Skip to main content
Log in

Filtered interpolation for solving Prandtl’s integro-differential equations

  • Original Paper
  • Published:
Numerical Algorithms Aims and scope Submit manuscript

Abstract

In order to solve Prandtl-type equations we propose a collocation-quadrature method based on de la Vallée Poussin (briefly VP) filtered interpolation at Chebyshev nodes. Uniform convergence and stability are proved in a couple of Hölder-Zygmund spaces of locally continuous functions. With respect to classical methods based on Lagrange interpolation at the same collocation nodes, we succeed in reproducing the optimal convergence rates of the L2 case and cut off the typical log factor which seemed inevitable dealing with uniform norms. Such an improvement does not require a greater computational effort. In particular, we propose a fast algorithm based on the solution of a simple 2-bandwidth linear system and prove that, as its dimension tends to infinity, the sequence of the condition numbers (in any natural matrix norm) tends to a finite limit.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

References

  1. Atkinson, K.E.: The Numerical Solution of Integral Equations of the Second Kind. Cambridge University Press, Cambridge (1997)

    Book  Google Scholar 

  2. Calió, F., Marchetti, E.: On an algorithm for the solution of generalized Prandtl equations. Numer. Algor. 28, 3–10 (2001)

    Article  MathSciNet  Google Scholar 

  3. Capobianco, M.R., Criscuolo, G., Junghanns, P.: A fast algorithm for Prandtl’s integro-differential equation. J. Comp. Appl. Math. 77, 103–128 (1997)

    Article  MathSciNet  Google Scholar 

  4. Capobianco, M.R., Criscuolo, G., Junghanns, P., Luther, U.: Uniform convergence of the collocation method for Prandtl’s integro-differential equation. ANZIAM J. 42, 151–168 (2000)

    Article  MathSciNet  Google Scholar 

  5. Capobianco, M.R., Themistoclakis, W.: Interpolating polynomial wavelets on [− 1, 1]. Adv. Comput. Math. 23(4), 353–374 (2005)

    Article  MathSciNet  Google Scholar 

  6. De Bonis, M.C.: Remarks on two integral operators and numerical methods for CSIE. J. Comput. Appl. Math. 260, 117–134 (2014)

    Article  MathSciNet  Google Scholar 

  7. De Bonis, M.C., Mastroianni, G.: Mapping properties of some singular operators in Besov type subspaces of C(− 1, 1). Integr. Equ. Oper. Theory 55, 387–413 (2006)

    Article  MathSciNet  Google Scholar 

  8. De Bonis, M.C., Occorsio, D.: Quadrature methods for integro-differential equations of Prandtl’s type in weighted uniform spaces. Appl. Math. Comput. 393, 125721 (2021)

    MATH  Google Scholar 

  9. Ditzian, Z., Totik, V.: Moduli of Smoothness. SCMG Springer–Verlag, New York (1987)

    Book  Google Scholar 

  10. Ditzian, Z., Totik, V.: Remarks on Besov spaces and best polynomial approximation. Proc. Amer. Math. Soc. 104(4), 1059–1066 (1988)

    Article  MathSciNet  Google Scholar 

  11. Dragos, L.: Integration of Prandtl’s equation with the aid of quadrature formulae of Gauss type. Quarterly of Applied Mathematics LII, 23–29 (1994)

  12. Dragos, L.: A collocation method for the integration of Prandtl’s equation. ZAMM-Z Angew Math Mech 74(7), 289–290 (1994)

    Article  MathSciNet  Google Scholar 

  13. Golub, G.H., Van Loan, C.F.: Matrix Computations, 2nd edn. The Johns Hopkins University Press, Baltimore (1989)

    MATH  Google Scholar 

  14. Junghanns, P., Luther, U.: Cauchy singular integral equations in spaces of continuous functions and methods for their numerical solution. J. Comp. Appl. Math. 77, 201–237 (1997)

    Article  MathSciNet  Google Scholar 

  15. Kress, R.: Linear Integral Equations, Vol. 82 of Applied. Mathematical Sciences. Springer, Berlin (1989)

    Book  Google Scholar 

  16. Koerniawan, B.: Numerical solution of Prandtl’s lifting. Thesis of degree in Master of Science, University of Adelaide (1992)

  17. Lifanov, I.K., Poltavskii, L.N., Vainikko, G.M: Hypersingular Integral Equations and their Applications. Chapman & Hall CRC (2003)

  18. Mastroianni, G., Russo, M.G., Themistoclakis, W.: The boundedness of the Cauchy singular integral operator in weighted Besov type spaces with uniform norms. Integr. Equ. Oper. Theory 42(1), 57–89 (2002)

    Article  MathSciNet  Google Scholar 

  19. Mastroianni, G., Russo, M.G., Themistoclakis, W.: Numerical methods for Cauchy singular integral equations in spaces of weighted continuous functions, Recent advances in operator theory and its applications. Oper. Theory Adv. Appl., Birkhauser, Basel 160, 311–336 (2005)

    MATH  Google Scholar 

  20. Mastroianni, G., Themistoclakis, W.: A numerical method for the generalized airfoil equation based on the de la Vallée Poussin interpolation. J. Comput. Appl. Math. 180, 71–105 (2005)

    Article  MathSciNet  Google Scholar 

  21. Mhaskar, H.N., Pai, D.V: Fundamentals of Approximation Theory. CRC Press, Boca Raton (2000)

    MATH  Google Scholar 

  22. Mkhitaryan, S.M., Mkrtchyan, M.S., Kanetsyan, E.G.: On a method for solving Prandtl’s integro-differential equation applied to problems of continuum mechanics using polynomial approximations. Z. Angew. Math. Mech. 97(6), 639–654 (2017)

    Article  MathSciNet  Google Scholar 

  23. Occorsio, D., Themistoclakis, W.: Uniform weighted approximation by multivariate filtered polynomials, in the Vol. Numerical Computations: Theory and Algorithms Chapter 9, Lecture Notes in Computer Science series, p. 11973 (2020), Chapter https://doi.org/10.1007/978-3-030-39081-5_9

  24. Occorsio, D., Themistoclakis, W.: Uniform weighted approximation on the square by polynomial interpolation at Chebyshev nodes, Applied Mathematics and Computation, 385 (2020) https://doi.org/10.1016/j.amc.2020.125457

  25. Occorsio, D., Themistoclakis, W.: On the filtered polynomial interpolation at Chebyshev nodes, arXiv:2008.00240 [math.NA]

  26. Occorsio, D., Themistoclakis, W.: Some remarks on filtered polynomial interpolation at Chebyshev nodes, To appear on Dolomites Research Notes on Approximation. ArXiv:2101.04551 [math.NA]

  27. Paul, S., Panja, M.M., Mandal, B.N.: Wavelet Based numerical solution of second kind hypersingular integral equation. Appl. Math. Sci. 10(54), 2687–2707 (2016)

    Google Scholar 

  28. Prössdorf, S., Silbermann, B.: Numerical Analysis for Integral and Related Operator Equations. Akademie, Berlin (1991)

    MATH  Google Scholar 

  29. Themistoclakis, W.: Weighted L1 approximation on [− 1, 1] via discrete de la Vallée Poussin means. Math. Comput. Simul. 147, 279–292 (2018)

    Article  Google Scholar 

  30. Themistoclakis, W.: Uniform approximation on [− 1, 1] via discrete de la Vallée Poussin means. Numer. Algorithms 60(4), 593–612 (2012)

    Article  MathSciNet  Google Scholar 

  31. Themistoclakis, W.: Some interpolating operators of de la Vallée-Poussin type. Acta Math. Hungar. 84(3), 221–235 (1999)

    Article  MathSciNet  Google Scholar 

  32. Themistoclakis, W.: Some error bounds for Gauss–Jacobi quadrature rules. Appl. Numer. Math. 116, 286–293 (2017)

    Article  MathSciNet  Google Scholar 

  33. Themistoclakis, W., Van Barel, M.: Generalized de la Vallée Poussin approximations on [− 1, 1]. Numer. Algorithms 75, 1–31 (2017)

    Article  MathSciNet  Google Scholar 

  34. Tvardovskii, V.V.: A pseudomacrocrack in an anisotropic body. J. Appl. Maths Mechs 55(4), 549–554 (1991)

    Article  MathSciNet  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to W. Themistoclakis.

Additional information

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix

Appendix

1.1 Proof of Theorem 4.2

Firstly note that, by (55) and (30), we have

$$ \tilde {H_{n}^{m}}f(y)={\int}_{-1}^{1} \mathcal{H} (x,y)f(x)\varphi(x)dx, \qquad \mathcal{H}(x,y):=\sum\limits_{k=1}^{n} h(x_{k},y){\Phi}_{n,k}^{m}(x). $$

Let us prove that the boundedness of \(\tilde {H_{n}^{m}}:C^{0}_{\varphi }\rightarrow Z_{s}(\varphi )\) follows from Proposition 2.3. Indeed, we recall that (cf. Theorem 3.2)

$$ \|{V_{n}^{m}}\|_{C^{0}_{\varphi}\rightarrow C^{0}_{\varphi}}:=\sup\limits_{\|f\varphi\|\le 1}\|({V_{n}^{m}}f)\varphi\|=\sup\limits_{|x|\le 1}\sum\limits_{k=1}^{n}\frac{|{\Phi}_{n,k}^{m}(x)|\varphi(x)}{\varphi(x_{k})}\le \mathcal{C}\ne\mathcal{C}(n,m). $$
(83)

Hence, from the assumptions on h, we deduce that

  1. (a)

    \({\mathscr{H}}(x,y)\varphi (x)\varphi (y)\) is a continuous function w.r.t. both the variables x, y ∈ [− 1, 1];

  2. (b)

    for all \(n\in {\mathbb N}\) and |x|≤ 1, taking into account that

    $$ E_{n}(\overline{h}_{x_{k}})_{\varphi}= \|(\overline{h}_{x_{k}}-P^{*}_{x_{k}})\varphi\|\le \frac {\mathcal{C}}{n^{s}}, \quad k=1,\ldots, n, \qquad {\mathcal{C}}\ne{\mathcal{C}} (n,k), $$

    we set

    $$ P^{*}(y):=\sum\limits_{k=1}^{n} P^{*}_{x_{k}}(y){\Phi}_{n,k}^{m}(x)\varphi(x),\qquad |y|\le 1. $$

    Hence, we have

    $$ \begin{array}{@{}rcl@{}} E_{n}(\overline{\mathcal{H}}_{x}\varphi(x))_{\varphi}&=&\inf\limits_{P\in{\mathbb P}_{n}} \sup\limits_{|y|\le 1} \left| \mathcal{H}(x,y)\varphi(x)-P(y)\right|\varphi(y)\\ &\le&\sup\limits_{|y|\le 1} \left| \mathcal{H}(x,y)\varphi(x)-P^{*}(y)\right|\varphi(y)\\ &=& \sup\limits_{|y|\le 1}\left|\sum\limits_{k=1}^{n} \left[ h(x_{k},y)-P^{*}_{x_{k}}(y)\right]{\Phi}_{n,k}^{m}(x)\varphi(x)\right|\varphi(y)\\ &\le& \sum\limits_{k=1}^{n}\frac{|{\Phi}_{n,k}^{m}(x)|\varphi(x)}{\varphi(x_{k})} \sup\limits_{|y|\le 1}\left|\overline{h}_{x_{k}}(y)-P^{*}_{x_{k}}(y)\right|\varphi(y)\\ &\le& \frac C{n^{s}}\sum\limits_{k=1}^{n}\frac{|{\Phi}_{n,k}^{m}(x)|\varphi(x)}{\varphi(x_{k})} \le \frac {\mathcal{C}}{n^{s}}, \qquad {\mathcal{C}}\ne {\mathcal{C}}(n,x). \end{array} $$

In conclusion, by virtue of (a) and (b), the kernel \({\mathscr{H}}\) satisfies the assumptions of Proposition 2.3 with v = φ, so that the operator defined by this kernel, namely \(\tilde {H_{n}^{m}}:C^{0}_{\varphi }\rightarrow Z_{s}(\varphi )\), is bounded.

Consequently, the map \({H_{n}^{m}}:C^{0}_{\varphi }\rightarrow Z_{s}(\varphi )\) is bounded too, since it is the following composition of bounded maps (cf. Theorem 3.2)

$$ H_{n}^{m}:C^{0}_{\varphi} \begin{array}{c} \tilde{H_{n}^{m}}\\ \longrightarrow \end{array} Z_{s}(\varphi) \begin{array}{c} {V_{n}^{m}}\\ \longrightarrow \end{array} Z_{s}(\varphi). $$

In order to prove (58), let us arbitrarily fix \(f\in C^{0}_{\varphi }\) and |y|≤ 1.

By using (30), we note that

$$ {H_{n}^{m}}f(y)=\sum\limits_{j=1}^{n}\tilde{H_{n}^{m}}f(x_{j}){\Phi}_{n,j}^{m}(y)= {\int}_{-1}^{1}\left( \sum\limits_{j=1}^{n}\sum\limits_{k=1}^{n} h(x_{k},x_{j}){\Phi}_{n,k}^{m}(x){\Phi}_{n,j}^{m}(y)\right)f(x)\varphi(x)dx, $$

that means \({H_{n}^{m}}f\) can be obtained by replacing h with its bivariate VP interpolation polynomial based on tensor-product Chebyshev nodes of the second kind (see [23, 24]).

Hence, we write

$$ \begin{array}{@{}rcl@{}} Hf(y){-}{H_{n}^{m}}f(y)\!&=&\!{\int}_{-1}^{1}\left( h(x,y){-}\!\sum\limits_{j=1}^{n}h(x,x_{j}){\Phi}_{n,j}^{m}(y)\right)f(x)\varphi(x) dx\\ &+&\!\!{\int}_{-1}^{1}\sum\limits_{j=1}^{n}{\Phi}_{n,j}^{m}(y)\left( \!h(x,x_{j}){-}\!\! \sum\limits_{k=1}^{n}\!h(x_{k},x_{j}){\Phi}_{n,k}^{m}(x)\!\right)f(x)\varphi(x)dx. \end{array} $$

Then, using (30), (26) and (cf. (36))

$$ \sum\limits_{k=1}^{n}{\Phi}_{n,k}^{m}(x)=1,\qquad \forall |x|\le 1, $$
(84)

we get

$$ \begin{array}{@{}rcl@{}} Hf(y)-{H_{n}^{m}}f(y)&=& {\int}_{-1}^{1}\bigg(\overline{h}_{x}(y)-{V_{n}^{m}} (\overline{h}_{x})(y)\bigg)f(x)\varphi(x) dx\\ &+&{\int}_{-1}^{1}\sum\limits_{j=1}^{n}{\Phi}_{n,j}^{m}(y)\sum\limits_{k=1}^{n}{\Phi}_{n,k}^{m}(x) \bigg(\overline{h}_{x}(x_{j})-\overline{h}_{x_{k}}(x_{j})\bigg)f(x)\varphi(x)dx. \end{array} $$

Consequently

$$ \begin{array}{@{}rcl@{}} &&|Hf(y)-{H_{n}^{m}}f(y)|\varphi(y)\le{\int}_{-1}^{1}\bigg|\overline{h}_{x}(y)-{V_{n}^{m}} (\overline{h}_{x})(y)\bigg| \varphi(y)|f(x)|\varphi(x) dx\\ &+&\sum\limits_{j=1}^{n} |{\Phi}_{n,j}^{m}(y)|\varphi(y){\int}_{-1}^{1}\sum\limits_{k=1}^{n}|{\Phi}_{n,k}^{m}(x)| \bigg|\overline{h}_{x}(x_{j})-\overline{h}_{x_{k}}(x_{j})\bigg| |f(x)|\varphi(x)dx\\ &=:&A+B. \end{array} $$

On the other hand, the hypothesis \(\overline {h}_{x}\in Z_{s}(\varphi )\) uniformly w.r.t. x ∈ [− 1, 1] implies that \(\forall n\in {\mathbb N}\) and ∀|x|≤ 1 we have

$$ E_{n}(\overline{h}_{x})_{\varphi}\le \frac {\mathcal{C}}{n^{s}},\qquad {\mathcal{C}}\ne {\mathcal{C}}(n,x), $$
(85)

and in particular, there exists \(P^{*}\in {\mathbb P}_{n}\) such that

$$ \sup\limits_{|y|\le 1}\bigg|\overline{h}_{x}(y)-P^{*}(y)\bigg|\varphi(y)\le \frac {\mathcal{C}}{n^{s}},\qquad {\mathcal{C}}\ne {\mathcal{C}}(n,x). $$
(86)

Hence, concerning A, by (48) and (85), recalling that m = ⌊𝜃n⌋ with a fixed 0 < 𝜃 < 1, we get

$$ A\le {\int}_{-1}^{1} E_{n-m}(\overline{h}_{x})_{\varphi} |f(x)|\varphi(x) dx\le \frac{\mathcal{C}}{(n-m)^{s}} {\int}_{-1}^{1} |f(x)|\varphi(x) dx\le \frac{\mathcal{C}}{n^{s}}\|f\varphi\|. $$

Moreover, regarding B, by (86) and (83), we have

$$ \begin{array}{@{}rcl@{}} B&\le&\|f\varphi\| \sum\limits_{j=1}^{n}\frac{|{\Phi}_{n,j}^{m}(y)|\varphi(y)}{\varphi(x_{j})} {\int}_{-1}^{1}\sum\limits_{k=1}^{n}\frac{|{\Phi}_{n,k}^{m}(x)|\varphi(x)}{\varphi(x_{k})} \bigg|(\overline{h}_{x}-P^{*})(x_{j}){-} (\overline{h}_{x_{k}}-P^{*})(x_{j})\bigg| \frac{\varphi(x_{j})}{\varphi(x)}dx\\ &\le&\frac {\mathcal{C}}{n^{s}} \|f\varphi\| \sum\limits_{j=1}^{n}\frac{|{\Phi}_{n,j}^{m}(y)|\varphi(y)}{\varphi(x_{j})} {\int}_{-1}^{1}\sum\limits_{k=1}^{n}\frac{|{\Phi}_{n,k}^{m}(x)|\varphi(x)}{\varphi(x_{k})} \frac{dx}{\varphi(x)}\\ &\le& \frac {\mathcal{C}}{n^{s}} \|f\varphi\|{\int}_{-1}^{1} \frac{dx}{\varphi(x)}\le\frac {\mathcal{C}}{n^{s}}\|f\varphi\|, \qquad\text{with} {\mathcal{C}}\ne{\mathcal{C}}(n,f,y). \end{array} $$

Finally, let us prove (57). Taking into account that \({H_{n}^{m}}f\in {\mathbb P}_{n+m-1}\) and recalling that \(H:C^{0}_{\varphi }\rightarrow Z_{s}(\varphi )\) is bounded (cf. Proposition 2.3), for any \(f\in C^{0}_{\varphi }\) we note that

$$ \begin{array}{@{}rcl@{}}\sup\limits_{k\ge n+m-1} (k+1)^{r} E_{k}(Hf-{H_{n}^{m}}f)_{\varphi}&=&\sup\limits_{k\ge n+m-1} (k+1)^{r} E_{k}(Hf)_{\varphi}\\ &\le& {\mathcal{C}} \sup\limits_{k\ge n+m-1} \frac{\|Hf\|_{Z_{s}(\varphi)}}{k^{s-r}}\le {\mathcal{C}}\frac{\|f\varphi\|}{n^{s-r}}. \end{array} $$

Moreover, by (58) we get

$$ \begin{array}{@{}rcl@{}} \sup\limits_{k< n+m-1} (k+1)^{r} E_{k}(Hf-{H_{n}^{m}}f)_{\varphi}&\le& \|(Hf-{H_{n}^{m}}f)\varphi\|\sup\limits_{k< n+m-1} (k+1)^{r} \\ &\le& {\mathcal{C}} \frac{\|f\varphi\|}{n^{s}}(n+m)^{r}\le {\mathcal{C}}\frac{\|f\varphi\|}{n^{s-r}}. \end{array} $$

Consequently, by (58) we have

$$ \|(H-{H_{n}^{m}})f \|_{Z_{r}(\varphi)}=\| (Hf-{H_{n}^{m}}f)\varphi \| + \sup\limits_{k\in{\mathbb N}} (k+1)^{r} E_{k}(Hf-{H_{n}^{m}}f)_{\varphi}\le {\mathcal{C}}\frac{\|f\varphi\|}{n^{s-r}} $$

and (57) follows.

1.2 Proof of Theorem 4.4

Note that we can write

$$ \begin{array}{@{}rcl@{}} f^{*}-\tilde {f_{n}^{m}}&=&(D+{U_{n}^{m}})^{-1}\left[(D+{U_{n}^{m}})f^{*}- {V_{n}^{m}}g \right]\\ &=&(D+{U_{n}^{m}})^{-1}\left[(g- {V_{n}^{m}}g)+ ({U_{n}^{m}}-U)f^{*}\right]. \end{array} $$

Hence, by the uniform boundedness of \((D+{U_{n}^{m}})^{-1}:Z_{r}(\varphi )\rightarrow Z_{r+1}(\varphi )\) (cf. Th. 4.3), for any 0 < rs we deduce

$$ \|f^{*}-\tilde {f_{n}^{m}}\|_{Z_{r+1}(\varphi)}\le {\mathcal{C}} \left[\|g- {V_{n}^{m}}g\|_{Z_{r}(\varphi)}+ \|({U_{n}^{m}}-U)f^{*}\|_{Z_{r}(\varphi)}\right] $$

and (62) follows from Th. 3.2, Th. 4.1 and Th. 4.2, taking into account that, by hypothesis, gZs(φ), fZs+ 1(φ) and \(\|g\|_{Z_{s}(\varphi )}\sim \|f^{*}\|_{Z_{s+1}(\varphi )}\).

Finally, (63) follows from (62) when \(r\rightarrow 0^{+}\).

1.3 Proof of Lemma 5.1

Let \(\displaystyle \tilde f(y)=\sum\limits_{j=0}^{n-1}\tilde f_{j} \tilde q_{j}(y)\). Recalling (41)–(42), we have \(\displaystyle {V_{n}^{m}}\tilde f(y)= \sum \limits_{r=0}^{n-1}q_{r}(y)c_{n,r}(\tilde f),\) where

$$ \begin{array}{@{}rcl@{}} c_{n,r}(\tilde f)&=&\sum\limits_{k=1}^{n}\lambda_{k} p_{r}(x_{k})\tilde f(x_{k}) \\ &=&\left( \sum\limits_{j=0}^{n-m}+ \sum\limits_{j=n-m+1}^{n-1}\right) \tilde f_{j} \sum\limits_{k=1}^{n}\lambda_{k} p_{r}(x_{k})\widetilde q_{j}(x_{k})\\ &=& \sum\limits_{j=0}^{n-m} \tilde f_{j} \sum\limits_{k=1}^{n}\lambda_{k} p_{r}(x_{k})\frac{p_{j}(x_{k})}{j+1}\\ &+& \sum\limits_{j=n-m+1}^{n -1} \tilde f_{j} \sum\limits_{k=1}^{n}\lambda_{k} p_{r}(x_{k})\left( \frac{m+n-j}{2m(j+1)}p_{j}(x_{k})- \frac{(j-n+m)}{2m(2n-j+1)}p_{2n-j}(x_{k})\right) . \end{array} $$

Consequently, by using [30]

$$ p_{2n-j}(x_{k})=- p_{j}(x_{k}), \qquad k=1,\ldots, n,\qquad n-m<j<n, $$
(87)

for any r = 0,…,n − 1, we get

$$ \begin{array}{@{}rcl@{}} c_{n,r}(\tilde f)&=&\sum\limits_{j=0}^{n-m}\tilde f_{j} < p_{r},p_{j}>\frac 1 {j+1}\\ &+& \sum\limits_{j=n-m+1}^{n -1}\tilde f_{j} < p_{r},p_{j}>\frac{m+n-j}{2m(j+1)}+ \sum\limits_{j=n-m+1}^{n -1}\tilde f_{j} <p_{r},p_{j}>\frac{(j-n+m)}{2m(2n-j+1)}. \end{array} $$

Hence, the orthogonality relation < pr,pj >= δr, j implies that

$${V_{n}^{m}}\tilde f(y)= \sum\limits_{r=0}^{n-m}q_{r}(y)\tilde f_{r}\left[\frac{1}{r+1}\right]+ \sum\limits_{r=n-m+1}^{n-1}q_{r}(y)\tilde f_{r}\left[\frac 1{2m} \left( \frac{m+n-r}{r+1}+\frac{r-n+m}{2n-r+1}\right)\right], $$

and the statement follows.

1.4 Proof of Lemma 5.2

Due to the assumption \(\displaystyle \tilde f(y)= \sum\limits_{\ell =0}^{n-1}\tilde f_{\ell }\ \tilde q_{\ell }(y)\), the core of the proof consists in stating that

$$ K \widetilde q_{\ell}(x_{k})= \alpha_{\ell} p_{\ell-2}(x_{k})+\beta_{\ell} p_{\ell}(x_{k})+\gamma_{\ell} p_{\ell+2}(x_{k}),\quad \ell=0,1,\dots,n -1. $$
(88)

This is an immediate consequence of (43), (22) and (23) in the case that = 0,…, (nm). In order to prove (88) also in the case (nm) < < n, we observe that in this case the previous equations yield

$$ \begin{array}{@{}rcl@{}} K \widetilde q_{\ell}(x_{k})&=& \frac{n+m-\ell}{2m(\ell+1)}K p_{\ell}(x_{k})-\frac{\ell-n+m}{2m(2n-\ell+1)}K p_{2 n-\ell}(x_{k})\\ &=& \frac{n+m-\ell}{8m(\ell+1)}\left[-\frac 1 \ell p_{\ell-2}(x_{k})+\left( \frac 1 \ell+\frac 1 {\ell+2}\right)p_{\ell}(x_{k})-\frac 1 {\ell+2}p_{\ell+2}(x_{k}) \right]\\ &-&\frac{\ell-n+m}{8m(2n-\ell+1)}\left[-\frac 1 {2n-\ell} p_{2n-\ell-2}(x_{k})+\left( \frac 1 {2n-\ell}+\frac 1 {2n-\ell+2}\right)p_{2n-\ell}(x_{k})\right. \\ &-& \left.\frac 1 {2n-\ell+2}p_{2n-\ell+2}(x_{k})\right], \end{array} $$

and using (87), we get

$$ \begin{array}{@{}rcl@{}} K \widetilde q_{\ell}(x_{k})= &-&\left[\frac{n+m-\ell}{8m\ell(\ell+1)}+\frac{\ell-n+m}{8m(2n-\ell+1)(2n-\ell+2)}\right] p_{\ell-2}(x_{k}) \\ &+&\left[\frac{n+m-\ell}{4m\ell(\ell+2)}+\frac{\ell-n+m}{4m(2n-\ell)(2n-\ell+2)}\right]p_{\ell}(x_{k})\\ & - &\left[ \frac{n+m-\ell}{8m(\ell+1)(\ell+2)}+\frac{\ell-n+m}{8m(2n-\ell+1)(2n-\ell)} \right]p_{\ell+2}(x_{k}) .\end{array} $$
(89)

By (89), the statement follows if (nm) < ≤ (n − 3) and the same holds if = (n − 2), being in this case p+ 2(xk) = pn(xk) = 0.

If = (n − 1) then (89) implies that

$$ \begin{array}{@{}rcl@{}} K \widetilde q_{n-1}(x_{k})= &-&\left[\frac{m+1}{8 m n (n-1)}+\frac{m-1}{8m(n+2)(n+3)}\right]p_{n-3}(x_{k}) \\ & +&\left[\frac{m+1}{4m(n-1)(n+1)}+\frac{m-1}{4m(n+1)(n+3)}\right]p_{n -1}(x_{k})\\ & - &\left[\frac{m+1}{8 m n (n+1)}+\frac{m-1}{8 m(n+2)(n+1)}\right]p_{n+1}(x_{k}), \end{array} $$

and taking into account that pn− 1(xk) = −pn+ 1(xk) (cf. (87)), we get

$$ \begin{array}{@{}rcl@{}} K \widetilde q_{n-1}(x_{k})= &-&\left[\frac{m+1}{8 m n (n-1)}+\frac{m-1}{8m(n+2)(n+3)}\right]p_{n-3}(x_{k}) \\ &+&\left[\frac{(m+1)(3n-1)}{8 m n (n-1)(n+1)}+\frac{(m-1)(3n+7)}{8m(n+1)(n+2)(n+3)}\right]p_{n-1}(x_{k}), \end{array} $$

which concludes the proof of (88).

Finally, by (88), we get the statement as follows

$$ \begin{array}{@{}rcl@{}} {K_{n}^{m}}\tilde f (y)&=&\sum\limits_{j=0}^{n-1}q_{j}(y)\left[\sum\limits_{k=1}^{n}\lambda_{k} p_{j}(x_{k})\sum\limits_{\ell=0}^{n-1} \tilde f_{\ell} K \widetilde q_{\ell}(x_{k})\right] \\ &=&\sum\limits_{j=0}^{n-1}q_{j}(y)\left[\sum\limits_{k=1}^{n}\lambda_{k} p_{j}(x_{k})\sum\limits_{\ell=0}^{n-1}\tilde f_{\ell} \bigg(\alpha_{\ell} p_{\ell-2}(x_{k})+\beta_{\ell} p_{\ell}(x_{k})+\gamma_{\ell} p_{\ell+2}(x_{k})\bigg)\right]\\ & =& \sum\limits_{j=0}^{n-1}q_{j}(y)\sum\limits_{\ell=0}^{n-1}\tilde f_{\ell} \bigg[\alpha_{\ell} <p_{\ell-2}, p_{j}>+\beta_{\ell}<p_{\ell}, p_{j}>+ \gamma_{\ell}<p_{\ell+2}, p_{j}>\bigg]\\ &=&\sum\limits_{j=0}^{n-1}q_{j}(y)\bigg[ \alpha_{j+2}\tilde f_{j+2}+\beta_{j}\tilde f_{j}+\gamma_{j-2}\tilde f_{j-2}\bigg]. \end{array} $$

1.5 Proof of Lemma 5.3

The statement can be deduced from (54)–(56) and (44), as follows

$$ \begin{array}{@{}rcl@{}} {H_{n}^{m}} \tilde f(y)&=&\sum\limits_{j=0}^{n-1}q_{j}(y)\left[\sum\limits_{k=1}^{n}\lambda_{k} p_{j}(x_{k})\widetilde H_{n} \tilde f(x_{k})\right] \\ &=&\frac 1 \pi\sum\limits_{j=0}^{n-1}q_{j}(y)\sum\limits_{k=1}^{n}\lambda_{k} p_{j}(x_{k})\left[\sum\limits_{i=0}^{n-1}\sum\limits_{s=1}^{n}\lambda_{s} p_{i}(x_{s})h(x_{s},x_{k}){\int}_{-1}^{1}\tilde f(x) q_{i}(x)\varphi(x)dx\right]\\ &=&\frac 1 \pi\sum\limits_{j=0}^{n-1}q_{j}(y)\left[\sum\limits_{i=0}^{n-1}\sum\limits_{k=1}^{n} \sum\limits_{s=1}^{n}\lambda_{k} \lambda_{s} p_{j}(x_{k})p_{i}(x_{s})h(x_{s},x_{k})\sum\limits_{\ell=0}^{n-1} \tilde f_{\ell}<\widetilde q_{\ell}, q_{i}>\right]\\ &=& \frac 1 \pi\sum\limits_{j=0}^{n-1}q_{j}(y)\left[\sum\limits_{i=0}^{n-1}\sum\limits_{k=1}^{n} \sum\limits_{s=1}^{n}\lambda_{k} \lambda_{s} p_{j}(x_{k})p_{i}(x_{s})h(x_{s},x_{k})\tilde f_{i}<\widetilde q_{i}, q_{i}>\right]. \end{array} $$

1.6 Proof of Theorem 5.4

Firstly, let us focus on the elements αk,γk,βk and wk that have been introduced by Lemmas 5.2 and 5.1, respectively. We note that they are decreasing in modulus and tend to 0 as \(n\to \infty \), |αi|,|γi|,βi with order 1/n2 and wi with order 1/n. Consequently, for any matrix norm ∥⋅∥ and for any ε > 0, there exists νε such that the matrix \(\mathcal {A}^{\prime }_{n}\) can be splitted into the following sum

$$ \mathcal{A}^{\prime}_{n}=\mathcal{R}_{n}+\mathcal{E}_{n}, \qquad \text{with}\ \|\mathcal{E}_{n}\|< \varepsilon, \quad\forall n>\nu_{\epsilon}. $$
(90)

More precisely, for any n > νε we represent the matrix \(\mathcal {A}^{\prime }_{n}\) in the following block form

$$ \mathcal{A}^{\prime}_{n} =: \begin{pmatrix}A_{1,1} & A_{1,2}\\ A_{2,1} & A_{2,2}\end{pmatrix},$$

where

$$ \begin{array}{@{}rcl@{}}A_{1,1}&:=&\begin{pmatrix} \delta_{0} & 0 & \alpha_{2} & & & & \\ 0 & \delta_{1} & 0 & \alpha_{3} & & \text{ 0} \\ \gamma_{0} & 0 & \delta_{2} & 0 & {\ddots} & \\ & {\ddots} & {\ddots} & {\ddots} & {\ddots} & \alpha_{\nu_{\varepsilon}} \\ &\text{ 0} & {\ddots} & {\ddots} & {\ddots} & 0\\ & & & \gamma_{\nu_{\varepsilon}-2} & 0 & \delta_{\nu_{\varepsilon}}\\ \end{pmatrix}\in {\mathbb R}^{(\nu_{\varepsilon}+1)\times (\nu_{\varepsilon}+1)} \\ [.15in] A_{1,2}&:=&\begin{pmatrix} 0 & 0 & {\ldots} & 0 & 0 \\ {\vdots} & 0 & 0 & 0 & 0 \\ 0 & {\vdots} & 0 & {\vdots} & \vdots\\ \alpha_{\nu_{\varepsilon}+1} & 0 & {\vdots} & 0 & 0 \\ 0 & \alpha_{\nu_{\varepsilon}+2} & 0 & {\ldots} & 0 \end{pmatrix} \in {\mathbb R}^{(\nu_{\varepsilon}+1)\times (n-\nu_{\varepsilon}-1)} \\ [.15in] A_{2,1}&:=& \begin{pmatrix} 0 & 0 & 0 & \gamma_{\nu_{\varepsilon}-1} & 0\\ 0 & {\vdots} & {\vdots} & 0 & \gamma_{\nu_{\varepsilon}} \\ {\vdots} & {\vdots} & 0 & {\ldots} & 0\\ 0 & 0 & 0 & {\vdots} & 0 & \\ 0 & 0 & {\ldots} & 0 & 0 \end{pmatrix} \in {\mathbb R}^{(n-\nu_{\varepsilon}-1)\times (\nu_{\varepsilon}+1) } \end{array} $$
$$ \begin{array}{@{}rcl@{}} \\ [.15in] A_{2,2}&:=&\begin{pmatrix} \delta_{\nu_{\varepsilon}+1} & 0 & \alpha_{\nu_{\varepsilon}+3} & & & & \\ 0 & \delta_{\nu_{\varepsilon}+2} & 0 & \alpha_{\nu_{\varepsilon}+4} & &\text{ 0} & \\ \gamma_{\nu_{\varepsilon}+1} & 0 & \delta_{\nu_{\varepsilon}+3} & 0 & & & \\ & & {\ddots} & {\ddots} & {\ddots} & & \\ & & {\ddots} & {\ddots} & {\ddots} & & \alpha_{n}\\ & \text{ 0} & & {\ddots} & {\ddots} & & 0 \\ & & & & \gamma_{n-2} & 0 & \delta_{n-1}\\ \end{pmatrix}\in {\mathbb R}^{(n-\nu_{\varepsilon}-1)\times (n-\nu_{\varepsilon}-1)} \end{array} $$

and we recall that δk = 1 + σwk + βk has been defined in (80).

Hence, for a suitable choice of ν𝜖 (depending on the choice of the matrix norm and 𝜖 > 0) we get (90) by taking

$$ \begin{array}{@{}rcl@{}} \mathcal{R}_{n}&:=&\begin{pmatrix}A_{1,1} & \mathbf{O}_{(\nu_{\varepsilon}+1)\times (n-\nu_{\varepsilon}-1)}\\ \mathbf{O}_{(n-\nu_{\varepsilon}-1)\times (\nu_{\varepsilon}+1)}& \mathcal{I}_{n-\nu_{\varepsilon}-1} \end{pmatrix}, \\ [.1in] \mathcal{E}_{n}&:=&\begin{pmatrix}\mathbf{O}_{(\nu_{\varepsilon}+1)\times (\nu_{\varepsilon}+1)} & A_{1,2}\\ A_{2,1} & A_{2,2}-\mathcal{I}_{n-\nu_{\varepsilon}-1} \end{pmatrix}, \end{array} $$

where Ok×h denotes the null matrix in \({\mathbb R}^{k\times h}\) and \(\mathcal {I}_{k}\) denotes the identity k × k matrix.

From (90), using \(\|\mathcal {E}_{n}\|<\varepsilon \), we easily deduce

$$ \|\mathcal{R}_{n}\|-\varepsilon\le \| \mathcal{A}^{\prime}_{n}\|\le \|\mathcal{R}_{n}\|+\varepsilon, \quad \forall \epsilon>0, \quad \forall n>\nu_{\epsilon}, $$
(91)

where we remark that \(\|\mathcal {R}_{n}\|\) is independent of n > νε, since increasing the order n of \(\mathcal {R}_{n}\), the order of the identity block increases, while A1,1 remains unchanged. Consequently, by (91) we get

$$\forall \varepsilon>0 \ \exists \nu_{\varepsilon}:\ \forall n_{1},n_{2}>\nu_{\varepsilon}, \ \left|\|\mathcal{A}^{\prime}_{n_{1}}\|-\|\mathcal{A}^{\prime}_{n_{2}}\|\right|< 2 \varepsilon,$$

i.e., the sequence \(\{\|\mathcal {A}^{\prime }_{n}\|\}_{n}\) converges since it is a Cauchy sequence.

In order to complete the proof, let us prove that also the sequence \(\{\|({\mathcal {A}^{\prime }}_{n})^{-1}\|\}_{n}\) is a Cauchy sequence.

Starting again from (90), we observe the matrix \(\mathcal {R}_{n}\) is nonsingular and, under the previous settings, we have

$$ \mathcal{R}_{n}^{-1}=\begin{pmatrix}A_{1,1}^{-1} & \mathbf{O}\\ \mathbf{O} & \mathcal{I} \end{pmatrix}, $$
(92)

where, for the sake of brevity, we omit to explicit the dimensions of the null and identity blocks.

Consequently, by (90) we get

$$\mathcal{A}^{\prime}_{n}=\mathcal{R}_{n}(\mathcal{I}_{n}+\mathcal{R}_{n}^{-1}\mathcal{E}_{n}).$$

Let us prove that ν𝜖 can be chosen in such a way that also \(\|\mathcal {R}_{n}^{-1}\mathcal {E}_{n}\|<\epsilon \) holds true for any n > ν𝜖. Indeed, we note that

$$ \mathcal{R}_{n}^{-1}\mathcal{E}_{n} =\begin{pmatrix} \mathbf{O} & A_{1,1}^{-1}A_{1,2}\\ A_{2,1} & A_{22}- \mathcal{I}\end{pmatrix} $$
(93)

differs from \(\mathcal {E}_{n}\) only for the block matrix in the position (1,2), which is \(A_{1,1}^{-1}A_{1,2}\) instead of A1,2. Nevertheless, in terms of the column vectors, denoted by \( \mathbf {e}_{i}\in {\mathbb R}^{\nu _{\varepsilon }+1}\) the ith vector of the canonical basis of \({\mathbb R}^{\nu _{\varepsilon }+1}\) and by \(\mathbf {0}\in {\mathbb R}^{\nu _{\epsilon } +1}\) the null vector, we note that

$$ \begin{array}{@{}rcl@{}} A_{1,2}&=&[\alpha_{\nu_{\varepsilon}+1} \mathbf{e}_{\nu_{\varepsilon}},\ \alpha_{\nu_{\varepsilon}+2} \mathbf{e}_{\nu_{\varepsilon}+1}, \ \mathbf{0}, \ldots, \mathbf{0}], \\ A_{1,1}^{-1}A_{1,2}&=&[\alpha_{\nu_{\varepsilon}+1} \mathbf{P}^{0}_{\nu_{\varepsilon}},\ \alpha_{\nu_{\varepsilon}+2} \mathbf{P}^{1}_{\nu_{\varepsilon}+1}, \ \mathbf{0}, \ldots, \mathbf{0}], \end{array} $$

where we set

$$\mathbf{P}^{0}:=A_{1,1}^{-1}\cdot \mathbf{e}_{\nu_{\varepsilon}}, \quad \mathbf{P}^{1}:=A_{1,1}^{-1}\cdot \mathbf{e}_{\nu_{\varepsilon}+1}.$$

In order to obtain an explicit expression of these vectors, we need only the last two columns of \(A_{1,1}^{-1}\), which we deduce from the following LU factorization of A1,1

$$A_{1,1}=LU=\begin{pmatrix} 1 & & & & & & \\ 0 & 1 & & & & \textrm{ 0} & \\ v_{0} & 0 & 1 & & & & \\ & v_{1} & 0 & 1 & & & \\ & & {\ddots} & {\ddots} & {\ddots} & & \\ & \textrm{ 0} & & v_{\nu_{\varepsilon}-2} & 0 & 1 & \end{pmatrix} \begin{pmatrix} d_{0} & 0 & \alpha_{2} & & & & \\ & d_{1} & 0 & \alpha_{3} & & \textrm{ 0} & \\ & & d_{2} & 0 & {\ddots} & & \\ & & & {\ddots} & {\ddots} & \alpha_{\nu_{\varepsilon}} \\ & & \textrm{ 0} & & d_{\nu_{\epsilon}-1}& 0 \\ & & & & & d_{\nu_{\varepsilon}} \end{pmatrix}, $$

where d0 = δ0, d1 = δ1, and for \(k=2,3,{\dots } \nu _{\varepsilon }\), we have

$$ \begin{array}{@{}rcl@{}}\ v_{k-2}&:=&\frac{\gamma_{k-2}}{d_{k-2}},\\ d_{k}&:=&\delta_{k}-v_{k-2}\alpha_{k}= 1+\sigma w_{k}+\beta_{k}-v_{k-2}\alpha_{k}.\end{array} $$

More precisely, it can be checked that the elements \(\mathbf {P}^{0}_{i}\) and \(\mathbf {P}^{1}_{i}\), i = 0,…,ν𝜖, of the previous vectors P0 and P1 can be determined by induction as follows

$$ \begin{array}{@{}rcl@{}} \mathbf{P}^0_{0}&=&\frac{1-\alpha_2 \mathbf{P}^0_2}{d_0},\\ \mathbf{P}^0_{2i+1}&=&0, \quad i=0,1,\ldots,\frac{\nu_\varepsilon}2-1,\\ \mathbf{P}^0_{2i}&=& \frac{(-1)^i{\prod}_{j=0}^{i-1}v_{2j}-\alpha_{2i+2}\mathbf{P}^0_{2i+2}}{d_{2i}} \quad i=\frac{\nu_\varepsilon}{2}-1, \frac{\nu_\varepsilon}{2}-2,\ldots,1,\\ \mathbf{P}^0_{\nu_\varepsilon}&=& \frac{(-1)^{\frac{\nu_\varepsilon}{2}}{\prod}_{j=0}^{\frac{\nu_\varepsilon}{2}-1}v_{2j}}{d_{\nu_\varepsilon}},\\ \mathbf{P}^1_{1}&=& \frac{1-\alpha_{3}\mathbf{P}^1_{3}}{d_{1}},\\ \mathbf{P}^1_{2i+1}&=& \frac{(-1)^{i}{\prod}_{j=0}^{i-1}v_{2j+1}-\alpha_{2i+3}\mathbf{P}^{1}_{2i+3}}{d_{2i+1}} \quad i=\frac{\nu_\varepsilon}{2}-3, \ldots,2,\\ \quad \mathbf{P}^1_{2i} & = & 0, \quad i=0,1,\ldots,\frac{\nu_\varepsilon}{2},\\ \mathbf{P}^{1}_{\nu_\varepsilon-1} & = & \frac{(-1)^{\frac{\nu_\varepsilon}{2}-1}{\prod}_{j=0}^{\frac{\nu_\varepsilon}{2}-2}v_{2j+1}}{d_{\nu_\varepsilon-1}}. \end{array} $$

Thus, recalling the behavior of the sequences αk,βk,γk,wk, we can conclude that, under a suitable choice of νε, all the entries of \(\mathcal {R}_{n}^{-1}\mathcal {E}_{n}\) are so small that we have

$$ \mathcal{A}^{\prime}_{n}=\mathcal{R}_{n}(\mathcal{I}_{n}+\mathcal{R}_{n}^{-1}\mathcal{E}_{n}), \qquad\text{with }\ \|\mathcal{R}_{n}^{-1}\mathcal{E}_{n}\|<\varepsilon, \qquad \forall n>\nu_{\epsilon} . $$

Consequently, taking 0 < 𝜖 < 1 and recalling that \(\|(\mathcal {I}_{n}+\mathcal {R}_{n}^{-1}\mathcal {E}_{n})^{-1}\|\le \frac 1 {1-\|\mathcal {R}_{n}^{-1}\mathcal {E}_{n}\|}\) (see, e.g., [13, Lemma 2.3.3, p. 59]) we have

$$ \|{(\mathcal{A}^{\prime}}_{n})^{-1}\|= \|(\mathcal{I}_{n}+\mathcal{R}_{n}^{-1}\mathcal{E}_{n})^{-1}\mathcal{R}_{n}^{-1}\| \le \frac{\|\mathcal{R}_{n}^{-1}\|}{1-\varepsilon},\qquad \forall n>\nu_{\varepsilon}, $$

as well as, we get

$$ \|\mathcal{R}_{n}^{-1}\|= \|(\mathcal{I}_{n}+\mathcal{R}_{n}^{-1}\mathcal{E}_{n})(\mathcal{A}^{\prime}_{n})^{-1}\|\le (1+\epsilon)\|(\mathcal{A}^{\prime}_{n})^{-1}\|,\qquad \forall n>\nu_{\varepsilon}. $$

Summing up, we have stated that

$$ \frac{\|\mathcal{R}_{n}^{-1}\|}{1+\varepsilon}\le \|{(\mathcal{A}^{\prime}}_{n})^{-1}\|\le \frac{\|\mathcal{R}_{n}^{-1}\|}{1-\varepsilon},\qquad \forall n>\nu_{\varepsilon}, \quad 0<\epsilon<1 . $$
(94)

On the other hand, in view of (92), we can say that \(\|\mathcal {R}_{n}^{-1}\|\) is independent of n > ν𝜖, like \(\|\mathcal {R}_{n}\|\), and in particular there exists a constant M independent of n s.t. \(\|\mathcal {R}_{n}^{-1}\|\le M\). Hence, by (94) we get for any n1,n2 > νε and 0 < ε < 1

$$ \left|\|({\mathcal{A}^{\prime}}_{n_{1}})^{-1}\|-\|({\mathcal{A}^{\prime}}_{n_{2}})^{-1}\|\right|\le \frac{2\varepsilon}{1-\varepsilon^{2}}\|\mathcal{R}_{n}^{-1}\|\le \frac{2M\varepsilon}{1-\varepsilon^{2}},$$

by which

$$\lim_{n_{1},n_{2}\to\infty}\left|\|({\mathcal{A}^{\prime}}_{n_{1}})^{-1}\| -\|({\mathcal{A}^{\prime}}_{n_{2}})^{-1}\|\right|=0,$$

and the statement follows.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

De Bonis, M.C., Occorsio, D. & Themistoclakis, W. Filtered interpolation for solving Prandtl’s integro-differential equations. Numer Algor 88, 679–709 (2021). https://doi.org/10.1007/s11075-020-01053-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11075-020-01053-x

Keywords

Mathematics Subject Classification (2010)

Navigation