1. Introduction
Highly oscillatory integrals (HOIs)—that is, integrals involving rapidly oscillating functions—frequently arise in mathematical physics, particularly in the context of dynamic problems in classical and quantum mechanics, electromagnetism, and related fields [
1]. These integrals typically do not admit closed-form solutions, necessitating the development of specialized numerical quadrature methods for their evaluation. Classical quadrature techniques, which rely on polynomial approximation of the integrand, generally fail to achieve the required accuracy in such cases.
The development of numerical methods for evaluating highly oscillatory integrals dates back to the work of Filon [
2], who was the first to propose an efficient quadrature rule for computing Fourier integrals. His method involves partitioning the integration interval into subintervals and applying a modified Simpson’s rule to each. Since then, numerous variants of Filon-type methods have been proposed [
3,
4,
5], each tailored to address specific characteristics of different classes of HOIs.
The continued development in this area underscores the relevance and challenge of accurately evaluating HOIs. For example, Wu and Sun [
6] presented an efficient algorithm for approximating integrals with highly oscillatory Hankel kernels using a Clenshaw–Curtis–Filon rule for fast computation. Gao and Chang [
7] developed a generalized bivariate Filon–Clenshaw–Curtis method capable of achieving high accuracy across a wide range of oscillatory parameters while maintaining low computational complexity. Kassid et al. [
8] proposed a Nystöm-type method for solving a class of nonlinear highly oscillatory Volterra integral equations with trigonometric kernels. Their approach leads to a nonlinear system involving oscillatory integrals, which is then efficiently handled using a two-point generalized quadrature rule to obtain a fully discretized scheme. Chen et al. [
9] derived a Filon-type method for computing finite-part integrals involving hypersingular and highly oscillatory terms. Similarly, Zhao et al. [
10] applied a generalized two-point quadrature method to Volterra integral equations with highly oscillatory Bessel kernels. By estimating appropriate quadrature weights, they ensured the solvability of the resulting discretized system.
A fundamentally different approach to evaluating highly oscillatory integrals (HOIs) was introduced by Levin [
11], who transformed the integration problem into the solution to an associated ordinary differential equation (ODE). This technique proved to be highly accurate for certain classes of HOIs and has since inspired numerous methods based on Levin’s original idea. For instance, Evans and Webster [
12] enhanced Levin’s method using Chebyshev polynomials and Chebyshev nodes to improve numerical stability and accuracy. Khan et al. [
13] further advanced the method by employing radial basis functions, resulting in a well-conditioned version of the Levin approach. Olver [
14] proposed a hybrid technique that combines Filon-type quadrature with Levin’s method to effectively handle integrals with stationary points. Zaman and Siraj-ul-Islam [
15] developed a stable algorithm based on reproducing kernel functions for the numerical evaluation of oscillatory integrals, both with and without a stationary phase. In their framework, reproducing kernel functions—defined on a real Hilbert space—serve as basis functions in the Levin formulation. Ma [
16] introduced a modified Levin quadrature to construct a class of composite quadrature rules for evaluating HOIs with weak singularities or stationary points. Wang and Xiang [
17] extended the method further by proposing augmented Levin methods for vector-valued HOIs involving exotic oscillators and turning points. In their approach, the original Levin ODE is reformulated into an augmented system, which is then solved using the spectral collocation method.
Since the Levin approach relies on solving differential equations, selecting an appropriate solution technique is crucial for achieving high accuracy in the evaluation of highly oscillatory integrals (HOIs). Among the widely used analytical and semi-analytical methods for solving differential equations, spectral methods, the Variational Iteration Method (VIM) [
18,
19,
20], and the Homotopy Perturbation Method (HPM) [
21,
22,
23] have received considerable attention due to their simplicity and effectiveness in handling a broad class of nonlinear problems. The VIM, originally introduced by He [
24], constructs correction functionals based on variational principles, enabling the iterative refinement of approximate solutions with relatively low computational cost. The HPM combines classical perturbation techniques with homotopy theory to produce rapidly converging series solutions without requiring small parameters, making it appealing in various applications in physics and engineering. Despite their advantages, both the VIM and HPM face limitations when applied to problems involving highly oscillatory behavior, such as those governed by Levin’s differential formulation. These methods typically require problem-specific constructions—such as tailored correction functionals in the VIM or embedding parameters in the HPM—limiting their generality and robustness. Moreover, Levin’s equation lacks initial conditions, which makes it an ill-posed problem in traditional formulations and complicates its resolution using standard iterative or perturbative techniques. In contrast, spectral methods naturally handle such formulations through global approximation and yield well-posed algebraic systems. Therefore, for HOIs arising from Levin’s formulation, spectral approaches offer a more robust, general, and highly accurate solution framework.
Spectral approximation methods—particularly Galerkin-type methods—are known for their high convergence rates and numerical stability across a wide range of applications. Chakraborty et al. [
25] established superconvergence rates for the multi-Galerkin method without relying on iterated formulations. Yang et al. [
26] applied a multistep pseudo-spectral continuous Galerkin method, based on orthogonal Legendre polynomials, to solve first-order Volterra integro-differential equations. Cai [
27] proposed a spectral approach for solving a class of second-kind Volterra integral equations with highly oscillatory kernels featuring stationary points. Qin and Cao [
28] developed a Legendre–Galerkin spectral collocation least-squares method for approximating Darcy flow problems in both homogeneous and heterogeneous media.
Ma and Liu [
29] proposed a well-conditioned spectral Levin method for evaluating HOIs. Their method involves expanding both the known and unknown functions into Chebyshev polynomial series (i.e., a spectral representation on the Chebyshev basis), followed by truncation of the series. However, a notable drawback of this approach is the lack of thorough analysis of the truncation error. To address this limitation, Ma and Liu [
30] introduced the Jacobi–Galerkin–Levin method, which employs fractional Jacobi polynomials to construct a Galerkin approximation of the Levin equation. In particular, the method is studied in detail for the case of Chebyshev polynomials. Another well-conditioned technique was proposed by Molabahrami [
31], who approximated the unknown function using monomials and then applied the Galerkin method. Although this work includes a detailed convergence analysis, it does not fully address the issue of numerical quadrature. Specifically, the method is formulated for analytic functions rather than their discrete nodal values, which limits its applicability in practical numerical computations.
Motivated by the high accuracy and stability of spectral methods, this paper proposes a spectral Galerkin–Levin approach, which combines Levin’s transformation of the integration problem into an ordinary differential equation with Galerkin orthogonalization using Legendre polynomials. The proposed method addresses key limitations of existing approaches and yields a highly efficient and straightforward quadrature rule for evaluating highly oscillatory integrals.
2. Levin’s Approach to the Evaluation of Highly Oscillatory Integrals
Consider the integral
where
and
are continuously differentiable functions on the interval
, and
denotes the imaginary unit,
.
The Levin approach to evaluating the integral (
1) is based on finding a function
that satisfies the differential equation [
11]:
Using this function, the integral (
1) can be evaluated as
Since
, Equation (
2) can be rewritten as
which is the differential equation satisfied by the sought function
(hereafter referred to as the Levin equation).
Levin [
11] proposed representing the function
as a finite series expansion in terms of a complete system of basis functions
(e.g., monomials or polynomials):
where the coefficients
are determined by solving a system of linear algebraic equations obtained from Equation (
4) via the collocation technique:
In his original work, Levin used a linear set of nodes
and monomials for
. Later, Evans and Webster [
12] adopted Chebyshev nodes and Chebyshev polynomials as basis functions. However, since no initial conditions are provided, this collocation approach often results in an ill-conditioned system of linear equations [
12,
29,
31].
3. Spectral Galerkin–Levin Approach
This section combines spectral methods—specifically the Galerkin and Levin approaches—to develop a spectral Galerkin–Levin quadrature for highly oscillatory integrals. Furthermore, it demonstrates that applying Galerkin orthogonalization enables the use of spectral methods with arbitrary polynomial bases, whether monomials or orthogonal polynomials.
3.1. Orthogonal Polynomial-Based Galerkin–Levin Approach
Consider approximating the sought function
with the Legendre polynomial,
, through the finite sum of Fourier–Legendre series
which converges uniformly on any compact domain within an ellipse with foci at
as
([
32], par. 18.18). The coefficients
are defined by
Due to Equation (
8), the Legendre polynomials satisfy the orthogonality relation
where
is the Kronecker delta. This orthogonality relation (
9) is fundamental for applying the Galerkin method to approximate the solution to the Levin Equation (
4).
The functions
and
can also be expressed in terms of Legendre polynomials as
According to Olver et al. [
32], the following differentiation rule holds for Legendre polynomials:
Using this, one can define the differentiation operator (matrix)
as
Then, the derivatives of functions (
8) and (
11) can be written in the following compact form
Substituting Equations (
7), (
10), (
11) and (
14) into the Levin Equation (
4) yields
According to Dougall [
33], the product of two Legendre polynomials can be expressed as
where
and
Using this result, Equation (
15) can be rewritten as
Applying the Galerkin orthogonality conditions
to Equation (
19)—that is, multiplying by
and integrating over
—leads to the following well-conditioned system for determining the coefficients
:
where
are the coefficients of the matrix
, which is the inverse of the matrix
with coefficients
and
According to Olver et al. [
32], the Legendre polynomials satisfy the endpoint conditions
Using these properties along with Equations (
3), (
7), (
10), (
11), and (
21), the sought integral
from Equation (
1) can be approximated as
In numerical analysis, it is often more convenient to work directly with the values of the functions
and
at sampling points rather than with their Legendre coefficients
and
, as used in Equation (
25). Therefore, to provide a more classical form for the resulting quadrature and enhance its practical implementation, we consider the following numerical approach to evaluate the Legendre coefficients
and
.
Consider a set of sampling points
for
. The functions
and
can be approximated by their values
and
at these points using Lagrange interpolation polynomials
:
where
According to Equation (
8), the coefficients
and
can then be approximated as
where
Since the integrand in Equation (
29) is a polynomial with a degree of at most
, the coefficients
can be computed exactly using an
-point Gaussian quadrature, which is exact for polynomials with a degree up to
.
The sampling points
can, in principle, be chosen arbitrarily. However, to achieve optimal approximation and avoid Runge’s phenomenon, it is advantageous to use Chebyshev–Gauss–Lobatto (CGL) points [
34]
In the present method, CGL points are used exclusively for approximating the amplitude and phase functions and , both of which are assumed to be smooth and non-oscillatory. The primary purpose of employing these nodes is to construct accurate polynomial interpolants of these smooth functions using global basis functions, thereby ensuring spectral accuracy and mitigating Runge’s phenomenon. The latter typically arises when interpolating smooth functions using high-degree polynomials at equidistant points, where increasing the number of nodes leads to large oscillations near the interval boundaries. In contrast, the clustering of CGL points near the endpoints suppresses such behavior and promotes numerical stability.
It is important to emphasize that these sampling points are not intended to resolve the high-frequency content of the full oscillatory integrand . Instead, the oscillatory behavior is handled analytically through the proposed Galerkin–Levin formulation, which decouples the resolution of the smooth core functions from the treatment of the oscillatory exponential term.
3.2. Convergence Properties of the Spectral Galerkin–Levin Approach
The convergence of the proposed spectral Galerkin–Levin quadrature is intrinsically tied to the smoothness and analytic properties of the solution
to the associated Levin differential Equation (
4). This subsection presents a theoretical analysis of the convergence rate of the approximation (see Equation (
7))
where
are Legendre polynomials and
are computed via Galerkin projection.
Consider a case where the amplitude function
and the phase function
, and hence, the solution
are analytic in a closed Bernstein ellipse
in the complex plane containing the interval
. This condition is satisfied when
on
, ensuring that the phase function has no stationary points and the transformation to the Levin Equation (
4) does not introduce singularities.
In this setting, the Legendre approximation
converges exponentially in the
-norm:
where
is determined by the size of the largest ellipse
in which
is analytic, and
C is a constant independent of
n [
35,
36]. The corresponding error in the final integral approximation
inherits this exponential convergence.
Furthermore, since Galerkin projection is an orthogonal projection in the Hilbert space , the Galerkin approximation achieves the best possible accuracy within the chosen polynomial subspace. Thus, the obtained convergence rate is not only guaranteed but also optimal for polynomial spaces of degree n.
When the phase function contains stationary points (i.e., ) at some , the solution may exhibit reduced smoothness or local singular behavior. This affects the decay rate of Legendre coefficients , and therefore the convergence of .
In such cases, the convergence of the Galerkin projection is algebraic [
35,
36]. If
, for some
, the approximation satisfies the bound:
where
s reflects the degree of smoothness, and
C is a constant [
35,
36]. In the worst case, convergence may be significantly slower, and more basis functions are required to achieve high precision.
The numerical results presented in the Numerical Examples section corroborate these theoretical findings.
3.3. Discussion on the Selection of Polynomial Basis Functions and Approximation
Points
Some authors [
12,
29,
34] highlight the advantages of Chebyshev polynomials over monomials or other classical orthogonal polynomials (such as Jacobi polynomials) for approximating solutions to certain equations, including the Levin Equation (
4). However, each approach has its own peculiarities, which may or may not be influenced by the choice of polynomial basis functions.
Consider three different polynomial bases:
In Equations (
32) and (
33), the index
.
When considering a finite set of
polynomial basis functions
, it can be shown that
where, according to Equations (
32) and (
33), the nonzero coefficients satisfy the recurrence relations
with
.
Moreover, the inverse coefficient matrices
and
exist such that
Therefore, the Galerkin Equation (
20) can be rewritten for monomial and Chebyshev basis functions as follows:
Therefore, regardless, the monomial, Legendre, or Chebyshev polynomial basis is selected, according to Equations (
34), (
37)–(
39). The resulting Galerkin–Levin system of equations is just a linear mapping of the system with the Legendre polynomial basis and vice versa. Thus, the main advantage of the proposed approach, as well as other Galerkin-type or spectral methods, can be achieved with the appropriate selection of approximation points
in Equations (
28) and (
29).
Numerical tests for selected highly oscillatory integrals, as well as those of Evans and Webster [
12], enable Chebyshev nodes to allow us to obtain as much as two times greater accuracy compared to the uniform distribution of sampling points for low numbers of these points. Increasing the number of sampling points with a simultaneous increase in the number of resulting equations significantly decreases the error of integration, and the deviation between different sampling becomes insufficient. Nevertheless, Chebyshev–Gauss–Lobatto (
30) points allow us to exclude Runge’s phenomenon; therefore, this sampling points set is chosen for numerical integration with the proposed spectral Galerkin–Levin approach.
4. Spectral Galerkin–Levin Quadrature for HOIs of Matrix Functions
Consider the integral
where
and
are square matrices, and
is positive definite. Following the Levin approach (
2), we assume there exists a matrix function
such that
Then, the integral (
40) evaluates to
By applying the product rule to the derivative in Equation (
41), one obtains
Since
, the latter equation can be rewritten as
Transposing the matrices, we have
Assuming the expansions
and following the spectral Galerkin–Levin approach presented above, the coefficients
are determined by solving the system
where the matrices
and
are defined by Equations (
13) and (
23), respectively.
This approach avoids the need to compute matrix eigenvectors and eigenvalues at each collocation point, which are usually required to evaluate the matrix exponential in Equation (
40). Instead, the matrix exponential is evaluated only at the endpoints. Accordingly, using Equations (
24), (
42), and (
46), the integral is approximated as
Equation (
48) can be effectively used to evaluate the 3D time-harmonic Green’s function. According to Pasternak et al. [
37], the Radon transform technique allows the expression of the 3D time-harmonic Green’s function (fundamental solution) for anisotropic multifield or quasicrystalline media in the following matrix form:
where
Here,
, where
is the extended tensor of material constants, and
is the diagonal matrix of inertial properties [
37]. The range of the capital indices depends on the type of material considered (e.g., crystalline, quasicrystalline, ferroelectric, etc.).
In the evaluation of the time-harmonic Green’s function (
49), its component (
50) is a highly oscillatory integral over the surface of the unit sphere, which can be reduced to the following double integral [
37]:
where
and
.
The inner integral in Equation (
52) can be efficiently evaluated using the proposed spectral Galerkin–Levin matrix quadrature (
48). However, Levin’s method is not suitable for the outer integral because the integrand is periodic over the integration domain, resulting in multiple stationary points of
in the general case. Consequently, the sought function
exhibits highly oscillatory behavior similar to that of the integrand, making the direct application of Levin’s method ineffective. Therefore, the outer integral requires alternative quadrature methods or a high number of integration nodes.
Nonetheless, applying the proposed approach to the inner integral substantially reduces the overall number of nodes required for numerical integration, which is a significant computational advantage.
5. Extension to Higher-Dimensional Oscillatory Integrals via Nested Quadrature
Although the proposed spectral Galerkin–Levin quadrature is formulated for one-dimensional integrals, it can be systematically extended to multidimensional highly oscillatory integrals (HOIs) defined over square or product-type domains. The key idea is to interpret the multidimensional integral as a sequence of nested one-dimensional integrals, allowing the Galerkin–Levin approach to be applied iteratively in each coordinate direction.
Consider the two-dimensional oscillatory integral over a square domain:
where
and
are smooth amplitude and phase functions, and
.
We begin by expressing the integral in a nested form:
To evaluate
, we apply the Galerkin–Levin technique in the
x-direction. We construct a function
satisfying:
According to the Levin identity, the inner integral reduces to
Substituting Equation (
56) into Equation (
54), the two-dimensional integral becomes
where
denotes 1D oscillatory integrals in
y with modified amplitudes
and phase functions
.
The above transformation demonstrates that the original two-dimensional HOI (
53) is reduced to th difference between two one-dimensional HOIs, where the highly oscillatory behavior in the
x-direction is handled by the Galerkin–Levin procedure. Importantly, the remaining
y-dependent integrals
retain a highly oscillatory structure and are likewise efficiently evaluated using the same spectral Galerkin–Levin technique. This recursive application of the method enables the systematic and accurate treatment of multidimensional oscillatory integrals over square or tensor-product domains.
In the multidimensional nested approach, the functions
and
, which appear as boundary values in the inner integral evaluation, are computed at a discrete set of points
. These nodal values are then approximated by expanding
in terms of Legendre polynomials as
where the coefficients
are determined via Equation (
28). This approach preserves the spectral accuracy in each coordinate direction by systematically approximating the boundary functions of the inner integrals, thereby enabling efficient and accurate computation of the overall multidimensional highly oscillatory integral.
The nested framework naturally extends the spectral Galerkin–Levin approach to multidimensional highly oscillatory integrals, enabling accurate and efficient numerical evaluation over tensor-product domains. Such multidimensional integrals commonly arise in various applied fields, including computational electromagnetics, wave propagation, and quantum mechanics, where handling oscillations in multiple variables is critical. By recursively applying the one-dimensional Galerkin–Levin quadrature in each coordinate direction, the method achieves spectral accuracy and computational efficiency in higher dimensions.
7. Conclusions
The presented spectral Galerkin–Levin approach defines a robust and efficient quadrature rule for the evaluation of highly oscillatory integrals of the form (
1), relying solely on the nodal values of the core functions
and
. Unlike the classical Levin collocation methods, this approach employs Galerkin orthogonalization combined with Legendre polynomial interpolation, resulting in a well-posed and numerically stable system of linear algebraic equations. The flexibility of the method allows the use of various interpolation nodes without loss of stability or accuracy. Furthermore, it is rigorously demonstrated that the Galerkin orthogonalization and the resulting solution are invariant with respect to the choice of polynomial basis—be it monomials or orthogonal polynomials—thus highlighting the generality and robustness of the approach.
The proposed method extends naturally to matrix-valued HOIs, significantly broadening its applicability to complex problems in mathematical physics, such as those involving anisotropic media and multifield models. This extension is achieved without incurring additional computational complexity related to matrix exponential evaluations, which are confined to boundary points.
Numerical experiments confirm that the method exhibits high convergence rates and excellent stability for HOIs both with and without stationary points in the oscillatory phase function. While the presence of stationary points naturally demands a higher density of sampling nodes to maintain accuracy, the approach remains computationally efficient and significantly reduces the overall complexity compared to traditional quadrature techniques.
This work provides a simple, flexible, and reliable numerical tool for tackling a broad class of highly oscillatory integrals, with potential for impactful applications in scientific computing and applied mathematics.