1 Introduction

Consider a positive measure dฯƒ on a real interval [a,b] having infinitely many points of increase and finite moments of all orders. It is well known that the corresponding monic orthogonal polynomials {ฯ€n} satisfy a three-term recurrence relation

$$ \begin{array}{rl} \pi_{-1}(t)&=\ \ 0,\quad \pi_{0}(t)=1,\\ \pi_{k+1}(t)&=\ \ (t-\alpha_{k})\pi_{k}(t)-\beta_{k}\pi_{k-1}(t),\quad k=0,1,2,\dots, \end{array} $$
(1)

where \(\alpha _{k}\in \mathbb {R}\), ฮฒk >โ€‰0, and by convention \(\beta _{0}={{\int \limits }_{a}^{b}} d\sigma (t)\). In [11] Gautschi and Li considered a modification of the original measure, for a fixed integer n โ‰ฅโ€‰1, given by

$$ d\widehat\sigma_{n}(t)=[\pi_{n}(t)]^{2} d\sigma(t)\quad\text{on}\quad [a,b], $$
(2)

and studied the corresponding (monic) orthogonal polynomials \(\widehat \pi _{m}=\widehat \pi _{m,n}\), \(m=0,1,2,\dots \). As pointed out by the authors in [11], this kind of modifications of measures are useful, for instance, when dealing with constrained polynomial least squares approximation (see, e.g., [8]), or to provide additional interpolation points (the zeros of the induced polynomial \( \{\widehat {\pi }_{n+1,n}\}\)) in the process of extending Lagrange interpolation at the zeros of ฯ€n (see [3]). Taking into account these and other applications, it seems natural to consider the numerical computation of integrals of the form

$$ I_{\sigma} (f) = I(f;\sigma,n) = \int f(t) d\widehat{\sigma}_{n} (t) $$

by means of quadrature formulae; in particular, Gauss type rules are our main subject of interest. It is well known that the zeros and nodes of the Gauss rule can be efficiently computed by means of the eigenvalues and eigenvectors of the related tridiagonal Jacobi matrix, whose entries are given in terms of the above mentioned recursion coefficients. Then, the following problem arises in a natural way: given the recursion coefficients ฮฑk,ฮฒk for dฯƒ, determine the recursion coefficients \(\hat \alpha _{k},\hat \beta _{k}\) for \(d\hat \sigma _{n}\). Unfortunately, in general it is not feasible to get closed analytic expressions of the entries of the Jacobi matrix for the induced measure \( d\widehat {\sigma }_{n}\) in terms of the corresponding for dฯƒ; in this sense, in [11] a stable numerical algorithm is given. But in the particular case of the well-known four Chebyshev weights dฯƒ[i],i =โ€‰1,2,3,4, where

$$ \begin{array}{ll} d\sigma^{[1]}(t)=\frac{1}{\sqrt{1-t^{2}}} dt,& \qquad d\sigma^{[2]}(t)=\sqrt{1-t^{2}} dt,\\[0.15in] d\sigma^{[3]}(t)=\sqrt{\frac{1+t}{1-t}} dt, & \qquad d\sigma^{[4]}(t)=\sqrt{\frac{1-t}{1+t}} dt \end{array} $$
(3)

the related induced orthogonal polynomials are easily expressible as combinations of Chebyshev polynomials of the first kind Tk, i.e., orthogonal polynomials with respect to the Chebyshev weight dฯƒ = dฯƒ[1] (see [11, ยง3]). These results are very useful for the analysis of the error of the related quadrature formulas.

In the present paper, we focus on the first modified Chebyshev measure, namely

$$ d\hat\sigma_{n}(t) = d\hat\sigma_{n}^{[1]}(t)=\left[\overset{\circ}{T}_{n}(t)\right]^{2} d\sigma(t),\quad -1<t<1, $$
(4)

where

$$ d\sigma(t)=\frac{1}{\sqrt{1-t^{2}}} dt \text{and} \overset{\circ}{T}_{n}(t)=2^{1-n}T_{n}(t) , $$
(5)

with \(\overset {\circ }{T}_{n}\) denoting the corresponding n th-degree monic Chebyshev polynomial. As we said above, for this, as well as for the other modified Chebyshev weights, it is feasible to get closed expressions of the entries of the Jacobi tridiagonal matrices in terms of the corresponding for the original Chebyshev ones. These are collected in previous papers as [11, Theorems 3.1โ€“3.7] and [21, Section 2]. Such results will be useful, among other things, for computing the actual (sharp) value of the quadrature error in the numerical examples.

In this paper, we aim to obtain accurate estimates of the error of the Gaussโ€“Kronrod quadrature formulas for analytic integrands related to this modification of the first kind Chebyshev measure; therefore, this partially completes the analysis started in [25], where those estimates were obtained for the ordinary Gauss quadrature formulas. In 1964, A. S. Kronrod, trying to estimate in a feasible way the error of the n-point Gauss-Legendre quadrature formula, developed the now called Gauss-Kronrod quadrature formula for the Legendre measure (cf.[15, 16]). For a general measure dฯƒ this formula has the form

$$ {{\int}_{a}^{b}}f(t)d\sigma(t)=\sum\limits_{\nu=1}^{n} W_{\nu} f(\tau_{\nu})+\sum\limits_{\mu=1}^{n+1} W_{\mu}^{*} f(\tau_{\mu}^{*})+R_{n}(f), $$
(6)

where ฯ„ฮฝ are the zeros of ฯ€n, and the \(\tau _{\mu }^{*}, W_{\nu }, W_{\mu }^{*}\) are chosen such that (6) has maximum degree of exactness. It turns out that a necessary and sufficient condition for this to happen is that \(\tau _{\mu }^{*}\) be the zeros of the polynomial \(\pi ^{*}_{n+1}\) (see [7, Corollary]), uniquely determined by the orthogonality relations

$$ {{\int}_{a}^{b}} \pi^{*}_{n+1}(t) t^{k} \pi_{n}(t) d\sigma(t)=0,\quad k=0,1,\dots,n. $$
(7)

Observe that (7) implies that \(\pi ^{*}_{n+1}\) is a polynomial orthogonal with respect to a variable-sign measure, from which the fact that its zeros be simple and belong to the interval (a,b) is not guaranteed in advance. Polynomials of this kind were considered for the first time by T. J. Stieltjes in 1894, for the Legendre measure dฯƒ(t) = dt on [โˆ’โ€‰1,1]. Stieltjes, in a letter to Hermite (see [1, vol 2, pp. 439โ€“441]), conjectured that \(\pi ^{*}_{n+1}\) has n +โ€‰1 real and simple zeros, all contained in (โˆ’โ€‰1,1), and interlacing with the zeros of the n th-degree Legendre polynomial. Stieltjesโ€™ conjectures were proved by Szegล‘ in 1935 (cf. [31]), not only for the Legendre but also for the Gegenbauer measure dฯƒ(t) = (1 โˆ’ t2)ฮปโˆ’โ€‰1/2dt on [โˆ’โ€‰1,1], when 0 < ฮป โ‰คโ€‰2. After that, the polynomials \(\pi ^{*}_{n+1}\), now appropriately called Stieltjes polynomials, have apparently gone unnoticed until Kronrodโ€™s papers in 1964 (cf.[15, 16]). The connection between Stieltjes polynomials and Gauss-Kronrod formulae was pointed out by Mysovskih in [20], and independently by Barrucand in [2]. A nice and detailed survey of Kronrod rules in the last 50 years is provided by Notaris [24]. Numerically stable and effective procedures for calculating Gauss-Kronrod formulas are proposed in [17] and [4].

We consider here Gauss-Kronrod quadrature formulas for the modified Chebyshev weight function of the first kind \(d\widehat \sigma _{n} = d\widehat \sigma _{n}^{[1]}\), that is,

$$ I_{\sigma}(f)=I_{n}(f)+R_{n}(f), $$
(8)

where

$$ I_{\sigma}(f)={\int}_{-1}^{1}f(t)d\widehat{\sigma}_{n}^{[1]}(t),\quad I_{n}(f)= \sum\limits_{\nu=1}^{n} W_{\nu} f(\tau_{\nu})+ \sum\limits_{\mu=1}^{n+1} W_{\mu}^{*} f(\tau_{\mu}^{*}), $$

under the assumption that all \(\tau _{\nu }, \tau _{\mu }^{*}\) belong to [โˆ’โ€‰1,1]. Our analysis of the error is based on its well-known representation in terms of an integral contour of a suitable kernel; namely, if we use a Gauss-Kronrod rule In(f) with 2n +โ€‰1 nodes to approximate the value of the integral Iฯƒ(f) for a certain positive measure ฯƒ (hereafter, we assume the absolute continuity of the measure ฯƒ and, hence, that dฯƒ(t) = w(t)dt) on the real interval [โˆ’โ€‰1,1] and an analytic integrand f in a neighborhood ฮฉ of this interval, the error of quadrature admits the following integral representation (see, e. g., [12])

$$ R_{n}(f) = I_{\sigma} (f) - I_{n}(f) = \frac{1}{2\pi i} \oint_{{\varGamma}} K_{n}(z) f(z) dz , $$
(9)

where the kernel Kn is given by

$$ K_{n} (z) = \frac{{\varrho}_{n} (z)}{\pi_{n} (z) \pi_{n+1}^{*}(z)} , {\varrho}_{n} (z) = {\int}_{-1}^{1} \frac{\pi_{n} (t) \pi_{n+1}^{*}(t)}{z-t} w(t) dt , $$
(10)

with ฯ€n denoting, as usual, the n th degree orthogonal polynomial with respect to w, \(\pi _{n+1}^{*}\) denoting the corresponding Stieltjes polynomial of degree n +โ€‰1 for the modified Chebyshev weight, ฯฑn is the commonly called 2nd kind function associated to the nodal polynomial, and ฮ“ โŠ‚ฮฉ is any closed smooth contour surrounding the real interval [โˆ’โ€‰1,1]. Elliptic contours \(\mathcal {E}_{\rho }\) with foci at the points ยฑโ€‰1 and semi-axes given by \( \frac {1}{2} (\rho + \rho ^{-1}) \) and \( \frac {1}{2} (\rho - \rho ^{-1}) ,\) with ฯ >โ€‰1, are often considered as contours of integration, in order to get suitable estimations of the error of quadrature; this is due to the fact that they are the level curves for the conformal function which maps the exterior of [โˆ’โ€‰1,1] onto the exterior of the unit circle |z| >โ€‰1 in the complex plane. In this sense, these elliptic level curves admit the expression

$$ \mathcal{E}_{\rho} = \{z\in \mathbb{C}: |\phi(z)| = |z+\sqrt{z^{2}-1}| = \rho\} , $$
(11)

where ฯ >โ€‰1 and the branch of \(\sqrt {z^{2}-1}\) is taken so that |ฯ•(z)| >โ€‰1 for |z| >โ€‰1. On the other hand, the inverse function of ฯ•, that is, the well-known Joukowsky transform, given by

$$ z = \frac{1}{2} \left( \xi + \frac{1}{\xi}\right),\ z\in \mathbb{C} \setminus [-1,1] , |\xi|>1 , $$
(12)

will also be used in the subsequent sections.

The outline of the current paper is as follows. In Sectionย 2 an explicit expression for the kernel (10) related to the induced Chebyshev weight \( d\widehat {\sigma }^{[1]}_{n}\) is provided, which will be useful to get appropriate bounds for the error of the corresponding Gauss-Kronrod rules, which represents the main contribution of the paper. In addition, the accuracy of the obtained bounds is checked by means of some illustrative numerical examples in Sectionย 3. Finally, and for the sake of completeness, similar computations for the kernels corresponding to the other modified Chebyshev measures \(d\sigma _{n}^{[i]} , i=2,3,4\), are gathered in the final appendix.

To end this introduction, let us say that the problem of estimating the quadrature error for Gaussโ€“type rules has been thoroughly studied in the literature; see the references [12, 18, 19, 22, 23], and [26,27,28,29,30], to only cite a few. See also [5] for a very recent survey of the error estimates of Gaussian type quadrature formulae for analytic functions on ellipses.

Let us finally point out that the seemingly restricted scope of our analysis is offset, in our opinion, by the extreme sharpness of the estimations shown in Sectionย 3.

2 Error bounds of Gaussโ€“Kronrod rules for the measure \(d\widehat {\sigma }_{n}^{[1]}\)

Hereafter, the (monic) orthogonal polynomials relative to the positive measure \(d\widehat \sigma ^{[1]}_{n}(t)=[\pi _{n}(t)]^{2} d\sigma ^{[1]}(t)\), defined in (2), will be denoted simply by \(\widehat \pi _{m,n},\ m=0,1,2,\dots \). For simplicity, here and in the next section we only consider what may be referred to as the โ€œdiagonalโ€ setting, that is, the case where m = n and we simply denote \(\widehat \pi _{n} = \widehat \pi _{n,n}\); on the other hand, \(\pi ^{*}_{n+1}\) will denote the corresponding Stieltjes polynomial. Our first result gives the explicit expression of the kernel Kn in this case.

Lemma 1

The kernel Kn is given by

$$ K_{n}(z)= -\frac{\pi(2\xi^{2n}+1)}{2^{2n-2}\xi^{2n-1}(\xi^{4n}-1)(\xi^{2}-1)}, $$
(13)

with ฮพ given by (12).

Proof

By (10) the corresponding kernel is given by

$$ K_{n}(z)= \frac{{\varrho}_{n}(z)}{\widehat\pi_{n}(z) \pi^{*}_{n+1}(z)},\quad z\notin[-1,1], $$

where

$$ \widehat\pi_{n}(t)=\mathring{T}_{n}(t),\quad \pi^{*}_{n+1}(t)=(t^{2}-1)\mathring{U}_{n-1}(t),\quad \mathring{T}_{n}(t)=\frac{1}{2^{n-1}}T_{n}(t), $$

with Uk denoting as usual the Chebyshev polynomial of the second kind of degree k, and ลฎk being the monic one, and

$${\varrho}_{n}(z)={\int}_{-1}^{1}\frac{\widehat\pi_{n}(z)\pi^{*}_{n+1}(z)}{z-t}\frac{\mathring{T}_{n}^{2}(t) dt}{\sqrt{1-t^{2}}} ={\int}_{-1}^{1}\frac{\mathring{T}_{n}(t)(t^{2}-1)\mathring{U}_{n-1}(t)}{z-t}\frac{\mathring{T}_{n}^{2}(t) dt}{\sqrt{1-t^{2}}},$$

that is,

$$ \begin{array}{@{}rcl@{}} {\varrho}_{n}(z)&=&{\int}_{-1}^{1}\frac{(t^{2}-1)\mathring{U}_{n-1}(t)}{z-t}\frac{\mathring{T}_{n}^{3}(t) dt}{\sqrt{1-t^{2}}}\\ &=& - {\int}_{0}^{\pi}\frac{\cos^{3}(n\theta)(-\sin^{2}\theta)\frac{\sin n\theta}{\sin \theta} d\theta }{2^{4n-4}(z-\cos \theta) }\\ &=& -\frac{1}{2^{4n-2}}{\int}_{0}^{\pi} \frac{(1+\cos 2n\theta) \sin 2n\theta \sin \theta}{z-\cos \theta}d\theta\\ &=& -\frac{1}{2^{4n-1}}{\int}_{0}^{\pi} \frac{\cos(2n-1)\theta - \cos(2n+1)\theta +\frac{1}{2}(\cos (4n-1)\theta -\cos(4n+1) \theta) }{z-\cos \theta}d\theta\\ &=& -\frac{1}{2^{4n-1}}\bigg({\int}_{0}^{\pi} \frac{\cos(2n-1)\theta}{z-\cos \theta}d\theta - {\int}_{0}^{\pi} \frac{\cos(2n+1)\theta}{z-\cos \theta}d\theta \\ &&+ \frac{1}{2}{\int}_{0}^{\pi} \frac{\cos(4n-1)\theta}{z-\cos \theta}d\theta -\frac{1}{2}{\int}_{0}^{\pi} \frac{\cos(4n+1)\theta}{z-\cos \theta}d\theta \bigg). \end{array} $$

Thus, by using the identity (see [12, p. 1176])

$$ {\int}_{0}^{\pi}\frac{\cos m\theta}{z-\cos \theta}d\theta=\frac{2\pi}{\xi^{m}(\xi-\xi^{-1})}, $$
(14)

we get

$$ \begin{array}{@{}rcl@{}} {\varrho}_{n}(z)&=& -\frac{\pi}{2^{4n-1}(\xi-\xi^{-1})}\bigg(\frac{2}{\xi^{2n-1}}-\frac{2}{\xi^{2n+1}}+\frac{1}{\xi^{4n-1}} -\frac{1}{\xi^{4n+1}} \bigg)\\ &=& -\frac{\pi}{2^{4n-1}(\xi-\xi^{-1})}\bigg(\frac{2(\xi^{2}-1)}{\xi^{2n+1}}+\frac{\xi^{2}-1}{\xi^{4n+1}} \bigg)\\ &=& -\frac{\pi(2\xi^{2n}+1)}{2^{4n-1}\xi^{4n}}. \end{array} $$
(15)

Next, to compute the denominator of the kernel the following representation will be used (see [12, pp. 1176โ€“1177])

$$ \mathring{T}_{n}(z)=\frac{1}{2^{n-1}}T_{n}(z)=\frac{1}{2^{n}}\bigg(\xi^{n}+\frac{1}{\xi^{n}} \bigg),\ \mathring{U}_{n-1}(z)=\frac{1}{2^{n-1}}U_{n-1}(z)=\frac{1}{2^{n-1}} \frac{\xi^{n}-\frac{1}{\xi^{n}}}{\xi-\frac{1}{\xi}}. $$

Therefore, we obtain

$$ \begin{array}{@{}rcl@{}}\pi^{*}_{n+1}(z)&=&(z^{2}-1)\mathring{U}_{n-1}(z)=\bigg[ \frac{1}{4}\bigg(\xi+\frac{1}{\xi}\bigg)^{2} -1 \bigg]\frac{1}{2^{n-1}} \frac{\xi^{n}-\frac{1}{\xi^{n}}}{\xi-\frac{1}{\xi}}\\ &=&\frac{1}{2^{n+1}}\bigg(\xi-\frac{1}{\xi} \bigg) \bigg(\xi^{n}-\frac{1}{\xi^{n}} \bigg), \end{array} $$

and

$$ \begin{array}{@{}rcl@{}} \pi^{*}_{n+1}(z) \widehat\pi_{n}(z)&=&\frac{1}{2^{n+1}}\bigg(\xi-\frac{1}{\xi} \bigg) \bigg(\xi^{n}-\frac{1}{\xi^{n}} \bigg) \frac{1}{2^{n}} \bigg(\xi^{n}+\frac{1}{\xi^{n}} \bigg)\\ &=&\frac{1}{2^{2n+1}} \bigg(\xi-\frac{1}{\xi} \bigg) \bigg(\xi^{2n}-\frac{1}{\xi^{2n}} \bigg). \end{array} $$
(16)

Then, the proof of (13) easily follows. โ–ก

Now, we are in a position to obtain bounds for the error of the Gaussโ€“Kronrod quadrature formula using (9) and (13). To do it, several methods will be employed.

2.1 \(L^{\infty }\)โ€“bound for the error

On the sequel, for a function g and a compact subset E of the complex plane, the \(L^{\infty }\)โ€“norm of g on E will be denoted by

$$\|g\|_{E} = \max_{z\in E} |g(z)| .$$

Now, from (9) and taking \({\varGamma } = \mathcal {E}_{\rho }\) for certain ฯ >โ€‰1, we easily get that if f is analytic on \(\mathcal {E}_{\rho }\) and its interior,

$$ |R_{m}(f)| \leq \frac{l(\mathcal{E}_{\rho})}{2\pi} \|K_{m}\|_{\mathcal{E}_{\rho}} \|f\|_{\mathcal{E}_{\rho}} , $$
(17)

where \(l(\mathcal {E}_{\rho })\) represents the length of the ellipse \(\mathcal {E}_{\rho }\). If we denote by Dฯ the closed interior of \(\mathcal {E}_{\rho }\), define

$$\rho_{max} = \sup \{\rho >1 : f \text{is analytic on} D_{\rho}\} .$$

Now, set

$$ a_{j}=\frac{\rho^{j}+\rho^{-j}}{2}, \ j \in \mathbb{N} . $$
(18)

Next, we have the following \(L^{\infty }\)โ€“bound for the error of the Gaussโ€“Kronrod quadrature formula.

Theorem 1

The error of the Gaussโ€“Kronrod quadrature formula for \(d\widehat {\sigma }_{n}^{[1]}\) is bounded by

$$ r_{1}(f)=\inf_{1<\rho< \rho_{\max}} \left[\frac{\pi(2\rho^{2n}+1) a_{1}\left( 1-\frac{1}{4}a_{1}^{-2}-\frac{3}{64}a_{1}^{-4}-\frac{5}{256}a_{1}^{-6}\right)\|f\|_{\mathcal{E}_{\rho}}} {2^{2n-2}\rho^{2n-1}(\rho^{4n}-1)(\rho^{2}-1)}\right] , $$
(19)

where the expression of aj is given in (18).

Proof

From (13) and using polar coordinates and the Joukowsky transform (12), the modulus of the kernel in this case may be expressed in the form

$$ |K_{n}(z)|= \frac{\pi\sqrt{4\rho^{4n}+4\rho^{2n}\cos2n\theta+1}}{2^{2n-1}\cdot\rho^{4n}\sqrt{(a_{2}-\cos 2\theta)(a_{4n}-\cos 4n\theta)}}, $$
(20)

with the aj given by (18), because

$$ K_{n}(z)= -\frac{\pi(2\xi^{2n}+1)}{2^{2n-2}\xi^{4n}\left( \xi^{2n}-1/\xi^{2n}\right)\left( \xi-1/\xi\right)} $$

and

$$ |\xi^{k}-1/\xi^{k}|=\sqrt{2}\sqrt{a_{2k}-\cos{2k\theta}}, \quad k\in\mathbb{N}, $$
$$ |2\xi^{2n}+1|=\sqrt{4\rho^{4n}+4\rho^{2n}\cos2n\theta+1}. $$

Since the numerator and denominator of this expression obviously reach its maximum and minimum, respectively, at ๐œƒ =โ€‰0 for all ฯ >โ€‰1, we can directly state that

$$\max \limits_{\theta \in[0, 2\pi]}\left|K_{n}(z)\right|=\left|K_{n}(0)\right|=\left|K_{n}(\pi)\right|, \quad \rho>1. $$

On the other hand, the length of the ellipse can be estimated by (cf. [28])

$$ l(\mathcal{E}_{\rho})\leqslant 2\pi a_{1}\left( 1-\frac{1}{4}a_{1}^{-2}-\frac{3}{64}a_{1}^{-4}-\frac{5}{256}a_{1}^{-6}\right), $$

and thus, (17) yields the bound (19). โ–ก

2.2 Error bounds based on an expansion of the remainder

If f is an analytic function in the interior of \(\mathcal {E}_{\rho }\), for some ฯ >โ€‰1, it admits the expansion

$$ f(z)=\sum\limits_{k=0}^{\infty}{\prime}\alpha_{k} T_{k}(z), $$
(21)

where ฮฑk are given by

$$ \alpha_{k}=\frac{1}{\pi}{\int}_{-1}^{1}(1-t^{2})^{-1/2}f(t)T_{k}(t)dt. $$

The prime in the corresponding sum denotes that the first term is taken with the factor 1/2. The series converges for each z in the interior of \(\mathcal {E}_{\rho }\). In general, the Chebyshev-Fourier coefficients ฮฑk in the expansion are unknown; however, Elliott [6] described a number of ways to estimate or bound them. In particular, under our assumptions the following upper bound will be useful,

$$ |\alpha_{k}|\leq \frac{2}{\rho^{k}} \|f\|_{\mathcal{E}_{\rho}} . $$
(22)

The following result provides the desired bound for the error of quadrature.

Theorem 2

The following bound for the error of the Gaussโ€“Kronrod quadrature formula based on the expansion of the remainder is obtained for \(d\widehat {\sigma }_{n}^{[1]} \):

$$ r_{2}(f)=\inf_{1<\rho<\rho_{\max}} \bigg[ \frac{\pi}{4^{n-1}}\cdot \frac{2\rho^{2n}+1}{2\rho^{2n}\left( \rho^{4n}-1\right)}\cdot \|f\|_{\mathcal{E}_{\rho}}\bigg]. $$
(23)

For the proof of this theorem, we need a result by D. B. Hunter [14, Lemma 5], which is included below, to make the paper selfโ€“contained.

Lemma 2

With ฮพ and z as in (12), we have:

$$ {\int}_{\mathcal{E}_{\rho}} \xi^{-k} T_{j}(z) dz = \begin{cases} i\pi, & j=0, k=1 ,\\ i\pi/2, & j>0, k=j+1 ,\\ -i\pi/2, & j>1, k=j-1 ,\\ 0, & \text{otherwise} . \end{cases} $$

Proof Proof of Theorem 2

In the current case, the kernel is given by \(K_{n}(z)= \frac {{\varrho }_{n}(z)}{\widehat \pi _{n}(z) \pi ^{*}_{n+1}(z) }, z\notin [-1,1]\), where we have (see (15))

$${\varrho}_{n}(z)=-\frac{2\pi}{4^{2n}}\left( 2\xi^{-2n}+\xi^{-4n}\right),$$

and (see (16))

$$ \begin{array}{@{}rcl@{}} \frac{1}{\widehat\pi_{n}(z) \pi^{*}_{n+1}(z)}&=& {2\cdot 4^{n}}\left( \xi^{2n}-\xi^{-2n}\right)^{-1} \left( \xi-\xi^{-1}\right)^{-1}\\ &=& 2\cdot 4^{n} \xi^{-2n-1} \frac{1}{1-\xi^{-4n}} \frac{1}{1-\xi^{-2}}\\ &=& 2\cdot 4^{n} \xi^{-2n-1} \sum\limits_{p=0}^{\infty}\xi^{-4np}\sum\limits_{q=0}^{\infty}\xi^{-2q}\\ &=&2\cdot 4^{n} \sum\limits_{p=0}^{\infty}\sum\limits_{q=0}^{\infty}\xi^{-4np-2q-2n-1}. \end{array} $$

Therefore,

$$ \begin{array}{@{}rcl@{}} K_{n}(z)&=& -\frac{2\pi}{4^{2n}}2\cdot 4^{n} \left( 2\xi^{-2n}+\xi^{-4n}\right)\sum\limits_{p=0}^{\infty}\sum\limits_{q=0}^{\infty}\xi^{-4np-2q-2n-1}\\ &=& -\frac{\pi}{4^{n-1}}\bigg[ 2\sum\limits_{p=0}^{\infty}\sum\limits_{q=0}^{\infty}\xi^{-4np-2q-4n-1} + \sum\limits_{p=0}^{\infty}\sum\limits_{q=0}^{\infty}\xi^{-4np-2q-6n-1} \bigg]\\ &=& -\frac{\pi}{4^{n-1}}\big[ 2\xi^{-4n-1}+2\xi^{-4n-3} +2\xi^{-4n-5} +{\dots} \\ &&+\ \xi^{-6n-1}+\xi^{-6n-3}+\xi^{-6n-5}+{\dots} \\ &&+\ 2\xi^{-8n-1}+2\xi^{-8n-3} +2\xi^{-8n-5}+{\dots} \\ &&+\ \xi^{-10n-1}+\xi^{-10n-3}+\xi^{-10n-5} +{\dots} \big]. \end{array} $$

This way, the following shorter expression for Kn may be written,

$$ K_{n}(z)= -\frac{\pi}{4^{n-1}}\sum\limits_{k=0}^{\infty} \omega_{n,k} \xi^{-4n-k-1}, $$
(24)

where we have

$$ \begin{array}{@{}rcl@{}} \omega_{n,4kn}&=&\omega_{n,4kn+2}=\dots=\omega_{n,(4k+2)n-2}=3k+2,\\ \omega_{n,(4k+2)n}&=&\omega_{n,(4k+2)n+2}=\dots=\omega_{n,(4k+4)n-2}=3k+3,\\ \omega_{n,k}&=&0 \ \text{for \ all \ other}\ k \in \mathbb{N}. \end{array} $$

The remainder term Rn(f) can be represented in the form

$$ R_{n}(f)=\frac{1}{2^{2n-2}}\sum\limits_{k=0}^{\infty}\alpha_{4n+k}\epsilon_{n,k}, $$
(25)

where the coefficients ๐œ–n,k are independent on f. Namely, using (21) and (24) in (9) we obtain

$$ \begin{array}{@{}rcl@{}} R_{n}(f)&=&\frac{1}{2^{2n-2}}\frac{1}{2\pi i}{\int}_{\mathcal{E}_{\rho}}\left( \sum\limits_{k=0}^{\infty}{\prime}\alpha_{k} T_{k}(z)\sum\limits_{k=0}^{+\infty}\omega_{n,k}\xi^{-4n-k-1}\right)dz \\ &=&\frac{1}{2^{2n-2}}\sum\limits_{k=0}^{+\infty}\left( \frac{1}{2\pi i}\sum\limits_{j=0}^{+\infty}{\prime}\alpha_{j} \right. \left.{\int}_{\mathcal{E}_{\rho}}T_{j}(z)\xi^{-4n-k-1}dz\right)\omega_{n,k}. \end{array} $$

Applying Lemma 2, this reduces to (25) with

$$ \begin{array}{@{}rcl@{}} \epsilon_{n,4kn}&=&\frac{1}{2} \ \text{for \ k}\in \mathbb{N}_{0}, \\ \epsilon_{n,(4k+2)n}&=&\frac{1}{4} \ \text{for\ k}\in \mathbb{N}_{0}, \\ \epsilon_{n,l}&=&0 \ \text{for \ all \ other}\ l \in \mathbb{N}. \end{array} $$

Now we easily reach the following expression for the error of quadrature

$$ R_{n}(f)= -\frac{1}{4^{n-1}}\bigg(\frac{1}{2} \cdot \sum\limits_{k=0}^{\infty} \alpha_{4nk+4n} + \frac{1}{4} \cdot \sum\limits_{k=0}^{\infty} \alpha_{4kn+6n} \bigg). $$

Then, inequality (22) yields

$$ \begin{array}{@{}rcl@{}} | R_{n}(f)|&\leq& \frac{\pi}{4^{n-1}}\cdot \|f\|_{\mathcal{E}_{\rho}}\cdot \bigg(\sum\limits_{k=0}^{\infty} \frac{1}{\rho^{4nk+4n}} + \frac{1}{2} \sum\limits_{k=0}^{\infty} \frac{1}{\rho^{4kn+6n}} \bigg). \\ &=& \frac{\pi}{4^{n-1}}\cdot \|f\|_{\mathcal{E}_{\rho}}\cdot\bigg(\frac{1}{\rho^{4n}}+\frac{1}{2\rho^{6n}} \bigg)\sum\limits_{k=0}^{\infty} \frac{1}{\rho^{4nk}} \\ &=& \frac{\pi}{4^{n-1}}\cdot \|f\|_{\mathcal{E}_{\rho}}\cdot \bigg(\frac{1}{\rho^{4n}}+\frac{1}{2\rho^{6n}} \bigg) \bigg(\frac{1}{1-\rho^{-4n}} \bigg)\\ &=& \frac{\pi}{4^{n-1}}\cdot \frac{2\rho^{2n}+1}{2\rho^{2n}\left( \rho^{4n}-1\right)}\cdot \|f\|_{\mathcal{E}_{\rho}}. \end{array} $$

Finally, the bound (23) is easily attained. โ–ก

2.3 L 1โ€“bound for the error

From the integral expression (9), the error of quadrature may be bounded in the form

$$ R_{n}(f)\leq r_{3}(f)=\inf_{1<\rho<\rho_{\max}} \big[ L^{[1]}(\mathcal{E}_{\rho}) \cdot \|f\|_{\mathcal{E}_{\rho}} \big], $$
(26)

where

$$ L^{[1]}(\mathcal{E}_{\rho})=\frac{1}{2\pi} {\int}_{\mathcal{E}_{\rho}}\left|K_{n}(z)\right||dz|. $$

Now, we shall prove the following result.

Theorem 3

In the case of the measure \(d\widehat {\sigma }_{n}^{[1]}\), the error bound (26) takes the form

$$ r_{3}(f)=\inf_{1<\rho<\rho_{\max}} \left( \frac{\pi}{2^{2n-1}\rho^{2n}} \sqrt{\frac{4\rho^{4n}+5}{\rho^{8n}-1}} \cdot \|f\|_{\mathcal{E}_{\rho}} \right). $$
(27)

Proof

From (20), the modulus of the kernel is given by

$$ |K_{n}(z)|= \frac{\pi\sqrt{4\rho^{4n}+4\rho^{2n}\cos2n\theta+1}}{2^{2n-1}\cdot\rho^{4n}\sqrt{(a_{2}-\cos 2\theta)(a_{4n}-\cos 4n\theta)}}. $$

It is easy to check that \(|dz|=(1/\sqrt {2})\cdot \sqrt {a_{2}-\cos \limits {2\theta }} d\theta \) (cf. [14]), which yields

$$ \begin{array}{@{}rcl@{}} L^{[1]}(\mathcal{E}_{\rho})&=&\frac{1}{\rho^{4n} 2^{2n}\sqrt{2}}{{\int}_{0}^{2\pi} \sqrt { \frac{4\rho^{4n}+4\rho^{2n}\cos2n\theta+1}{a_{4n}-\cos 4n\theta} } } d\theta\\ &=&\frac{1}{\rho^{4n} 2^{2n-1}\sqrt{2}}{{\int}_{0}^{\pi} \sqrt { \frac{4\rho^{4n}+4\rho^{2n}\cos2n\theta+1}{a_{4n}-\cos 4n\theta} } } d\theta. \end{array} $$

Applying the Cauchy-Schwarz inequality in L2 to the last expression, we obtain

$$ \begin{array}{rl} L^{[1]}(\mathcal{E}_{\rho})\leq&\frac{\sqrt{{\int}_{0}^{\pi} d\theta} }{\rho^{4n}2^{2n-1} \cdot \sqrt{2} }\sqrt{(4\rho^{4n}+1)\cdot I_{0}+4\rho^{2n} \cdot I_{1}}, \end{array} $$

where, using from [13, 3.613] that

$$ \begin{array}{rl} &{\int}_{0}^{\pi}\frac{\cos{mx} dx}{a^{2}-2a\cos{x}+1}=\frac{\pi}{a^{m}(a^{2}-1)}, \end{array} $$

we obtain the explicit expressions for the integrals

$$ \begin{array}{@{}rcl@{}} I_{0}&=&{\int}_{0}^{\pi}\frac{d \theta}{a_{4n}-\cos{4n\theta}}={\int}_{0}^{\pi}\frac{d \theta}{\frac{1}{2}\big(\rho^{4n}+\frac{1}{\rho^{4n}} \big)-\cos{4n\theta}}=\frac{2\pi\rho^{4n}}{\rho^{8n}-1},\\ I_{1}&=&{\int}_{0}^{\pi}\frac{\cos{2n\theta} d\theta}{a_{4n}-\cos{4n\theta}}={\int}_{0}^{\pi}\frac{\cos{2n\theta} d\theta}{\frac{1}{2}\big(\rho^{4n}+\frac{1}{\rho^{4n}} \big)-\cos{4n\theta}}=\frac{2\pi\rho^{2n}}{\rho^{8n}-1}. \end{array} $$

Then, we get

$$ \begin{array}{@{}rcl@{}} L^{[1]}(\mathcal{E}_{\rho})&\leq&\frac{\sqrt{\pi}}{\rho^{2n}\sqrt{2}\cdot 2^{2n-1}}\sqrt{ (4\rho^{4n}+1) \frac{2\pi}{\rho^{8n}-1} +4 \cdot \frac{2\pi}{\rho^{8n}-1} }\\ &=& \frac{\pi}{2^{2n-1}\rho^{2n}} \sqrt{\frac{4\rho^{4n}+5}{\rho^{8n}-1}}. \end{array} $$

Finally, the bound (27) easily follows. โ–ก

3 Numerical results

Throughout this section, several numerical experiments are displayed to illustrate the results in previous section. In this sense, the obtained error bounds r1(f),r2(f) and r3(f) have been tested for the three following characteristic examples (commonly used in the literature on numerical integration):

$$f_{0}(z)=e^{\omega z^{2}}, \ \omega>0; f_{1}(z)=e^{\cos (\omega z)},\ \omega>0; f_{2}(z)\!=\frac{e^{e^{z}}} {(a+z)^{k}(b+z)^{l}(c+z)^{m}},$$

where a < โˆ’โ€‰1, c โ‰ค b โ‰ค a and \(k\in \mathbb {N} \), \(l,\ m \in \mathbb {N}_{0} \), and it is easy to check that the following properties are satisfied,

$$ \max_{z\in \mathcal{E}_{\rho}}|f_{0}(z)|=e^{\omega {a_{1}^{2}}}, \quad \max_{z\in \mathcal{E}_{\rho}}|f_{1}(z)|=e^{\cosh{\omega b_{1}} }, $$
$$ \max_{z\in \mathcal{E}_{\rho}}|f_{2}(z)|=\frac{e^{e^{a_{1}}}} {|a+a_{1}|^{k}|b+a_{1}|^{l}|c+a_{1}|^{m}}, $$

with \(a_{1}=\frac {\rho +\rho ^{-1}}{2}\) and \(b_{1}=\frac {\rho -\rho ^{-1}}{2}\).

It is clear that the functions f0(z) and f1(z) are entire, so \(\rho _{\max \limits }=\infty \) in both cases. Otherwise, for f2(z) the condition a < โˆ’โ€‰1, c โ‰ค b โ‰ค a means that the function f is analytic inside the elliptical contour \(\mathcal {E}_{\rho _{\max \limits }}\), for a certain \(\rho _{\max \limits }>1\), where \(|a|=\frac {1}{2}(\rho _{\max \limits }+\rho _{\max \limits }^{-1})\). We use some values of parameters a,b,c which have been used in literature (see, e. g., [30]); in particular, a = โˆ’โ€‰1.408333333333333, b = โˆ’โ€‰1.892857142857143, c = โˆ’โ€‰2.408695652173913, k =โ€‰1, l =โ€‰5, m =โ€‰10, which means that \(\rho _{\max \limits }=2.4\).

In order to compute the actual (sharp) error bound for the quadrature formula

$$ {\int}_{-1}^{1}f(t)d\widehat{\sigma}_{n}^{[1]}(t)\approx\sum\limits_{\nu=1}^{n} W_{\nu} f(\tau_{\nu})+\sum\limits_{\mu=1}^{n+1} W_{\mu}^{*} f(\tau_{\mu}^{*}), $$
(28)

we use [21, Theorem 4.1], which provides explicit formulas for all coefficients \(W_{\nu }, W_{\mu }^{*} \) and nodes \(\tau _{\nu }, \ \tau _{\mu }^{*},\ \nu =1, \dots , n,\ \mu =1, \dots , n+1.\) To proceed analogously with the other Chebyshev measures \(d\sigma _{n}^{[i]}\), i =โ€‰2,3,4, it is possible to use the numerically stable and effective methods [4, 17] (see also [9] along with [10]).

First of all and though it is a well-known fact that the Gaussโ€“Kronrod quadrature formula is a refinement of the classical Gauss rule (up to the extent that the value given by the former is commonly used to estimate the error of the latter), let us previously include a small table (Tableย 1) comparing the actual estimations of the error of both quadrature rules for some examples corresponding to the integrand f0. We denote by โ€œError GFโ€ and โ€œError GKFโ€ the errors of the classical Gauss formula and the Gaussโ€“Kronrod rule, respectively. The numerical examples displayed below clearly show that the number of precision digits of the latter is approximately the double of that of the former. The results corresponding to the Gauss rule are taken from [25, Table 4.3].

Table 1 The values of the actual error of Gauss and Gaussโ€“Kronrod rules for the integrand f0 and some values of n and ฯ‰ in the case of \(d\widehat {\sigma }_{n}^{[1]}\)

Now, we are concerned with showing the sharpness of the error estimations (19), (23), and (27). The results are displayed in Tablesย 2,ย 3, andย 4, where โ€œErrorโ€ means the actual (sharp) error and Iฯƒ(f) represents the exact value of the integral \({\int \limits }_{-1}^{1} f(t) d\widehat \sigma _{n}(t)\).

Table 2 The values of the derived bounds r1(f0),r2(f0),r3(f0), compared with the actual error for some values of n, ฯ‰
Table 3 The values of the derived bounds r1(f1),r2(f1),r3(f1), compared with the actual error for some values of n, ฯ‰
Table 4 The values of the derived bounds r1(f2),r2(f2),r3(f2), compared with the actual error for some values of n

Tablesย 2โ€“4 above show how sharp the bounds of the quadrature error obtained in Sectionย 2 are; namely, the average deviation from the actual value of the error does not exceed one precision digit. At the same time, the high accuracy of the Gaussโ€“Kronrod rules, especially in the case of the entire integrands f0 and f1, is clearly shown.