equation*} \sqrt{\frac{z(z-c)}{(z-a)(z-b)}}\, \mathcal{V}'(z)= \frac{1}{2\pi i} \oint \sqrt{\frac{t(t-c)}{(t-a)(t-b)}}\, \frac{\mathcal{V}'(t)\,dt}{z-t} \end{equation*} \notag
on a contour shrinking to the two sides of the cut F. Let \mathcal{V}'(t) be the value on the right-hand side of F. When we proceed with the integral from b to a on the left-hand side, the square root of (t-a)(t-b) changes sign, so that we have to consider the sum of the values of \mathcal{V}'(t) =-1/2\pm \pi i\mu_{\mathrm p}'(t) (see (2.5)) on the two sides of F. Then (3.1) follows.

The values of \mathcal{V} near E and F give \mu_{\mathrm z} and \mu_{\mathrm p} from the Sokhotskyi–Plemelj relations seen above, and \mathcal{V}_{\mathrm z} and \mathcal{V}_{\mathrm p} can then be reconstructed. However, the curious quadratic relations (6.8) in § 6.3 do the job for our particular problem!

Remark that we do not need to know where F is: as \Phi is analytic in a region containing F (here, \Phi(z)=-z), the right-hand side of (3.1) is the same for any arc joining a and b without crossing E=[0,c]. Later on, the true set F will be found as a part of the locus where 2V(z)+x=2V(a)+\operatorname{Re}a=2V(b)+2\operatorname{Re}b; see Figure 3.

It will also be necessary to find a contour C such that 2V(z)-\phi(z)=2V(z)+x is larger on C\setminus F than its constant value on F.

3.2.2. The first equation for the parameters

From

\begin{equation*} \begin{aligned} \, \mathcal{V}(z) &= \int_E \log(z-t)\,d\mu_{\mathrm z}(t) -\int_F \log(z-t)\,d\mu_{\mathrm p}(t) \\ &= -\int_{E\cup F} \log(z-t)\,d\mu(t)= \frac{\mu_1}{z}+\frac{\mu_2}{2z^2}+\dotsb, \end{aligned} \end{equation*} \notag
as \mathcal{V}'(z) is only O(z^{-2}) at \infty, we have
\begin{equation} \int_a^b \sqrt{\frac{t(t-c)}{(t-a)(t-b)}}\,dt =0, \end{equation} \tag{3.6}
as a bonus!! (3.6) gives our first equation for a and b, knowing c, and another equation will be worked later on, from the unit charge condition
\begin{equation*} \int_E d\mu_{\mathrm z}(t)= \int_F d\mu_{\mathrm p}(t)=1. \end{equation*} \notag

When c=0, (3.6) is solved for b=-a, as we integrate the odd function t/\sqrt{t^2-a^2} over (a,-a). Moreover, as we know that F must be symmetric with respect to the real axis, we deduce that a and b are two opposite purely imaginary numbers. There is no more information in (3.6), but (3.1) yields the explicit formula

\begin{equation*} \frac{z}{\sqrt{z^2-a^2}}\, \mathcal{V}'(z) =\frac{-1}{2\pi i}\int_a^{-a}\frac{t}{\sqrt{t^2-a^2}}\, \frac{dt}{z-t}= \frac{1}{2} -\frac{z}{2\sqrt{z^2-a^2}} \end{equation*} \notag
from the Chebyshev–Markov example (see (5.7) below). So \mathcal{V}'(z) = -{1}/{2}+{\sqrt{z^2-a^2}}/({2z}) shows a potential well of charge -ia/2 at the origin, whence a=-2i, and therefore, for c=0,
\begin{equation} \mathcal{V}(z)= -\frac z2 +\frac{\sqrt{z^2+4}}2 -\log\frac{\sqrt{z^2+4}+2}z \end{equation} \tag{3.7}
has indeed a potential well \operatorname{const}+\log z, coming from the negative unit charge at z=0. Moreover, the derivative shows that \mathcal{V}'(z)+1/2 takes indeed opposite values on the two sides of any cut joining -2i to 2i and avoiding the origin. The actual line of poles is the locus where the two definitions of the square root of z^2+4 yield the same real part in the formula for \mathcal{V}(z) [24], so, where \mathcal{V}'(z)\,dz is purely imaginary, the quadratic differential (z^2+4)\,dz^2/z^2 is negative. For real negative z=-2\sinh\theta we have \cosh\theta-\log(\cosh\theta+1)=-\cosh\theta-\log(\cosh\theta-1), so 2\cosh\theta=2\log\coth(\theta/2), \theta=0.6219\dotsc, and z=\alpha=-1.3255\dotsc .

In general, given c\neq 0, we have a nonhomogeneous equation expressing that E carries a negative unit charge and F a positive unit charge. From \operatorname{Re}\mathcal{V}= constant on E and the Sokhotskyi–Plemelj relations (see [39], § 14.1) we have \mathcal{V}'({x\pm i0}) = \mp \pi i \mu'_i(x) on x\in E, and we will need formulae for \mathcal{V} to use the equality \mathcal{V}(c+ i0) - \mathcal{V}(i0)=-\pi i.

Near F a positive test particle is repelled, so the gradient of V(z)+x/2 is directed towards F, and so is the complex derivative \mathcal{V}'+1/2. On the right-hand side of F, \mathcal{V}'+1/2 has a negative real part, and the integral with respect to dz from a to b has a negative imaginary part.

3.2.3. The second derivative of the complex potential

We need more formulae for \mathcal{V} and its derivatives.

We build differential equation (3.2) for \mathcal{V} from (3.1): first we multiply by (z- a)(z-b), and let R(z)=\sqrt{z(z-c)(z-a)(z-b)}:

\begin{equation*} R(z) \mathcal{V}'(z) = \frac{1}{2\pi i}\int_a^b \frac{R(t)}{(t-a)(t-b)}\, \frac{t(z-t)+(t-a)(t-b)\,dt}{z-t}. \end{equation*} \notag
From (z-a)(z-b)= (t-a)(t-b)+(z-t)(z+t-a-b), using (3.6) as
\begin{equation*} \int_a^b \frac{R(t)\,dt}{(t-a)(t-b)}=0, \end{equation*} \notag
we divide by R(z) and differentiate and integrate by parts:
\begin{equation*} \begin{gathered} \, \mathcal{V}''(z)= \frac{1}{2\pi i}\int_a^b\biggl[ \frac{-tR'(z)R(t)}{(t-a)(t-b)R^2(z)} +\frac{[R'(t)/R(z)-R'(z)R(t)/R^2(z)] }{z-t}\biggr]dt, \\ R(z)\mathcal{V}''(z)= \frac{1}{2\pi i}\int_a^b R(t)\biggl[\frac{-tR'(z)}{(t-a)(t-b)R(z)} +\frac{[R'(t)/R(t)-R'(z)/R(z)] }{z-t}\biggr]dt, \end{gathered} \end{equation*} \notag
which is therefore a rational function of the poles 0, c, a and b, given that 2R'(z)/R(z)=1/z+1/(z-c)+1/(z-a)+1/(z-b). More precisely,
\begin{equation*} \begin{aligned} \, R(z)\mathcal{V}''(z) &= \frac{1}{2\pi i}\int_a^b R(t)\biggl[\frac{-t}{(t-a)(t-b)} \biggl( \frac{1}{2z}+\frac{1}{2(z-c)}+\frac{1}{2(z-a)}+\frac{1}{2(z-b)} \biggr) \\ &\qquad+\frac{1}{2tz}+\frac{1}{2(t-c)(z-c)}+\frac{1}{2(t-a)(z-a)} +\frac{1}{2(t-b)(z-b)} \biggr]\,dt \\ &=\frac{1}{2\pi i}\int_a^b \frac{R(t)}{(t-a)(t-b)}\biggl[ -\frac{(a+b)t-ab}{2tz}-\frac{(a+b-c)t-ab}{2(t-c)(z-c)} \\ &\qquad-\frac{b}{2(z-a)}-\frac{a}{2(z-b)} \biggr]\,dt \\ &=\frac{1}{4\pi i}\int_a^b \frac{R(t)}{(t-a)(t-b)}\biggl[ \frac{ab}{tz}+\frac{(a-c)(b-c)}{(t-c)(z-c)}\biggr]\,dt, \end{aligned} \end{equation*} \notag
using again (3.6). There is therefore no pole at z=a and b, the residues at 0 and c are the integrals of ab/t and (a-c)(b-c)/(t-c) divided by 4\pi i, and this leads to (3.2) and (3.3).

The final result (3.2) for \mathcal{V}''(z) is as in Gonchar and Rakhmanov (see [36], p. 323).

Here \nu_0 is the sum of the two residues, and \nu_1 is -c times the residue at 0; \nu_0 is also equal to 2\mu_1 from the behaviour of \mathcal{V}(z)=\mu_1/z+\dotsb for large z in § 3.2.2.

From the usual definition of the square root as positive for large positive z and continuous outside the cuts, the denominator of (3.2) is negative imaginary on the upper side of the cut [0,c], \mathcal{V}''(z)=-\pi i \mu''(z), where \mathcal V=\mathcal V_{\mathrm z}-\mathcal V_{\mathrm p} and \mu=\mu_{\mathrm z}-\mu_{\mathrm p}, behaves like ((\nu_0 c+\nu_1)/\sqrt{abc^3})(z-c)^{-3/2} near z=c,

\begin{equation*} \mathcal{V}'(z)=-\pi i \mu'(z) \sim -2\frac{\nu_0 c+\nu_1}{\sqrt{abc^3}}(z-c)^{-1/2} \ \ \text{and}\ \ \mathcal{V}(z)\sim \operatorname{const} -4\frac{\nu_0 c+\nu_1}{\sqrt{abc^3}}(z-c)^{1/2}. \end{equation*} \notag
As \mu' must be positive, \nu_0c+\nu_1<0. Near z=0,
\begin{equation*} \begin{gathered} \, \mathcal{V}''(z)=-\pi i \mu''(z) \sim \frac{\nu_1 i}{\sqrt{abc^3}}z^{-3/2}, \qquad \mathcal{V}'(z)=-\pi i \mu'(z) \sim -2\frac{\nu_1 i}{\sqrt{abc^3}}z^{-1/2}, \\ \mathcal{V}(z)\sim \operatorname{const} -4\frac{\nu_1 i}{\sqrt{abc^3}}z^{1/2} \end{gathered} \end{equation*} \notag
and \nu_1>0.

Finally, \mathcal{V}' and \mathcal{V} are

\begin{equation} \mathcal{V}'(z)= \int_\infty^z \frac{(\nu_0 x +\nu_1)\,dx}{\sqrt{x^3 (x-c)^3 (x-a)(x-b) }}, \end{equation} \tag{3.8}
and
\begin{equation} \mathcal{V}(z)= \int_\infty^z \frac{(\nu_0 x +\nu_1)(z-x)\,dx}{\sqrt{x^3 (x-c)^3 (x-a)(x-b)}} =z\mathcal{V}'(z) -\int_\infty^z \frac{(\nu_0 x +\nu_1)\,dx}{\sqrt{x (x-c)^3 (x-a)(x-b) }}. \end{equation} \tag{3.9}

The behaviour of \mathcal{V}(z) and its derivatives near a and b: let

\begin{equation*} \frac{\nu_0 b +\nu_1 }{\sqrt{b^3 (b-c)^3 (b-a) }}=R e^{i\Phi}, \end{equation*} \notag
so \mathcal{V}''(z) \sim R e^{i\Phi}(z-b)^{-1/2} near b, \mathcal{V}'(z)\sim -1/2+2R e^{i\Phi}(z-b)^{1/2}, as we know that \mathcal{V}'(a)=\mathcal{V}'(b)=-1/2, and \mathcal{V}(z) \sim \operatorname{const} -z/2+(4R/3) e^{i\Phi}(z-b)^{3/2}, where the constant is \mathcal{V}(b)+b/2. As the square roots must remain continuous outside the cut F, we follow z-b=|z-b|e^{i\theta} from \theta=\theta_0 on the right-hand side of F, to \theta=\theta_0+2\pi on the left-hand side. The real part of \mathcal{V}(z)+z/2-\mathcal{V}(b)-b/2 must vanish at the starting point, so \Phi+3\theta_0/2=\pi/2 and the real part of \exp(i\Phi+3\theta/2) is \cos(\pi/2+3(\theta-\theta_0)/2)=-\sin(3(\theta-\theta_0)/2) and is indeed negative for \theta-\theta_0<2\pi/3, positive between 2\pi/3 and 4\pi/3, and negative again between 4\pi/3 and 2\pi, explaining the pattern of signs in Figure 3.

Some values are in Table 1, their formulae will be established in § 4 of the present study.

Table 1.The values of c, \rho, a, k, \nu_0 and \nu_1

capproximation name\rhoa=\overline{b}k\nu_0\nu_1
0Padé0-2i0-20
c\to 0c^2 e^{2-c/2}/256c/2-2ic/4-2c
0.51/177.9340.234 - 1.992i0.1245-2.01550.565894
11/57.0700 0.437 - 1.970i0.2458-2.06081.27279
21/23.2287 0.746 - 1.890i0.4626-2.22483.11261
51/12.43301.088 - 1.646i0.7696-2.931411.1496
201/9.860481.183 - 1.447i0.8865-5.534483.6351
\infty‘1/9’1/9.289031.195-1.389 i0.9089-\infty\infty

3.2.4. The second equation for the parameters

On the two sides of the cut F we have

\begin{equation} \mathcal{V}'(z)=-\frac12 \pm \pi i \mu_{\mathrm p}'(z) \end{equation} \tag{3.10}
(from (2.5), where the Sokhotskyi–Plemelj relations are used again), or near any arc joining a and b and not crossing E=[0,c].

So, as seen above, the integral of \mathcal{V}'(z)+1/2 from a to b on the right-hand side of the cut has a negative imaginary part, and

\begin{equation} \mathcal{V}(b)-\mathcal{V}(a)= \frac{a-b}2 -\pi i \end{equation} \tag{3.11}
on the right-hand side of the cut F. From (3.9),
\begin{equation*} \mathcal{V}(b)-\mathcal{V}(a)= \underbrace{b\mathcal{V}'(b)-a\mathcal{V}'(a)}_{(a-b)/2} -\int_a^b \frac{(\nu_0 x +\nu_1) \,dx}{\sqrt{x (x-c)^3 (x-a)(x-b) }} =\frac{a-b}2-\pi i, \end{equation*} \notag
knowing that \mathcal{V}'(a)=\mathcal{V}'(b)=-1/2, our second equation is (3.4).

Remark that we could have taken in (3.9)

\begin{equation*} \mathcal{V}(z)= (z-c)\mathcal{V}'(z)-\int_\infty^z \frac{(\nu_0 x +\nu_1)\,dx}{\sqrt{x^3 (x-c) (x-a)(x-b) }}, \end{equation*} \notag
leading to
\begin{equation*} \int_a^b \frac{(\nu_0 x +\nu_1)\, dx}{\sqrt{x^3 (x-c) (x-a)(x-b) }} = -\pi i \end{equation*} \notag
as well, so that one must have
\begin{equation} \int_a^b \frac{(\nu_0 x+\nu_1)\,dx}{\sqrt{x^3(x-c)(x-a)(x-b)}} =\int_a^b \frac{(\nu_0 x +\nu_1)\,dx}{\sqrt{x (x-c)^3 (x-a)(x-b) }}, \end{equation} \tag{3.12}
and \mathcal{V}'(a)=\mathcal{V}'(b) means that
\begin{equation*} \int_a^b \frac{(\nu_0 x+\nu_1)\,dx}{\sqrt{x^3(x-c)^3(x-a)(x-b)}}=0. \end{equation*} \notag
But this is a consequence of (3.12): for
\begin{equation*} R(x)=\sqrt{x (x-c) (x-a)(x-b)}, \end{equation*} \notag
as before,
\begin{equation*} \int_a^b \frac{\nu_0 x+\nu_1}{xR(x)}\, dx -\int_a^b \frac{\nu_0 x+\nu_1}{(x-c)R(x)}\, dx=0 = -c\int_a^b \frac{\nu_0 x+\nu_1}{x(x-c)R(x)}\, dx=0. \end{equation*} \notag

As c\to +0, the right-hand side of (3.3) tends to

\begin{equation*} \frac{1}{4\pi i}\int_a^{-a} \frac{t}{\sqrt{t^2-a^2}}\,\frac{-2a^2x}{t}\,dt= \frac{a^2x}2; \end{equation*} \notag
and the left-hand side of (3.4) is half of the contour integral about the cut F of (\nu_0/x+\nu_1/x^2)/\sqrt{(x-a)(x-b)} resulting in \pi i times the residue \nu_0/\sqrt{ab} at x=0, so, \nu_0\sim -\sqrt{ab}=-|a|. This leaves \nu_0\sim a^2/2 \sim -|a| \to -2 as c\to 0, in agreement with numerical tests.

The rate of decrease of error norm is (3.5).5

§ 4. Elliptic integrals of first, second, and third kind

4.1. Change of variables

We follow Byrd and Friedman (see [20], p. 133).

A convenient transformation sending the four branch points z=0, c, a and b= \overline{a} to and from the symmetric set \{\mp 1, \mp i k'/k\} is

\begin{equation*} \begin{aligned} \, & v=\operatorname{cn} u = \frac{Az+B(z-c)}{Az-B(z-c)} \\ &\quad\Longleftrightarrow\quad z = -\frac{Bc}{A-B}+\frac{2ABc/(A-B)}{A+B-(A-B)v} =\frac{Bc(1+v)}{A+B-(A-B)v}, \end{aligned} \end{equation*} \notag
where A and B are the absolute values |a-c| and |a|. So,
\begin{equation} v=\biggl(\frac{z}{|a|}+\frac{z-c}{|a-c|} \biggr) \biggl(\frac{z}{|a|}-\frac{z-c}{|a-c|}\biggr)^{-1}. \end{equation} \tag{4.1}
In particular, at z=a and b,
\begin{equation*} \begin{aligned} \, v&=\frac{\exp(\pm i\arg a)+\exp(\pm i\arg(a-c))}{\exp(\pm i\arg a)-\exp(\pm i\arg(a-c))} \\ & = \mp i\cot \biggl[\arg\frac{a/(a-c)}2\biggr] = \mp i \cot\biggl(\frac\theta2\biggr), \end{aligned} \end{equation*} \notag
where \theta is the angle at a or b in Figure 8. These values for v= \operatorname{cn} u must be \pm i k'/k, so that s= \operatorname{sn} u = \pm \sqrt{1+k'^2/k^2}=\pm 1/k there. We also have e^{i\theta/2}= k'+ik.

As neither a nor b is known, we may as well take k and \zeta=(A-B)/(A+B). Then

\begin{equation*} \begin{gathered} \, A=\frac{(A+B)(1+\zeta)}2, \qquad B=\frac{(A+B)(1-\zeta)}2, \\ \begin{aligned} \, c^2&=A^2+B^2-2AB\cos\theta \\ &= \frac{(A+B)^2[1+\zeta^2-(1-\zeta^2)\cos\theta]}2 =(A+B)^2(k^2+\zeta^2k'^2), \end{aligned} \\ a,b = \frac{(1-\zeta)c (1\mp ik'/k)/2 }{1+\pm i\zeta k'/k}, \qquad a+b=\frac{(1-\zeta)c(k^2-k'^2\zeta)}{k^2+\zeta^2 k'^2}, \\ c-a-b= \frac{c\zeta}{k^2+\zeta^2 k'^2}, \qquad ab=\frac{(1-\zeta)^2c^2 }{4(k^2 +\zeta^2 k'^2) }. \end{gathered} \end{equation*} \notag

Also,

\begin{equation*} \begin{gathered} \, a,b= \frac{c}{1 -A(k'\pm i k)/B(k'\mp i k)} =\frac{c}{1 -|a-c|e^{\pm i\theta}/|a| }, \\ z=\frac{(1-\zeta)c(v+1)}{2(1-\zeta v)} = \frac{(1-\zeta^{-1})c}{2}+\frac{(1-\zeta^2)c}{2\zeta(1-\zeta v)}, \qquad z-c=\frac{(1+\zeta)c(v-1)}{2(1-\zeta v)}, \\ z-a,b = \frac{(1-\zeta^2)c(kv\pm i\zeta k')}{2(1-\zeta v)(k\pm i\zeta k')}, \qquad (z-a)(z-b)=\frac{(1-\zeta^2)^2 c^2 (k^2 v^2 +k'^2) }{4(1-\zeta v)^2 (k^2+\zeta^2 k'^2)} \end{gathered} \end{equation*} \notag
(see [20], p. 215, 361.54).

4.2.

Theorem 4.1. The best rational approximation problem of \exp(-(n+\nu)z) involves the sets E=[0,c] and F=[a,b], which are mapped by (4.1) onto [-1,1] and an arc joining -ik'/k to ik'/k avoiding [-1,1]; see Figure 4.

The first equation for the modulus k and \zeta=(|a-c|-|a|)/(|a-c|+|a|) is

\begin{equation} {\mathsf E} +(k^2-\alpha^2)\frac{{\mathsf K}-\Pi}{\alpha^2} =0, \end{equation} \tag{4.2}
where k'=\sqrt{1-k^2}, \alpha^2= -k^2(1-\zeta^2)/\zeta^2, \zeta^2= k^2/(k^2-\alpha^2) and
\begin{equation} \begin{gathered} \, \notag {\mathsf K}= \int_0^{\pi/2} \frac{d\varphi}{\sqrt{1-k^2\sin^2\varphi}}, \qquad {\mathsf E}=\int_0^{\pi/2}\sqrt{1-k^2\sin^2\varphi}\, d\varphi, \\ \Pi= \int_0^{\pi/2} \frac{d\varphi}{(1-\alpha^2\sin^2\varphi)\sqrt{1-k^2\sin^2\varphi}} = \int_0^{\mathsf K} \frac{du}{1-\alpha^2\operatorname{sn}^2\,u} \end{gathered} \end{equation} \tag{4.3}
are the complete elliptic integrals of the first, second and third kind.6 The second equation for k and \alpha (or k and \zeta) is
\begin{equation} {\mathsf E}({\mathsf K}-{\mathsf E})=\frac{\pi^2\zeta}{c(1-\zeta^2)} = \frac{\pi^2 k\sqrt{k^2-\alpha^2} }{-\alpha^2 c}. \end{equation} \tag{4.4}

Finally, the rate of decrease is given by

\begin{equation} \log\rho= \frac{\pi}{1+\zeta} \biggl[ -\frac{\zeta {\mathsf E}'}{{\mathsf K}-{\mathsf E}} +\frac{{\mathsf E}'-{\mathsf K}'}{{\mathsf E}} \biggr], \end{equation} \tag{4.5}
where K' and E' are defined by (4.3) for k replaced by k'.

4.3. Proof

4.3.1. The first equation

The integral (3.6) is a constant times

\begin{equation*} \int_{ik'/k}^{-ik'/k} \sqrt{\frac{1-v^2}{k^2 v^2+k'^2} } \frac{dv}{(1-\zeta v)^2}, \end{equation*} \notag
where integration goes from ik'/k to -ik'/k through \pm i\infty, or, for v=\operatorname{cn} u,
\begin{equation*} \begin{aligned} \, &\int_{-{\mathsf K}+i{\mathsf K}'}^{{\mathsf K}+i{\mathsf K}'} \frac{\operatorname{sn}^2\,u}{(1-\zeta\operatorname{cn}\,u)^2}\,du = \int_{-{\mathsf K}}^{\mathsf K} \frac{1}{(k\operatorname{sn}\,u -i\zeta\operatorname{dn}\,u)^2}\,du \\ &\qquad = 2\int_0^{\mathsf K} \frac{\zeta^2\operatorname{dn}^2\,u-k^2\operatorname{sn}^2\,u }{(k^2\operatorname{sn}^2\,u+\zeta^2\operatorname{dn}^2\,u)^2}\,du = 2\int_0^{\mathsf K} \frac{\zeta^2-k^2(1+\zeta^2)\operatorname{sn}^2\,u }{(\zeta^2+k^2(1-\zeta^2)\operatorname{sn}^2\,u)^2}\,du \\ &\qquad = \text{a constant times (4.2)}. \end{aligned} \end{equation*} \notag
Here we have used the formulae \operatorname{sn}(u+i{\mathsf K}')=1/(k\operatorname{sn} u) and \operatorname{cn}(u+i{\mathsf K}')=\operatorname{dn} u/(ik\operatorname{sn} u) (see [42], § 6, A42, and also [20], 362.15–16 and 410.07–08).

Equation (4.2) has exactly one root \alpha^2\in(-\infty, k^2) for -1<k<1, as the left-hand side of (4.2) is

\begin{equation*} {\mathsf E}-\int_0^{\pi/2} \frac{k^2-\alpha^2}{1-\alpha^2\sin^2\varphi}\,\frac{\sin^2\varphi\,d\varphi}{\sqrt{1-k^2\sin^2\varphi}}, \end{equation*} \notag
which is an increasing function of \alpha^2, starting with \mathsf{E}-\mathsf{K}<0 at \alpha^2=-\infty and reaching \mathsf{E}>0 at \alpha^2=k^2.

Incidentally (!), at \alpha=0 we have

\begin{equation*} {\mathsf E}-\int_0^{\pi/2} \frac{k^2\sin^2\varphi\,d\varphi}{\sqrt{1-k^2\sin^2\varphi}} ={\mathsf E}-({\mathsf K}-{\mathsf E})=2{\mathsf E}-{\mathsf K}, \end{equation*} \notag
known to vanish when c=\infty (the ‘1/9’ case). The value of the modulus is then
\begin{equation*} k_\infty= 0.9089085575485414782361189087447935049010139693404\dots \end{equation*} \notag
(see [53]).

When c is small, (4.2) is \alpha^2\pi/2 times (2/\pi)(\mathsf{E}-\mathsf{K})+2\Pi/\pi \sim -k^2/2 +1/|\alpha| (see [42], § 5, C1) and

\begin{equation*} \int \frac{du}{1-\alpha^2 \operatorname{sn}^2\,u} \sim \int_0^\infty \frac{du}{1-\alpha^2 u^2}= \frac{\pi}{2|\alpha|} \end{equation*} \notag
(for \alpha^2 strongly negative), so the left-hand side of (4.2) is about \alpha^2\pi/2 times {-k^2/2 +1/|\alpha|} whence \alpha^2 \sim -4/k^4, or \zeta\sim k^3/2.

We could not resist looking ‘beyond \infty’, when \alpha^2>0 in Table 2.

Table 2.\alpha^2, \mathsf K, \mathsf E and \Pi as functions of k from (4.2)

kapproximation name\alpha^22{\mathsf K}/\pi2{\mathsf E}/\pi2\Pi/\pi
0Padé-\infty110
k\to 0-4/k^41+k^2/41-k^2/4k^2/2
0.25-972.901.0161990.9841870.032076
0.50-49.1309391.0731820.9342150.143696
0.75-4.787741.2165740.8393650.465456
k_\infty=0.908909\dots‘1/9’01.4776260.7388131.477626
0.950.550371.6488520.7020142.746061
0.990.9284472.136880.65474813.9059

4.3.2. The second equation

We now keep track of all the constants in (3.4). From (3.3) for

\begin{equation*} t=\frac{c(1-\zeta)(v+1)}{2(1-\zeta v)} =\frac{(1-\zeta^{-1})c}{2}+\frac{(1-\zeta^2)c}{2\zeta(1-\zeta v)} \end{equation*} \notag
we have
\begin{equation*} \begin{aligned} \, &\nu_0 x+\nu_1= \frac{1}{4\pi i}\int_{ik'/k}^{-ik'/k} \sqrt{\frac{(v^2-1)(k^2+\zeta^2k'^2)}{(1-\zeta^2)(k^2 v^2+k'^2)} } \\ &\qquad\qquad \times \biggl[\frac{2ab(1-\zeta v)(x-c)}{(1-\zeta)c(v+1)} +\frac{2(a-c)(b-c)(1-\zeta v)x}{(1+\zeta)c(v-1)}\biggr] \frac{(1-\zeta^2)c\, dv}{2(1-\zeta v)^2}, \end{aligned} \end{equation*} \notag
where integration goes from ik'/k to -ik'/k through \pm i\infty, or, for v=\operatorname{cn} u and \chi=(-1/(4\pi)) \sqrt{(k^2+\zeta^2k'^2)/(1-\zeta^2)},
\begin{equation*} \begin{aligned} \, &\nu_0 x +\nu_1= \chi \int_{-{\mathsf K}+i{\mathsf K}'}^{{\mathsf K}+i{\mathsf K}'} [ (1+\zeta)ab(x-c)(1-\operatorname{cn} u) \\ &\qquad\qquad\qquad - (1-\zeta) (a-c)(b-c)x(1+\operatorname{cn} u)] \frac{du}{1-\zeta\operatorname{cn} u} \\ &= \chi\int_{-{\mathsf K}}^{{\mathsf K}} \frac{(1\,{+}\,\zeta)ab(x-c)(ik\operatorname{sn} u\,{-}\operatorname{dn} u)\,{-}\, (1\,{-}\,\zeta) (a\,{-}\,c)(b\,{-}\,c)x(ik\operatorname{sn} u\,{+}\operatorname{dn} u)}{ik\operatorname{sn} u-\zeta\operatorname{dn} u}\,du \\ &=2\chi\int_0^{\mathsf K} \!\frac{ (1{\kern1pt}{+}\,\zeta)ab(x{\kern1pt}{-}{\kern1pt}c)(k^2\!\operatorname{sn}^2 \!u{\kern1pt}{+}{\kern1pt}\zeta \!\operatorname{dn}^2 \!u){\kern1pt}{-}{\kern1pt}(1\!-\!\zeta) (a{\kern1pt}{-}{\kern1pt}c)(b{\kern1pt}{-}{\kern1pt}c)x(k^2\!\operatorname{sn}^2 \!{-}{\kern1pt}\zeta\!\operatorname{dn}^2 \!u)}{k^2\operatorname{sn}^2 u+\zeta^2\operatorname{dn}^2 u = \zeta^2(1-\alpha^2\operatorname{sn}^2 u)}\,du \\ &=2\chi\biggl[ (1+\zeta)ab(x-c)\biggl(\frac{{\mathsf K}}{1+\zeta}+\frac{\Pi}{\zeta(1+\zeta)}\biggr) \\ &\qquad\qquad - (1-\zeta) (a-c)(b-c)x\biggl(\frac{{\mathsf K}}{1-\zeta} -\frac{\Pi}{\zeta(1-\zeta)}\biggr)\biggr] \\ &=\frac{-c^2}{8\pi \sqrt{(1-\zeta^2)(k^2+\zeta^2k'^2)}}\biggl[ (1-\zeta)^2(x-c)\biggl({\mathsf K}+\frac{\Pi}{\zeta}\biggr) - (1+\zeta)^2 x\biggl({\mathsf K} -\frac{\Pi}{\zeta}\biggr)\biggr] \\ &=\frac{-c^2}{8\pi\zeta} \sqrt{\frac{1-\zeta^2}{k^2+\zeta^2k'^2}}\biggl[ (1-\zeta)(x-c)({\mathsf K}-(1-\zeta){\mathsf E}) + (1+\zeta) x({\mathsf K}-(1+\zeta){\mathsf E})\biggr], \end{aligned} \end{equation*} \notag
where we use the representations
\begin{equation*} ab=\frac{(1-\zeta)^2 c^2 }{4 (k^2+\zeta^2 k'^2)} \quad\text{and}\quad (a-c)(b-c)=\frac{(1+\zeta)^2 c^2 }{4 (k^2+\zeta^2 k'^2)} \end{equation*} \notag
from above and equality (4.2) in the form \Pi=\mathsf{K}-(1-\zeta^2)\mathsf{E}, so
\begin{equation} \begin{aligned} \, \nu_0& =\frac{-c^2}{4\pi\zeta} \sqrt{\frac{1-\zeta^2}{k^2+\zeta^2k'^2}}[{\mathsf K}-(1+\zeta^2){\mathsf E}], \\ \nu_1& =\frac{c^3}{8\pi\zeta} \sqrt{\frac{1-\zeta^2}{k^2+\zeta^2k'^2}}(1-\zeta)l[ {\mathsf K}-(1-\zeta){\mathsf E}]. \end{aligned} \end{equation} \tag{4.6}

We now need (3.4) and (3.12) to obtain

\begin{equation*} \begin{aligned} \, \pi i &=\int_a^b \frac{ (\nu_0 +\nu_1/x)\,dx}{\sqrt{x (x-c) (x-a)(x-b) }} \\ &= \int_{ik'/k}^{-ik'/k} \frac{\bigl[\nu_0 +\nu_1 2(1-\zeta v)/((1-\zeta)c(v+1))\bigr](1-\zeta^2)c/(2(1-\zeta v)^2)} {\sqrt{((1-\zeta^2)^3 c^4 (v^2-1) (k^2 v^2 +k'^2))/(16 (1-\zeta v)^4 (k^2+\zeta^2 k'^2))}} \,dv \end{aligned} \end{equation*} \notag
and
\begin{equation*} \begin{aligned} \, \pi&=2\sqrt{\frac{k^2+\zeta^2k'^2}{1-\zeta^2}} \\ &\qquad\times\int_{ik'/k}^{-ik'/k}\frac{\bigl[\nu_0 +\nu_1 2(1-\zeta v)/((1-\zeta)c(v+1))\bigr] (\,dv=-\operatorname{sn}\,u \operatorname{dn}\,u\,du) } {c\sqrt{ (1-v^2) (k^2 v^2 +k'^2) }} \\ &= \frac{-2}{c}\sqrt{\frac{k^2+\zeta^2k'^2}{1-\zeta^2}}\ \int_{-{\mathsf K}+i{\mathsf K}'}^{{\mathsf K}+i{\mathsf K}'} \biggl[\nu_0 +\nu_1 \frac{2(1-\zeta \operatorname{cn}\,u)}{(1-\zeta)c(1+\operatorname{cn}\, u)}\biggr]du \\ &= \frac{-2}{c}\sqrt{\frac{k^2+\zeta^2k'^2}{1-\zeta^2}}\ \int_{-{\mathsf K}}^{{\mathsf K}} \biggl[\nu_0 +\nu_1 \frac{2(k\operatorname{sn}\, u+i\zeta \operatorname{dn}\,u)}{(1-\zeta)c(k\operatorname{sn}\, u-i\operatorname{dn}\, u)}\biggr]du \\ &= \frac{-2}{c}\ \int_0^{\mathsf K} \biggl[ \frac{-c^2}{2\pi\zeta} ({\mathsf K}-(1+\zeta^2){\mathsf E}) +\frac{c^2}{2\pi\zeta} ({\mathsf K}-(1-\zeta){\mathsf E}) (-\zeta+(1+\zeta)k^2\operatorname{sn}^2u)\biggr]du \\ &= \frac{c}{\pi\zeta}\ [{\mathsf K} ({\mathsf K}-(1+\zeta^2){\mathsf E}) - ( {\mathsf K}-(1-\zeta){\mathsf E}) (-\zeta {\mathsf K}+(1+\zeta)({\mathsf K}-{\mathsf E}))] \\ &= \frac{c}{\pi\zeta}\ (1-\zeta^2){\mathsf E}({\mathsf K}-{\mathsf E}) \end{aligned} \end{equation*} \notag
by using \displaystyle\int_0^{\mathsf K} \operatorname{sn}^2u\,du=({\mathsf K}-{\mathsf E})/k^2 (see [1], formula 16.26.1), and (4.4) follows, yielding with (4.2) the two equations for the two real unknowns k and \alpha^2 (or k and \zeta). Table 1 was done using numerical solutions of equations (4.2) and (4.4) (and (4.8) for \rho).

As c is a continuous function of k and \zeta, the existence of a solution follows. It is even possible to look at k>0.9089\dotsc, where c<0, the meaning of this region being a small enigma (see § 8).

It is also strongly suspected, but without proof here, that c is an increasing function of k and \zeta, implying the uniqueness of the solution.

When c\to 0, we expect k and \zeta to tend to 0 and \mathsf{K}-\mathsf{E} to be very small \sim \pi k^2/4, and we know that \zeta\sim k^3/2, so, c\sim 4k.

A curious consequence of (4.4) is the quadratic relation

\begin{equation} \frac{(\nu_0 c+\nu_1)^2}{(c-a)(c-b)} -\frac{\nu_1^2}{ab} = -\frac{c^3}{4}. \end{equation} \tag{4.7}

Indeed, from

\begin{equation*} ab=\frac{(1-\zeta)^2 c^2 }{4 (k^2+\zeta^2 k'^2)}, \qquad (a-c)(b-c)=\frac{(1+\zeta)^2 c^2 }{4 (k^2+\zeta^2 k'^2)} \end{equation*} \notag
and (4.6),
\begin{equation*} \frac{(\nu_0 c+\nu_1)^2}{(c-a)(c-b)} = \frac{c^4(1-\zeta^2)}{16\pi^2\zeta^2} [-{\mathsf K}+(1+\zeta){\mathsf E}]^2\quad\text{and} \quad \frac{\nu_1^2}{ab}=\frac{c^4(1-\zeta^2)}{16\pi^2\zeta^2} [{\mathsf K}-(1-\zeta){\mathsf E}]^2, \end{equation*} \notag
so that the difference is (c^4(1-\zeta^2)/(16\pi^2\zeta^2)4\zeta) {\mathsf E}({\mathsf E}-{\mathsf K}) and the result follows from (4.4).

4.3.3. The rate of the error norm decrease

From (3.5)

\begin{equation} \begin{aligned} \, \notag \log\rho &=2 \operatorname{Re}\int_0^a \frac{ (\nu_0 t +\nu_1)\, dt}{\sqrt{t (t-c)^3 (t-a)(t-b) }} \\ \notag &=2\operatorname{Re}\frac{4(1-\zeta)}{c^2}\sqrt{\frac{k^2+\zeta^2k'^2}{(1-\zeta^2)^3}} \\ \notag &\qquad\times \int_{-1}^{-ik'/k} \frac{ (1-\zeta v) \bigl[\nu_0 (1-\zeta)c(v+1)/(2(1-\zeta v))+\nu_1\bigr]\,dv} {(v-1)\sqrt{(v^2-1)(k^2v^2+k'^2)}} \\ &=2\operatorname{Re}\frac{4(1-\zeta)}{c^2}\sqrt{\frac{k^2+\zeta^2k'^2}{(1-\zeta^2)^3}} \notag \\ \notag &\qquad\times \int_{-1}^{-ik'/k} \frac{ 2(1-\zeta)(\nu_0 c+\nu_1) +((1-\zeta)\nu_0 c-2\zeta\nu_1)(v-1) } {2(v-1)\sqrt{(v^2-1)(k^2v^2+k'^2)}}\,dv \\ \notag &=2\operatorname{Re}\frac{4i(1-\zeta)}{c^2} \biggl[ \int_{2{\mathsf K}}^{-{\mathsf K}+i{\mathsf K}'}\frac{2c^3((1+\zeta){\mathsf E}-{\mathsf K})}{8\pi\zeta(\operatorname{cn}\ u-1)}\,du -\frac{c^3({\mathsf K}-{\mathsf E})}{4\pi\zeta} (-3{\mathsf K}+i{\mathsf K}')\biggr] \\ \notag &=2\operatorname{Re}4ic(1-\zeta) \\ \notag &\qquad\times \biggl[ -\frac{(1+\zeta){\mathsf E}-{\mathsf K}}{4\pi\zeta} \underbrace{[\Delta(u)-E(u)]_{2{\mathsf K}}^{-{\mathsf K}+i{\mathsf K}'}}_{ -3{\mathsf K}+i{\mathsf K}' +3{\mathsf E}-i({\mathsf K}'-{\mathsf E}') } -\frac{{\mathsf K}-{\mathsf E}}{4\pi\zeta} (-3{\mathsf K}+i{\mathsf K}')\biggr], \\ \log\rho &=-\frac{c(1-\zeta)}{\pi\zeta} [ ((1+\zeta){\mathsf E}-{\mathsf K}){\mathsf E}' + ({\mathsf K}-{\mathsf E}) {\mathsf K}' ]. \end{aligned} \end{equation} \tag{4.8}
From [20], 119.02, p. 26, Figure 12, and 361.51, and [5], pp. 81 and 82, equality (4.5) follows.

When c is small, k\sim c/4, \mathsf{E} and \mathsf{K}\sim\pi/2, \mathsf{K}-\mathsf{E}\sim \pi k^2/4\sim \pi c^2/64, \zeta\sim k^3/2 \sim c^3/128, \mathsf{E}'\sim 1 and \mathsf{K}'\sim \log(4/k) \sim \log(16/c), so,

\begin{equation*} \log\rho\sim -c\biggl[\frac12 -\frac{\pi c^2/64}{\pi c^3/128} \biggl[1-\log\frac{16}c\biggr]\biggr]= -\frac c2 +2-2\log\frac{16}c, \end{equation*} \notag
the same as the Meinardus–Braess results in § 1.2.

And when c\to\infty, the ’1/9’ case, it is better to use (4.4):

\begin{equation*} \log\rho= \frac{\pi}{1+\zeta} \biggl[ \frac{{\mathsf K}-(1+\zeta){\mathsf E}}{{\mathsf E}({\mathsf K}-{\mathsf E})}{\mathsf E}' -\frac{{\mathsf K}'}{{\mathsf E}} \biggr]. \end{equation*} \notag

If c\to\infty and \zeta\to 1, and we know that \mathsf{K}-2\mathsf{E}\to 0, then the relation \log\rho\sim -\pi \mathsf{K}'/\mathsf{K} remains!

§ 5. Strong asymptotics

5.1. Introduction

We considered expressions like X_n=A_n e^{B_n}, where A_n is convergent, and where B_n increases linearly, say, B_n=nB^*+O(1). The root asymptotics deals only with e^{B^*}= the limit of X_n^{1/n}. We now try the more accurate asymptotic expression X_n\sim A_\infty e^{nB^*+B^{**}}, meaning that

\begin{equation*} \lim_{n\to\infty} \frac{X_n}{A_\infty \exp(nB^*+B^{**})}=1, \end{equation*} \notag
where A_\infty and B^{**} are the limits of A_n and B_n-nB^* if these limits exist.

Consider

\begin{equation*} f_n(z)= \int_C \frac{\rho_n(t)\,dt}{z-t}, \end{equation*} \notag
where \rho_n(t)= \exp(-(n+\nu)t), so that, as in (2.1), f_n(z)=-2\pi i \exp(-(n+\nu)z) for z inside the contour C containing the F-cut.

We look for more than just the main behaviour with respect to n of the rational approximation, that is why the parameter \nu must not be neglected.

We proceed with accurate asymptotic descriptions of p_n, q_n, p_nf_n-q_n and \|f_n-q_n/p_n\|_E/\rho^n.

5.2. Theorem

Theorem 5.1. The best rational approximation q_n/p_n of degree n to \exp(-(n+ \nu)x) on [0,c] satisfies

\begin{equation} \begin{aligned} \, \notag &p_n(z) \sim [(z-a)(z-b)]^{-1/4}\exp\biggl((n+\nu)\mathcal{V}_{\mathrm p}(z)-\biggl(\nu-\frac12\biggr)\mathcal{V}_{*,{\mathrm p}}(z) \biggr), \\ \notag &p_n(z)\exp(-(n+\nu)z)-q_n(z) \sim C_n [(z-a)(z-b)]^{-1/4} \\ &\qquad\times \exp\biggl[(n+\nu)(2 \mathcal{V}_{\mathrm z}(z)- \mathcal{V}_{\mathrm p}(z)) -\biggl(\nu-\frac12\biggr)(2\mathcal{V}_{*,{\mathrm z}}(z)-\mathcal{V}_{*,{\mathrm p}}(z)) \biggr] \end{aligned} \end{equation} \tag{5.1}
as n\to\infty, where \mathcal{V} is the complex potential introduced in § 3 and \mathcal{V}_* =\mathcal{V}_{*,{\mathrm z}}-\mathcal{V}_{*,{\mathrm p}} is the auxiliary complex potential
\begin{equation} \mathcal{V}_*(z)= C \int_\infty^z \frac{dt}{\sqrt{t(t-c)(t-a)(t-b)}}, \qquad C=\sqrt{\frac{1-\zeta^2}{k^2+\zeta^2k'^2}}\, \frac{\pi c}{4{\mathsf K}}, \end{equation} \tag{5.2}
and where C_n=\exp(-(n+\nu)(2\mathcal{V}(a)+a)+(2\nu-1)\mathcal{V}_*(a)).

On F one must add the contributions from the two sides of F:

\begin{equation*} \begin{aligned} \, p_n(z) &\sim [(z-a)(z-b)]^{-1/4}\exp\biggl[(n+\nu)\mathcal{V}_{\mathrm z}(z)-\biggl(\nu-\frac12\biggr)\mathcal{V}_{*,{\mathrm z}}(z)\biggr] \\ &\qquad \times \biggl\{\exp\biggl[-(n+\nu) \mathcal{V}_{+}(z) +\biggl(\nu-\frac12\biggr)\mathcal{V}_{*,+}(z)\biggr] \\ &\qquad\qquad +\exp\biggl[-(n+\nu) \mathcal{V}_{-}(z) +\biggl(\nu-\frac12\biggr)\mathcal{V}_{*,-}(z)\biggr]\biggr\},\qquad z\in F. \end{aligned} \end{equation*} \notag

Finally, the error norm satisfies

\begin{equation} \biggl\| \exp(-(n+\nu)z)-\frac{q_n(z)}{p_n(z)}\biggr\|_E \sim 2\rho^{n+\nu}\rho_*^{1/2-\nu}, \end{equation} \tag{5.3}
where \rho=\exp(-2V(a)-a+2V(0)) is the main rate of decrease (see (4.5)(4.8)) and
\begin{equation} \rho_*= \exp(2(V_*(0)-V_*(a)))= \exp\biggl(-\frac{\pi {\mathsf K}'}{{\mathsf K}}\biggr). \end{equation} \tag{5.4}

5.3. Lines of poles

The polynomial p_n(z) has strong asymptotics of the form \exp(nB^*(z)+B^{**}(z)), which is discontinuous on any cut joining a to b; on the other hand p_n makes no jumps, so the asymptotic behaviour of p_n is satisfactorily described by a linear combination of the values of \exp(nB_+^*(z)+B_+^{**}(z)) and \exp(nB_-^*(z)+B_-^{**}(z)) on the two sides of the cut, just as in the case of the Liouville–Green–Steklov–WKB problem ([11], § 10.1). Poles of the rational approximant, that is, zeros of P_n occur when the two exponentials have the same absolute value, so when nB_+^*(z)+B_+^{**}(z) and nB_-^*(z)+B_-^{**}(z) have the same real part. As n\to\infty, the limit locus is given by B_+^*(z)-B_-^*(z) being purely imaginary. This limit locus is reached especially fast if B_+^{**}(z)-B_-^{**}(z) is also purely imaginary on the limit locus. This happens obviously in (5.5) and also above when \nu=1/2. This explains the n+1/2 phenomenon of Figure 1 for c=0. The same conclusion holds for c\neq 0 too, as seen from the second formula in (5.1).

5.4. Proof of Theorem 5.1

5.4.1. Padé case

Consider first the limit case c=0.

The Padé approximant of degree n to \exp(-(n+\nu)z) is known to be

\begin{equation*} \frac{ {}_1F_1(-n,-2n,-(n+\nu)z) }{ {}_1F_1(-n,-2n,(n+\nu)z) } \end{equation*} \notag
(see [9], (5.39), and [69], § 75), where the {}_1F_1 expansions are limited to their first n+1 terms.

The monic denominator is

\begin{equation*} p_n(z)=\frac{(2n)!}{n!\,(n+\nu)^n}{}_1F_1(-n,-2n,(n+\nu)z) = z^n y_n\biggl(\frac{2}{(n+\nu)z}\biggr), \end{equation*} \notag
where y_n(u)={}_2F_0(-n,n+1;-u/2) is a Bessel polynomial and satisfies
\begin{equation*} \begin{gathered} \, u^2\, \frac{d^2 y_n(u)}{du^2} +2(u+1)\,\frac{dy_n(u)}{du}-n(n+1)y_n(u)=0, \\ \frac{d^2[ue^{-1/u} y_n(u)]}{du^2} +\biggl(-\frac{1}{u^4} -\frac{n(n+1)}{u^2}\biggr)ue^{-1/u} y_n(u)=0, \\ \begin{split} &\frac{d^2[x\exp(-(n+\nu)/(2x)) y_n(2x/(n+\nu))]}{dx^2} \\ &\qquad\qquad -\biggl(\frac{(n+\nu)^2}{4x^4} +\frac{n(n+1)}{x^2}\biggr)x \exp\biggl(-\frac{n+\nu}{2x}\biggr) y_n\biggl(\frac{2x}{n+\nu}\biggr)=0 \end{split} \end{gathered} \end{equation*} \notag
(see [67], § 18.34 for a=2), where x=(n+\nu)u/2=1/z. It also satisfies the differential equation with large parameter
\begin{equation*} \frac{d^2y_n(x)}{dx^2} +q(x,n)y_n(x)=0 \end{equation*} \notag
(see [26], § 4.2), where we have
\begin{equation*} q(x,n)= -\biggl(\frac{1}{x^2}+\frac{1}{4x^4}\biggr)n^2 - \biggl(\frac{1}{x^2} +\frac{\nu}{2x^4}\biggr)n+\dotsb \end{equation*} \notag
with turning points near x=\pm i/2 or z=\pm 2i.

The first approximation of the solution of this Liouville–Green–Steklov-WKB problem is (-q(x,n))^{-1/4} times a combination of the two exponential functions \exp\biggl(\displaystyle\pm \int^x \sqrt{-q(\xi,n)}\,d\xi\biggr) (see [11], § 10.1), amounting for y_n(u) to be a combination of

\begin{equation*} x^{-1}\exp\biggl(\frac{n+\nu}{2x}\biggr) \biggl(\frac{1}{x^2}+\frac{1}{4x^4}\biggr)^{-1/4} \exp\biggl(\pm \int^x \sqrt{\frac{n(n+1)}{\xi^2}+\frac{(n+\nu)^2}{4\xi^4}}\,d\xi\biggr) \end{equation*} \notag
or
\begin{equation*} z^n \exp\biggl(\frac{(n+\nu)z}{2}\biggr) \biggl(\frac{1}{4}+\frac{1}{z^2}\biggr)^{-1/4} \exp\biggl(\pm \biggl[ \int^z \sqrt{n(n+1)\eta^2+\frac{(n+\nu)^2\eta^4}4}\,\frac{d\eta}{\eta^2}\biggr]\biggr) \end{equation*} \notag
for P_n(z). The main behaviour with respect to n is z^n \exp[n (z/2 \pm (\mathcal{V}(z)+z/2) ], where \mathcal{V} is from the equality \mathcal{V}'(z) = -{1}/{2}+{\sqrt{z^2+4}}/({2z}) used in (3.7). From the z^n behaviour of P_n(z) for large z we keep only the minus sign, so for the monic denominator,
\begin{equation} \begin{aligned} \, \notag p_n(z) &\sim z^n \biggl(1+\frac{4}{z^2}\biggr)^{-1/4} \exp\biggl(\int_\infty^z \biggl[\frac{n+\nu}2 - \sqrt{\frac{n(n+1)}{\eta^2}+\frac{(n+\nu)^2}4}\biggr]\,d\eta\biggr), \\ \notag p_n(z) &\sim z^n \biggl(1+\frac{4}{z^2}\biggr)^{-1/4} \exp\biggl(\frac{n+\nu}{2}\biggl[ z -\sqrt{z^2+4} +2\log\frac{\sqrt{z^2+4}+2}{z}\biggr] \\ &\qquad -\biggl(\nu-\frac12\biggr) \log\frac{\sqrt{z^2+4}+2}{z}\biggr). \end{aligned} \end{equation} \tag{5.5}

From (3.7) we see that the expression under the exponential sign is

\begin{equation*} \begin{aligned} \, & -(n+\nu) \mathcal{V}(z) -\biggl(\nu-\frac12\biggr) \log\frac{\sqrt{z^2+4}+2}{z} \\ &\qquad =-\biggl(n+\frac12\biggr) \mathcal{V}(z) +\frac12\biggl(\nu-\frac12\biggr) \Bigl[ z -\sqrt{z^2+4}\Bigr]. \end{aligned} \end{equation*} \notag

Remark that the coefficients of z^n, z^{n-1} and z^{n-2} of the monic denominator are

\begin{equation} \begin{aligned} \, \notag p_n(z) &=z^ny_n\biggl(\frac{2}{(n+\nu)z}\biggr) =z^n {}_2F_0\biggl(-n,n+1;-\frac{1}{(n+\nu)z}\biggr) \\ &= z^n +\frac{n(n+1)z^{n-1}}{n+\nu} +\frac{(n-1)n(n+1)(n+2)z^{n-2}}{2(n+\nu)^2}+\dotsb\,. \end{aligned} \end{equation} \tag{5.6}

When z is on the F-cut between -2i and 2i, one must consider the two possible square roots in (5.5). This will be discussed after the second method.

Wong and Zhang [92] use a generating function yielding here

\begin{equation*} y_n\biggl(\frac{2}{(n+\nu)z}\biggr)= \frac{-4^n\ n!}{\pi i(n+\nu)^n z^n} \oint \exp\biggl(\frac{(n+\nu)(1-\zeta)z}{2} -(n+1)\log(1-\zeta^2)\biggr)d\zeta \end{equation*} \notag
on a contour enclosing \zeta=1 but excluding \zeta=0. An accurate asymptotic estimate can then be achieved with saddle points analysis when z is not close to \pm 2i, as
\begin{equation*} \begin{aligned} \, y_n\biggl(\frac{2}{(n+\nu)z}\biggr)\ &\sim -\frac{n!}{\pi i}\frac{4^n}{(n+\nu)^n z^n} \\ &\qquad\times \Biggl[\sqrt{-\frac{2\pi}{f''(\zeta_+)}}\, \exp(f(\zeta_+))+\sqrt{-\frac{2\pi}{f''(\zeta_-)}}\,\exp(f(\zeta_-))\Biggr], \end{aligned} \end{equation*} \notag
with f(\zeta)=(n+\nu)(1-\zeta)z/{2} -(n+1)\log(1-\zeta^2), the saddle points \zeta_{\pm} are the two roots of f'(\zeta)=0, that is, of the equation \zeta^2 +4(n+1)/((n+\nu)z)\zeta- 1= 0:
\begin{equation*} \begin{aligned} \, \zeta_\pm & = \frac{ -2(n+1) \pm \sqrt{4(n+1)^2 +(n+\nu)^2z^2} }{(n+\nu)z} \\ &\sim \frac{ -2 \pm \sqrt{z^2+4} }{z}+2(\nu-1)\frac{1\mp 2/\sqrt{z^2+4}}{(n+\nu)z} = \zeta_{\pm,\infty}\biggl[1\mp \frac{2(\nu-1)}{n\sqrt{z^2+4} }\biggr], \end{aligned} \end{equation*} \notag
where \zeta_{\pm,\infty}=(-2\pm \sqrt{z^2+4})/{z} are the limits of \zeta_\pm as n\to\infty; f(\zeta_\pm)=f(\zeta_{\pm,\infty})+o(1) since f'(\zeta_\pm)=0, so
\begin{equation*} \begin{aligned} \, f(\zeta_\pm) &\sim \frac{(n+\nu)(1-\zeta_{\pm,\infty})z}2-(n+1)\log(1-\zeta_{\pm,\infty}^2) \\ &= (n+\nu)\biggl[\frac{(1-\zeta_{\pm,\infty})z}2-\log\frac{4\zeta_{\pm,\infty}}z\biggr] +(\nu-1)\log\frac{4\zeta_{\pm,\infty}}z. \end{aligned} \end{equation*} \notag
Using the equality 1-\zeta_{\pm,\infty}^2=4\zeta_{\pm,\infty}/z, it follows that
\begin{equation*} \begin{aligned} \, f''(\zeta_\pm) & = 2(n+1)\frac{1+\zeta_\pm^2}{(1-\zeta_\pm^2)^2}= \frac{(n+\nu)^2z^2 (\zeta_\pm^{-1}+\zeta_\pm)}{8(n+1)\zeta_\pm} \\ & =\pm \frac{(n+\nu)z \sqrt{(n+\nu)^2z^2+4(n+1)^2}}{4(n+1)\zeta_\pm} \sim \pm \frac{ nz\sqrt{z^2+4} }{4\zeta_{\pm,\infty} }. \end{aligned} \end{equation*} \notag

Thus, we have at our disposal accurate estimates for f and f'' at the points \zeta_\pm, from which we can extract strong asymptotics of p_n (and similar quantities) as a sum of at most two terms of the form

\begin{equation*} \begin{aligned} \, &-\frac{n!}{\pi i}\frac{4^n}{(n+\nu)^n }\sqrt{-2\pi \frac{4\zeta_{\pm,\infty}}{ nz\sqrt{z^2+4}}} \\ &\qquad\times\exp\biggl\{ (n+\nu)\biggl[\frac{(1-\zeta_{\pm,\infty})z}2-\log\frac{4\zeta_{\pm,\infty}}z\biggr] +(\nu-1)\log\frac{4\zeta_{\pm,\infty}}z \biggl\} \\ &\sim 2^{2n+1}e^{-n-\nu}(z^2+4)^{-1/4} \\ &\qquad\times\exp\biggl\{ (n+\nu)\biggl[\frac{(1-\zeta_{\pm,\infty})z}2-\log\frac{4\zeta_{\pm,\infty}}z\biggr] +\biggl(\nu-\frac12\biggr)\log\frac{4\zeta_{\pm,\infty}}z \biggr\} \\ &\sim e^{-n-\nu}\biggl(1+\frac{4}{z^2}\biggr)^{-1/4}z^n \\ &\qquad\times\exp\biggl\{ \biggl(n+\frac12\biggr)\biggl[\frac{(1-\zeta_{\pm,\infty})z}2 -\log(\zeta_{\pm,\infty})\biggr]+\biggl(\nu-\frac12\biggr)\frac{(1-\zeta_{\pm,\infty})z}2 \biggr\}, \end{aligned} \end{equation*} \notag
confirming (5.5).

When z is large, we only keep the term with \zeta_{+,\infty}= (\sqrt{z^2+4}-2)/{z}= 1-2/z+ \dotsb . In other regions the two terms must be considered, as carefully established by Wong and Zhang in [92], Theorem A. A neighbourhood of \pm 2i is also considered by the same authors; see Theorem B in [92].

The asymptotic estimate for p_n,

\begin{equation*} \biggl(1+\frac{4}{z^2}\biggr)^{-1/4}z^n \exp\biggl[ -(n+\nu)\mathcal{V}_{\pm}(z) +\biggl(\nu-\frac12\biggr)\log\frac{-2\pm\sqrt{z^2+4}}{z}\biggr], \end{equation*} \notag
also follows from the formula for the complex potential \mathcal{V} in (3.7). The potential \mathcal{V} is created by a unit positive charge spread on F joining -2i to 2i and a negative unit charge concentrated at the origin; it satisfies \mathcal{V}_+(z)+\mathcal{V}_-(z)=-z+\pi i. The function \mathcal{V}_0(z)= \log[(-2+\sqrt{z^2+4})/z] is a potential function too, created by the same charges and satisfying \mathcal{V}_{0,+}(z)+\mathcal{V}_{0,-}(z)=\pi i on the two sides of any arc joining -2i to 2i avoiding the origin (a potential without an ‘external field’). This will be discussed in a more general setting in (5.8).

5.4.2. Formal orthogonal polynomials

For general c>0, the Padé approximation is useless; we have no more differential equations, and have to turn to the formal orthogonality property from § 2.2.

We follow Aptekarev [6], § 1.3.

We add to the hypotheses already given the symmetry conditions with respect to the real axis, F made of a single connected arc with endpoints a and b, and a positive density of poles in the interior of F. Then we obtain a strong asymptotic of the denominators p_n as orthogonal polynomials on F.

First, we consider the Joukowsky map

\begin{equation*} z=\frac{a+b}2+(b-a)\frac{\Phi_0(z)+\Phi_0(z)^{-1}}2 \end{equation*} \notag
defining the algebraic function
\begin{equation*} \Phi_0(z)=\frac{2z-a-b+2\sqrt{(z-a)(z-b)}}{b-a}. \end{equation*} \notag
As long as we consider symmetric expressions in \Phi_0 and \Phi_0^{-1}, we need not worry about the sign of the square root, but the current convention is to take \Phi_0(z)\sim 4z/(b-a) for large z, and \Phi_0 continuous outside the F-cut joining a to b. One often writes \Phi_0=e^{i\theta}, where \theta is not limited to real values.

For a useful relation with integrals on arcs, let us define as above f(z)=[(z- a)(z-b)]^{-1/2} as a continuous function outside F such that \lim zf(z)=1 as z\to\infty. Then the Cauchy formula gives f(z) outside a contour containing [a,b] as the integral of [2\pi i(z-t)]^{-1} f(t)\,dt over the contour, which we shrink on the two sides of the cut, so it becomes the integral from a to b of (2\pi i)^{-1}[f_-(t)-f_+(t)], which is here [(t-a)(b-t)]^{-1/2}/\pi, the Chebyshev weight:

\begin{equation} \begin{gathered} \, \frac{1}{\sqrt{(z-a)(z-b)}}= \frac{1}{\pi}\int_a^b \frac{dt}{(z-t)\sqrt{(t-a)(b-t)}}, \\ -\kern-10.5pt\int _a^b \frac{dt}{(z-t)\sqrt{(t-a)(b-t)}} =0, \qquad z\in (a,b) \end{gathered} \end{equation} \tag{5.7}
(Chebyshev’s example).

The zero principal value follows from the opposite values of f=\Phi'_0/\Phi_0=\mathcal{V}'_0 on the two sides of [a,b].

Next, let p_n be a polynomial of degree n orthogonal with respect to a possibly complex-valued weight function w on an arc joining a to b. Actually, we need w to be analytic in a convenient region, so that several arcs joining a to b may be considered as the support of w. This also requires formal orthogonality \displaystyle\langle f,g\rangle =\int_a^b f(t)g(t)w(t)\,dt, where no complex conjugate appears. The polynomial p_n(z) is also the denominator of the (n,n) Padé approximant to the Stieltjes-type or Markov-type, function \displaystyle\int_a^b \frac{w(t)\,dt}{z-t}; see [35], [49]–[51], [58] and [66].

Then the Bernstein–Szegő estimate, extended to a general arc [a,b], is an accurate asymptotic formula for p_n, which involves \Phi_0^{\pm n} and factors whose product reconstructs w^{-1}. These ideas will be used in (5.8).

This is already seriously at variance with the classical Markov–Bernstein–Szegő theory limited to positive weight functions on real supports.

The extension to complex weights has a long history, summarized in [6], § 2.1; see also [58] and [55].

Here p_n is not orthogonal with respect to a weight function independent of n, but with respect to

\begin{equation*} \frac{\rho_n(t)}{ (t-z_1^{(n)})\dotsb(t-z_{2n+1}^{(n)})} \sim \rho_n(t)\exp(-(2n+1)\mathcal{V}_{\mathrm z}(t)), \end{equation*} \notag
depending on n, as seen in § 2.2. Actually, the interpolation points z_1^{(n)},\dots,z_{2n+1}^{(n)} will have to be described more accurately than through their limit distribution \mu_{\mathrm z} on E. Let (z-z_1^{(n)})\dotsb (z-z_{2n+1}^{(n)}) \sim \exp(\mathcal{V}_{{\mathrm z},n}(z)), for \mathcal{V}_{{\mathrm z},n}(z)-2n\mathcal{V}_{\mathrm z}(z) bounded but left undefined for the moment.

We return to the settings of § 2.4, the problem of the best rational approximation to

\begin{equation*} f_n(z)=(2\pi i)^{-1}\oint \frac{\rho_n(t)\,dt}{z-t} \end{equation*} \notag
for \rho_n(t)=\exp(n\Phi(t)+\Psi(t)).

The denominator p_n must be orthogonal to all polynomials of degree <n with respect to \rho_n(t)\exp(-(2n+1)\mathcal{V}_{\mathrm z}(t)). As in § 2.4, we try

\begin{equation*} \begin{aligned} \, p_n(t) &=A_+(t)\exp\biggl(n\biggl(\mathcal{V}_{\mathrm z}(t)-\frac{\Phi(t)}2+i\theta_{\mathrm q}(t)\biggr)\biggr) \\ &\qquad+A_-(t)\exp\biggl(n\biggl(\mathcal{V}_{\mathrm z}(t)-\frac{\Phi(t)}2-i\theta_{\mathrm q}(t)\biggr)\biggr) \end{aligned} \end{equation*} \notag
(from [6], §§ 2.2 and 2.3), and check the orthogonality of p_n and p_n(t)/(t-p), amounting essentially to the principal value of the integral of 2A_+(t)A_-(t)\exp(\Psi(t)-\mathcal{V}_{\mathrm z}(t))/(t-p) over an arc (a,b). This latter principal value must vanish for each p\in F.

We consider now a new complex potential \mathcal{V}_\Psi satisfying conditions (I)–(III), (V) and (VI) in § 2.4, but for 2\Psi instead of \Phi.

In our case \mathcal{V}'_{\Psi,+}(z)+\mathcal{V}'_{\Psi,-}(z) =2\Psi'(z), and each A_{\pm}(t) contains an \exp(-\Psi_{\pm}(t)/2) factor. Thus, let A_{\pm}(t)= B(t)\exp(-\Psi_{\pm}(t)/2), where the function B(t) (not known yet) has the property B_+(t)=B_-(t) at the interior points of F. Then the vanishing principal value of an integral relates to the integral of 2B^2(t)\exp(-\mathcal{V}_z(t))/(t-p), and B(t) must contain the factor [(t-a)(b-t)]^{-1/4}. Using (5.7) and the results on p. 148 in [39], we obtain a Privalov problem of quite special form (see [39], § 19.4).

In summary, if q_n/p_n is the best rational approximation of degree n to

\begin{equation*} f_n(z)=A \int_C \exp(n\Phi(t)+\Psi(t)) \frac{dt}{z-t} \end{equation*} \notag
on E=[0,c], with \Phi and \Psi analytic, then
\begin{equation} \begin{gathered} \, p_n(z) \sim [(z-a)(z-b)]^{-1/4}\exp\biggl[\frac{\mathcal{V}_{{\mathrm z},n}(z)}2 -n \mathcal{V}(z) -\frac{\mathcal{V}_{\Psi}(z)}2 \biggr], \\ f_n(z)p_n(z)-q_n(z) \sim 2\pi i AC_n [(z-a)(z-b)]^{-1/4} \exp\biggl[ \frac{\mathcal{V}_{{\mathrm z},n}(z)}2 +n \mathcal{V}(z) +\frac{\mathcal{V}_{\Psi}(z)}2 \biggr], \end{gathered} \end{equation} \tag{5.8}
where \mathcal{V} and \mathcal{V}_\Psi are the complex potentials related to a unit negative charge on E, and a unit positive charge on F= an arc joining a to b, with constant real part on E and real parts of -\Phi/2 and -\Psi on F, and where
\begin{equation*} C_n=\exp(-n(2\mathcal{V}(a)-\Phi(a))-\mathcal{V}_\Psi(a)+\Psi(a)) \end{equation*} \notag
and \mathcal{V}_{{\mathrm z},n}=2n\mathcal{V}_{\mathrm z}+\mathcal{V}_{\Psi,{\mathrm z}} (see the end of the present section).

On F we must sum the contributions of the two sides:

\begin{equation*} \begin{aligned} \, p_n(z) &\sim [(z-a)(z-b)]^{-1/4}\exp\biggl(\biggl(n+\frac12\biggr)\mathcal{V}_{\mathrm z}(z)\biggr) \\ &\qquad\times\biggl\{\exp\biggl[-n \mathcal{V}_{+}(z) -\frac{\mathcal{V}_{\Psi,+}(z)}2 \biggr] -\exp\biggl[-n \mathcal{V}_{-}(z) -\frac{\mathcal{V}_{\Psi,-}(z)}2 \biggr]\biggr\}, \end{aligned} \end{equation*} \notag
where \mathcal{V}_{+}(z)+ \mathcal{V}_{-}(z)= \Phi(z)+\operatorname{const}= \Phi(z)+2\mathcal{V}(a)-\Phi(a) and \mathcal{V}_{\Psi,+}(z)+ \mathcal{V}_{\Psi,-}(z)= 2\Psi(z)+\operatorname{const}= 2\Psi(z)+2\mathcal{V}_\Psi(a)-2\Psi(a).

The last line of (5.8) follows from (2.3) where we take the integral of p_n^2(t)\exp(n\Phi(t)+\Psi(t)-(2n+1)\mathcal{V}_{\mathrm z}(t))/(z-t), dominated as above on F by

\begin{equation*} \begin{aligned} \, &2((t-a)(t-b))^{-1/2}\exp\frac{n\Phi(t)+\Psi(t) -n [\mathcal{V}_{+}(t) +\mathcal{V}_{-}(t) ]-[\mathcal{V}_{\Psi,+}(t) +\mathcal{V}_{\Psi,-}(t)]/2}{z-t} \\ &\qquad=\exp\frac{-n(2\mathcal{V}(a)-\Phi(a))-\mathcal{V}_\Psi(a)+\Psi(a)}{(z-t)\sqrt{(t-a)(t-b)}}. \end{aligned} \end{equation*} \notag
There is no principal value now, as z in not on F and we apply the first formula of (5.7).

Aptekarev’s great paper (see [6], §§ 2.2 and 2.3) contains accurate statements on the meaning of ‘\sim’: it may be uniform convergence on compact sets in the interior of F, but sometimes in the whole of F, when the endpoint singularities of [(z-a)(z-b)]^{-1/4} and \mathcal{V}(z) cancel each other.

Why does this play with 2\Psi and \mathcal{V}_\Psi/2? Because all complex potential functions \mathcal{V} and \mathcal{V}_\Psi considered here are related to unit charges on E and F, [(z-a)(z- b)]^{-1/4} represents a charge 1/2 on F (the principle of argument for the logarithm on a contour circling F), which is cancelled by -\mathcal{V}_\Psi(z)/2 in (5.8); the 2n+1 interpolation points on E correspond to a charge -(2n+1) thanks to the \mathcal{V}_\Psi(z)/2 term in the second line of (5.8).

Even if \Psi(t)\equiv 0, the identity \mathcal{V}_\Psi(t)\equiv 0 would be wrong, as the condition on the charges would not be fulfilled. We then need \mathcal{V}_{*}(z), the potential of (E,F) without an external field. When E and F are the two arcs [0,c] and [a,b], we have

\begin{equation*} \mathcal{V}_*(z)=C \int_\infty^z \frac{dt}{\sqrt{t(t-c)(t-a)(t-b)}} \end{equation*} \notag
as encountered in the Zolotarev problems (see [4], p. 285, Appendix E, [12], [25], [30] and [84]), as also in investigations on orthogonal polynomials on two intervals (see [3] and [5], Ch. 10). The constant C is precisely such that E and F carry unit charges, that is, \mathcal{V}_*(b)-\mathcal{V}_*(a)=\pi i, so
\begin{equation*} C= \frac{ \pi i}{\displaystyle \int_a^b[t(t-c)(t-a)(t-b)]^{-1/2}\,dt } = \frac{(a-c)(b-c)c}{4(\nu_0+\nu_1/x^*)} \end{equation*} \notag
(see (3.3)), where 1/x^*=[(a-c)(b-c)+ab]/(abc).

Using the elliptic integrals notation (4.6),

\begin{equation*} \frac{1}{x^*}=\frac{(1+\zeta)^2}{(1-\zeta)^2 c}+\frac{1}{c}=2\frac{1+\zeta^2}{(1-\zeta)^2 c} \end{equation*} \notag
and
\begin{equation*} C=\sqrt{\frac{k^2+\zeta^2k'^2}{1-\zeta^2}} \frac{\pi(a-c)(b-c)(1-\zeta)}{c^2(1+\zeta){\mathsf K}} = \sqrt{\frac{1-\zeta^2}{k^2+\zeta^2k'^2}}\, \frac{\pi c}{4{\mathsf K}}, \end{equation*} \notag
and (5.2) follows.

We will also need the formulae

\begin{equation*} \begin{aligned} \, \mathcal{V}_{*,{\mathrm p}}(z) &=\int_F \log(z-t)\,d\mu_{*,{\mathrm p}}(t) = \int_F \frac{ \mathcal{V'}_{*,+}(t)-\mathcal{V'}_{*,-}(t)}{2\pi i} \log(z-t)\,dt \\ &= \int_F \frac{C\,\log(z-t)\,dt}{\pi i \sqrt{t(t-c)(t-a)(t-b)}} =\log(z) - \frac{\omega}{z}+o\biggl(\frac 1z\biggr), \end{aligned} \end{equation*} \notag
where
\begin{equation*} \begin{aligned} \, \omega &= \int_a^b \frac{C t\,dt}{\pi i\sqrt{t(t-c)(t-a)(t-b)}} \\ &= \frac{\displaystyle \int_a^b \frac{t\,dt}{\sqrt{t(t-c)(t-a)(t-b)}}}{\displaystyle\int_a^b \frac{dt}{\sqrt{t(t-c)(t-a)(t-b)}}} =\frac{ \nu_0 c+\nu_1}{\nu_0 +\nu_1/x^*}, \end{aligned} \end{equation*} \notag
so
\begin{equation} \mathcal{V}_{*,{\mathrm p}}(z)= \log(z) +c(1-\zeta)\frac{ {\mathsf K} -(1+\zeta){\mathsf E}}{2\zeta{\mathsf K} z}+o\biggl(\frac 1z\biggr), \end{equation} \tag{5.9}
from (3.3) and (4.6).

When z is not close to F, we keep \mathcal{V}(z) and \mathcal{V}_{\Psi}(z), which remain bounded for large z, in the first line of (5.8). On the other hand \mathcal{V}_{\mathrm z}(z) (which is regular on F) behaves like \log z for large z, so the product [(z-a)(z-b)]^{-1/4} \exp((n+ 1/2)\mathcal{V}_{\mathrm z}(z)) behaves as z^n, as it should!

Finally, f_n(z)-q_n(z)/p_n(z) \sim 2\pi i C_n \exp[2n \mathcal{V}(z) +\mathcal{V}_{\Psi}(z)] on the cut E, where the two potential functions have a constant real part and opposite imaginary parts. On E we sum the contributions from the two sides:

\begin{equation*} f_n(z)-\frac{q_n(z)}{p_n(z)} \sim 2\pi i C_n \{ \exp[ 2n \mathcal{V}_+(z) +\mathcal{V}_{\Psi,+}(z) ] +\exp[ 2n \mathcal{V}_-(z) +\mathcal{V}_{\Psi,-}(z) ]\}. \end{equation*} \notag

The common real part yields the strong asymptotics of the error norm

\begin{equation} \biggl\| f_n-\frac{q_n}{p_n}\biggr\|_E \sim 4\pi A\exp[ n(2V(0)-2V(a)+\Phi(a)) +V_{\Psi}(0) -V_{\Psi}(a)+\Psi(a) ]. \end{equation} \tag{5.10}

For the imaginary parts, recall that \mathcal{V} and \mathcal{V}_\Psi have opposite purely imaginary derivatives \pm \pi\mu'_{\mathrm z} and \pm \pi\mu'_{{\mathrm z},\Psi} on the two sides of E=[0,c], so that the error function oscillates like \cos(2n\pi\mu_{\mathrm z} +\pi\mu_{{\mathrm z},\Psi}) on E. The corrected formula for the distribution of the 2n+1 interpolation points is therefore \mu_{{\mathrm z},n}=2n\mu_{\mathrm z} +\mu_{{\mathrm z},\Psi} instead of (2n+1)\mu_{\mathrm z}, and the corresponding potential is

\begin{equation} \mathcal{V}_{{\mathrm z},n}=2n\mathcal{V}_{\mathrm z} +\mathcal{V}_{{\mathrm z},\Psi}. \end{equation} \tag{5.11}

5.4.3. Returning to the approximation to the exponential function

With our problem of rational interpolation to

\begin{equation*} f_n(z)=\exp(-(n+\nu)z) =-\frac{1}{2\pi i} \int_C \exp(n\Phi(t)+\Psi(t))\frac{dt}{z-t}, \end{equation*} \notag
where \Phi(t)=-t and \Psi(t)=-\nu t, the discontinuity of \mathcal{V}_{\Psi} is the same as for 2\nu\mathcal{V}, and we recover a unit charge from the combination \mathcal{V}_{\Psi}=2\nu\mathcal{V} +(1-2\nu)\mathcal{V}_* (see [62], equation (10)). We have now
\begin{equation*} \begin{aligned} \, p_n(z) &\sim [(z-a)(z-b)]^{-1/4}\exp\biggl[\frac{\mathcal{V}_{{\mathrm z},n}(z)}2 -(n+\nu) \mathcal{V}(z) +\biggl(\nu-\frac12\biggr)\mathcal{V}_{*}(z)\biggr] \\ &\sim [(z-a)(z-b)]^{-1/4}\exp\biggl((n+\nu)\mathcal{V}_{\mathrm p}(z)-\biggl(\nu-\frac12\biggr)\mathcal{V}_{*,{\mathrm p}}(z) \biggr) \end{aligned} \end{equation*} \notag
and
\begin{equation*} \begin{aligned} \, &p_n(z)\exp(-(n+\nu)z)-q_n(z) \sim C_n [(z-a)(z-b)]^{-1/4} \\ &\qquad\qquad \times\exp\biggl[\frac{\mathcal{V}_{{\mathrm z},n}(z)}2 +(n+\nu) \mathcal{V}(z) -\biggl(\nu-\frac12\biggr)\mathcal{V}_{*}(z) \biggr], \end{aligned} \end{equation*} \notag
with C_n=\exp(-(n+\nu)(2\mathcal{V}(a)+a)+(2\nu-1)\mathcal{V}_*(a)), \mathcal{V}_{{\mathrm z},n}= 2(n+\nu)\mathcal{V}_{\mathrm z}-(2\nu-1)\mathcal{V}_{*,{\mathrm z}} (see (5.11)) and \mathcal{V}_{{\mathrm z},\Psi}=2\nu\mathcal{V}_{\mathrm z} +(1-2\nu)\mathcal{V}_{*,{\mathrm z}}.

On F one must add the contributions from the two sides of F:

\begin{equation*} \begin{aligned} \, p_n(z) &\sim [(z-a)(z-b)]^{-1/4}\exp\biggl(\frac{\mathcal{V}_{{\mathrm z},n}(z)}2\biggr) \\ &\qquad\times \biggl\{ \exp\biggl[-(n+\nu) \mathcal{V}_{+}(z) +\biggl(\nu-\frac12\biggr)\mathcal{V}_{*,+}(z)\biggr] \\ &\qquad\qquad +\exp\biggl[-(n+\nu) \mathcal{V}_{-}(z) +\biggl(\nu-\frac12\biggr)\mathcal{V}_{*,-}(z)\biggr]\biggr\},\qquad z\in F. \end{aligned} \end{equation*} \notag

The error norm satisfies

\begin{equation*} \begin{aligned} \, &\biggl\| \exp(-(n+\nu(z))-\frac{q_n(z)}{p_n(z)})\biggr\|_E \\ &\qquad \sim2 \exp\biggl[-2(n+\nu) \biggl(V(a)+\frac a2-V(0)\biggr) - (2\nu-1)(V_{*}(a)-V_*(0))\biggr], \end{aligned} \end{equation*} \notag
so (5.3) follows
\begin{equation*} \biggl\| \exp(-(n+\nu(z))-\frac{q_n(z)}{p_n(z)}\biggr\|_E \sim 2\rho^{n+\nu}\rho_*^{1/2-\nu}, \end{equation*} \notag
where \rho=\exp(-2V(a)-a+2V(0)) is the main rate of decrease discussed in (4.5)(4.8), and \rho_*= \exp(2(V_*(0)-V_*(a))).

From

\begin{equation*} \mathcal{V}_*(z)=C \int_\infty^z \frac{dt}{\sqrt{t(t-c)(t-a)(t-b)}} = \pi i \frac{\displaystyle\int_\infty^z \frac{dt}{\sqrt{t(t-c)(t-a)(t-b)}}} {\displaystyle\int_a^b \frac{dt}{\sqrt{t(t-c)(t-a)(t-b)}}} \end{equation*} \notag
(see § 1) we have
\begin{equation*} \rho_*= \exp\left(2 \pi i \frac{\displaystyle\int_a^0 \frac{dt}{\sqrt{t(t-c)(t-a)(t-b)}}}{\displaystyle\int_a^b \frac{dt}{\sqrt{t(t-c)(t-a)(t-b)}}}\right)= \exp\biggl(-\frac{\pi {\mathsf K}'}{{\mathsf K}}\biggr), \end{equation*} \notag
where \mathsf{K}' is related to the integral from 0 to a and 2\mathsf{K} to the integral from a to b.

Theorem 5.1 is proved.

5.5. Some numerical checks

We now proceed with various numerical checks of the validity of (5.1).

5.5.1. The real pole position

We look at the real pole of the approximation when n is odd. By (5.1) poles occur when -(n+\nu) \mathcal{V}_{+}(z) +(\nu-1/2)\mathcal{V}_{*,+}(z) and -(n+ \nu) \mathcal{V}_{-}(z) +(\nu-1/2)\mathcal{V}_{*,-}(z) have the same real part. Let \alpha be the real root of \operatorname{Re}(\mathcal{V}_{+}(z)- \mathcal{V}_{-}(z))=0. We test the O(n^{-1}) correction by the Newton step

\begin{equation} \alpha +\operatorname{Re}\frac{ (\nu-1/2) [\mathcal{V}_{*,+}(\alpha)- \mathcal{V}_{*,-}(\alpha)] }{ (2\mathcal{V}'(\alpha)+1)n}, \end{equation} \tag{5.12}
where we have used that \mathcal{V}'_{-}(z)=-1-\mathcal{V}'(z); we write \mathcal{V} instead of \mathcal{V}_+ in what follows.

1. When c=0, the real zeros of the Padé denominators {}_1F_1(-n,-2n,z) to e^{-z} are -2, -4.644, -7.293, -9.944 and -12.594 for n=1,3,\dots, 9, and come very close to (n+1/2)\alpha for \alpha=-1.325\dotsc, so the empirical formula for the real pole of the nth degree approximation to \exp(-(n+\nu)z) is (n+1/2)\alpha/(n+\nu) = \alpha -(\nu- 1/2)\alpha/n+o(1/n). Does it fit with (5.12)? From (3.7), 2\mathcal{V}'(z)+1= \sqrt{z^2+4}/z and

\begin{equation*} \mathcal{V}_*(z)= \int_\infty^z \frac{2\,dt}{t\sqrt{t^2+4}}= -\log \biggl[\frac{\sqrt{z^2+4}+2}z\biggr] =\mathcal{V}(z)+\frac z2 -\frac{\sqrt{z^2+4}}2, \end{equation*} \notag
and (5.12) follows, knowing that \alpha is the real root of \mathcal{V}_+(z)-\mathcal{V}_-(z)=0.

2. When c=1, the real pole is very stable for \nu=1/2 and the value \alpha=-0.9315 follows. For various values of \nu and n, one finds the empirical formula for the real pole \alpha +1.2(\nu-1/2)/(n+\nu). One has \mathcal{V}'(\alpha)=0.4326 as found by integrating \mathcal{V}'' in (3.2) from -\infty to \alpha; the values of \mathcal{V}_{*,\pm}(\alpha) needed in (5.12) are obtained by integrating (5.2) between a and \alpha on a half of F. The value is \pm 1.1468 +\pi i/2 (the parameter C in (5.2) is precisely such that the imaginary part is \pi/2, that is, C=2.0019). One then finds that the factor of (\nu-1/2) in (5.12) is 1.230.

3. When c=5, we find that \alpha=-0.5045, and the empirical formula for the real pole is \alpha +0.63(\nu-1/2)/(n+\nu). In addition, 2\mathcal{V}'(\alpha)+1=2.295 and \mathcal{V}_{*,\pm}(\alpha)=\pm 0.735+\pi i/2 (C=2.3416), so

\begin{equation*} \operatorname{Re} \frac{ \mathcal{V}_{*,+}(\alpha)- \mathcal{V}_{*,-}(\alpha)}{ (2\mathcal{V}'(\alpha)+1)}= 0.6405. \end{equation*} \notag

5.5.2. The second coefficient of the monic denominator

Let p_n(z)=z^n+\alpha_n z^{n-1}+ \dotsb. What is \alpha_n? From the second line of (5.1) one must find the first two terms of the expansions of \mathcal{V}_{\mathrm p} and \mathcal{V}_{*,\mathrm{p}}. It will be seen in (6.9) that \exp((n+\nu)\mathcal{V}_{\mathrm p}(z))=[z -c(1-\lambda)+O(1/z)]^{n+\nu}=z^{n+\nu}[1-(n+\nu)c(1-\lambda)/z+ \dotsb], where \lambda=4\nu_1^2/(abc^3), and by (5.9) \mathcal{V}_{*,\mathrm{p}}(z)= \log(z) -\omega/z+o(1/z), where \omega=-c(1-\zeta)({\mathsf K} -(1+\zeta){\mathsf E})/(2\zeta{\mathsf K}), so \exp(-(\nu-1/2)\mathcal{V}_{*,{\mathrm p}}(z))= z^{1/2-\nu}[1 +(\nu-1/2)\omega/z+o(1/z)]. Finally, [(z-a)(z-b)]^{-1/4}=z^{-1/2}[1+(a+b)/(4z)+o(1/z)] and p_n(z)=z^n[1 +(-(n+\nu)(1-\lambda)+(\nu-1/2)\omega+(a+b)/4)/z+o(1/z):

\begin{equation*} \begin{gathered} \, \alpha_n =(n+\nu)\beta+\biggl(\nu-\frac12\biggr)\gamma+\delta,\qquad \beta=-c(1-\lambda)= -c\biggl(1-4\frac{\nu_1^2}{abc^3}\biggr), \\ \gamma=\omega=-c(1-\zeta)\frac{ {\mathsf K} -(1+\zeta){\mathsf E}}{2\zeta{\mathsf K} } \quad\text{and}\quad \delta=\frac{a+b}4. \end{gathered} \end{equation*} \notag

1. For instance, when c=0, \alpha_n=n(n+1)/(n+\nu)= n+1-\nu+O(1/n) by (5.6), so \beta=1, \gamma=-2 and \delta=0. We also check that \mathcal{V}_{\mathrm z}(z)=\log z, so from (3.7), \mathcal{V}_{\mathrm p}(z)= \mathcal{V}_{\mathrm z}(z)- \mathcal{V}(z)= z/2 -\sqrt{z^2+4}/2 +\log(\sqrt{z^2+4}+2) = \log z +1/z+o(1/z) for large z, whence \beta=1 again; \mathcal{V}_*(z)= -\log [(\sqrt{z^2+4}+2)/z] as seen above, \mathcal{V}_{*,\mathrm{p}}(z)= \log z -\mathcal{V}_*(z)= \log (\sqrt{z^2+4}+2) = \log z +2/z+o(1/z), confirming that \gamma=-2. Finally, the limits as c\to 0 of the quantities - c(1-\lambda)= 4\nu_1^2/(abc^2) and c({\mathsf K} -{\mathsf E})/(2\zeta{\mathsf K}) are 1 and

\begin{equation*} \lim_{c\to0} \frac{ ck^2 = c^3/16}{2\zeta= 2c^3/64}=2 \end{equation*} \notag
from Table 1, as they should be.

2. Numerical approximations of \beta, \gamma and \delta from actually computed denominators for c=1 are 0.59, -1.51 and 0.19, whereas \beta=-c(1-4\nu_1^2/(abc^3))=0.591401 by (6.8), (6.9) and Table 1, \zeta=0.0076133, 2\mathsf{K}/\pi= 1.015644 and 2\mathsf{E}/\pi= 0.984716, which yields \gamma=-1.50380 and (a+b)/4=0.218 there (Table 1).

3. (\beta,\gamma,\delta)=(0.11, -0.37, 0.50) when c=5, instead of \beta=0.109136, \zeta= 0.36530, 2\mathsf{K}/\pi= 1.23593 and 2\mathsf{E}/\pi= 0.829238, which yield \gamma=-0.36471 and (a+b)/4=0.544 there: a slightly satisfactory match.

4. From the very accurate data of Carpenter, Ruttan and Varga [21] for the approximation of e^{-z} when c=\infty we have p_{10}(z)=z^{10}+5.9426z^{9}+\dotsb, p_{20}(z)=z^{20}+11.9158z^{19}+\dotsb and p_{30}(z)=z^{30}+17.8898z^{29}+\dotsb, so, after division by n, needed by the translation to f_n(z)=e^{-nz}, we obtain \beta=\gamma=0 and \delta=0.597, which is indeed very close to (a+b)/4 of Table 1.

Example. We show that \beta=c(4\nu_1^2/(abc^3)-1) and \gamma=-c(1-\zeta)({\mathsf K} -(1+\zeta){\mathsf E})/(2\zeta{\mathsf K})\to 0 as c\to\infty.

Indeed, ab=(1-\zeta)^2 c^2/(4 (k^2+\zeta^2 k'^2)) and from (4.6),

\begin{equation*} \frac{\nu_1}{c^{3/2}\sqrt{ab}} =\frac{\sqrt{c(1-\zeta^2)}}{4\pi\zeta} [{\mathsf K}-(1-\zeta){\mathsf E}], \end{equation*} \notag
so, \zeta\to 1 and c\to\infty in such a way that c(1-\zeta)\to 2\sqrt{ab}=2|a| at c=\infty. Then7 \nu_1/(c^{3/2}\sqrt{ab})\to\sqrt{4|a|}/(4\pi) {\mathsf K} =1/2. Finally, we use (3.3) as
\begin{equation*} \begin{aligned} \, \frac{\nu_1}{c^{3/2}\sqrt{ab}} &=-\frac{\sqrt{ab}}{4\pi i}\int_a^b \sqrt{ \frac{t/c-1}{t(t-a)(t-b)}} \,dt = \frac{\sqrt{ab}}{4\pi }\int_a^b \frac{1-{t}/(2c)+O(1/c^2) }{\sqrt{ t(t-a)(t-b)} }\, dt \\ &= \frac{\sqrt{ab}}{4\pi }\int_a^b \frac{1-\sqrt{t^2(1-t/c)}/(2c)+O(1/c^2) }{\sqrt{ t(t-a)(t-b)} }\,dt=\frac 12 +O\biggl(\frac1{c^2}\biggr), \end{aligned} \end{equation*} \notag
and use the limit 1/2 just found and (3.6) to get rid of O(1/c). The result for \gamma is much easier to obtain: as we know that c(1-\zeta) remains bounded, the limit of \mathsf{K}-2\mathsf{E} is known to vanish at c=\infty.

5.5.3. The error norm

The check of (5.3) and (5.4). For instance, for c=5, n=5 and \nu=0, 1/2, 1, the error norms are [1.528\cdot 10^{-6}, 1.877\cdot 10^{-6}, 2.214\cdot 10^{-6}]= 2\rho^{n+\nu}\cdot [0.227, 0.982, 4.083] ; for n=10 they are [5.248\cdot 10^{-12}, 6.369\cdot 10^{-12}, 7.570\cdot 10^{-12}]= 2\rho^{10+\nu}\cdot [0.231, 0.989, 4.142]. Indeed, \sqrt{\rho_*}=\exp(-\pi \mathsf{K}'/(2\mathsf{K}))= 0.237 = 1/4.226 (Table 3).

Table 3

cn=5n=10n=15\nu_0/c
12.2365i2.1530i2.1233i-2.0608
21.1922i1.1544i1.1409i-1.1124
50.6085i0.5980i0.5943i-0.5863

When c=\infty, it was remarked in (4.5) that \rho=\exp(-\pi \mathsf{K}'/\mathsf{K}) too, so that \|\exp(-(n+\nu)z)-r_n(z)\|_E \sim 2 \rho^{n+1/2}. Stated as a conjecture in [52] and [53], this was proved by Aptekarev as a very small by-product, in [6]!

For small c we have k\sim c/4 and \rho\sim (c^2/256)\exp(2-c/2), as seen at the end of § 4.3.3, \mathsf{K}'\sim \log(4/k) and \mathsf{K}\sim \pi/2, so, \rho_*\sim k^2/16 \sim c^2/256, the error norm being \sim 2 (c^2/256)^{n+1/2} \exp((2-c/2)(n+\nu)).

Now, by the Meinardus–Braess estimate of § 1.2, as adapted to \exp(-(n+\nu)x) on [0,c] (so that L=(n+\nu)c), we have the error norm

\begin{equation*} \begin{aligned} \, & \sim 2\exp\biggl(-\frac{(n+\nu)c}2\biggr) \biggl(\frac{(n+\nu)ce}{16n+8}\biggr)^{2n+1} \\ & \sim 2\biggl(\frac{c^2}{256}\biggr)^{n+1/2} \exp\biggl[-\frac{(n+\nu)c}2+2n+1 +(2n+1) \underbrace{\log\frac{n+\nu}{n+1/2}}_{ \sim(\nu-1/2)/n}\biggr]. \end{aligned} \end{equation*} \notag

§ 6. Adamyan–Arov–Krein theory

6.1. Rational approximation through Chebyshev expansions

Consider first the Chebyshev expansion of the function

\begin{equation} F(t)=\exp\biggl(-\frac{(n+\nu)c(t+1)}2\biggr) =\frac{c_0}2 +\sum_1^\infty c_k T_k(t), \end{equation} \tag{6.1}
where x=c(t+1)/2 sends t\in [-1,1] to x\in [0,c]. This does not work for c=\infty, when x=\alpha_n (1+t)/(1-t) for \alpha_n>0 is used (see [21], p. 392). It is known that the coefficients in (6.1) are c_k = 2\exp(-(n+\nu)c/2) I_k(-(n+\nu)c/2), where I_k is the kth modified Bessel function (see [68] and [75], Ch. 3, § 4).

The recurrence relation satisfied by the c_k can be found from Bessel functions identities, or also from the differential equation satisfied by F(t) in (6.1), in the form

\begin{equation*} F(t)= -\frac{(n+\nu)c}2\int F(t)\,dt +\mathrm{const}, \end{equation*} \notag
using formulae for the integrals of Chebyshev polynomials:
\begin{equation} c_k +(n+\nu)c \frac{c_{k-1} - c_{k+1} }{4k} =0, \qquad k=1,2,\dotsc \end{equation} \tag{6.2}
(see [27], § 5.7).

It is therefore extremely easy and cheap to compute the numerical values of a large sequence of the coefficients. However, stability requires the coefficients to be computed in a particular order; see Miller’s algorithm in the introduction and §§ 9 and 19 of Abramowitz and Stegun8 [1]; see also Fox and Parker [27], § 5.10.

One considers now the approximation of \sum_1^\infty c_k z^k by meromorphic functions in |z|>1 with exactly n poles in that region. Such functions can always be written in the form

\begin{equation*} \widetilde{r}(z) = \frac{p(z)}{q(z)} =\frac{\sum_{-\infty}^n d_kz^k}{\sum_0^n e_kz^k}, \end{equation*} \notag
where the n zeros of q must have modulus larger than 1. With respect to the supremum norm on the unit circle, the best approximation \widetilde{r}^* to \sum_1^\infty c_k z^k in this class is characterized by the property that, except in degenerate cases, the error function \sum_1^\infty c_k z^k-\widetilde{r}^*(z) describes a precise circle of winding number 2n+1 centred at the origin, as z describes the unit circle [2], [60], [61], [82], [85]. Let \sigma_n be the radius of that circle. A consequence is that the error function can then be written
\begin{equation} \sum_1^\infty c_k z^k-\widetilde{r}^*(z)= b(z) = \frac{\sigma_n b_1(z)}{b_2(z)} = \sigma_n \frac{ \overline{u_1} +\overline{u_2} z +\dotsb}{u_1 z^{-1} + u_2 z^{-2} +\dotsb }, \end{equation} \tag{6.3}
where the denominator b_2 of b is holomorphic in |z|>1 and must have exactly n zeros in 1<|z|<\infty, which will precisely be the zeros of the denominator q of \widetilde{r}^*:
\begin{equation*} b_2(z)=u_1 z^{-1} +u_2 z^{-2} +\dotsb = (e_n+e_{n-1} z^{-1} +\dotsb + e_0z^{-n})v(z) = z^{-n} q(z)v(z), \end{equation*} \notag
where v is still holomorphic (and without zeros in 1<|z|<\infty). Therefore, multiplying the two sides of (6.3) by b_2(z),
\begin{equation*} \biggl(\sum_1^\infty c_k z^k\biggr)b_2(z) -z^{-n}p(z)v(z) = \sigma_n b_1(z)\,, \end{equation*} \notag
or
\begin{equation*} \biggl(\sum_1^\infty c_k z^k\biggr)(u_1z^{-1}+u_2z^{-2}+\dotsb) = \sigma_n (\overline{u_1} +\overline{u_2} z +\dotsb)+\text{negative powers of } z, \end{equation*} \notag
that is,
\begin{equation*} \boldsymbol{H} U=\sigma_n \overline{U}, \end{equation*} \notag
where U is the vector [u_1, u_2,\dots]^\top, and \boldsymbol{H} is the infinite Hankel matrix
\begin{equation} \boldsymbol{H} = [c_{k+m-1}] , \qquad k,m=1,2,\dotsc\,. \end{equation} \tag{6.4}

Numerically, one considers a large finite section k,m=1,2,\dots, N.

Now let \sigma_n be the nth singular value (s-number) of \boldsymbol{H}; see [60], § 4. Thus, \overline{\boldsymbol{H}U} = \sigma_n U, that is, \overline{\boldsymbol{H}} \boldsymbol{H} U=\sigma_n^2U. The fact that the nth singular value of \boldsymbol{H} (\sigma_0 \geqslant \sigma_1 \geqslant \dotsb) is indeed related to a vector U such that the function u_1 z^{-1} + u_2 z^{-2}+\dotsb has precisely n zeros in 1<|z|<\infty requires a deeper understanding of Hankel matrix theory [2], [60], [61].

If the coefficients c_k happen to be real, then \sigma_{n} =|\lambda_{n}|, the absolute value of the nth eigenvalue of \boldsymbol{H} (starting with n=0).

The negative powers add to (u_1 z^{-1} +u_2 z^{-2}+\dotsb)p(z)/q(z)= z^{-n}v(z)p(z), allowing us to retrieve the numerator p.

For instance, for \nu=1/2 and c=1 the coefficients c_k are 0.51090, -0.40344, 0.21749, -0.08709, 0.02749, -0.00713, 0.00157, -0.00030, 0.00005, \dots, and for n=5 we have \sigma_5=4.2433\cdot 10^{-10} and U=i[0.00014, 0.00307, 0.03144, 0.17926, 0.54605, 0.59020, -0.52349, 0.20747, -0.05596, 0.01172, -0.00203, 0.00030,\dots]. The sixth eigenvalue \lambda_5 of \boldsymbol{H} happens to be negative.

The last step in Carathéodory–Fejér approximation is to approximate \widetilde{R}(x) = (c_0+\widetilde{r}^*(z)+ \widetilde{r}^*(z^{-1}))/2 by a rational function of degree n of x=(z+z^{-1})/2. One naturally chooses Q(x)=q(z)q(z^{-1}) as the denominator and determines the numerator P(x) such that the Chebyshev expansion of R_{\rm CF}(x) = P(x)/Q(x) and \widetilde{R}(x) agree through the T_n(x)=(z^n+z^{-n})/2 term. This operation destroys the exact equioscillation of the error function, but the perturbation is usually very much smaller than \sigma_{n} [85].

It may be useful to introduce a free parameter \alpha in the making of the function F, which is here

\begin{equation*} F(t)= \exp\biggl(\frac{(n+\nu)\alpha (t+1) }{t-1-2\alpha/c}\biggr). \end{equation*} \notag

We now use

\begin{equation*} F'(t)= \frac{2(n+\nu)\alpha(1+\alpha/c)}{(t-1-2\alpha/c)^2}F(t) \end{equation*} \notag
as
\begin{equation*} \biggl(t-1-\frac{2\alpha}c\biggr)^2 F(t) - \int 2\biggl(t-1-\frac{2\alpha}c\biggr) F(t)\,dt =2(n+\nu)\alpha\biggl(1+\frac{\alpha}c\biggr) \int F(t)\,dt+\operatorname{const}, \end{equation*} \notag
so
\begin{equation*} \begin{aligned} \, &\frac{c_{k-2}+2c_k+c_{k+2}}{4} -\biggl(1+\frac{2\alpha}c\biggr)(c_{k-1}+c_{k+1}) +\biggl(1+\frac{2\alpha}c\biggr)^2c_k \\ &\qquad\qquad -\frac{c_{k-2}-c_{k+2}}{2k} +\biggl[1+\frac{2\alpha}c-(n+\nu)\alpha\biggl(1+\frac\alpha c\biggr) \biggr] \frac{c_{k-1}-c_{k+1}}{k}=0, \end{aligned} \end{equation*} \notag
k=1,2,\dots, or
\begin{equation} \begin{aligned} \, \notag &\frac{k-2}{4k}c_{k-2} -\frac{ (k-1)(1+2\alpha/c) -(n+\nu)\alpha(1+\alpha/c) }{k} c_{k-1} +\biggl[\frac12+\biggl(1+\frac{2\alpha}c\biggr)^2\biggr]c_k \\ &\qquad\qquad -\frac{ (k+1)(1+2\alpha/c) +(n+\nu)\alpha(1+\alpha/c) }{k} c_{k+1} +\frac{k+2}{4k}c_{k+2}=0 , \end{aligned} \end{equation} \tag{6.5}
k=1,2,\dots, c_{-1}=c_1. These recurrence relations are numerically solved by a ‘compact method’ [27], producing three-term relations \xi_k c_k +\eta_k c_{k+1}+\sigma_k c_{k+2}=\tau_k c_0 (upper triangular matrix relations), k=1,2,\dots, solved themselves for k=N,{N-1},\dots,1 with the initial data c_{N+1}=c_{N+2}=0. The last unknown c_0 follows from the equation F(-1)=c_0/2+\sum_1^N (-1)^kc_k=1.

The parameter \alpha may be useful if one looks for a faster decrease of the |c_k| with k. If we expect the ratios \rho_k=c_{k+1}/c_k to be slowly varying with k (see theorems of Poincaré and Perron in [64]), then

\begin{equation*} \frac{(\rho+\rho^{-1})^2}4 -\biggl(1+\frac{2\alpha}c\biggr)(\rho+\rho^{-1}) -\alpha\biggl(1+\frac\alpha c\biggr)\biggl(\frac nk\biggr)(\rho-\rho^{-1})+ \biggl(1+\frac{2\alpha}c\biggr)^2=0 \end{equation*} \notag
follows from (6.5) for large k and n. We only look at the roots such that |\rho|<1. Most effective \alpha should be such that \rho is minimum for k close to 2N. Numerical tests suggest that \rho is minimum in the case of a double root.

As \alpha\to\infty, it all follows from (6.2) that (n/k)(\rho^{-1}-\rho)/c+4/c^2=0. If k is close to 2N and is much larger than nc, then \rho is close to -nc/(8N).

This may also be discussed using a Fourier coefficient formula, or a contour integral:

\begin{equation*} \begin{aligned} \, c_k &= \frac{1}{\pi}\int_{-\pi}^\pi \exp\biggl(\frac{(n+\nu)\alpha (\cos\theta+1) }{\cos\theta-1-2\alpha/c}\biggr) e^{-ik\theta}\,d\theta \\ &=\frac{1}{\pi i}\oint \exp\biggl(\frac{(n+\nu)\alpha (u+2+u^{-1}) }{u-2-4\alpha/c+u^{-1}}\biggr) \frac{du}{u^{k+1}} \end{aligned} \end{equation*} \notag
estimated at a saddle point
\begin{equation} -\frac{4(n+\nu)\alpha (1+\alpha/c) }{(u-2-4\alpha/c+u^{-1})^2}-\frac{k+1}{u}=0. \end{equation} \tag{6.6}

Change of variables: if i\mu r= (u^{1/2}+u^{-1/2})/{u^{1/2}-u^{-1/2}}, that is, u+u^{-1}= 2(\mu^2 r^2-1)/(\mu^2 r^2+1), then

\begin{equation*} z=-\frac{\alpha (u+2+u^{-1}) }{u-2-4\alpha/c+u^{-1}}= \frac{ \alpha \mu^2 r^2}{1+\alpha/c +(\alpha/c)\mu^2 r^2} = \frac{c r^2}{1+r^2} \quad \text{for } \mu^2=1+\frac c\alpha. \end{equation*} \notag

6.2. Method of electrostatic images

The change of variable z=c r^2/(1+r^2) (that is, r=\pm\sqrt{1/(c/z-1)}) suggested just above and sending z\in E=[0,c] to r\in\{\text{the whole real line}\ \mathbb{R}\} sheds new light on the potential function V of our problem. Indeed, \mathcal{V}(c r^2/(1+r^2)) is a harmonic function of two real variables which happens to be constant on the real line. A valid formula is \mathrm{const} + \sum_k w_k \log|(r-r_k)/(r-\overline{r_k})|, as |r-r_k|=|r-\overline{r_k}| if and only if r is real. The potential is therefore created by the charges -w_k at the points r_k and the charges w_k at their images \overline{r_k}= the complex conjugate of r_k (see [18], p. 485, and [45], Ch. IX).

There may be a relation with § 7.3.2 (‘reflected sets’) of [10].

Considering point distribution where -\overline{r_k} and r_k have the same charge, we have a combination of the functions \log\bigl|(r-r_k)(r+\overline{r_k})/((r-\overline{r_k})(r+r_k))\bigr|, or, keeping an analytic form for the complex potential function,

\begin{equation*} \mathcal{V}\biggl(\frac{c r^2}{1+r^2}\biggr)=\operatorname{const} +\int_\Gamma \log\frac{r-s}{r+s}\,d\widetilde{\mu}(s), \end{equation*} \notag
where \Gamma is the set of r-values corresponding to the cut F in, say, the upper half r-plane. This complex potential \mathcal{V} must be the same as encountered before in the z-plane outside E=[0,c].

So d\mathcal{V}(z)/{dz} +{1}/{2} takes opposite values on the two sides of F, the same holds for {d\mathcal{V}}/{dr} +{dz}/(2\,dr)= {d\mathcal{V}}/{dr} +{cr}/{(1+r^2)^2} on the two sides of \Gamma. More precisely, as {d\mathcal{V}(z)}/{dz} +{1}/{2}= \pm \pi i {d\mu_{\mathrm p}(z)}/{dz} by (3.10) on the two sides of F, where d\mu_{\mathrm p} is the continuous limit measure of the poles,

\begin{equation*} \frac{d\mathcal{V}}{dr} +\frac{cr}{(1+r^2)^2}= \pm \pi i \frac{2cr}{(1+r^2)^2}\frac{d\mu_{\mathrm p}(z)}{dz} =\pm \pi i \frac{d\mu_{\mathrm p}(cr^2/(1+r^2))}{dr} \end{equation*} \notag
on the two sides of \Gamma. This means that
\begin{equation*} \frac{d\mathcal{V}(cr^2/(1+r^2))}{dr} = \int_\Gamma \biggl(-\frac{1}{r-s}+\mathrm{\ reg.\ }\biggr) \, d\mu_{\mathrm p}\biggl(\frac{cs^2}{1+s^2}\biggr), \end{equation*} \notag
where the term ‘reg.’ is regular on \Gamma. But we just saw that the conditions for \mathcal{V}(cr^2/(1+r^2)) on E are fulfilled by an integral of \log((r-s)/(r+s)) on \Gamma, hence the derivative9 -1/(r-s)+1/(r+s) with respect to r yields
\begin{equation*} \frac{d\mathcal{V}(cr^2/(1+r^2))}{dr} = \int_\Gamma \biggl(-\frac{1}{r-s}+\frac{1}{r+s}\biggr) d\mu_{\mathrm p}\biggl(\frac{cs^2}{1+s^2}\biggr). \end{equation*} \notag
Let p_k, k=1,\dots,n, be the poles of the nth degree approximant. Then the integral on \Gamma is approximated by the discrete sum \frac{1}{n}\sum_{k=1}^n \bigl(-1/(r-r_k)+1/(r+r_k)\bigr), where r_k is the root in the upper r-plane of p_k=cr_k^2/(1+r_k^2).

For c=1 and n=5 let us check that the poles are r_k=\pm 0.695i, \pm 0.116 \pm 0.721i and \pm 0.215 \pm 0.809i (the endpoints a and b are taken to \pm 0.244 \pm 0.962 i).

In fact, for large z we know from (3.2) that \mathcal{V}'\sim -\nu_0/(2z^2), so near r=i,

\begin{equation*} \frac{d\mathcal{V}}{dr} \sim -\frac{\nu_0}{2z^2} \frac{dz}{dr} = -\frac{\nu_0(1+r^2)^2}{2c^2r^4} \frac{2cr}{(1+r^2)^2} \to -\frac{i\nu_0}c. \end{equation*} \notag
We compare with ({2}/{n})\sum_1^n r_k/(1+r_k^2) in some cases (see Table 3).

Interesting confirmation, as the r_k were estimated by the Gutknecht and Trefethen’s application of the Adamyan–Arov–Krein theory [85], and \nu_0 was computed in terms of elliptic integrals in (4.6).

Also, as {d\mathcal{V}}/{dr}+{cr}/{(1+r^2)^2} takes opposite values on the two sides of \Gamma,

\begin{equation*} \frac{d\mathcal{V}}{dr} = -\kern-10.5pt\int _\Gamma \biggl(-\frac{1}{r-s}+\frac{1}{r+s}\biggr) d\mu_{\mathrm p}\biggl(\frac{cs^2}{1+s^2}\biggr) = -\frac{cr}{(1+r^2)^2} \end{equation*} \notag
on \Gamma. The discretization at points r_k\in\Gamma is
\begin{equation*} \frac{1}{n}\sum_{\substack{j=1\\j\neq k}}^n \biggl(-\frac{1}{r_k-r_j}+\frac{1}{r_k+r_j}\biggr)= -\frac{cr_k}{(1+r_k^2)^2}, \qquad k=1,\dots, n, \end{equation*} \notag
which are equilibrium conditions on \Gamma, easily seen as the vanishing of the resultant of forces acting on a positive particle at r_k, which is repelled by its neighbours at the r_j, j\neq k, attracted by the images -r_j, j=1,\dots, n, and submitted to the supplementary force cr_k/(1+r_k^2). So the particle at r_k is submitted to the three forces F_1, F_2 and F_3, which are10 the complex conjugates of
\begin{equation*} \frac{1}{n}\sum_{\substack{j=1\\j\neq k}}^n\frac{1}{r_k-r_j}, \quad -\frac{1}{n}\sum_{j=1}^n \frac{1}{r_k+r_j} \quad\text{and}\quad-\frac{cr_k}{1+r_k^2}. \end{equation*} \notag

Figure 5, (b), shows an example of these forces acting on the points corresponding to a set of poles, and how they are close, but not exactly, to an equilibrium. A set of points in equilibrium is also shown nearby. This is an illustration of how the limit distribution of poles as n\to\infty is the required equilibrium continuous distribution.

6.3. Quadratic relations

We return to the original variable:

\begin{equation*} \begin{aligned} \, \frac{d\mathcal{V}(z)}{dz} &=\frac{dr}{dz}\int_a^b \frac{-2s}{r^2-s^2}\, d\mu_{\mathrm p}(t)\ \biggl(\text{with } r=\sqrt{\frac{1}{c/z-1}},\quad s=\sqrt{\frac{1}{c/t-1}}\biggr) \\ &= \frac{c}{2z^2(c/z-1)^{3/2}} \int_a^b \frac{-2(z-c)(t-c)(c/t-1)^{-1/2} }{c(z-t)}\,d\mu_{\mathrm p}(t), \end{aligned} \end{equation*} \notag
so,
\begin{equation} \frac{d\mathcal{V}(z)}{dz} =-\int_a^b \sqrt{\frac{t(t-c)}{z(z-c)}} \,\frac{d\mu_{\mathrm p}(t)}{z-t}, \end{equation} \tag{6.7}
relating the total potential \mathcal{V} to the limit distribution of poles \mu_{\mathrm p} only, whereas we remind that we have \mathcal{V}=\mathcal{V}_{\mathrm z}-\mathcal{V}_{\mathrm p}, and
\begin{equation*} \frac{d\mathcal{V}_{\mathrm p}(z)}{dz} =\int_a^b \,\frac{d\mu_{\mathrm p}(t)}{z-t}. \end{equation*} \notag

We look now at what happens on the two sides of F: from (2.5), \mathcal{V}'(z)=P_1(z) \pm \pi i \mu'_{\mathrm p}(z), where P_1 is an appropriate principal value. Actually, P_1(z)\equiv -1/2 in our case. Of course, the same pattern holds for \mathcal{V}_{\mathrm p} itself: \mathcal{V}_{\mathrm p}'(z)=P_2(z) \mp \pi i \mu'_{\mathrm p}(z), with \mp instead of \pm because \mathcal{V}_{\mathrm p}'=\mathcal{V}_{\mathrm z}'-\mathcal{V}'. The same discontinuity \mp \pi i \mu'_{\mathrm p}(z) appears in \mathcal{V}'^2(z)=1/4 \mp \pi i \mu'_{\mathrm p}(z) - \pi^2 (\mu'_{\mathrm p}(z))^2 on the two sides of F, so that \mathcal{V}'_{\mathrm p}(z)-(\mathcal{V}'(z))^2 has no more discontinuity on F. Remark also that \mathcal{V}'_{\mathrm p} has no singularity on E, and the same is true of \mathcal{V}'^2 (as \mathcal{V}' has opposite values on the two sides of E), so \mathcal{V}'_{\mathrm p}-(\mathcal{V}')^2 must be a meromorphic function in the whole complex plane. This function will be determined from its derivative.

Differentiate \mathcal{V}'_{\mathrm p} as

\begin{equation*} \mathcal{V}''_{\mathrm p}(z) =-\int_a^b \frac{\mu'_{\mathrm p}(t)\,dt}{(z-t)^2} = \int_a^b \frac{\mu''_{\mathrm p}(t)\,dt}{z-t}. \end{equation*} \notag
Now, \mu''_{\mathrm p}(t)=\mathcal{V}''(t)/(\pi i) on F, so by (3.2),
\begin{equation*} \mathcal{V}''_{\mathrm p}(z)= \int_a^b \frac{ (\nu_0 t +\nu_1)\,dt/(\pi i) }{(z-t)\sqrt{t^3 (t-c)^3 (t-a)(t-b) }}. \end{equation*} \notag
Multiply by z^2(z-c)^2=t^2(t-c)^2+(z-t)(z^3+z^2t+zt^2+t^3 -2z^2c-2ztc-2t^2c+zc^2+tc^2):
\begin{equation*} z^2(z-c)^2\mathcal{V}''_{\mathrm p}(z) = P(z)+\int_a^b (\nu_0 t +\nu_1)\sqrt{\frac{t(t-c)}{(t-a)(t-b)}}\frac{dt} {\pi i(z-t)}, \end{equation*} \notag
where P is the third degree polynomial
\begin{equation*} P(z) =\int_a^b \frac{ (\nu_0 t +\nu_1)(z^3+z^2t+zt^2+t^3-2z^2c-2ztc-2t^2c+zc^2\,{+}\,tc^2 )\,dt/(\pi i) }{\sqrt{t^3 (t-c)^3 (t-a)(t-b) }}. \end{equation*} \notag

From (3.1) and (3.6),

\begin{equation*} z^2(z-c)^2\mathcal{V}''_{\mathrm p}(z) =P(z)+2(\nu_0 z+\nu_1)\sqrt{\frac{z(z-c)}{(z-a)(z-b)}}\, \mathcal{V}'(z). \end{equation*} \notag
New division by z^2(z-c)^2 and integration yield at last
\begin{equation*} \mathcal{V}'_{\mathrm p}(z)= (\mathcal{V}'(z))^2+ \int^z_\infty \frac{P(t)\,dt}{t^2(t-c)^2} \end{equation*} \notag
(a very painful derivation, inspired by, but surprisingly not using (6.7)).

Let us look at P more closely. Let m_r be the moment

\begin{equation*} \int_a^b \frac{ t^rdt/(\pi i) }{\sqrt{t^3 (t-c)^3 (t-a)(t-b) }}. \end{equation*} \notag
We have
\begin{equation*} \begin{aligned} \, P(z)&=(\nu_0 m_1+\nu_1 m_0)z^3 +(\nu_0 m_2+\nu_1 m_1)z^2 +(\nu_0 m_3+\nu_1 m_2)z + \nu_0 m_4+\nu_1 m_3 \\ &\qquad -2c(\nu_0 m_1+\nu_1 m_0)z^2 -2c(\nu_0 m_2+\nu_1 m_1)z -2c(\nu_0 m_3+\nu_1 m_2) \\ &\qquad+c^2(\nu_0 m_1+\nu_1 m_0)z +c^2(\nu_0 m_2+\nu_1 m_1). \end{aligned} \end{equation*} \notag
The actual degree of P is 2 instead of 3, as the coefficient of z^3 is
\begin{equation*} \nu_0 m_1+\nu_1 m_0=\int_a^b \frac{\mathcal{V}''(t)\,dt}{\pi i} = \frac{\mathcal{V}'(b)-\mathcal{V}'(a)}{\pi i}=0. \end{equation*} \notag
The coefficient of z^2 is
\begin{equation*} \begin{aligned} \, &\nu_0(m_2-2cm_1)+\nu_1(m_1-2cm_0)=\nu_0 m_2+\nu_1 m_1= \int_a^b \frac{t\mathcal{V}''(t)\,dt}{\pi i} \\ &\qquad= \frac{b\mathcal{V}'(b)-a\mathcal{V}'(a)}{\pi i} -\int_a^b \frac{\mathcal{V}'(t)\,dt}{\pi i} = -\frac{b-a}{2\pi i}+\frac{\mathcal{V}(b)-\mathcal{V}(a)}{\pi i} =-1 \end{aligned} \end{equation*} \notag
by (3.11), as it should be, as \mathcal{V}'_{\mathrm p}(z)= 1/z+O(1/|z|^2) for large z. The last coefficients \lambda_1 and \lambda_0 of P(t)=-t^2+\lambda_1 t+\lambda_0 are the integrals of (t-c)^2 and t(t-c)^2, so \lambda_1=\nu_0(m_3-2cm_2+c^2m_1)+\nu_1(m_2-2cm_1+c^2m_0) and \lambda_0=\nu_0(m_4-2cm_3+c^2m_2)+\nu_1(m_3-2cm_2+c^2m_1).

For m_3 the best combination is

\begin{equation*} m_3-cm_2=\int_a^b \frac{ t \,dt}{\pi i\sqrt{t (t-c) (t-a)(t-b) }} = \frac{4(\nu_0 c+\nu_1)}{(a-c)(b-c)c} \end{equation*} \notag
by (3.3) at x=c.

Next, m_2-cm_1 is the complete integral of the first kind

\begin{equation*} \frac{1}{\pi i}\int_a^b \frac{dt} {\sqrt{(t-a)(t-b)t(t-c)}} =\frac{4(\nu_0+\nu_1/x^*)}{(a-c)(b-c)c}; \end{equation*} \notag
the above formula follows from (3.3) at x=x^*=abc/((a-c)(b-c)+ab), hence ab(x^*-c)=-(a-c)(b-c)x^*.

Note the beautiful relation

\begin{equation*} m_3-2cm_2+c^2m_1= \frac{4(1-c/x^*)\nu_1 }{(a-c)(b-c)c}=-\frac{4\nu_1 }{abc}. \end{equation*} \notag

The even more beautiful equality m_4-2cm_3+c^2m_2=0 follows from (3.6)! Moreover,

\begin{equation*} \begin{aligned} \, \lambda_1 & =\nu_0(m_3-2cm_2+c^2m_1)+\nu_1(m_2-2cm_1+c^2m_0) \\ &=\nu_0(m_3-cm_2) +\nu_1(m_2-cm_1)+c =4\frac{\nu_0(\nu_0 c+\nu_1) +\nu_1(\nu_0+\nu_1/x^*)}{(a-c)(b-c)c}+c \\ &=\frac{4c\nu_0^2+8\nu_0\nu_1 +4((c-a)(c-b)+ab)/(abc)\nu_1^2}{(a-c)(b-c)c}+c \\ &= \frac{4(\nu_0 c+\nu_1)^2}{(a-c)(b-c)c^2}+4\frac{\nu_1^2}{abc^2}+c = \frac{8\nu_1^2}{abc^2} \end{aligned} \end{equation*} \notag
from (4.7). The final formula for \lambda_0 is
\begin{equation*} \lambda_0=\nu_0(m_4-2cm_3+c^2m_2)+\nu_1(m_3-2cm_2+c^2m_1) = -\frac{4\nu_1^2 }{abc}, \end{equation*} \notag
so for P(t)/(t^2(t-c)^2) we obtain
\begin{equation*} \begin{aligned} \, \frac{P(t)}{t^2(t-c)^2} & = \frac{\lambda_0+\lambda_1 t-t^2}{t^2(t-c)^2} = \frac{\lambda_0}{c^2t^2} +\frac{\lambda_0+\lambda_1 c-c^2}{c^2(t-c)^2}-\frac{2\lambda_0+ c\lambda_1}{c^2t(t-c)} \\ &= \frac{\lambda_0/c^2}{t^2}+\frac{-\lambda_0/c^2-1}{(t-c)^2}. \end{aligned} \end{equation*} \notag
Thus,
\begin{equation} \begin{aligned} \, \mathcal{V}'_{\mathrm p}(z) &=(\mathcal{V}'(z))^2 +\frac{\lambda}{z} +\frac{1-\lambda}{z-c} , \\ \mathcal{V}'_{\mathrm z}(z) &=(\mathcal{V}'(z))^2 +\mathcal{V}'(z)+\frac{\lambda}{z} +\frac{1-\lambda}{z-c} , \end{aligned} \end{equation} \tag{6.8}
where \lambda=4\nu_1^2/(abc^3). For large z,
\begin{equation} \exp(\mathcal{V}_{\mathrm p}(z))=z -c(1-\lambda)+O\biggl(\frac 1z\biggr) \quad\text{and}\quad \exp(\mathcal{V}_{\mathrm z}(z))=z +\frac{\nu_0}2-c(1-\lambda)+O\biggl(\frac 1z\biggr) \end{equation} \tag{6.9}
since \mathcal{V}'(z)\sim -\nu_0/(2z^2) in (3.2).

6.4. Integral Hankel operator

When \alpha\to 0, the c_k vary so slowly that \boldsymbol{H} turns into an integral operator! Indeed, write (6.5) (times 4k) as \Delta^4 ((k-2)c_{k-2}) -(\alpha/c)\Delta^2((k-1)c_{k-1}) -4(n+\nu)\alpha(1+\alpha/c)(c_{k+1}-c_{k-1})+16(\alpha/c)^2kc_k=0.

Consider kc_k as a function \gamma(k\sqrt{\alpha}). Then

\begin{equation*} \begin{aligned} \, \sum_{m=0}^\infty c_{k+m}f(y_m = m\sqrt{\alpha}) &= \sum_{m=0}^\infty \underbrace{(k+m)c_{k+m}}_{ \gamma(x+y)} \frac{f(y) [\sqrt{\alpha}=\Delta y]}{(k+m)\sqrt{\alpha}=x+y } \\ &\sim \int_0^\infty \frac{\gamma(x+y)}{x+y} f(y)\,dy. \end{aligned} \end{equation*} \notag
Also, (k+j)c_{k+j} \sim \gamma((k+j)\sqrt{\alpha}) =\gamma(k\sqrt{\alpha})+ j\sqrt{\alpha} \gamma'(k\sqrt{\alpha})+(\alpha j^2/2)\gamma''(k\sqrt{\alpha})+\dotsb, \Delta^r kc_k \sim \alpha^{r/2} d^r\gamma(x)/dx^k at x=k\sqrt{\alpha}, and
\begin{equation*} \frac{d^4\gamma(x)}{dx^4} -\frac{1}{c}\,\frac{d^2 \gamma(x)}{dx^2} -8(n+\nu)\frac{\alpha(1+\alpha/c)}{k\alpha^{3/2}= x\alpha}\,\frac{d\gamma(x)}{dx} +16\frac{1}{c^2}\gamma(x)=0. \end{equation*} \notag

We return to (6.5) and the equation for \rho=c_{k+1}/c_k for small \alpha with \rho such that |\rho|<1 and (\rho+1/\rho)^2 -4(1+2\alpha/c)(\rho+1/\rho)+4(1+2\alpha/c)^2=0, so \rho\sim 1-2\sqrt{\alpha/c} and \rho^k = (\rho^{1/\sqrt{\alpha}})^x \to \exp(-2x/\sqrt{c}).

When \alpha is small, the saddle-point equation (6.6) has a solution near u=1, as u-2-4\alpha/c+u^{-1} \sim \pm iA/\sqrt{k}, where A=\sqrt{4(n+\nu)\alpha (1+\alpha/c) }, so u\sim 1+2\alpha/c \pm iA/(2\sqrt{k}) +\sqrt{ 4\alpha/c \pm iA/\sqrt{k} } \sim 1+2\sqrt{\alpha/c} for large k. Then the asymptotic behaviour for large n and k of

\begin{equation*} c_k =\frac{1}{\pi i}\oint \exp\biggl(\frac{(n+\nu)\alpha (u+2+u^{-1}) }{u-2-4\alpha/c+u^{-1}}\biggr) \frac{du}{u^{k+1}} \end{equation*} \notag
is dominated by
\begin{equation*} \exp\biggl( \pm i 4(n+\nu)\frac{\alpha\sqrt{k}}A\biggr) \biggl(1+2\sqrt{\frac{\alpha}c}\,\biggr)^{-k} \sim \exp\biggl( \pm i 4(n+\nu)\frac{\alpha\sqrt{k}}A -\frac{2x}{\sqrt{c}}\biggr) . \end{equation*} \notag

Figure 6 shows that an \exp(-2x/\sqrt{c}) envelope is an oversimplification for large c. There must be a quadratic term, as an \exp(-(x+y)^2) kernel was found for c=\infty; see [53], § 3, Theorem 1.

§ 7. Divided differences and B-splines

The denominator of rational interpolation can be interpreted as an orthogonal polynomial with respect to a scalar product related to the interpolation points z_1^{(n)},\dots, z_{2n+1}^{(n)}, as in § 2.2 for functions defined by a contour integral.

The z_j^{(n)} are unknowns here, their limit distribution \mu_{\mathrm z} and the related complex potential \mathcal{V}_{\mathrm z} were found so as to fulfill the Gonchar–Rakhmanov–Stahl conditions of §§ 2.3 and 2.4. Later on, a more precise estimate was needed for establishing strong asymptotics formulae, and a modified potential called here \mathcal{V}_{\mathrm{z},n} was introduced in (5.8) and (5.11).

Since the numerator q_n interpolates p_n f at the above 2n+1 nodes, qq_n interpolates the function p_n qf at the same nodes for each polynomial q of degree <n by elementary polynomial interpolation theory, and as the degree of q_nq is still smaller than 2n, the divided difference satisfies [z_1^{(n)},\dots, z_{2n+1}^{(n)}]_{p_nqf}=0, whence the orthogonality of p_n and q for the formal scalar product

\begin{equation} \begin{aligned} \, \notag \langle p,q\rangle_n &= [z_1^{(n)},\dots, z_{2n+1}^{(n)}]_{pqf} \\ &=\sum_{j=1}^{2n+1} \frac{p(z_j^{(n)})q(z_j^{(n)})f(z_j^{(n)})} {\prod_{m\neq j} (z_j^{(n)}-z_m^{(n)})} =\frac{1}{2\pi i} \int_{C} \frac{p(t)q(t)f(t)\,dt}{(t-z_1^{(n)})\dotsb (t-z_{2n+1}^{(n)}) }, \end{aligned} \end{equation} \tag{7.1}
as seen in (2.2).

Consider now the B-spline formula

\begin{equation} \langle p,q\rangle_n = \int_{z_1^{(n)} }^{z_{2n+1}^{(n)}} \frac{B(x)}{(2n)!} \frac{d^{2n}}{dx^{2n}}(p(x)q(x)f(x))\,dx , \end{equation} \tag{7.2}
where B(x) is actually the Curry–Schoenberg B-spline (see [13], Ch. IX, and [14])
\begin{equation*} \begin{aligned} \, B(x) &= 2n[z_1^{(n)},\dots, z_{2n+1}^{(n)}]_{(u-x)_+^{2n-1}} \\ &= M(x;z_1^{(n)},\dots, z_{2n+1}^{(n)}) = 2n\frac{B(x;z_1^{(n)},\dots, z_{2n+1}^{(n)})}{z_{2n+1}^{(n)}-z_1^{(n)}}, \end{aligned} \end{equation*} \notag
where [z_1^{(n)},\dots, z_{2n+1}^{(n)}]_{(u-x)_+^{2n-1}} means the divided difference at u=z_1^{(n)}, \dots, {u=z_{2n+1}^{(n)}} of the function of u whose value is (u-x)^{2n-1} for u>x and 0 for {u<x}.

See in Figure 7 some instances of the splines B(x) approximating \exp(-(n+1/2)x) for n=5,10,15,20.

Open Problem 1. Is there a clear limit behaviour for B? When the points z_j^{(n)} are equidistant (the cardinal case), the (scaled) limit is a Gaussian function [87], [41], [88].

§ 8. Beyond infinity? Exploring off-limits modulus

The approximation scheme stops at k=k_\infty=0.9089\dotsc when c=\infty (Figure 8). Of course, we want to know what happens further, and a possible explanation is that we have to look at c<0. The parameter A is redefined as |a-c|\operatorname{sign}(c) in order to keep the formula c^2=A^2+B^2-2AB\cos\theta. The parameters c, \alpha^2, \zeta, \mathsf{K}, \mathsf{E} and so on vary smoothly when k becomes > k_\infty. For \nu_0 and \nu_1 convenient combinations are \nu_0/\nu_1 and c^3/\nu_1^2 (see Tables 4 and 5).

Table 4.The quantities c, 1/\rho, a, k, \alpha^2, \nu_0, \nu_1, {\mathsf K}, {\mathsf E}, \Pi, {\mathsf K}' and {\mathsf E}'

c1/\rhoak\alpha_2\nu_0\nu_1{\mathsf K}{\mathsf E}\Pi{\mathsf K}'{\mathsf E}'
1.057.070.44-1.97i0.247-1043-2.061.271.601.550.052.821.07
2.518.750.85-1.84i0.550-31.1-2.334.221.721.440.282.071.24
3.315.140.97-1.76i0.657-11.8-2.536.301.801.380.471.92 1.32
5.012.431.09-1.65i0.769-3.85-2.9311.11.941.300.811.781.40
10.010.551.16-1.51i0.857-1.01-3.9830.22.131.221.421.691.46
\infty9.291.19-1.39i0.9090\infty\infty2.321.162.321.651.50
-10.08.411.20-1.29i0.9380.413.79i27.9i2.491.123.481.621.52
-5.07.771.20-1.21i0.9560.612.64i9.6i2.641.094.861.611.54
-3.337.291.19-1.15i0.9660.722.13i5.08i2.771.086.431.601.54
-2.56.911.17-1.10i0.9730.791.83i3.23i2.891.068.181.591.55
-1.05.611.10-0.92i0.9900.931.14i0.76i3.361.0321.81.581.56
-0.54.721.02-0.79i0.9960.970.81i0.26i3.801.0154.21.581.57

Table 5.The quantities \nu_0/\nu_1 and c^3/\nu_1^2

c12.53.3510\infty-10-5-3.3-2.5-1-0.5
\nu_0/\nu_1-1.62-0.55-0.40-0.26-0.1300.140.280.420.571.503.17
c^3/\nu_1^20.620.880.931.011.101.191.281.361.431.501.741.92

So, we can still build the potential, starting with (3.2).

Curiously, \mathcal{V}(z) is now purely imaginary on the real axis. It is not even clear what the cut F may be. We have a solution without a problem.

Open Problem 2. Find a scheme involving a modulus larger than k_\infty.

§ 9. Algorithms

Elliptic integrals of the first and second kind are easily computed using the Gauss–Landen transformations, also related to the arithmetic-geometric mean of two numbers M(a,b)=\lim_{n\to\infty} a_n, where a_0=a, b_0=b, a_{n+1}=(a_n+b_n)/2 and b_{n+1}=\sqrt{a_nb_n}; then \mathsf{K}=\pi/(2M(1,k')) (see [1], Ch. 17, [15], pp. 14–18; [20], p. 39; [46] and [67]).

Indeed, we build an auxiliary sequence c_n=\sqrt{a_n^2-b_n^2} computed11 for n>0 as c_n=c_{n-1}^2/(4a_n); then

\begin{equation*} \frac{2d\varphi_n/a_n}{\sqrt{1-(c_n/a_n)^2 \sin^2\varphi_n}}= \frac{d\varphi_{n+1}/a_{n+1}}{\sqrt{1-(c_{n+1}/a_{n+1})^2 \sin^2\varphi_{n+1}}} \quad \text{when } \varphi_0=\varphi, \end{equation*} \notag
so that \tan(\varphi_{n+1}-\varphi_n)=(b_n/a_n)\tan\varphi_n, or \tan\varphi_{n+1}=\dfrac{2a_{n+1}\tan\varphi_n}{a_n-b_n\tan^2\varphi_n}, or also \sin\varphi_{n+1}=a_{n+1}\sin 2\varphi_n/\sqrt{a_n^2-c_n^2\sin^2\varphi_n}. The new variable \varphi_{n+1} runs from 0 to \pi as \varphi_n runs from 0 to \pi/2 (a complete integral), so that K(c_n/a_n)/a_n=K(c_{n+1}/a_{n+1})/a_{n+1}= \dotsb = K(0)/M=\pi/(2M).

For the complete integral of the second kind, \mathsf{E}/\mathsf{K}=1-\sum_0^\infty 2^{n-1}c_n^2, still with a=1 and b=k'. Convergence is reached extremely fast (quadratically).

For the complete integral of the third kind (4.3) the algorithm is hardly longer (see Bulirsch [17], Algorithms 1 and 2, or Byrd and Friedman [20], § 164.02); we add three sequences d_n, h_n and p_n, where p_0=\sqrt{1-\alpha^2}, d_0=1/p_0 and h_0=1, and p_{n+1}=4^n a_n b_n/p_n+p_n, h_{n+1}=d_n/p_n+h_n and d_{n+1}=2(4^n a_n b_n h_n/p_n+d_n); then \Pi= \lim_{n\to\infty} \pi d_n/(2^{2n+1} a_n^2).

These sophisticated algorithms were used for the computation of the tables. The algorithms are extended to incomplete integrals as well (see [17], Algorithm 3), but simpler (and more robust) algorithms have been preferred when high accuracy is not needed.

Most graphs (Figures 13 and 57) used the incredibly efficient CF-algorithm of Trefethen [86], building very fast excellent near-best rational approximations.

Acknowledgements

Many thanks to S. P. Suetin and the referees.

Just as Chebyshev benefited from lessons of Liouville, the first author had the privilege to receive lessons from A. Aptekarev on 14 February 2001 in Louvain-la-Neuve and 28 May 2001 in Leuven. Spassiba!

Were elliptic functions such a dead subject (before Brent, Salamin, and the Borweins)? Not so in Leuven and Louvain: Georges Lemaître [48], of big bang fame, taught analytical mechanics with elliptic functions examples [38], [44] (to often bewildered students), and Vitold Belevitch used them in filtering problems (see papers by Todd [84]) in MBLE Research Lab (Brussels) (see [63] and [83]). And now, the subject is found in cryptography (Joye and Quisquater [43]).


Bibliography

1. Handbook of mathematical functions with formulae, graphs and mathematical tables, National Bureau of Standards Applied Mathematics Series, 55, eds. M. Abramowitz and I. A. Stegun, Superintendent of Documents, U.S. Government Printing Office, Washington, DC, 1964, xiv+1046 pp.  mathscinet  zmath
2. V. M. Adamjan (Adamyan), D. Z. Arov and M. G. Kreĭn, “Analytic properties of Schmidt pairs for a Hankel operator and the generalized Schur–Takagi problem”, Math. USSR-Sb., 15:1 (1971), 31–73  mathnet  crossref  mathscinet  zmath
3. N. Achyeser (Akhiezer), Zap. Kharkov. Mat. Obshch., 13 (1936), 3–14  zmath
4. N. I. Achieser (Akhiezer), Theory of approximation, Dover Publ., New York, 1992, x+307 pp.  mathscinet  zmath
5. N. I. Akhiezer, Elements of the theory of elliptic functions, Transl. Math. Monogr., 79, Amer. Math. Soc., Providence, RI, 1990, viii+237 pp.  crossref  mathscinet  zmath
6. A. I. Aptekarev, “Sharp constants for rational approximations of analytic functions”, Sb. Math., 193:1 (2002), 1–72  mathnet  crossref  mathscinet  zmath  adsnasa
7. Th. Bagby, “The modulus of a plane condenser”, J. Math. Mech., 17 (1967), 315–329  crossref  mathscinet  zmath
8. Th. Bagby, “On interpolation by rational functions”, Duke Math. J., 36 (1969), 95–104  crossref  mathscinet  zmath
9. G. A. Baker, Jr., Essentials of Padé approximants, Academic Press, New York–London, 1975, xi+306 pp.  mathscinet  zmath
10. L. Baratchart, H. Stahl and M. Yattselev, “Weighted extremal domains and best rational approximation”, Adv. Math., 229:1 (2012), 357–407  crossref  mathscinet  zmath
11. C. M. Bender and S. A. Orszag, Advanced mathematical methods for scientists and engineers, Internat. Ser. Pure Appl. Math., McGraw-Hill Book Co., New York, 1978, xiv+593 pp.  mathscinet  zmath
12. A. Bogatyrev, Extremal polynomials and Riemann surfaces, Springer Monogr. Math., Springer, Heidelberg, 2012, xxvi+150 pp.  crossref  mathscinet  zmath
13. C. de Boor, A practical guide to splines, Appl. Math. Sci., 27, Rev. ed., Springer-Verlag, New York, 2001, xviii+346 pp.  mathscinet  zmath
14. C. de Boor, “B(asic)-spline basics”, Fundamental developments of computer-aided geometric modeling, Academic Press, London, 1993, 27–49; MRC 2952, 1986 https://www.cs.unc.edu/~dm/UNC/COMP258/Papers/bsplbasic.pdf
15. J. M. Borwein and P. B. Borwein, Pi and the AGM. A study in analytic number theory and computational complexity, Canad. Math. Soc. Ser. Monogr. Adv. Texts, Wiley-Intersci. Publ., John Wiley & Sons, Inc., New York, 1987, xvi+414 pp.  mathscinet  zmath
16. D. Braess, “On the conjecture of Meinardus on rational approximation of e^x”, J. Approx. Theory, 36:4 (1982), 317–320  crossref  mathscinet  zmath; II, 40:4 (1984), 375–379  crossref  mathscinet  zmath
17. R. Bulirsch, “Numerical calculation of elliptic integrals and elliptic functions”, Numer. Math., 7 (1965), 78–90  crossref  mathscinet  zmath; II, 353–354  crossref  mathscinet  zmath; III, 13 (1969), 305–315  crossref  mathscinet  zmath
18. H. Burkhardt and W. F. Meyer, “Potentialtheorie (Theorie der Laplace–Poissonschen Differentialgleichung)”, Encyclopädie der mathematischen Wissenschaften, v. 2 A (Analysis), § 7.b, B. G. Teubner, Leipzig, 1899–1916, 464–503
19. V. I. Buslaev and S. P. Suetin, “On the existence of compacta of minimal capacity in the theory of rational approximation of multi-valued analytic functions”, J. Approx. Theory, 206 (2016), 48–67  crossref  mathscinet  zmath
20. P. F. Byrd and M. D. Friedman, Handbook of elliptic integrals for engineers and scientists, Grundlehren Math. Wiss., 67, 2nd rev. ed., Springer-Verlag, New York–Heidelberg, 1971, xvi+358 pp.  crossref  mathscinet  zmath
21. A. J. Carpenter, A. Ruttan and R. S. Varga, “Extended numerical computations on the “1/9” conjecture in rational approximation theory”, Rational approximation and interpolation (Tampa, FL 1983), Lecture Notes in Math., 1105, Springer-Verlag, Berlin, 1984, 383–411  crossref  mathscinet  zmath
22. W. J. Cody, G. Meinardus and R. S. Varga, “Chebyshev rational approximations to e^{-x} on [0,+\infty) and applications to heat-conduction problems”, J. Approx. Theory, 2 (1969), 50–65  crossref  mathscinet  zmath
23. P. D. Dragnev, B. Fuglede, D. P. Hardin, E. B. Saff and N. Zorii, “Minimum Riesz energy problems for a condenser with touching plates”, Potential Anal., 44:3 (2016), 543–577  crossref  mathscinet  zmath
24. K. A. Driver and N. M. Temme, “Zero and pole distribution of diagonal Padé approximants to the exponential function”, Quaest. Math., 22:1 (1999), 7–17  crossref  mathscinet  zmath
25. V. Druskin, S. Güttel and L. Knizhnerman, “Near-optimal perfectly matched layers for indefinite Helmholtz problems”, SIAM Rev., 58:1 (2016), 90–116  crossref  mathscinet  zmath
26. A. Erdélyi, Asymptotic expansions, Dover Publications, Inc., New York, 1956, vi+108 pp.  mathscinet  zmath
27. L. Fox and I. B. Parker, Chebyshev polynomials in numerical analysis, Oxford Univ. Press, London–New York–Toronto, 1968, ix+205 pp.  mathscinet  zmath
28. D. Gaier, Vorlesungen über Approximation im Komplexen, Birkhäuser Verlag, Basel–Boston, Mass., 1980, 174 pp.  crossref  mathscinet  zmath; English. transl., D. Gaier, Lectures on complex approximation, Birkhäuser Boston Inc., Boston, MA, 1987, xvi+196 pp.  crossref  mathscinet  zmath
29. T. Ganelius, “Degree of rational approximation”, Lectures on approximation and value distribution, Sém. Math. Sup., 79, Presses Univ. Montréal, Montreal, QC, 1982, 9–78  mathscinet  zmath
30. A. A. Gončar, “Zolotarev problems connected with rational functions”, Math. USSR-Sb., 7:4 (1969), 623–635  mathnet  crossref  mathscinet  zmath  adsnasa
31. A. A. Gončar, “The rate of rational approximation and the property of single-valuedness of an analytic function in the neighborhood of an isolated singular point”, Math. USSR-Sb., 23:2 (1974), 254–270  mathnet  crossref  mathscinet  zmath  adsnasa
32. A. A. Gončar, “On the speed of rational approximation of some analytic functions”, Math. USSR-Sb., 34:2 (1978), 131–145  mathnet  crossref  mathscinet  zmath  adsnasa
33. A. A. Gonchar, “5.6. Rational approximation of analytic functions”, Studies in linear operators and function theory. 99 Unsolved problems in linear and complex analysis. Ch. 6. Approximation, J. Soviet Math., 26:5 (1984), 2218–2220  mathnet  crossref
34. A. A. Gonchar, “Rational approximation of analytic functions”, Problem 12.17 (Ch. 12. Approximations and capacities), Linear and complex analysis problem. Problem book 3. Part II, Lecture Notes in Math., 1574, Springer-Verlag, Berlin–Heidelberg, 1994, 73–176  crossref  mathscinet  zmath
35. A. A. Gončar and G. López Lagomasino, “On Markov's theorem for multipoint Padé approximants”, Math. USSR-Sb., 34:4 (1978), 449–459  mathnet  crossref  mathscinet  zmath  adsnasa
36. A. A. Gonchar and E. A. Rakhmanov, “Equilibrium distributions and degree of rational approximation of analytic functions”, Math. USSR-Sb., 62:2 (1989), 305–348  mathnet  crossref  mathscinet  zmath  adsnasa
37. A. A. Gončar and E. A. Rakhmanov, “On the rate of rational approximation of analytic functions”, Approximation and optimization (Havana 1987), Lecture Notes in Math., 1354, Springer-Verlag, Berlin, 1988, 25–42  crossref  mathscinet  zmath
38. L. Haine, “Geodesic flow on \operatorname{SO}(4) and Abelian surfaces”, Math. Ann., 263:4 (1983), 435–472  crossref  mathscinet  zmath
39. P. Henrici, Applied and computational complex analysis, v. 3, Pure Appl. Math. (N. Y.), Discrete Fourier analysis–Cauchy integrals–construction of conformal maps–univalent functions, Wiley-Intersci. Publ., John Wiley & Sons, Inc., New York, 1986, xvi+637 pp.  mathscinet  zmath
40. E. Hille, Analytic function theory, v. I, Ginn and Co., Boston, MA, 1959, xi+308 pp.  mathscinet  zmath; 2nd ed., Chelsea Publ. Co., New York, 1982, xi+308 pp.  zmath
41. Z. G. Ignatov and V. K. Kaishev, B-splines and linear combinations of uniform order statistics, MRC tech. summary rep. # 2817, Math. Res. Center Univ. Wisconsin–Madison, 1985, 23 pp.
42. E. Jahnke, F. Emde and F. Lösch, Tafeln höherer Funktionen (Tables of higher functions), ed. 6. Aufl., B. G. Teubner Verlagsgesellschaft, Stuttgart; McGraw-Hill Book Co., Inc., New York–Toronto–London, 1960, xii+318 pp.  mathscinet  zmath
43. M. Joye and J.-J. Quisquater, “Hessian elliptic curves and side-channel attacks”, Cryptographic hardware and embedded systems – CHES 2001 (Paris 2001), Lecture Notes in Comput. Sci., 2162, Springer-Verlag, Berlin, 2001, 402–410  crossref  mathscinet  zmath
44. M. Kac and P. van Moerbeke, “A complete solution of the periodic Toda problem”, Proc. Nat. Acad. Sci. U.S.A., 72:8 (1975), 2879–2880  crossref  mathscinet  zmath  adsnasa
45. O. D. Kellogg, Foundations of potential theory, Grundlehren Math. Wiss., 31, J. Springer, Berlin, 1929, ix+334 pp.  mathscinet  zmath
46. L. V. King, On the direct numerical calculation of elliptic functions and integrals, Cambridge Univ. Press, Cambridge, 1924, viii+42 pp.  zmath
47. M. A. Komarov, “A note on rational approximations to e^x, constructed by Németh and Newman”, J. Approx. Theory, 240 (2019), 126–128  crossref  mathscinet  zmath
48. G. Lemaître, “Calcul des intégrales elliptiques”, Acad. Roy. Belgique. Bull. Cl. Sci. (5), 33 (1947), 200–211  mathscinet
49. G. López Lagomasino, “Survey of multipoint Padé approximation to Markov type meromorphic functions and asymptotic properties of the orthogonal polynomials generated by them”, Polynômes orthogonaux et applications (Bar-le-Duc 1984), Lecture Notes in Math., 1171, Springer-Verlag, Berlin, 1985, 309–316  crossref  mathscinet  zmath
50. G. L. Lopes, “On the asymptotics of the ratio of orthogonal polynomials and convergence of multipoint Padé approximants”, Math. USSR-Sb., 56:1 (1987), 207–219  mathnet  crossref  mathscinet  zmath  adsnasa
51. G. López and E. A. Rakhmanov, “Rational approximations, orthogonal polynomials and equilibrium distributions”, Orthogonal polynomials and their applications (Segovia 1986), Lecture Notes in Math., 1329, Springer-Verlag, Berlin, 1988, 125–157  crossref  mathscinet  zmath
52. A. P. Magnus, “On the use of Carathéodory–Fejér method for investigating ‘1/9’ and similar constants”, Nonlinear numerical methods and rational approximation (Wilrijk 1987), Math. Appl., 43, D. Reidel Publ. Co., Dordrecht, 1988, 105–132  crossref  mathscinet  zmath
53. A. P. Magnus, “Asymptotics and super asymptotics for best rational approximation error norms to the exponential function (the ‘1/9’ problem) by the Carathéodory–Fejér's method”, Nonlinear methods and rational approximation, II (Wilrijk 1993), Math. Appl., 296, Kluwer Acad. Publ., Dordrecht, 1994, 173–185  crossref  mathscinet  zmath
54. A. P. Magnus and J. Meinguet, “The elliptic functions and integrals of the “1/9” problem”, Numer. Algorithms, 24:1–2 (2000), 117–139  crossref  mathscinet  zmath  adsnasa
55. F. Marcellán and A. Martínez-Finkelshtein, “Guillermo López Lagomasino: mathematical life”, Recent trends in orthogonal polynomials and approximation theory, Contemp. Math., 507, Amer. Math. Soc., Providence, RI, 2010, 1–24  crossref  mathscinet  zmath
56. A. Martínez-Finkelshtein, “Trajectories of quadratic differentials and approximations of exponents on the semiaxis”, Complex methods in approximation theory (Almería 1995), Monogr. Cienc. Tecnol., 2, Univ. Almería, Serv. Publ., Almería , 1997, 69–84  mathscinet  zmath
57. A. Martínez-Finkelshtein, “Equilibrium problems of potential theory in the complex plane”, Orthogonal polynomials and special functions, Lecture Notes in Math., 1883, Springer-Verlag, Berlin, 2006, 79–117  crossref  mathscinet  zmath
58. A. Martínez-Finkelshtein and E. A. Rakhmanov, “Do orthogonal polynomials dream of symmetric curves?”, Found. Comput. Math., 16:6 (2016), 1697–1736  crossref  mathscinet  zmath
59. G. Meinardus, Approximation of functions: theory and numerical methods, Exp. transl. of the German ed., Springer Tracts Nat. Philos., 13, Springer-Verlag New York, Inc., New York, 1967, viii+198 pp.  crossref  mathscinet  zmath
60. J. Meinguet, “A simplified presentation of the Adamjan–Arov–Krein approximation theory”, Computational aspects of complex analysis (Braunlage 1982), NATO Adv. Sci. Inst. Ser. C: Math. Phys. Sci., 102, D. Reidel Publ. Co., Dordrecht, 1983, 217–248  crossref  mathscinet  zmath
61. J. Meinguet, “Once again: the Adamjan–Arov–Krein approximation theory”, Nonlinear numerical methods and rational approximation (Wilrijk 1987), Math. Appl., 43, D. Reidel Publ. Co., Dordrecht, 1988, 77–91  crossref  mathscinet  zmath
62. J. Meinguet, “An electrostatics approach to the determination of extremal measures”, Math. Phys. Anal. Geom., 3:4 (2000), 323–337  crossref  mathscinet  zmath  adsnasa
63. J. Meinguet and V. Belevitch, “On the realizability of ladder filters”, IRE Trans. Circuit Theory, 5:4 (1958), 253–255  crossref
64. L. M. Milne-Thomson, The calculus of finite differences, Macmillan & Co., Ltd., London, 1933, xxiii+558 pp.  mathscinet  zmath
65. Z. Nehari, Conformal mapping, McGraw-Hill Book Co., Inc., New York–Toronto–London, 1952, viii+396 pp.  mathscinet  zmath
66. J. Nuttall, “Location of poles of Padé approximants to entire functions”, Rational approximation and interpolation (Tampa, FL 1983), Lecture Notes in Math., 1105, Springer-Verlag, Berlin, 1984, 354–363  crossref  mathscinet  zmath
67. NIST handbook of mathematical functions, eds. F. W. J. Olver, D. W. Lozier, R. F. Boisvert and C. W. Clark, U.S. Department of Commerce, National Institute of Standards and Technology, Washington, DC; Cambridge Univ. Press, Cambridge, 2010, xvi+951 pp.  mathscinet  zmath
68. S. Paszkowski, Zastosowania numeryczne wielomianów i szeregów Czebyszewa [Numerical applications of Čebyšev polynomials and series], Państwowe Wydawnictwo Naukowe, Warsaw, 1975, 481 pp.  mathscinet  zmath
69. O. Perron, Die Lehre von den Kettenbrüchen, B. G. Teubner, Leipzig–Berlin, 1913, xiii+520 pp.  zmath
70. 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  crossref  mathscinet  zmath
71. E. A. Rakhmanov, “The Gonchar–Stahl \rho^2-theorem and associated directions in the theory of rational approximations of analytic functions”, Sb. Math., 207:9 (2016), 1236–1266  mathnet  crossref  mathscinet  zmath  adsnasa
72. E. A. Rakhmanov, “Zero distribution for Angelesco Hermite–Padé polynomials”, Russian Math. Surveys, 73:3 (2018), 457–518  mathnet  crossref  mathscinet  zmath  adsnasa
73. E. B. Saff, “Logarithmic potential theory with applications to approximation theory”, Surv. Approx. Theory, 5 (2010), 165–200  mathscinet  zmath
74. E. B. Saff and V. Totik, Logarithmic potentials with external fields, Grundlehren Math. Wiss., 316, Springer-Verlag, Berlin, 1997, xvi+505 pp.  crossref  mathscinet  zmath
75. M. A. Snyder, Chebyshev methods in numerical approximation, Prentice-Hall, Inc., Englewood Cliffs, NJ, 1966, x+114 pp.  mathscinet  zmath
76. H. Stahl, “Orthogonal polynomials with complex-valued weight function. I”, Constr. Approx., 2:3 (1986), 225–240  crossref  mathscinet  zmath
77. H. Stahl, “Orthogonal polynomials with complex-valued weight function. II”, Constr. Approx., 2:3 (1986), 241–251  crossref  mathscinet  zmath
78. H. Stahl, “A note on three conjectures by Gonchar on rational approximation”, J. Approx. Theory, 50:1 (1987), 3–7  crossref  mathscinet  zmath
79. H. Stahl, “Convergence of rational interpolants”, Bull. Belg. Math. Soc. Simon Stevin, 1996, Suppl., 11–32  mathscinet  zmath
80. H. Stahl, Calculating rational best approximants on (-\infty,0], 3ièmes journées approximation 15–16 mai 2008, Lille, unpubl. manuscript
81. H. Stahl and Th. Schmelzer, “An extension of the '1/9'-problem”, J. Comput. Appl. Math., 233:3 (2009), 821–834  crossref  mathscinet  zmath
82. H. Stahl, L. Baratchart and M. Yattselev, AAK approximants to functions with branch points, Workshop ‘Approximation days 2012’ (Leuven 2012) https://wis.kuleuven.be/events/ad12/talks/stahl0.pdf
83. J.-P. Thiran and C. Detaille, “Chebyshev polynomials on circular arcs in the complex plane”, Progress in approximation theory, Academic Press, Inc., Boston, MA, 1991, 771–786  mathscinet  zmath
84. J. Todd, “Applications of transformation theory: a legacy from Zolotarev (1847–1878)”, Approximation theory and spline functions (St. John's, NL 1983), NATO Adv. Sci. Inst. Ser. C: Math. Phys. Sci., 136, D. Reidel Publishing Co., Dordrecht, 1984, 207–245  crossref  mathscinet  zmath; “A legacy from E. I. Zolotarev (1847–1878)”, Math. Intelligencer, 10:2 (1988), 50–53  crossref  mathscinet  zmath
85. L. N. Trefethen and M. H. Gutknecht, “The Carathéodory–Fejér method for real rational approximation”, SIAM J. Numer. Anal., 20:2 (1983), 420–436  crossref  mathscinet  zmath  adsnasa
86. L. N. Trefethen, “MATLAB programs for CF approximation”, Approximation theory V, Academic Press, Inc., Boston, MA, 1986, 599–602  mathscinet  zmath
87. L. N. Trefethen and M. Javed, Central limit theorem, 2012 https://www.chebfun.org/examples/stats/CentralLimitTheorem.html
88. M. Unser, A. Aldroubi and M. Eden, “On the asymptotic convergence of B-spline wavelets to Gabor functions”, IEEE Trans. Inform. Theory, 38:2 (1992), 864–872  crossref  mathscinet  zmath
89. R. S. Varga, Scientific computation on mathematical problems and conjectures, CBMS-NSF Regional Conf. Ser. in Appl. Math., 60, SIAM, Philadelphia, PA, 1990, vi+122 pp.  crossref  mathscinet  zmath
90. J. L. Walsh, Interpolation and approximation by rational functions in the complex domain, Amer. Math. Soc. Colloq. Publ., XX, Amer. Math. Soc., New York, 1935, ix+383 pp.  crossref  mathscinet  zmath
91. J. L. Walsh, The location of critical points of analytic and harmonic functions, Amer. Math. Soc. Colloq. Publ., 34, Amer. Math. Soc., New York, 1950, viii+384 pp.  mathscinet  zmath
92. R. Wong and J.-M. Zhang, “Asymptotic expansions of the generalized Bessel polynomials”, J. Comput. Appl. Math., 85:1 (1997), 87–112  crossref  mathscinet  zmath

Citation: A. P. Magnus, J. Meinguet, “Strong asymptotics of the best rational approximation to the exponential function on a bounded interval”, Sb. Math., 215:12 (2024), 1666–1719
Citation in format AMSBIB
\Bibitem{MagMei24}
\by A.~P.~Magnus, J.~Meinguet
\paper Strong asymptotics of the best rational approximation to the exponential function on a~bounded interval
\jour Sb. Math.
\yr 2024
\vol 215
\issue 12
\pages 1666--1719
\mathnet{http://mi.mathnet.ru/eng/sm8815}
\crossref{https://doi.org/10.4213/sm8815e}
\mathscinet{http://mathscinet.ams.org/mathscinet-getitem?mr=4868566}
\adsnasa{https://adsabs.harvard.edu/cgi-bin/bib_query?2024SbMat.215.1666M}
\isi{https://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=Publons&SrcAuth=Publons_CEL&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=001443898100004}
\scopus{https://www.scopus.com/record/display.url?origin=inward&eid=2-s2.0-85219622443}
Linking options:
  • https://www.mathnet.ru/eng/sm8815
  • https://doi.org/10.4213/sm8815e
  • https://www.mathnet.ru/eng/sm/v215/i12/p89
  • This publication is cited in the following 2 articles:
    1. S. P. Suetin, “O skalyarnykh podkhodakh k izucheniyu predelnogo raspredeleniya nulei mnogochlenov Ermita–Pade dlya sistemy Nikishina”, UMN, 80:1(481) (2025), 85–152  mathnet  crossref
    2. S. P. Suetin, “Asimptoticheskie svoistva mnogochlenov, opredelyaemykh sdvinutymi usloviyami ortogonalnosti”, UMN, 80:2(482) (2025), 169–170  mathnet  crossref
    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:358
    Russian version PDF:17
    English version PDF:13
    Russian version HTML:43
    English version HTML:70
    References:26
    First page:13
     
      Contact us:
     Terms of Use  Registration to the website  Logotypes © Steklov Mathematical Institute RAS, 2025