This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Computation of highly oscillatory integrals
using a Fourier extension approximation

Akash Anand Akash Anand, Department of Mathematics and Statistics, Indian Institute of Technology, Kanpur, UP 208016 [email protected]  and  Damini Dhiman Damini Dhiman, Department of Mathematics and Statistics, Indian Institute of Technology, Kanpur, UP 208016 [email protected]
Abstract.

The numerical evaluation of integrals of the form

abf(x)eikg(x)𝑑x\displaystyle\int_{a}^{b}f(x)e^{ikg(x)}\,dx

is an important problem in scientific computing with significant applications in many branches of applied mathematics, science and engineering. The numerical approximation of such integrals using classical quadratures can be prohibitively expensive at high oscillation frequency (k1k\gg 1) as the number of quadrature points needed for achieving a reasonable accuracy must grow proportionally to kk. To address this significant computational challenge, starting with Filon in 1930, several specialized quadratures have been developed to compute such oscillatory integrals efficiently. A crucial element in such Filon-type quadrature is the accurate evaluation of certain moments which poses a significant challenge when non-linear phase functions gg are involved.

In this paper, we propose an equispaced-grid Filon-type quadrature for computing such highly oscillatory integrals that utilizes a Fourier extension of the slowly varying envelope ff. This strategy is primarily aimed at significantly simplifying the moment calculations, even when the phase function has stationary points. Moreover, the proposed approach can also handle certain integrable singularities in the integrand. We analyze the scheme to theoretically establish high-order convergence rates. We also include a wide variety of numerical experiments, including oscillatory integrals with algebraic and logarithmic singularities, to demonstrate the performance of the quadrature.

Key words and phrases:
oscillatory integrals, integrable singularities, Fourier extension, Filon quadrature

1. Introduction

The evaluation of highly oscillatory integrals plays a significant role in many branches of applied mathematics. The accuracy of their numerical approximation using a classical quadrature suffers significantly when the oscillation frequency is much more than the number of quadrature points. Since the initial attempt by Filon [1] to solve this difficulty, numerous approaches have been developed [2, 3, 4, 5, 6]. The Filon quadrature for approximating the integral of the form

(1) abf(x)eikg(x)𝑑x,\displaystyle\int_{a}^{b}f(x)e^{ikg(x)}\,dx,

relies on the interpolating polynomial pp satisfying p(xj)=f(xj)p(x_{j})=f(x_{j}) for interpolation nodes x0,x1,,xn,x_{0},x_{1},\ldots,x_{n}, to obtain the approximation

(2) abf(x)eikg(x)𝑑xabp(x)eikg(x)𝑑x.\displaystyle\int_{a}^{b}f(x)e^{ikg(x)}\,dx\approx\int_{a}^{b}p(x)e^{ikg(x)}\,dx.

Both the Filon method and its generalisation [2, 3, 4] where ff is approximated by splines, are shown to be efficient for suitably smooth functions provided the moments abxjeikg(x)𝑑x\int_{a}^{b}x^{j}e^{ikg(x)}\,dx can be explicitly calculated. While obtaining these moments is easy for linear oscillators, that is, g(x)=αx+β,α,βg(x)=\alpha x+\beta,\ \alpha,\beta\in\mathbb{R}, in general, the accurate moment calculations pose a significant challenge. For example, The Filon-Clenshaw-Curtis quadrature [7], that uses cos(jπ/N),j=0,1,,N\cos(j\pi/N),j=0,1,\dots,N as nodes and Chebyshev approximations for ff, obtain the underlying moments recursively using multiple recurrence relations. The Filon-Clenshaw-Curtis and its variations [8, 9, 10, 11] have been developed to efficiently handle integrable algebraic and logarithmic singularities in the integrand.

Several Filon-like approaches use the change of variable y=g(x)y=g(x) to transform the integral to the one with linear oscillations

(3) abf(x)eikg(x)𝑑x=g(a)g(b)1g(g1(y))f(g1(y))eiky𝑑y\displaystyle\int_{a}^{b}f(x)e^{ikg(x)}\,dx=\int_{g(a)}^{g(b)}\frac{1}{g^{\prime}(g^{-1}(y))}f(g^{-1}(y))e^{iky}\,dy

under the assumption that g(x)>0g^{\prime}(x)>0 for all x(a,b)x\in(a,b) or g(x)<0g^{\prime}(x)<0 for all x(a,b)x\in(a,b). Indeed, if g(x)g^{\prime}(x) has finitely many zeros in (a,b)(a,b), say xj,j=1,2,,Jx_{j},j=1,2,\ldots,J, then, by setting x0=ax_{0}=a and xJ+1=bx_{J+1}=b, we have

abf(x)eikg(x)𝑑x=j=0Jxjxj+1f(x)eikg(x)𝑑x\displaystyle\int_{a}^{b}f(x)e^{ikg(x)}\,dx=\sum_{j=0}^{J}\int_{x_{j}}^{x_{j+1}}f(x)e^{ikg(x)}\,dx

where, for each jj, the integral over [xj,xj+1][x_{j},x_{j+1}] can be transformed into the one with linear oscillations.

In this paper, we propose a Filon-type quadrature for oscillatory integrals with equispaced quadrature nodes. We show that this approach not only handles algebraic and logarithmic singularities in the integrand but also significantly simplifies the solution of the underlying moment problem. In fact, for several special singular weights, the proposed approach yields moment problems that can be solved analytically. Moreover, our approach requires only discrete functional data on a uniform grid. The fact that approximating the integral to high order does not require explicit knowledge of the derivatives is a significant strength of this approach.

The method to yield a more tractable moment problem relies on the interpolating trigonometric polynomial approximation of the smooth non-oscillatory part of the integrand. However, in general, such an approximation is inflicted with Gibb’s oscillations that manifest due to the discontinuous periodic extension of the function being approximated. To overcome this difficulty, we first construct a periodic extension of the function, which is then approximated by the interpolating trigonometric polynomial using FFT. We discuss the periodic extension employed in the paper in section 2.2. In section 2.4, we delve into the moment problem and discuss the solution of several important instances. In section 3, we present a convergence analysis. A variety of numerical experiments are presented in section 4 to demonstrate the performance of the proposed quadrature.

2. Methodology

Motivated by the need to compute the integrals such as those that appear on the right-hand side of the eq. 3, we consider the problem of approximation of the integrals

(4) Ik[a,b](w,f)=abw(x)f(x)eikx𝑑x,\displaystyle I^{[a,b]}_{k}(w,f)=\int_{a}^{b}w(x)f(x)e^{ikx}\,dx,

where wL1(a,b)w\in L^{1}(a,b) is the weight function and ff is a smooth function on [a,b][a,b].

2.1. The basic idea

For clarity, we introduce the main ideas assuming that the ff is known everywhere in the interval [a,b][a,b] and its derivatives are available at x=ax=a and x=bx=b. At the heart of this procedure is the step where we approximate ff by a trigonometric polynomial

=nn1c(f)e2πi(xa)/ξ\displaystyle\sum_{\ell=-n}^{n-1}c_{\ell}(f)e^{2\pi i\ell(x-a)/\xi}

of period ξba\xi\geq b-a. The primary motivation for doing so is to simplify the underlying moment calculations. Indeed, we see that

(5) Ik[a,b](w,f)abw(x)(=nn1c(f)e2πi(xa)/ξ)eikx𝑑x=eika=nn1c(f)Wk,[a,b](w)\displaystyle I^{[a,b]}_{k}(w,f)\approx\int_{a}^{b}w(x)\left(\sum_{\ell=-n}^{n-1}c_{\ell}(f)e^{2\pi i\ell(x-a)/\xi}\right)e^{ikx}\,dx=e^{ika}\sum_{\ell=-n}^{n-1}c_{\ell}(f)W^{[a,b]}_{k,\ell}(w)

where

(6) c(f)=1ξaa+ξf(x)e2πi(xa)/ξ𝑑xandWk,[a,b](w)=abw(x)ei(kξ+2π)(xa)/ξ𝑑x.\displaystyle c_{\ell}(f)=\frac{1}{\xi}\int_{a}^{a+\xi}f(x)e^{-2\pi i\ell(x-a)/\xi}\,dx\quad\text{and}\quad W^{[a,b]}_{k,\ell}(w)=\int_{a}^{b}w(x)e^{i(k\xi+2\pi\ell)(x-a)/\xi}\,dx.

For many important weights ww, the moments Wk,[a,b](w)W^{[a,b]}_{k,\ell}(w) can be obtained analytically as we see later in section 2.4.

The natural choice for the approximating trigonometric polynomial of ff is to use its truncated Fourier series. However, if we use the period ξ=ba\xi=b-a, it fails to converge uniformly if the (ba)(b-a)-periodic extension of ff is discontinuous, that is, if f(a)f(b)f(a)\neq f(b), due to Gibb’s phenomenon. Instead of directly approximating ff, we utilize the Fourier series of a periodic function with a larger period that coincides with ff on the interval [a,b][a,b] to eliminate Gibb’s oscillations. In the next sub-section, we discuss one such construction that we use for numerical experiments in this paper.

2.2. A periodic extension:

Among several possible periodic extensions for a given function, for simplicity of implementation and analysis, we utilize the construction based on the two-point Hermite polynomial [12, 13]. Given ff on [a,b][a,b], we introduce a periodic function fef^{e} of period ξ=2(ba)\xi=2(b-a) given by

(7) fe(x)={f(x),x[a,b],p(x),x[b,2ba),\displaystyle f^{e}(x)=\begin{dcases}f(x),&x\in[a,b],\\ p(x),&x\in[b,2b-a),\end{dcases}

where the polynomial pp is chosen as

(8) p(x)=m=0rf(m)(b)pm,r[b,2ba](x)+m=0rf(m)(a)pm,r[2ba,b](x)\displaystyle p(x)=\sum_{m=0}^{r}f^{(m)}(b)p_{m,r}^{[b,2b-a]}(x)+\sum_{m=0}^{r}f^{(m)}(a)p_{m,r}^{[2b-a,b]}(x)

with

pm,r[t1,t2](t)\displaystyle p_{m,r}^{[t_{1},t_{2}]}(t) =1m!(tt1)m(tt2t1t2)r+1n=0rm(r+nr)(tt1t2t1)n.\displaystyle=\frac{1}{m!}(t-t_{1})^{m}\left(\frac{t-t_{2}}{t_{1}-t_{2}}\right)^{r+1}\sum_{n=0}^{r-m}\binom{r+n}{r}\left(\frac{t-t_{1}}{t_{2}-t_{1}}\right)^{n}.

2.3. The Discrete Problem:

In many practical applications, only discrete functional data is available. In this regard, we consider the equispaced grid of size n+1n+1 on the interval [a,b][a,b] with grid points xj=a+(ba)j/nx_{j}=a+(b-a)j/n where the functional data fj,j=0,1,,nf_{j},j=0,1,\dots,n is assumed to be available. We introduce the grid function fr,qef_{r,q}^{e} given as

(9) (fr,qe)j={fj,j=0,,n,pr,q(a+(ba)j/n),j=n+1,,2n,\displaystyle(f_{r,q}^{e})_{j}=\begin{dcases}f_{j},&j=0,\dots,n,\\ p_{r,q}(a+(b-a)j/n),&j=n+1,\dots,2n,\\ \end{dcases}

where

(10) pr,q(x)=m=0r𝒟n,qm,(f)(xn)pm,r[b,2ba](x)+m=0r𝒟n,qm,+(f)(x0)pm,r[2ba,b](x),\displaystyle p_{r,q}(x)=\sum_{m=0}^{r}\mathcal{D}^{m,-}_{n,q}(f)(x_{n})p_{m,r}^{[b,2b-a]}(x)+\sum_{m=0}^{r}\mathcal{D}^{m,+}_{n,q}(f)(x_{0})p_{m,r}^{[2b-a,b]}(x),

for 1mr1\leq m\leq r, the operators 𝒟n,qm,+\mathcal{D}^{m,+}_{n,q} and 𝒟n,qm,\mathcal{D}^{m,-}_{n,q} respectively denote the forward and backward finite difference derivative operators for approximating f(m)f^{(m)} of the form

𝒟n,qm,±(f)(xj)=(±n)m(=0m+q1(aqm)fj±)\displaystyle\mathcal{D}^{m,\pm}_{n,q}(f)(x_{j})=(\pm n)^{m}\left(\sum_{\ell=0}^{m+q-1}(a_{q}^{m})_{\ell}f_{j\pm\ell}\right)

with appropriately chosen constants (aqm)(a_{q}^{m})_{\ell} so that the order of accuracy is qq. The discrete Fourier coefficients are obtained as

(11) d,n(fr,qe)=12nj=02n1(fr,qe)jeπij/n\displaystyle d_{\ell,n}(f_{r,q}^{e})=\frac{1}{2n}\sum_{j=0}^{2n-1}(f_{r,q}^{e})_{j}e^{-\pi i\ell j/n}

for =n,,n1\ell=-n,\dots,n-1. Finally, the integration scheme for approximation of Ik[a,b](w,f)I_{k}^{[a,b]}(w,f) is given by

(12) Ik[a,b](w,f)Ik,n[a,b](w,f):=eika=nn1d,n(fr,qe)Wk,[a,b](w).\displaystyle I_{k}^{[a,b]}(w,f)\approx I_{k,n}^{[a,b]}(w,f):=e^{ika}\sum_{\ell=-n}^{n-1}d_{\ell,n}(f_{r,q}^{e})W_{k,\ell}^{[a,b]}(w).

2.4. The moment problem:

The numerical integration scheme requires obtaining the moments given by

Wk,[a,b](w)\displaystyle W^{[a,b]}_{k,\ell}(w) =abw(x)eπi(xa)/(ba)eik(xa)𝑑x=(ba)01w(a+(ba)u)ei((ba)k+π)u𝑑u\displaystyle=\int_{a}^{b}w(x)e^{\pi i\ell(x-a)/(b-a)}e^{ik(x-a)}\,dx=(b-a)\int_{0}^{1}w(a+(b-a)u)e^{i((b-a)k+\pi\ell)u}\,du
=ba2ei((ba)k+π)/211w(a+b2+ba2v)ei((ba)k+π)v/2𝑑v\displaystyle=\frac{b-a}{2}e^{i((b-a)k+\pi\ell)/2}\int_{-1}^{1}w\left(\frac{a+b}{2}+\frac{b-a}{2}v\right)e^{i((b-a)k+\pi\ell)v/2}\,dv

The following are solutions to some instances of the moment problem corresponding to important classes of weights of interest.

Example 2.1.

For the simplest case where w(x)=1w(x)=1, we have

Wk,[a,b]\displaystyle W^{[a,b]}_{k,\ell} ={ba,=(ba)k/π,2(ba)(ba)k+πei((ba)k+π)/2sin(((ba)k+π)/2),(ba)k/π.\displaystyle=\begin{cases}b-a,&\ell=-(b-a)k/\pi,\\ \dfrac{2(b-a)}{(b-a)k+\pi\ell}e^{i((b-a)k+\pi\ell)/2}\sin(((b-a)k+\pi\ell)/2),&\ell\neq-(b-a)k/\pi.\end{cases}
Example 2.2.

For one sided algebraic singularity where w(x)=(xa)β,w(x)=(x-a)^{\beta}, β(1,0)\beta\in(-1,0), the moments are given by

Wk,[a,b]\displaystyle W^{[a,b]}_{k,\ell} ={(ba)1+β1+β,=(ba)k/π,(i(ba)(ba)k+π)1+β(Γ(1+β)Γ(1+β,i((ba)k+π))),(ba)k/π.\displaystyle=\begin{cases}\dfrac{(b-a)^{1+\beta}}{1+\beta},&\ell=-(b-a)k/\pi,\\ \left(\dfrac{i(b-a)}{(b-a)k+\pi\ell}\right)^{1+\beta}(\Gamma(1+\beta)-\Gamma(1+\beta,-i((b-a)k+\pi\ell))),&\ell\neq-(b-a)k/\pi.\end{cases}
Example 2.3.

For w(x)=(xa)β(bx)βw(x)=(x-a)^{\beta}(b-x)^{\beta} with β(1,0)\beta\in(-1,0), the weak algebraic singularity at both end-points of the same order β\beta, we have

Wk,[a,b]=πΓ(1+β)(ba)1+2βei((ba)k+π)/2((ba)k+π)1/2βJ1/2+β(((ba)k+π)/2)\displaystyle W^{[a,b]}_{k,\ell}=\sqrt{\pi}\ \Gamma(1+\beta)(b-a)^{1+2\beta}e^{i((b-a)k+\pi\ell)/2}((b-a)k+\pi\ell)^{-1/2-\beta}J_{1/2+\beta}(((b-a)k+\pi\ell)/2)

provided (ba)k/π\ell\neq-(b-a)k/\pi, whereas

Wk,[a,b]=(ba2)1+2βπΓ(1+β)Γ(32+β),\displaystyle W^{[a,b]}_{k,\ell}=\left(\frac{b-a}{2}\right)^{1+2\beta}\frac{\sqrt{\pi}\hskip 2.84544pt\Gamma(1+\beta)}{\Gamma\left(\frac{3}{2}+\beta\right)},

if =(ba)k/π\ell=-(b-a)k/\pi. In particular, for β=1/2\beta=-1/2, we get

Wk,[a,b]=πei((ba)k+π)/2J0(((ba)k+π)/2).\displaystyle W^{[a,b]}_{k,\ell}=\pi e^{i((b-a)k+\pi\ell)/2}J_{0}(((b-a)k+\pi\ell)/2).
Example 2.4.

For w(x)=(xa)α(bx)βw(x)=(x-a)^{\alpha}(b-x)^{\beta}, for α,β(1,0)\alpha,\beta\in(-1,0), when the singularity is present at both end-points with different exponents, then

Wk,[a,b]\displaystyle W^{[a,b]}_{k,\ell} =ba2ei((ba)k+π)/211(ba2)α+β(1+v)α(1v)βei((ba)k+π)v/2𝑑v\displaystyle=\frac{b-a}{2}e^{i((b-a)k+\pi\ell)/2}\int_{-1}^{1}\left(\frac{b-a}{2}\right)^{\alpha+\beta}(1+v)^{\alpha}(1-v)^{\beta}e^{i((b-a)k+\pi\ell)v/2}\,dv
=(ba)α+β+1Γ(2+α+β)Γ(1+α)Γ(1+β)1F1(1+α,2+α+β,i((ba)k+π)).\displaystyle=\frac{(b-a)^{\alpha+\beta+1}}{\Gamma(2+\alpha+\beta)}\Gamma(1+\alpha)\Gamma(1+\beta)_{1}F_{1}(1+\alpha,2+\alpha+\beta,i((b-a)k+\pi\ell)).
Example 2.5.

For w(x)=log(xa)w(x)=\log(x-a), when the logarithmic singularity is present at the left end-point, then

Wk,[a,b]=i(ba)(ab)kπ(\displaystyle W^{[a,b]}_{k,\ell}=\frac{i(b-a)}{(a-b)k-\pi\ell}( γ+Γ(0,i((ba)k+π))+(1+ei((ba)k+π))log(ba)\displaystyle\gamma+\Gamma(0,-i((b-a)k+\pi\ell))+(-1+e^{i((b-a)k+\pi\ell)})\log(b-a)
+log(i((ba)k+π)),\displaystyle+\log(-i((b-a)k+\pi\ell)),

if (ba)k+π0(b-a)k+\pi\ell\neq 0 and Wk,[a,b]=(ba)(log(ba)1)W^{[a,b]}_{k,\ell}=(b-a)(\log(b-a)-1) otherwise.

3. Error Analysis

We begin by noting that, using the Fourier series of the periodic extension of ff, that is,

f(x)==c(fe)eπi(xa)/(ba),\displaystyle f(x)=\sum_{\ell=-\infty}^{\infty}c_{\ell}(f^{e})e^{\pi i\ell(x-a)/(b-a)},

where

c(fe)=1ξaa+ξfe(x)e2πi(xa)/ξ𝑑x=12(ba)a2bafe(x)eπi(xa)/(ba)𝑑x,\displaystyle c_{\ell}(f^{e})=\frac{1}{\xi}\int_{a}^{a+\xi}f^{e}(x)e^{-2\pi i\ell(x-a)/\xi}\,dx=\frac{1}{2(b-a)}\int_{a}^{2b-a}f^{e}(x)e^{-\pi i\ell(x-a)/(b-a)}\,dx,

the exact integral can be expressed as

Ik[a,b](w,f)=eika=c(fe)Wk,[a,b](w).\displaystyle I^{[a,b]}_{k}(w,f)=e^{ika}\sum_{\ell=-\infty}^{\infty}c_{\ell}(f^{e})W^{[a,b]}_{k,\ell}(w).

The error in the numerical integration, therefore, is given by

Ik[a,b](w,f)Ik,n[a,b](w,f)=eika(=c(fe)Wk,[a,b](w)=nn1d,n(fr,qe)Wk,[a,b](w))=\displaystyle I^{[a,b]}_{k}(w,f)-I^{[a,b]}_{k,n}(w,f)=e^{ika}\left(\sum_{\ell=-\infty}^{\infty}c_{\ell}(f^{e})W_{k,\ell}^{[a,b]}(w)-\sum_{\ell=-n}^{n-1}d_{\ell,n}(f_{r,q}^{e})W_{k,\ell}^{[a,b]}(w)\right)=
(13) eika(=nn1m=m0c+2mn(fe)Wk,+2mn[a,b](w)+=nn1(c(fe)d,n(fr,qe))Wk,[a,b](w))\displaystyle e^{ika}\left(\sum_{\ell=-n}^{n-1}\sum_{\begin{subarray}{c}m=-\infty\\ m\neq 0\end{subarray}}^{\infty}c_{\ell+2mn}(f^{e})W_{k,\ell+2mn}^{[a,b]}(w)+\sum_{\ell=-n}^{n-1}\left(c_{\ell}(f^{e})-d_{\ell,n}(f_{r,q}^{e})\right)W_{k,\ell}^{[a,b]}(w)\right)

Using

c(fe)d,n(fr,qe)=c(fefr,qe)m=m0c+2mn(fr,qe),\displaystyle c_{\ell}(f^{e})-d_{\ell,n}(f_{r,q}^{e})=c_{\ell}(f^{e}-f_{r,q}^{e})-\sum_{\begin{subarray}{c}m=-\infty\\ m\neq 0\end{subarray}}^{\infty}c_{\ell+2mn}(f_{r,q}^{e}),

we rewrite the second summation in eq. 13 as

(14) =nn1c(fefr,qe)Wk,[a,b](w)=nn1m=m0c+2mn(fr,qe)Wk,[a,b](w),\displaystyle\sum_{\ell=-n}^{n-1}c_{\ell}(f^{e}-f_{r,q}^{e})W_{k,\ell}^{[a,b]}(w)-\sum_{\ell=-n}^{n-1}\sum_{\begin{subarray}{c}m=-\infty\\ m\neq 0\end{subarray}}^{\infty}c_{\ell+2mn}(f_{r,q}^{e})W_{k,\ell}^{[a,b]}(w),

and, therefore, combining the expressions in eq. 13 and eq. 14, the error reads

Ik[a,b](w,f)Ik,n[a,b](w,f)=eika(=nn1m=m0c+2mn(fe)(Wk,+2mn[a,b](w)Wk,[a,b](w))+\displaystyle I^{[a,b]}_{k}(w,f)-I^{[a,b]}_{k,n}(w,f)=e^{ika}\left(\sum_{\ell=-n}^{n-1}\sum_{\begin{subarray}{c}m=-\infty\\ m\neq 0\end{subarray}}^{\infty}c_{\ell+2mn}(f^{e})\left(W_{k,\ell+2mn}^{[a,b]}(w)-W_{k,\ell}^{[a,b]}(w)\right)+\right.
(15) =nn1m=m0c+2mn(fefr,qe)Wk,[a,b](w)+=nn1c(fefr,qe)Wk,[a,b](w))\displaystyle\left.\sum_{\ell=-n}^{n-1}\sum_{\begin{subarray}{c}m=-\infty\\ m\neq 0\end{subarray}}^{\infty}c_{\ell+2mn}(f^{e}-f_{r,q}^{e})W_{k,\ell}^{[a,b]}(w)+\sum_{\ell=-n}^{n-1}c_{\ell}(f^{e}-f_{r,q}^{e})W_{k,\ell}^{[a,b]}(w)\right)

To see that

(16) =nn1c(fefr,qe)Wk,[a,b](w)==nn1m=m0c+2mn(fefr,qe)Wk,+2mn[a,b](w)\displaystyle\sum_{\ell=-n}^{n-1}c_{\ell}(f^{e}-f_{r,q}^{e})W_{k,\ell}^{[a,b]}(w)=-\sum_{\ell=-n}^{n-1}\sum_{\begin{subarray}{c}m=-\infty\\ m\neq 0\end{subarray}}^{\infty}c_{\ell+2mn}(f^{e}-f_{r,q}^{e})W_{k,\ell+2mn}^{[a,b]}(w)

we introduce

A(g)(z):=a2baw0(y)g(z+y)eik(ya)𝑑y\displaystyle A(g)(z):=\int_{a}^{2b-a}w_{0}(y)g(z+y)e^{ik(y-a)}\,dy

where w0w_{0} is the 2(ba)2(b-a)-periodic function with w0(x)=w(x)w_{0}(x)=w(x) for x[a,b]x\in[a,b] and w0(x)=0w_{0}(x)=0 for x(b,2ba)x\in(b,2b-a). As

A(g)(z)==c(g)eπiz/(ba)Wk,[a,b](w),\displaystyle A(g)(z)=\sum_{\ell=-\infty}^{\infty}c_{\ell}(g)e^{\pi i\ell z/(b-a)}W^{[a,b]}_{k,\ell}(w),

we have

=c(fefr,qe)Wk,[a,b](w)=A(fefr,qe)(0)=0.\displaystyle\sum_{\ell=-\infty}^{\infty}c_{\ell}(f^{e}-f_{r,q}^{e})W^{[a,b]}_{k,\ell}(w)=A(f^{e}-f_{r,q}^{e})(0)=0.

Therefore, using eq. 16, and after simplification, the error in section 3 is expressed as

(17) Ik[a,b](w,f)Ik,n[a,b](w,f)=eika=nn1m=m0c+2mn(fr,qe)(Wk,+2mn[a,b](w)Wk,[a,b](w)).\displaystyle I^{[a,b]}_{k}(w,f)-I^{[a,b]}_{k,n}(w,f)=e^{ika}\sum_{\ell=-n}^{n-1}\sum_{\begin{subarray}{c}m=-\infty\\ m\neq 0\end{subarray}}^{\infty}c_{\ell+2mn}(f_{r,q}^{e})\left(W_{k,\ell+2mn}^{[a,b]}(w)-W_{k,\ell}^{[a,b]}(w)\right).

Thus, to obtain an error bound, we need an estimate on |c(fr,qe)|\left|c_{\ell}(f_{r,q}^{e})\right|. In this direction, it is important to note that fr,qe(x)=fe(x)+nqg(x)f_{r,q}^{e}(x)=f^{e}(x)+n^{-q}g(x) where g(x)=0g(x)=0 for x[a,b]x\in[a,b], gC2(b,2ba)g\in C^{2}(b,2b-a) and g(b)=g(2ba)=0g(b)=g(2b-a)=0. Thus, for 0\ell\neq 0, we have

(18) c(fr,qe)=c(fe)+nqba2(πi)2((1)g(b)g(2ba)+b2bag′′(x)eπi(xa)/(ba)𝑑x)\displaystyle c_{\ell}(f_{r,q}^{e})=c_{\ell}(f^{e})+n^{-q}\frac{b-a}{2(\pi i\ell)^{2}}\left((-1)^{\ell}g^{\prime}(b)-g^{\prime}(2b-a)+\int_{b}^{2b-a}g^{\prime\prime}(x)e^{-\pi i\ell(x-a)/(b-a)}\,dx\right)

and

(19) c(fe)=(ba)r+12(πi)r+2((1)(p(r+1)(b)f(r+1)(b))+f(r+1)(a)p(r+1)(2ba)\displaystyle c_{\ell}(f^{e})=\frac{(b-a)^{r+1}}{2(\pi i\ell)^{r+2}}\Bigg{(}(-1)^{\ell}\left(p^{(r+1)}(b)-f^{(r+1)}(b)\right)+f^{(r+1)}(a)-p^{(r+1)}(2b-a)
+abf(r+2)(x)eπi(xa)/(ba)dx+b2bap(r+2)(x)eπi(xa)/(ba)dx).\displaystyle\quad\quad\quad\quad\quad+\int_{a}^{b}f^{(r+2)}(x)e^{-\pi i\ell(x-a)/(b-a)}\,dx+\int_{b}^{2b-a}p^{(r+2)}(x)e^{-\pi i\ell(x-a)/(b-a)}\,dx\Bigg{)}.

Therefore, using eq. 18 and eq. 19, we conclude that there exits a positive constant CC such that

(20) |c(fr,qe)|C(||(r+2)+nq||2),{0}.\displaystyle\left|c_{\ell}(f_{r,q}^{e})\right|\leq C\left(|\ell|^{-(r+2)}+n^{-q}|\ell|^{-2}\right),\quad\ell\in\mathbb{Z}\setminus\{0\}.

Finally, combining eq. 17 and eq. 20, we have

(21) |Ik[a,b](w,f)Ik,n[a,b](w,f)|Cn(min{r,q}+2)=nn1m=|Wk,+2mn[a,b](w)Wk,[a,b](w)|(2|m|1)2.\displaystyle\left|I^{[a,b]}_{k}(w,f)-I^{[a,b]}_{k,n}(w,f)\right|\leq Cn^{-(\min\{r,q\}+2)}\sum_{\ell=-n}^{n-1}\sum_{m=-\infty}^{\infty}\frac{\left|W_{k,\ell+2mn}^{[a,b]}(w)-W_{k,\ell}^{[a,b]}(w)\right|}{\left(2|m|-1\right)^{2}}.

We, therefore, see that the convergence rate for the approximations does depend on the weight function ww. In this context, using the notation

𝒲k,n,M[a,b](w):==nn1m=MM|Wk,+2mn[a,b](w)Wk,[a,b](w)|(2|m|1)2\displaystyle\mathcal{W}^{[a,b]}_{k,n,M}(w):=\sum_{\ell=-n}^{n-1}\sum_{m=-M}^{M}\frac{\left|W_{k,\ell+2mn}^{[a,b]}(w)-W_{k,\ell}^{[a,b]}(w)\right|}{\left(2|m|-1\right)^{2}}

we introduce a class of weight functions defined as

𝕎γ(a,b):={wL1(a,b):n0,C>0,𝒲k,n,[a,b](w)Cnγ, for all nn0}\displaystyle\mathbb{W}_{\gamma}(a,b):=\left\{w\in L^{1}(a,b):\exists\ n_{0}\in\mathbb{N},\ C>0,\mathcal{W}^{[a,b]}_{k,n,\infty}(w)\leq Cn^{\gamma},\text{ for all }n\geq n_{0}\right\}

for which the approximations converge at the same rate for a given ff. Obviously, 𝕎γ1(a,b)𝕎γ2(a,b)\mathbb{W}_{\gamma_{1}}(a,b)\subseteq\mathbb{W}_{\gamma_{2}}(a,b) for γ1γ2\gamma_{1}\leq\gamma_{2}. The numerical calculations plotted in fig. 1 indicate that the the weight w=(xa)β,β(1,0)w=(x-a)^{\beta},\beta\in(-1,0) belongs to 𝕎β(a,b)\mathbb{W}_{-\beta}(a,b) whereas w=(xa)α(bx)β,α,β(1,0)w=(x-a)^{\alpha}(b-x)^{\beta},\alpha,\beta\in(-1,0) belongs to 𝕎γ(a,b)\mathbb{W}_{\gamma}(a,b) with γ=max{α,β}\gamma=\max\{-\alpha,-\beta\}.

The following theorem captures the essence of the error analysis of the proposed integration scheme.

Refer to caption
(a) w(x)=xβ,k=10w(x)=x^{\beta},\ k=10
Refer to caption
(b) w(x)=xα(1x)β,k=10w(x)=x^{\alpha}(1-x)^{\beta},\ k=10
Refer to caption
(c) w(x)=xβ,k=100w(x)=x^{\beta},\ k=100
Refer to caption
(d) w(x)=xα(1x)β,k=100w(x)=x^{\alpha}(1-x)^{\beta},\ k=100
Figure 1. The plot of log10Wk,n,300[a,b](w)\log_{10}W_{k,n,300}^{[a,b]}(w) against log10(1/n)\log_{10}\left(1/n\right)
Theorem 3.1.

Let w𝕎γ(a,b)w\in\mathbb{W}_{\gamma}(a,b) and fC([a,b])f\in C^{\infty}([a,b]). Let

Ik,n[a,b](w,f):=eika=nn1d,n(fr,qe)Wk,[0,1](w),q,n,r{0},\displaystyle I_{k,n}^{[a,b]}(w,f):=e^{ika}\sum_{\ell=-n}^{n-1}d_{\ell,n}(f_{r,q}^{e})W_{k,\ell}^{[0,1]}(w),\quad q,n\in\mathbb{N},\ \ r\in\mathbb{N}\cup\{0\},

where

fr,qe(x)={f(x),x[a,b],m=0r𝒟n,qm,(f)(b)pm,r[b,2ba](x)+m=0r𝒟n,qm,+(f)(a)pm,r[2ba,b](x),x(b,2ba),\displaystyle f_{r,q}^{e}(x)=\begin{dcases}f(x),&x\in[a,b],\\ \sum_{m=0}^{r}\mathcal{D}^{m,-}_{n,q}(f)(b)p_{m,r}^{[b,2b-a]}(x)+\sum_{m=0}^{r}\mathcal{D}^{m,+}_{n,q}(f)(a)p_{m,r}^{[2b-a,b]}(x),&x\in(b,2b-a),\end{dcases}
d,n(fr,qe)=12nj=02n1fr,qe(a+(ba)j/n)eπij/n,Wk,[a,b](w)=abw(x)ei(k(ba)+π)(xa)/(ba)𝑑x,\displaystyle d_{\ell,n}(f_{r,q}^{e})=\frac{1}{2n}\sum_{j=0}^{2n-1}f_{r,q}^{e}(a+(b-a)j/n)e^{-\pi i\ell j/n},\ W^{[a,b]}_{k,\ell}(w)=\int_{a}^{b}w(x)e^{i(k(b-a)+\pi\ell)(x-a)/(b-a)}\,dx,

be the approximation to the oscillatory integral Ik[a,b](w,f)I_{k}^{[a,b]}(w,f). Then, there exist n0n_{0}\in\mathbb{N} and a constant C>0C>0 such that

|Ik[a,b](w,f)Ik,n[a,b](w,f)|Cn(min{r,q}+2γ),\displaystyle\left|I_{k}^{[a,b]}(w,f)-I_{k,n}^{[a,b]}(w,f)\right|\leq Cn^{-(\min\{r,q\}+2-\gamma)},

for all nn0n\geq n_{0}.

The optimal choice q=rq=r is, therefore, recommended in practice for approximating the derivatives that results in order r+2γr+2-\gamma rate of convergence for integral approximations.

4. Numerical Results

In this section, we present a variety of numerical experiments to validate the performance of the proposed method. Unless otherwise stated, we use q=rq=r in quadrature approximations.

Example 4.1.

To begin with, we test our method on an example used by David Levin in [14] and consider the integral

Ik=01sin(t)eik(t+t2)𝑑t.\displaystyle I_{k}=\int_{0}^{1}\sin(t)e^{ik(t+t^{2})}\,dt.

The change of variable x=t+t2x=t+t^{2} yields Ik=02f(x)eikx𝑑xI_{k}=\int_{0}^{2}f(x)e^{ikx}\,dx where

f(x)=sin(1+4x+12)14x+1\displaystyle f(x)=\sin\left(\frac{-1+\sqrt{4x+1}}{2}\right)\frac{1}{\sqrt{4x+1}}

is smooth on the interval [0,2][0,2]. Thus, the quadrature with weight function w(x)=1w(x)=1 produces rapidly convergent approximations Ik,nI_{k,n} of IkI_{k} where the convergence rate can be raised arbitrarily by increasing the parameter rr. We studied the convergence at k=100k=100, k=500k=500 and k=1000k=1000 for r=1,2,3,4,r=1,2,3,4, and presented the results in fig. 2 where we plot log10|Ik,nIk|\log_{10}|I_{k,n}-I_{k}| against log10(1/n)\log_{10}(1/n) to gauge the speed of convergence through the slope of respective lines. The plots clearly show that approximations converge with the rate r+2r+2.

Refer to caption
(a) k=100k=100
Refer to caption
(b) k=500k=500
Refer to caption
(c) k=1000k=1000
Figure 2. The plot of log10|Ik,nIk|\log_{10}|I_{k,n}-I_{k}| against log10(1/n)\log_{10}\left(1/n\right) for integral in 4.1
Example 4.2.

For studying the convergence behaviour of the numerical integration scheme on a second example without any singularity, that is, w(x)=1w(x)=1, we consider the integral

Ik=01sin(cost)sinteik(cost)dt\displaystyle I_{k}=\int_{0}^{1}\sin(\cos t)\sin t\,e^{ik(\cos t)}\,dt

The results in fig. 3 show that the order of convergence r+2r+2 does not degrade with increasing oscillation frequency.

Refer to caption
(a) k=10k=10
Refer to caption
(b) k=100k=100
Refer to caption
(c) k=1000k=1000
Figure 3. The plot of log10|Ik,nIk|\log_{10}|I_{k,n}-I_{k}| against log10(1/n)\log_{10}(1/n) for integral in 4.2
Example 4.3.

For the next example, we consider the integral

Ikp=01eiktpdt=1p01x(p1)/peikxdx\displaystyle I_{k}^{p}=\int_{0}^{1}e^{ikt^{p}}\,dt\quad=\frac{1}{p}\int_{0}^{1}x^{-(p-1)/p}e^{ikx}\,dx

that has been studied in literature for measuring the performance of quadratures in dealing with end-point algebraic singularities. For instance, the results for this example is discussed in [15] where, for k=10000k=10000 and p=2p=2, a relative error of the order 10710^{-7} is achieved with n=100n=100. For k=3000k=3000 and p=10p=10, a relative error of 101210^{-12} is obtained with n=100n=100. Approximations to these integrals using Filon-Clenshaw-Curtis have also been reported in [8] for p=2/3p=2/3 and p=4/3p=4/3 under experiment 1. Our approach, by construction, captures these integrals exactly upto the machine precision through analytic moment computations as seen in table 1 where we report the errors |Ik,2pIkp||I^{p}_{k,2}-I^{p}_{k}| for k=10m,m=3,4,5,6,7k=10^{m},\ m=3,4,5,6,7 and p=2/3,4/3,2,10p=2/3,4/3,2,10. The results shown are for n=2n=2 and r=0r=0.

kk p=2/3p=2/3 p=4/3p=4/3 p=2p=2 p=10p=10
10310^{3} 7.4325×10177.4325\times 10^{-17} 1.7110×10161.7110\times 10^{-16} 1.2337×10161.2337\times 10^{-16} 4.6653×10164.6653\times 10^{-16}
10410^{4} 0.0000×10+000.0000\times 10^{+00} 2.7730×10162.7730\times 10^{-16} 9.7618×10179.7618\times 10^{-17} 5.8885×10165.8885\times 10^{-16}
10510^{5} 2.2818×10172.2818\times 10^{-17} 2.2485×10162.2485\times 10^{-16} 1.5455×10161.5455\times 10^{-16} 5.5786×10165.5786\times 10^{-16}
10610^{6} 0.0000×10+000.0000\times 10^{+00} 2.9916×10162.9916\times 10^{-16} 0.0000×10+000.0000\times 10^{+00} 8.1510×10168.1510\times 10^{-16}
10710^{7} 0.0000×10+000.0000\times 10^{+00} 6.9014×10166.9014\times 10^{-16} 1.3676×10161.3676\times 10^{-16} 4.4208×10164.4208\times 10^{-16}
Table 1. The approximation errors |Ik,npIkp||I_{k,n}^{p}-I_{k}^{p}| for the integral in 4.3 with n=2n=2 and r=0r=0
Example 4.4.

Next, we study the performance of our method in approximating the integral

Ik=0πeik(cost1)2+sin2t𝑑t,\displaystyle I_{k}=\int_{0}^{\pi}e^{ik\sqrt{(\cos t-1)^{2}+\sin^{2}t}}\,dt,

Using the change of variable x=(cost1)2+sin2tx=\sqrt{(\cos t-1)^{2}+\sin^{2}t}, we have

Ik=0224x2eikx𝑑x=0212x22+xeikx𝑑x,\displaystyle I_{k}=\int_{0}^{2}\frac{2}{\sqrt{4-x^{2}}}e^{ikx}\,dx=\int_{0}^{2}\frac{1}{\sqrt{2-x}}\frac{2}{\sqrt{2+x}}e^{ikx}\,dx,

where we have an algebraic singularity of order β=1/2\beta=-1/2 at the upper end point of the integration interval. Moreover, the function f(x)=1/2+xf(x)=1/\sqrt{2+x} is smooth on [0,2][0,2]. The results in fig. 4 show that the approximations convergence at the expected rate r+3/2r+3/2.

Refer to caption
(a) k=100k=100
Refer to caption
(b) k=500k=500
Refer to caption
(c) k=1000k=1000
Figure 4. The plot of log10|Ik,nIk|\log_{10}|I_{k,n}-I_{k}| against log10(1/n)\log_{10}(1/n) for integral in 4.4
Example 4.5.

For the following integral

Ik=01xβ|x12|eikx𝑑x\displaystyle I_{k}=\int_{0}^{1}x^{\beta}\left|x-\frac{1}{2}\right|e^{ikx}\,dx

where f(x)=|x1/2|f(x)=|x-1/2| is only piecewise smooth, high-order approximation can still be obtained by breaking the integral appropriately as seen in the results reported in fig. 8 where numerical integrals exhibit the convergence rate r+2+βr+2+\beta.

Refer to caption
(a) β=1/2\beta=-1/2
Refer to caption
(b) β=1/4\beta=-1/4
Refer to caption
(c) β=2/3\beta=-2/3
Figure 5. The plot of log10|Ik,nIk|\log_{10}|I_{k,n}-I_{k}| against log10(1/n)\log_{10}(1/n) for integral in 4.5
Example 4.6.

We now consider an integral with a weight of the form w(x)=(xa)α(bx)βw(x)=(x-a)^{\alpha}(b-x)^{\beta}. In particular, we approximate the following integral

Ik1=01x1/2(1x)1/3exeikx𝑑x\displaystyle I_{k}^{1}=\int_{0}^{1}x^{-1/2}(1-x)^{-1/3}e^{x}e^{ikx}\,dx

using the proposed scheme for k=10k=10 and k=100k=100. The convergence rates for r=1,2,3,4r=1,2,3,4, shown in fig. 6, are consistent with the theoretical rate r+2max{α,β}r+2-\max\{-\alpha,-\beta\}.

Refer to caption
(a) k=10k=10
Refer to caption
(b) k=100k=100
Refer to caption
(c) k=500k=500
Figure 6. The plot of log10|Ik,n1Ik1|\log_{10}|I_{k,n}^{1}-I_{k}^{1}| against log10(1/n)\log_{10}(1/n) for integral in 4.6

The the results for a second integral of this type

Ik2=23(x2)1/4(3x)2/3sin(x)eikx𝑑x,\displaystyle I_{k}^{2}=\int_{2}^{3}(x-2)^{-1/4}(3-x)^{-2/3}\sin(x)e^{ikx}\,dx,

presented in fig. 7, confirm the convergence rates.

Refer to caption
(a) k=10k=10
Refer to caption
(b) k=100k=100
Refer to caption
(c) k=500k=500
Figure 7. The plot of log10|Ik,n2Ik2|\log_{10}|I_{k,n}^{2}-I_{k}^{2}| against log10(1/n)\log_{10}(1/n) for integral in 4.6
Example 4.7.

Finally, we study the performance of the proposed method in approximating the following integral

Ik=0π/2log((cost1)2+sin2t)eik(cost1)2+sin2t𝑑t,\displaystyle I_{k}=\int_{0}^{\pi/2}\log\left(\sqrt{(\cos t-1)^{2}+\sin^{2}t}\right)e^{ik\sqrt{(\cos t-1)^{2}+\sin^{2}t}}\,dt,

that has a logarithmic singularity at t=0t=0. As in 4.4, using the change of variable x=(cost1)2+sin2tx=\sqrt{(\cos t-1)^{2}+\sin^{2}t}, yields a linear oscillator

Ik=02log(x)24x2eikx𝑑x,\displaystyle I_{k}=\int_{0}^{\sqrt{2}}\log(x)\frac{2}{\sqrt{4-x^{2}}}e^{ikx}\,dx,

where w(x)=log(x)w(x)=\log(x) and f(x)=2/4x2f(x)=2/\sqrt{4-x^{2}} is smooth on [0,2][0,\sqrt{2}].

Refer to caption
(a) k=100k=100
Refer to caption
(b) k=500k=500
Refer to caption
(c) k=1000k=1000
Figure 8. The plot of log10|Ik,nIk|\log_{10}|I_{k,n}-I_{k}| against log10(1/n)\log_{10}(1/n) for integral in 4.7

5. Conclusion

In this paper, we introduce a new Filon-type quadrature for numerical of highly oscillatory integrals of the form eq. 4 that can be employed to approximate more general oscillatory integrals of the form eq. 1 with algebraic singularities and stationary points. The proposed method utilises an interpolating trigonometric polynomial to approximate the smooth and non-oscillatory part of the integrand. Normally, when a function is not periodic, such trigonometric polynomial approximations suffer from the Gibb’s phenomenon. In order that the unwanted Gibb’s oscillations do not adversely impact the accuracy of numerical integration, the method constructs a periodic extension of the underlying function before its approximation. The main advantage of using trigonometric interpolation lies in obtaining a relatively simpler moment problem that can be solved analytically for many important classes of integrands. This approach can easily handle one and two sided integrable algebraic singularities as well as logarithmic singularities, The accompanying numerical analysis of the method reveals that the approximations converge rapidly to the exact integral. The scheme is tested on a variety of highly oscillatory integrals and these numerical experiments confirm that the approximations converge at the rates predicted by the theory.

References

  • [1] L. N. G. Filon, Iii.—on a quadrature formula for trigonometric integrals, Proceedings of the Royal Society of Edinburgh 49 (1930) 38–47.
  • [2] Y. L. Luke, On the computation of oscillatory integrals, in: Mathematical Proceedings of the Cambridge Philosophical Society, Vol. 50, Cambridge University Press, 1954, pp. 269–277.
  • [3] E. Flinn, A modification of filon’s method of numerical integration, Journal of the ACM (JACM) 7 (2) (1960) 181–184.
  • [4] E. Tuck, A simple “filon-trapezoidal” rule, Mathematics of Computation 21 (98) (1967) 239–241.
  • [5] A. Iserles, On the numerical quadrature of highly-oscillating integrals i: Fourier transforms, IMA Journal of Numerical Analysis 24 (3) (2004) 365–391.
  • [6] A. Iserles, S. Nørsett, On quadrature methods for highly oscillatory integrals and their implementations, Tech. rep. (2004).
  • [7] V. Dominguez, I. Graham, V. Smyshlyaev, Stability and error estimates for filon–clenshaw–curtis rules for highly oscillatory integrals, IMA journal of numerical analysis 31 (4) (2011) 1253–1280.
  • [8] V. Dominguez, I. G. Graham, T. Kim, Filon–clenshaw–curtis rules for highly oscillatory integrals with algebraic singularities and stationary points, SIAM Journal on Numerical Analysis 51 (3) (2013) 1542–1566.
  • [9] V. Domínguez, Filon–clenshaw–curtis rules for a class of highly-oscillatory integrals with logarithmic singularities, Journal of computational and applied mathematics 261 (2014) 299–319.
  • [10] B. Li, S. Xiang, Efficient methods for highly oscillatory integrals with weakly singular and hypersingular kernels, Applied Mathematics and Computation 362 (2019) 124499.
  • [11] H. Kang, S. Xiang, Efficient integration for a class of highly oscillatory integrals, Applied Mathematics and Computation 218 (7) (2011) 3553–3564.
  • [12] A. Anand, A. K. Tiwari, A fourier extension based numerical integration scheme for fast and high-order approximation of convolutions with weakly singular kernels, SIAM Journal on Scientific Computing 41 (5) (2019) A2772–A2794.
  • [13] D. Huybrechs, S. Vandewalle, On the evaluation of highly oscillatory integrals by analytic continuation, SIAM Journal on Numerical Analysis 44 (3) (2006) 1026–1048.
  • [14] D. Levin, Procedures for computing one-and two-dimensional integrals of functions with rapid irregular oscillations, Mathematics of Computation 38 (158) (1982) 531–538.
  • [15] S. Zaman, et al., New quadrature rules for highly oscillatory integrals with stationary points, Journal of Computational and Applied Mathematics 278 (2015) 75–89.