Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Approximation of rogue waves using Malmquist-Takenaka functions

Justin T. Cole Troy I. Johnson
Abstract

Rogue waves are fascinating large amplitude coherent structures that abruptly appear and then disappear soon after. In certain partial differential equations these waves are modeled by rational solutions. In this work we discuss approximating rogue wave solutions in a basis of orthogonal functions known as the Malmquist-Takenaka (MT) functions. This family of rational functions can be directly mapped to a modified Fourier series, allowing the fast Fourier transform computation of the spectral MT coefficients. Spectral differentiation matrices are derived. The approximation of the various rogue wave solutions in the nonlinear Schrödinger (NLS) equation is explored. The unstable nature of the NLS equation on a constant background and its effect on destabilizing and generating rogue waves is studied. Perturbing the constant solution with certain localized functions is found to generate rogue wave-like structures.

keywords:
rogue waves , Malmquist-Takenaka functions , rational functions
journal: Wave Motion
\affiliation

[label1]organization=University of Colorado, Colorado Springs,addressline=1420 Austin Bluffs Parkway, city=Colorado Springs, postcode=80918, state=CO, country=USA

{graphicalabstract}
{highlights}

Approximation of rational solutions to partial differential equations in terms of a family of orthogonal rational functions called the Malmquist-Takenaka functions. For certain rational functions, these approximations converge spectrally (exponentially) fast. Spectral differentiation matrices are derived.

Using this technique, different models for rogue waves in the nonlinear Schrödinger equation are studied. The well-known Peregrine soliton is observed to be exponentially unstable to localized perturbations. A perturbation of the constant background generates rogue wave-type structures which appear to be rational in structure.

1 Introduction

Rogue, or freak, waves are brief, but intense wave formations that have the capability to inflict serious harm with little notice. A commonly accepted description is that of a coherent structure that is localized in both space and time. Since their initial detection on the Draupner platform in 1995 [1], several observations have been reported around the world [2, 3, 4]. As a result, the understanding and detection of rogue waves is important for the safety of maritime vessels and coastal infrastructure.

A canonical model for nonlinear dispersive waves is the nonlinear Schrödinger (NLS) partial differential equation (PDE). The NLS equation has been derived to describe a wide variety of physical systems, such as water waves [5], fiber optics [6], and plasmas [7], to name a few. The derivation of the model typically relies on a slowly-varying enveloped approximation, where a monochromatic linear wave is modulated by a slowly-varying amplitude function that is governed by an NLS-type equation [8]. While these models typically assume weak nonlinearity, the NLS equation often captures leading-order effects and is a reasonable place to begin a discussion on rogue waves in deep water.

Within the NLS equation, there are two well-known paradigms for describing the physical generation of rogue waves [2]. The first is rational solutions [9, 10] that are localized in space and time; the most well-known is the Peregrine soliton [11]. The second is modulational instability (MI), where perturbations with a small sideband wavenumber are linearly unstable [12, 13]. Recently, several works have highlighted the unstable nature of the Peregrine soliton [14, 15, 16], throwing into question how this state could be realized in the tempestuous ocean. The MI route is worth pursuing, but it should be performed on the real line, and not on a periodic domain, to truly capture it’s nature in large bodies of water. The numerical simulation of these two rogue wave models is the motivation of this paper.

The industry standard for numerically approximating wave-related (that is, periodic) phenomena is spectral Fourier methods. For relatively smooth solutions, Fourier spectral methods are an attractive option due to their rapid convergence rates [17, 18]. Fourier approximations can also be useful when approximating exponentially decaying solutions on the real line. In the latter case, one typically truncates the infinite line to a large but finite domain, on which an exponentially localized solution can be treated as a periodic function [19, 20]. However, this approach requires the function reach its (constant) boundary conditions well within the computation window. As such, secant hyperbolic-type soliton solutions are an ideal candidate for this approach.

However, the Peregrine soliton, a rogue wave model, is a rational function that does not decay exponentially fast. The Fourier method described above struggles with rational functions due to their (slow) algebraic decay rate. On most feasible computational domains, these rational solutions are nowhere near zero, or machine precision, in practice. Hence, the periodic extension of these functions on finite domains are not sufficiently smooth and so, in general, their Fourier coefficients decay algebraically (slow).

In order to better approximate rational functions several ideas have been explored. One idea is to use the Hermite functions [21, 22] as a basis. This is a natural approach since the Hermite functions form an orthogonal basis on the real line. However, there is no known Hermite fast transform and the truncated series only converges spectrally fast if the function decays exponentially fast at ±plus-or-minus\pm\infty± ∞ [18]. Another approach is that of coordinate mapping [23, 24]. The idea here is to map the (infinite) real line to a finite interval by some coordinate transformation. One such mapping is the so-called algebraic type, e.g. x=y1y2𝑥𝑦1superscript𝑦2x=\frac{y}{\sqrt{1-y^{2}}}italic_x = divide start_ARG italic_y end_ARG start_ARG square-root start_ARG 1 - italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG for y(1,1)𝑦11y\in(-1,1)italic_y ∈ ( - 1 , 1 ) [25]. A method that uses said mapping and approximates rational solutions of the nonlinear Schrödinger equation is the Chebyshev pseudo-spectral method given in [26]. Often scaling parameters are included in the transformation and chosen to optimize the accuracy of the method.

In this work we explore the idea of using a rational basis to approximate rational functions. A family of orthogonal rational functions was considered in [27] that consists of the algebraic mapping mentioned above substituted into the argument of the Chebyshev polynomials, i.e. ψn(x)=Tn(x1+x2),n=0,1,formulae-sequencesubscript𝜓𝑛𝑥subscript𝑇𝑛𝑥1superscript𝑥2𝑛01\psi_{n}(x)=T_{n}\left(\frac{x}{\sqrt{1+x^{2}}}\right),n=0,1,\dotsitalic_ψ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) = italic_T start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( divide start_ARG italic_x end_ARG start_ARG square-root start_ARG 1 + italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG ) , italic_n = 0 , 1 , … for x(,)𝑥x\in(-\infty,\infty)italic_x ∈ ( - ∞ , ∞ ) where Tnsubscript𝑇𝑛T_{n}italic_T start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT are the Chebyshev polynomials. This orthogonal basis yields exponential convergence for functions that are analytic on the real line as well as banded differentiation matrices. A related candidate for rational basis is called the Malmquist-Takenaka (MT) functions [28, 29, 30]. It was shown in [31] that the MT functions are actually equivalent to the Chebyshev polynomials through a linear combination. As such, they inherit the many desirable properties mentioned above. Moreover, this family of functions are mutually orthogonal and the coefficients can be computed via Fourier transform.

Our method highlights the unstable nature of the Peregrine soliton. This instability is an existential threat to the formation of the Pergrine soliton solution over long time scales. We then explore on to perturbing constant background solutions. Localized perturbations are found to generate rational Peregrine-like structures that are localized in both space and time. The latter finding suggests that localized disturbances to periodic wave trains e.g. a strong wind gust, are capable of forming rogue waves.

The outline of the work is as follows. In Section 2 we introduce the MT functions and some of their properties. We discuss an expansion in terms of these functions and the rapid decay of their coefficients. In Section 3 spectral Galerkin differentiation matrices are constructed to approximate derivatives. In Section 4 we discuss a spectral Galerkin method for approximating PDEs. In particular, we establish the split-step method for approximating the dynamics of the NLS equation. In Section 5 we explore instability in the Peregrine soliton and mechanisms for generating rogue waves on a constant background.

2 The Malmquist-Takenaka functions

In this section, some properties of the MT functions are reviewed. Define the family of Malmquist–Takenaka rational functions

ϕn(x)=2πin(1+2ix12ix)n112ix,subscriptitalic-ϕ𝑛𝑥2𝜋superscript𝑖𝑛superscript12𝑖𝑥12𝑖𝑥𝑛112𝑖𝑥\phi_{n}(x)=\sqrt{\frac{2}{\pi}}i^{n}\left(\frac{1+2ix}{1-2ix}\right)^{n}\frac% {1}{1-2ix},italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) = square-root start_ARG divide start_ARG 2 end_ARG start_ARG italic_π end_ARG end_ARG italic_i start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ( divide start_ARG 1 + 2 italic_i italic_x end_ARG start_ARG 1 - 2 italic_i italic_x end_ARG ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 1 - 2 italic_i italic_x end_ARG , (1)

for n𝑛n\in{\mathbb{Z}}italic_n ∈ blackboard_Z. First, observe that ϕn(x)subscriptitalic-ϕ𝑛𝑥\phi_{n}(x)italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) is well-defined for x𝑥x\in\mathbb{R}italic_x ∈ blackboard_R with poles at x=i/2𝑥𝑖2x=-i/2italic_x = - italic_i / 2 (x=i/2𝑥𝑖2x=i/2italic_x = italic_i / 2) when n𝑛nitalic_n is even (odd). Several MT functions are plotted in Fig. 1. Below is a list of some properties of these functions.

Refer to caption
Figure 1: The real and imaginary parts of ϕn(x)subscriptitalic-ϕ𝑛𝑥\phi_{n}(x)italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) for n{3,,2}𝑛32n\in\{-3,\dots,2\}italic_n ∈ { - 3 , … , 2 } indicated by blue (solid) and red (dashed) curves, respectively.
  • 1.

    Boundedness and decay

    The MT functions are uniformly bounded on the real line, that is

    |ϕn(x)|=2π11+4x22π.subscriptitalic-ϕ𝑛𝑥2𝜋114superscript𝑥22𝜋|\phi_{n}(x)|=\sqrt{\frac{2}{\pi}}\frac{1}{\sqrt{1+4x^{2}}}\leq\sqrt{\frac{2}{% \pi}}.| italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) | = square-root start_ARG divide start_ARG 2 end_ARG start_ARG italic_π end_ARG end_ARG divide start_ARG 1 end_ARG start_ARG square-root start_ARG 1 + 4 italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG ≤ square-root start_ARG divide start_ARG 2 end_ARG start_ARG italic_π end_ARG end_ARG . (2)

    The algebraic decay rate of the MT functions as x±𝑥plus-or-minusx\rightarrow\pm\inftyitalic_x → ± ∞ is linear since

    |ϕn(x)|=2π11+4x212π1|x|.subscriptitalic-ϕ𝑛𝑥2𝜋114superscript𝑥212𝜋1𝑥|\phi_{n}(x)|=\sqrt{\frac{2}{\pi}}\frac{1}{\sqrt{1+4x^{2}}}\rightarrow\frac{1}% {\sqrt{2\pi}}\frac{1}{|x|}.| italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) | = square-root start_ARG divide start_ARG 2 end_ARG start_ARG italic_π end_ARG end_ARG divide start_ARG 1 end_ARG start_ARG square-root start_ARG 1 + 4 italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG → divide start_ARG 1 end_ARG start_ARG square-root start_ARG 2 italic_π end_ARG end_ARG divide start_ARG 1 end_ARG start_ARG | italic_x | end_ARG . (3)
  • 2.

    Oscillations

    The frequency of oscillations is directly related to the magnitude of the modal value |n|𝑛|n|| italic_n |. As |n|(0)𝑛annotatedabsent0|n|\rightarrow\infty~{}(\rightarrow 0)| italic_n | → ∞ ( → 0 ), the functions become more (less) oscillatory near the origin. Express the rational function that lies on the unit circle by eiθ=1+2ix12ixsuperscript𝑒𝑖𝜃12𝑖𝑥12𝑖𝑥e^{i\theta}=\frac{1+2ix}{1-2ix}italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT = divide start_ARG 1 + 2 italic_i italic_x end_ARG start_ARG 1 - 2 italic_i italic_x end_ARG. Then ϕn(x)subscriptitalic-ϕ𝑛𝑥\phi_{n}(x)italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) contains a factor of the form einθ=cos(nθ)+isin(nθ)superscript𝑒𝑖𝑛𝜃𝑛𝜃𝑖𝑛𝜃e^{in\theta}=\cos(n\theta)+i\sin(n\theta)italic_e start_POSTSUPERSCRIPT italic_i italic_n italic_θ end_POSTSUPERSCRIPT = roman_cos ( italic_n italic_θ ) + italic_i roman_sin ( italic_n italic_θ ) whose angular frequency increases with |n|𝑛|n|| italic_n |.

  • 3.

    Symmetry

    When n𝑛nitalic_n is even, the real part of ϕn(x)subscriptitalic-ϕ𝑛𝑥\phi_{n}(x)italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) is even and the imaginary part is odd. On the other hand, if n𝑛nitalic_n is an odd integer, then the real part of ϕn(x)subscriptitalic-ϕ𝑛𝑥\phi_{n}(x)italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) is odd and the imaginary part is even. That is,

    neven:𝑛even:\displaystyle n~{}\text{even:}italic_n even: (ϕn(x))=(ϕn(x)),(ϕn(x))=(ϕn(x))formulae-sequencesubscriptitalic-ϕ𝑛𝑥subscriptitalic-ϕ𝑛𝑥subscriptitalic-ϕ𝑛𝑥subscriptitalic-ϕ𝑛𝑥\displaystyle~{}~{}~{}~{}~{}~{}~{}\Re(\phi_{n}(-x))=\Re(\phi_{n}(x)),~{}~{}~{}% ~{}\Im(\phi_{n}(-x))=-\Im(\phi_{n}(x))roman_ℜ ( italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( - italic_x ) ) = roman_ℜ ( italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) ) , roman_ℑ ( italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( - italic_x ) ) = - roman_ℑ ( italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) )
    nodd:𝑛odd:\displaystyle n~{}\text{odd:}italic_n odd: (ϕn(x))=(ϕn(x)),(ϕn(x))=(ϕn(x))formulae-sequencesubscriptitalic-ϕ𝑛𝑥subscriptitalic-ϕ𝑛𝑥subscriptitalic-ϕ𝑛𝑥subscriptitalic-ϕ𝑛𝑥\displaystyle~{}~{}~{}~{}~{}~{}~{}\Re(\phi_{n}(-x))=-\Re(\phi_{n}(x)),~{}~{}~{% }~{}\Im(\phi_{n}(-x))=\Im(\phi_{n}(x))roman_ℜ ( italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( - italic_x ) ) = - roman_ℜ ( italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) ) , roman_ℑ ( italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( - italic_x ) ) = roman_ℑ ( italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) )

    This symmetry is evident in Fig. 1.

  • 4.

    Orthogonality

    Define the complex L2()superscript𝐿2L^{2}(\mathbb{R})italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( blackboard_R ) inner product

    f,g=f(x)g(x)𝑑x,𝑓𝑔superscriptsubscriptsuperscript𝑓𝑥𝑔𝑥differential-d𝑥\langle f,g\rangle=\int_{-\infty}^{\infty}f^{*}(x)g(x)dx,⟨ italic_f , italic_g ⟩ = ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_f start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_x ) italic_g ( italic_x ) italic_d italic_x , (4)

    where denotes complex conjugation. Then for distinct modes, the MT functions are mutually orthogonal in L2()superscript𝐿2L^{2}({\mathbb{R}})italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( blackboard_R ), that is

    ϕm,ϕn=ϕm(x)ϕn(x)𝑑x=δmn,subscriptitalic-ϕ𝑚subscriptitalic-ϕ𝑛superscriptsubscriptsubscriptsuperscriptitalic-ϕ𝑚𝑥subscriptitalic-ϕ𝑛𝑥differential-d𝑥subscript𝛿𝑚𝑛\langle\phi_{m},\phi_{n}\rangle=\int_{-\infty}^{\infty}\phi^{*}_{m}(x)\phi_{n}% (x)dx=\delta_{mn},⟨ italic_ϕ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT , italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ⟩ = ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_ϕ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ( italic_x ) italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) italic_d italic_x = italic_δ start_POSTSUBSCRIPT italic_m italic_n end_POSTSUBSCRIPT , (5)

    where δmnsubscript𝛿𝑚𝑛\delta_{mn}italic_δ start_POSTSUBSCRIPT italic_m italic_n end_POSTSUBSCRIPT denotes the Kronecker delta function. The coefficients in (1) are chosen to also ensure unit norm, i.e. ϕn=1normsubscriptitalic-ϕ𝑛1||\phi_{n}||=1| | italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT | | = 1, for each n𝑛nitalic_n.

  • 5.

    Derivative recurrence relation

    The MT functions satisfy the following skew-symmetric recurrence relation [30, 34]

    ddxϕn(x)=nϕn1(x)+i(2n+1)ϕn(x)+(n+1)ϕn+1(x).𝑑𝑑𝑥subscriptitalic-ϕ𝑛𝑥𝑛subscriptitalic-ϕ𝑛1𝑥𝑖2𝑛1subscriptitalic-ϕ𝑛𝑥𝑛1subscriptitalic-ϕ𝑛1𝑥\frac{d}{dx}\phi_{n}(x)=-n\phi_{n-1}(x)+i(2n+1)\phi_{n}(x)+(n+1)\phi_{n+1}(x).divide start_ARG italic_d end_ARG start_ARG italic_d italic_x end_ARG italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) = - italic_n italic_ϕ start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT ( italic_x ) + italic_i ( 2 italic_n + 1 ) italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) + ( italic_n + 1 ) italic_ϕ start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT ( italic_x ) . (6)

    This formula naturally leads to tridiagonal differentiation matrices of the first derivative. Contrast this with Fourier methods which are diagonal, i.e. ddxφn(x)=iknφn(x),𝑑𝑑𝑥subscript𝜑𝑛𝑥𝑖𝑘𝑛subscript𝜑𝑛𝑥\frac{d}{dx}\varphi_{n}(x)=ikn\varphi_{n}(x),divide start_ARG italic_d end_ARG start_ARG italic_d italic_x end_ARG italic_φ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) = italic_i italic_k italic_n italic_φ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) , where φn(x)=eiknxsubscript𝜑𝑛𝑥superscript𝑒𝑖𝑘𝑛𝑥\varphi_{n}(x)=e^{iknx}italic_φ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) = italic_e start_POSTSUPERSCRIPT italic_i italic_k italic_n italic_x end_POSTSUPERSCRIPT. Higher-order derivatives can be obtained by differentiating this formula.

  • 6.

    Hilbert transform eigenfunction

    The MT functions are eigenfunctions of the Hilbert transform. The Hilbert transform is defined by

    H[u](x)=1πu(y)xy𝑑y,𝐻delimited-[]𝑢𝑥1𝜋superscriptsubscript𝑢𝑦𝑥𝑦differential-d𝑦H[u](x)=\frac{1}{\pi}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}% {{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle% -$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!% \int_{-\infty}^{\infty}\frac{u(y)}{x-y}dy,italic_H [ italic_u ] ( italic_x ) = divide start_ARG 1 end_ARG start_ARG italic_π end_ARG - ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_u ( italic_y ) end_ARG start_ARG italic_x - italic_y end_ARG italic_d italic_y ,

    where the integral is defined in the Cauchy principal values sense

    f(x)𝑑x=limϵ0+[bϵf(x)𝑑x+b+ϵf(x)𝑑x].superscriptsubscript𝑓𝑥differential-d𝑥italic-ϵsuperscript0delimited-[]superscriptsubscript𝑏italic-ϵ𝑓𝑥differential-d𝑥superscriptsubscript𝑏italic-ϵ𝑓𝑥differential-d𝑥\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$% \scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.291% 66pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{-\infty}^{% \infty}f(x)dx=\underset{\epsilon\to 0^{+}}{\lim}\left[\int_{-\infty}^{b-% \epsilon}f(x)dx\hskip 3.61371pt+\hskip 3.61371pt\int_{b+\epsilon}^{\infty}f(x)% dx\right].- ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_f ( italic_x ) italic_d italic_x = start_UNDERACCENT italic_ϵ → 0 start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT end_UNDERACCENT start_ARG roman_lim end_ARG [ ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_b - italic_ϵ end_POSTSUPERSCRIPT italic_f ( italic_x ) italic_d italic_x + ∫ start_POSTSUBSCRIPT italic_b + italic_ϵ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_f ( italic_x ) italic_d italic_x ] .

    The MT functions satisfy the eigenvalue problem

    H[ϕn](x)=isgn(n)ϕn(x)wheresgn(n)={1ifn01ifn<0.𝐻delimited-[]subscriptitalic-ϕ𝑛𝑥𝑖sgn𝑛subscriptitalic-ϕ𝑛𝑥wheresgn𝑛cases1if𝑛01if𝑛0H[\phi_{n}](x)=-i\hskip 3.61371pt\text{sgn}(n)\phi_{n}(x)~{}~{}{\rm where}~{}~% {}\text{sgn}(n)=\begin{cases}1&\text{if}\hskip 21.68121ptn\geq 0\\ -1&\text{if}\hskip 21.68121ptn<0\\ \end{cases}.italic_H [ italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ] ( italic_x ) = - italic_i sgn ( italic_n ) italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) roman_where sgn ( italic_n ) = { start_ROW start_CELL 1 end_CELL start_CELL if italic_n ≥ 0 end_CELL end_ROW start_ROW start_CELL - 1 end_CELL start_CELL if italic_n < 0 end_CELL end_ROW . (7)

    A natural application of the MT functions has been in the approximation of solutions to the Benjamin-Ono equation which includes the Hilbert transform [37, 41].

Relationship to Fourier series

Throughout this work we consider expanding a function in terms of the MT functions, that is

f(x)=n=fˇnϕn(x),wherefˇn=ϕn,f=ϕn(x)f(x)𝑑x,formulae-sequence𝑓𝑥superscriptsubscript𝑛subscriptˇ𝑓𝑛subscriptitalic-ϕ𝑛𝑥wheresubscriptˇ𝑓𝑛subscriptitalic-ϕ𝑛𝑓superscriptsubscriptsubscriptsuperscriptitalic-ϕ𝑛𝑥𝑓𝑥differential-d𝑥f(x)=\sum_{n=-\infty}^{\infty}\check{f}_{n}\phi_{n}(x),\hskip 7.22743pt\text{% where}\hskip 7.22743pt\check{f}_{n}=\langle\phi_{n},f\rangle=\int_{-\infty}^{% \infty}\phi^{*}_{n}(x)f(x)~{}dx,italic_f ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_n = - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) , where overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = ⟨ italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , italic_f ⟩ = ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_ϕ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) italic_f ( italic_x ) italic_d italic_x , (8)

for the MT functions given in (1). As can be seen, the coefficients are obtained by projecting on an arbitrary mode and exploiting the orthogonality (5). Through a natural change of variable, this MT expansion can be related to the Fourier series of a periodic function. This is useful because it allows us to use the Fast Fourier transform (FFT) to compute the MT coefficients.

Start from coordinate transformation

eiθ=1+2ix12ixx=12tan(θ2),superscript𝑒𝑖𝜃12𝑖𝑥12𝑖𝑥𝑥12𝜃2e^{i\theta}=\frac{1+2ix}{1-2ix}~{}~{}~{}\Leftrightarrow~{}~{}~{}x=\frac{1}{2}% \tan\left(\frac{\theta}{2}\right),italic_e start_POSTSUPERSCRIPT italic_i italic_θ end_POSTSUPERSCRIPT = divide start_ARG 1 + 2 italic_i italic_x end_ARG start_ARG 1 - 2 italic_i italic_x end_ARG ⇔ italic_x = divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_tan ( divide start_ARG italic_θ end_ARG start_ARG 2 end_ARG ) , (9)

which defines a map from the real line x(,)𝑥x\in(-\infty,\infty)italic_x ∈ ( - ∞ , ∞ ) to the finite interval θ(π,π)𝜃𝜋𝜋\theta\in(-\pi,\pi)italic_θ ∈ ( - italic_π , italic_π ). Applying this change of variable to the MT functions (1) yields

ϕn(12tan(θ2))=2πineinθ11itan(θ2).subscriptitalic-ϕ𝑛12𝜃22𝜋superscript𝑖𝑛superscript𝑒𝑖𝑛𝜃11𝑖𝜃2\phi_{n}\left(\frac{1}{2}\tan\left(\frac{\theta}{2}\right)\right)=\sqrt{\frac{% 2}{\pi}}i^{n}e^{in\theta}\frac{1}{1-i\tan\left(\frac{\theta}{2}\right)}.italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_tan ( divide start_ARG italic_θ end_ARG start_ARG 2 end_ARG ) ) = square-root start_ARG divide start_ARG 2 end_ARG start_ARG italic_π end_ARG end_ARG italic_i start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_n italic_θ end_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 1 - italic_i roman_tan ( divide start_ARG italic_θ end_ARG start_ARG 2 end_ARG ) end_ARG .

Next, observe that

11itan(θ2)=cos(θ/2)cos(θ/2)isin(θ/2)=eiθ2cos(θ2),11𝑖𝜃2𝜃2𝜃2𝑖𝜃2superscript𝑒𝑖𝜃2𝜃2\frac{1}{1-i\tan\left(\frac{\theta}{2}\right)}=\frac{\cos(\theta/2)}{\cos(% \theta/2)-i\sin(\theta/2)}=e^{i\frac{\theta}{2}}\cos\left(\frac{\theta}{2}% \right),divide start_ARG 1 end_ARG start_ARG 1 - italic_i roman_tan ( divide start_ARG italic_θ end_ARG start_ARG 2 end_ARG ) end_ARG = divide start_ARG roman_cos ( italic_θ / 2 ) end_ARG start_ARG roman_cos ( italic_θ / 2 ) - italic_i roman_sin ( italic_θ / 2 ) end_ARG = italic_e start_POSTSUPERSCRIPT italic_i divide start_ARG italic_θ end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT roman_cos ( divide start_ARG italic_θ end_ARG start_ARG 2 end_ARG ) ,

so that

ϕn(θ)=2πineiθ(n+12)cos(θ2).subscriptitalic-ϕ𝑛𝜃2𝜋superscript𝑖𝑛superscript𝑒𝑖𝜃𝑛12𝜃2\phi_{n}(\theta)=\sqrt{\frac{2}{\pi}}i^{n}e^{i\theta\left(n+\frac{1}{2}\right)% }\cos\left(\frac{\theta}{2}\right).italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_θ ) = square-root start_ARG divide start_ARG 2 end_ARG start_ARG italic_π end_ARG end_ARG italic_i start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_θ ( italic_n + divide start_ARG 1 end_ARG start_ARG 2 end_ARG ) end_POSTSUPERSCRIPT roman_cos ( divide start_ARG italic_θ end_ARG start_ARG 2 end_ARG ) . (10)

Notice that the MT function in (10) is 2π2𝜋2\pi2 italic_π-periodic in the θ𝜃\thetaitalic_θ. Through the change of variable (9), the projection integral in (8) becomes

fˇn=ϕn(x)f(x)𝑑x=(i)n22πππf(12tan(θ2))(1itan(θ2))einθ𝑑θ.subscriptˇ𝑓𝑛superscriptsubscriptsubscriptsuperscriptitalic-ϕ𝑛𝑥𝑓𝑥differential-d𝑥superscript𝑖𝑛22𝜋superscriptsubscript𝜋𝜋𝑓12𝜃21𝑖𝜃2superscript𝑒𝑖𝑛𝜃differential-d𝜃\check{f}_{n}=\int_{-\infty}^{\infty}\phi^{*}_{n}(x)f(x)dx=\frac{(-i)^{n}}{2% \sqrt{2\pi}}\int_{-\pi}^{\pi}f\left(\frac{1}{2}\tan\left(\frac{\theta}{2}% \right)\right)\left(1-i\tan\left(\frac{\theta}{2}\right)\right)e^{-in\theta}d\theta.overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_ϕ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) italic_f ( italic_x ) italic_d italic_x = divide start_ARG ( - italic_i ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG 2 square-root start_ARG 2 italic_π end_ARG end_ARG ∫ start_POSTSUBSCRIPT - italic_π end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_π end_POSTSUPERSCRIPT italic_f ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_tan ( divide start_ARG italic_θ end_ARG start_ARG 2 end_ARG ) ) ( 1 - italic_i roman_tan ( divide start_ARG italic_θ end_ARG start_ARG 2 end_ARG ) ) italic_e start_POSTSUPERSCRIPT - italic_i italic_n italic_θ end_POSTSUPERSCRIPT italic_d italic_θ . (11)

Notice that the MT coefficients are precisely the Fourier coefficients of the (periodic) function f(12tan(θ2))(1itan(θ2))𝑓12𝜃21𝑖𝜃2f\left(\frac{1}{2}\tan\left(\frac{\theta}{2}\right)\right)\left(1-i\tan\left(% \frac{\theta}{2}\right)\right)italic_f ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_tan ( divide start_ARG italic_θ end_ARG start_ARG 2 end_ARG ) ) ( 1 - italic_i roman_tan ( divide start_ARG italic_θ end_ARG start_ARG 2 end_ARG ) ).

The MT coefficients of the function f(x)𝑓𝑥f(x)italic_f ( italic_x ) are computed via

fˇn=[f(x)]=(i)nπ2[f(12tan(θ2))(1itan(θ2))],nformulae-sequencesubscriptˇ𝑓𝑛delimited-[]𝑓𝑥superscript𝑖𝑛𝜋2delimited-[]𝑓12𝜃21𝑖𝜃2𝑛\check{f}_{n}=\mathcal{MF}[f(x)]=(-i)^{n}\sqrt{\frac{\pi}{2}}\mathcal{F}\left[% f\left(\frac{1}{2}\tan\left(\frac{\theta}{2}\right)\right)\left(1-i\tan\left(% \frac{\theta}{2}\right)\right)\right],~{}~{}~{}~{}~{}~{}n\in\mathbb{Z}overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = caligraphic_M caligraphic_F [ italic_f ( italic_x ) ] = ( - italic_i ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT square-root start_ARG divide start_ARG italic_π end_ARG start_ARG 2 end_ARG end_ARG caligraphic_F [ italic_f ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_tan ( divide start_ARG italic_θ end_ARG start_ARG 2 end_ARG ) ) ( 1 - italic_i roman_tan ( divide start_ARG italic_θ end_ARG start_ARG 2 end_ARG ) ) ] , italic_n ∈ blackboard_Z (12)

and the function is summed (see Eqs. (8) and (10)) by

f(x)=1[fˇn]=2π1[(i)nfˇn]1itan(θ2),𝑓𝑥superscript1delimited-[]subscriptˇ𝑓𝑛2𝜋superscript1delimited-[]superscript𝑖𝑛subscriptˇ𝑓𝑛1𝑖𝜃2f(x)=\mathcal{MF}^{-1}[\check{f}_{n}]=\sqrt{\frac{2}{\pi}}\frac{\mathcal{F}^{-% 1}\left[(i)^{n}\check{f}_{n}\right]}{1-i\tan\left(\frac{\theta}{2}\right)},italic_f ( italic_x ) = caligraphic_M caligraphic_F start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT [ overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ] = square-root start_ARG divide start_ARG 2 end_ARG start_ARG italic_π end_ARG end_ARG divide start_ARG caligraphic_F start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT [ ( italic_i ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ] end_ARG start_ARG 1 - italic_i roman_tan ( divide start_ARG italic_θ end_ARG start_ARG 2 end_ARG ) end_ARG , (13)

where \mathcal{F}caligraphic_F and 1superscript1\mathcal{F}^{-1}caligraphic_F start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT denote the Fourier and inverse Fourier (semi-discrete) transforms, respectively. We refer to \mathcal{MF}caligraphic_M caligraphic_F and 1superscript1\mathcal{MF}^{-1}caligraphic_M caligraphic_F start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT as the modified Fourier and inverse modified Fourier transforms, respectively, which we will utilize to compute the MT coefficients of a given function. To be clear, we compute the MT coefficients directly through Fourier transforms. Note that the divisor, 1itan(θ2)1𝑖𝜃21-i\tan\left(\frac{\theta}{2}\right)1 - italic_i roman_tan ( divide start_ARG italic_θ end_ARG start_ARG 2 end_ARG ), in (13), is included because (12) is computing the Fourier transform of the function f(12tan(θ2))(1itan(θ2))=f(x)(12ix)𝑓12𝜃21𝑖𝜃2𝑓𝑥12𝑖𝑥f\left(\frac{1}{2}\tan\left(\frac{\theta}{2}\right)\right)\left(1-i\tan\left(% \frac{\theta}{2}\right)\right)=f\left(x\right)\left(1-2ix\right)italic_f ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_tan ( divide start_ARG italic_θ end_ARG start_ARG 2 end_ARG ) ) ( 1 - italic_i roman_tan ( divide start_ARG italic_θ end_ARG start_ARG 2 end_ARG ) ) = italic_f ( italic_x ) ( 1 - 2 italic_i italic_x ), so one recovers the function f(x)𝑓𝑥f(x)italic_f ( italic_x ) by dividing out the term, 1itan(θ2)1𝑖𝜃21-i\tan\left(\frac{\theta}{2}\right)1 - italic_i roman_tan ( divide start_ARG italic_θ end_ARG start_ARG 2 end_ARG ). The discrete version of the Fourier transform are naturally computed using the fast Fourier transform (FFT) algorithm (see B).

Decacy rate of MT coefficients

In this work we focus on rational functions that are square integrable on the real line, i.e. L2()superscript𝐿2L^{2}(\mathbb{R})italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( blackboard_R ). Equivalently, we consider rational functions of the form p(x)/q(x)𝑝𝑥𝑞𝑥p(x)/q(x)italic_p ( italic_x ) / italic_q ( italic_x ), where p(x)𝑝𝑥p(x)italic_p ( italic_x ) and q(x)𝑞𝑥q(x)italic_q ( italic_x ) are polynomials of degree r𝑟ritalic_r and s𝑠sitalic_s, respectively, where r<s𝑟𝑠r<sitalic_r < italic_s. We assume q(x)𝑞𝑥q(x)italic_q ( italic_x ) has no zeros in \mathbb{R}blackboard_R. We can also consider the case r=s𝑟𝑠r=sitalic_r = italic_s when p(x)q(x)𝑝𝑥𝑞𝑥\frac{p(x)}{q(x)}divide start_ARG italic_p ( italic_x ) end_ARG start_ARG italic_q ( italic_x ) end_ARG approaches the finite value pqsubscript𝑝subscript𝑞\frac{p_{\infty}}{q_{\infty}}divide start_ARG italic_p start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT end_ARG start_ARG italic_q start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT end_ARG as x±𝑥plus-or-minusx\rightarrow\pm\inftyitalic_x → ± ∞. This latter class of rational of functions is clearly not in L2()superscript𝐿2L^{2}(\mathbb{R})italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( blackboard_R ), however the shifted function p(x)q(x)pq𝑝𝑥𝑞𝑥subscript𝑝subscript𝑞\frac{p(x)}{q(x)}-\frac{p_{\infty}}{q_{\infty}}divide start_ARG italic_p ( italic_x ) end_ARG start_ARG italic_q ( italic_x ) end_ARG - divide start_ARG italic_p start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT end_ARG start_ARG italic_q start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT end_ARG decays to zero and is in L2()superscript𝐿2L^{2}(\mathbb{R})italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( blackboard_R ). Moreover, the functions pq𝑝𝑞\frac{p}{q}divide start_ARG italic_p end_ARG start_ARG italic_q end_ARG and pq𝑝𝑞\frac{p}{q}divide start_ARG italic_p end_ARG start_ARG italic_q end_ARG - pqsubscript𝑝subscript𝑞\frac{p_{\infty}}{q_{\infty}}divide start_ARG italic_p start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT end_ARG start_ARG italic_q start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT end_ARG have the same derivatives. Hence, we can indirectly consider rational functions even when the degree of the number and denominator are equal.

One essential ingredient for this spectral method, or any method for that matter, to be effective is the coefficients (see (8)) decay rapidly. In the case of Fourier approximations on the real line, the coefficient decay rate is well-known to be related to the distance between the real line and poles [35]. We can apply this approach to this method too by locating the poles of the periodic function in (11).

Refer to caption
Figure 2: (a) Numerically computed MT coefficients for the function f(x)=(1+x2)1𝑓𝑥superscript1superscript𝑥21f(x)=(1+x^{2})^{-1}italic_f ( italic_x ) = ( 1 + italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT in blue, the exact MT coefficients (14) are in red. (b) Numerically computed MT coefficients for the function f(x)=(1+x4)1𝑓𝑥superscript1superscript𝑥41f(x)=(1+x^{4})^{-1}italic_f ( italic_x ) = ( 1 + italic_x start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT in blue, the decay rate of MT coefficients in red. Notice for the slowly decaying rational function, the coefficients decay exponentially fast.

As an example consider the function f(x)=(1+x2)1𝑓𝑥superscript1superscript𝑥21f(x)=(1+x^{2})^{-1}italic_f ( italic_x ) = ( 1 + italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT on the real line. Under transformation (9), the poles of the function are located at x=±i𝑥plus-or-minus𝑖x=\pm iitalic_x = ± italic_i or θ=±2itanh1(2)±(π+1.09861i)𝜃plus-or-minus2𝑖superscript12plus-or-minus𝜋1.09861𝑖\theta=\pm 2i\tanh^{-1}(2)\approx\pm(\pi+1.09861i)italic_θ = ± 2 italic_i roman_tanh start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( 2 ) ≈ ± ( italic_π + 1.09861 italic_i ). Since the distance from the real line to the pole is approximately 1.09861, the Fourier-MT coefficients defined in (11) are predicted to decay exponentially with the form |fˇn|e1.09861|n|similar-tosubscriptˇ𝑓𝑛superscript𝑒1.09861𝑛|\check{f}_{n}|\sim e^{-1.09861|n|}| overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT | ∼ italic_e start_POSTSUPERSCRIPT - 1.09861 | italic_n | end_POSTSUPERSCRIPT as n±𝑛plus-or-minusn\rightarrow\pm\inftyitalic_n → ± ∞. Actually, e1.09861|n|=3|n|superscript𝑒1.09861𝑛superscript3𝑛e^{-1.09861|n|}=3^{-|n|}italic_e start_POSTSUPERSCRIPT - 1.09861 | italic_n | end_POSTSUPERSCRIPT = 3 start_POSTSUPERSCRIPT - | italic_n | end_POSTSUPERSCRIPT which agrees with the analytical formula for this series

11+x2=2πn=1in3nϕn(x)+2πn=0in3n+1ϕn(x).11superscript𝑥22𝜋superscriptsubscript𝑛1superscript𝑖𝑛superscript3𝑛subscriptitalic-ϕ𝑛𝑥2𝜋superscriptsubscript𝑛0superscript𝑖𝑛superscript3𝑛1subscriptitalic-ϕ𝑛𝑥\frac{1}{1+x^{2}}=-\sqrt{2\pi}\sum_{n=-\infty}^{-1}\frac{i^{n}}{3^{-n}}\phi_{n% }(x)+\sqrt{2\pi}\sum_{n=0}^{\infty}\frac{i^{n}}{3^{n+1}}\phi_{n}(x).divide start_ARG 1 end_ARG start_ARG 1 + italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = - square-root start_ARG 2 italic_π end_ARG ∑ start_POSTSUBSCRIPT italic_n = - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT divide start_ARG italic_i start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG 3 start_POSTSUPERSCRIPT - italic_n end_POSTSUPERSCRIPT end_ARG italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) + square-root start_ARG 2 italic_π end_ARG ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_i start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG 3 start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT end_ARG italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) . (14)

Another example is the function f(x)=(1+x4)1𝑓𝑥superscript1superscript𝑥41f(x)=(1+x^{4})^{-1}italic_f ( italic_x ) = ( 1 + italic_x start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT which has poles at x=eiπ4(1+2m)𝑥superscript𝑒𝑖𝜋412𝑚x=e^{i\frac{\pi}{4}(1+2m)}italic_x = italic_e start_POSTSUPERSCRIPT italic_i divide start_ARG italic_π end_ARG start_ARG 4 end_ARG ( 1 + 2 italic_m ) end_POSTSUPERSCRIPT for m=0,1,2,3𝑚0123m=0,1,2,3italic_m = 0 , 1 , 2 , 3 or θ=2tan1(2eiπ4(1+2m))𝜃2superscript12superscript𝑒𝑖𝜋412𝑚\theta=2\tan^{-1}\left(2e^{i\frac{\pi}{4}(1+2m)}\right)italic_θ = 2 roman_tan start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( 2 italic_e start_POSTSUPERSCRIPT italic_i divide start_ARG italic_π end_ARG start_ARG 4 end_ARG ( 1 + 2 italic_m ) end_POSTSUPERSCRIPT ). Regardless of which root one takes, the distance from the real line to the imaginary part of the poles is approximately 0.641155. Similar to the previous case, the MT-Fourier coefficients decay like |fˇn|e0.641155|n|similar-tosubscriptˇ𝑓𝑛superscript𝑒0.641155𝑛|\check{f}_{n}|\sim e^{-0.641155|n|}| overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT | ∼ italic_e start_POSTSUPERSCRIPT - 0.641155 | italic_n | end_POSTSUPERSCRIPT as n±𝑛plus-or-minusn\rightarrow\pm\inftyitalic_n → ± ∞. Note we predict a slower convergence rate here.

We compare these decay rates with the numerically computed values in Fig. 2 for the functions (1+x2)1superscript1superscript𝑥21(1+x^{2})^{-1}( 1 + italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and (1+x4)1superscript1superscript𝑥41(1+x^{4})^{-1}( 1 + italic_x start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, respectively. Importantly, the MT coefficients of each function decay exponentially fast, reaching round-off error at around N=80𝑁80N=80italic_N = 80 and N=120𝑁120N=120italic_N = 120, a modest amount of grid points. The rapid decay of the MT coefficients for slowly decaying rational functions in L2()superscript𝐿2L^{2}(\mathbb{R})italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( blackboard_R ) suggest they could form the basis of an efficient spectral method. The next section considers spectral differentiation using the MT expansion. This, as with all spectral methods, benefits from exponentially decaying coefficients.

3 Spectral differentiation

Let us now explore differentiation via MT spectral methods. To begin, consider a function f(x)𝑓𝑥f(x)italic_f ( italic_x ) expressed in terms of MT functions as in (8) whose derivative is

f(x)=n=fˇnϕn(x).superscript𝑓𝑥superscriptsubscript𝑛subscriptˇ𝑓𝑛subscriptsuperscriptitalic-ϕ𝑛𝑥f^{\prime}(x)=\sum_{n=-\infty}^{\infty}\check{f}_{n}\phi^{\prime}_{n}(x).italic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_n = - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_ϕ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) .

We wish to express this derivative function in terms of MT functions itself, that is

f(x)=n=fˇnϕn(x)=n=fˇnϕn(x),superscript𝑓𝑥superscriptsubscript𝑛subscriptˇ𝑓𝑛subscriptsuperscriptitalic-ϕ𝑛𝑥superscriptsubscript𝑛subscriptsuperscriptˇ𝑓𝑛subscriptitalic-ϕ𝑛𝑥f^{\prime}(x)=\sum_{n=-\infty}^{\infty}\check{f}_{n}\phi^{\prime}_{n}(x)=\sum_% {n=-\infty}^{\infty}\check{f}^{\prime}_{n}\phi_{n}(x),italic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_n = - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_ϕ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_n = - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT overroman_ˇ start_ARG italic_f end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) ,

where fˇnsubscriptsuperscriptˇ𝑓𝑛\check{f}^{\prime}_{n}overroman_ˇ start_ARG italic_f end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT denotes the derivative function coefficients. Recall that the derivative of the MT functions satisfies the recurrence relation (6). As a result, the MT coefficients of the derivative can be expressed solely in terms of MT functions

f(x)=n=fˇnϕn(x)=n=fˇn[nϕn1(x)+i(2n+1)ϕn(x)+(n+1)ϕn+1(x)].superscript𝑓𝑥superscriptsubscript𝑛subscriptsuperscriptˇ𝑓𝑛subscriptitalic-ϕ𝑛𝑥superscriptsubscript𝑛subscriptˇ𝑓𝑛delimited-[]𝑛subscriptitalic-ϕ𝑛1𝑥𝑖2𝑛1subscriptitalic-ϕ𝑛𝑥𝑛1subscriptitalic-ϕ𝑛1𝑥f^{\prime}(x)=\sum_{n=-\infty}^{\infty}\check{f}^{\prime}_{n}\phi_{n}(x)=\sum_% {n=-\infty}^{\infty}\check{f}_{n}\left[-n\phi_{n-1}(x)+i(2n+1)\phi_{n}(x)+(n+1% )\phi_{n+1}(x)\right].italic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_n = - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT overroman_ˇ start_ARG italic_f end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_n = - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT [ - italic_n italic_ϕ start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT ( italic_x ) + italic_i ( 2 italic_n + 1 ) italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) + ( italic_n + 1 ) italic_ϕ start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT ( italic_x ) ] .

After a shift of indices the MT coefficients of the derivative are expressed in terms of the the coefficients of the function itself

f(x)=n=fˇnϕn(x)=n=(nfˇn1+i(2n+1)fˇn(n+1)fˇn+1)ϕn(x).superscript𝑓𝑥superscriptsubscript𝑛subscriptsuperscriptˇ𝑓𝑛subscriptitalic-ϕ𝑛𝑥superscriptsubscript𝑛𝑛subscriptˇ𝑓𝑛1𝑖2𝑛1subscriptˇ𝑓𝑛𝑛1subscriptˇ𝑓𝑛1subscriptitalic-ϕ𝑛𝑥f^{\prime}(x)=\sum_{n=-\infty}^{\infty}\check{f}^{\prime}_{n}\phi_{n}(x)=\sum_% {n=-\infty}^{\infty}\left(n\check{f}_{n-1}+i(2n+1)\check{f}_{n}-(n+1)\check{f}% _{n+1}\right)\phi_{n}(x).italic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_n = - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT overroman_ˇ start_ARG italic_f end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_n = - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ( italic_n overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT + italic_i ( 2 italic_n + 1 ) overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - ( italic_n + 1 ) overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT ) italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) . (15)

Notice that the MT derivative involves self {n}𝑛\{n\}{ italic_n } and nearest {n1,n+1𝑛1𝑛1n-1,n+1italic_n - 1 , italic_n + 1} modes; this is a direct consequence of recurrence relation (6). On the other hand, many spectral methods often involve completely dense interactions [40]. So, the MT differentiation matrices are relatively sparse, and as we shall see below, have the ability to converge exponentially fast for appropriate rational functions.

The derivative MT coefficients given in (15) are related to the coefficients of the original function by

fˇ=(,fˇ1,fˇ0,fˇ1,)T,fˇ=𝔻1fˇ,fˇ=(,fˇ1,fˇ0,fˇ1,)T,formulae-sequencesuperscriptˇfsuperscriptsubscriptsuperscriptˇ𝑓1subscriptsuperscriptˇ𝑓0subscriptsuperscriptˇ𝑓1𝑇formulae-sequencesuperscriptˇfsubscript𝔻1ˇfˇfsuperscriptsubscriptˇ𝑓1subscriptˇ𝑓0subscriptˇ𝑓1𝑇\check{\textbf{f}}^{\prime}=\left(\dots,\check{f}^{\prime}_{-1},\check{f}^{% \prime}_{0},\check{f}^{\prime}_{1},\dots\right)^{T},~{}~{}~{}~{}\check{\textbf% {f}}^{\prime}=\mathbb{D}_{1}\check{\textbf{f}},~{}~{}~{}~{}\check{\textbf{f}}=% \left(\dots,\check{f}_{-1},\check{f}_{0},\check{f}_{1},\dots\right)^{T},overroman_ˇ start_ARG f end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = ( … , overroman_ˇ start_ARG italic_f end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT , overroman_ˇ start_ARG italic_f end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , overroman_ˇ start_ARG italic_f end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT , overroman_ˇ start_ARG f end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = blackboard_D start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT overroman_ˇ start_ARG f end_ARG , overroman_ˇ start_ARG f end_ARG = ( … , overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT , overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT , (16)

where 𝔻1subscript𝔻1\mathbb{D}_{1}blackboard_D start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is the banded matrix

𝔻1=[35i223i11i00i113i225i3].subscript𝔻1delimited-[]missing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpression35𝑖2missing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpression23𝑖1missing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpression1𝑖0missing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpression0𝑖1missing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpression13𝑖2missing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpression25𝑖3missing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpression\mathbb{D}_{1}=\left[\begin{array}[]{*{10}c}\ddots&\ddots&&&&&&&&\\ \ddots&\ddots&\ddots&&&&&&&\\ &-3&-5i&2&&&&&&\\ &&-2&-3i&1&&&&&\\ &&&-1&-i&0&&&&\\ &&&&0&i&-1&&&\\ &&&&&1&3i&-2&&\\ &&&&&&2&5i&-3&\\ &&&&&&&\ddots&\ddots&\ddots\\ &&&&&&&&\ddots&\ddots\end{array}\right].blackboard_D start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = [ start_ARRAY start_ROW start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL - 3 end_CELL start_CELL - 5 italic_i end_CELL start_CELL 2 end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL - 2 end_CELL start_CELL - 3 italic_i end_CELL start_CELL 1 end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL - 1 end_CELL start_CELL - italic_i end_CELL start_CELL 0 end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL 0 end_CELL start_CELL italic_i end_CELL start_CELL - 1 end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL 1 end_CELL start_CELL 3 italic_i end_CELL start_CELL - 2 end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL 2 end_CELL start_CELL 5 italic_i end_CELL start_CELL - 3 end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL end_ROW end_ARRAY ] .

Notice that this matrix is tridiagonal and skew-Hermitian, that is, 𝔻1=𝔻1superscriptsubscript𝔻1subscript𝔻1\mathbb{D}_{1}^{\dagger}=-\mathbb{D}_{1}blackboard_D start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT = - blackboard_D start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, where denotes the conjugate transpose.

To approximate derivatives via MT expansions in practice, we first truncate the series in (15) to a finite range of N𝑁Nitalic_N modes: N/2,N/2+1,,N/21𝑁2𝑁21𝑁21-N/2,-N/2+1,\dots,N/2-1- italic_N / 2 , - italic_N / 2 + 1 , … , italic_N / 2 - 1, where N𝑁Nitalic_N is a positive even integer. Then one computes the MT coefficients of the function f(x)𝑓𝑥f(x)italic_f ( italic_x ) using the discrete version of (12), given in (32). Next, the truncated differentiation matrix in (16) is applied. The discrete and finite version of the series (15), given in (33), is summed to give the derivative approximation. This process is summarized in A.

Second-order differentiation matrix

The differentiation matrix for approximating the second derivative can be derived in a similar fashion. Consider a function f𝑓fitalic_f expressed in basis of MT functions, i.e. (8), whose second derivative is clearly

f′′(x)=n=fˇnϕn′′(x).superscript𝑓′′𝑥superscriptsubscript𝑛subscriptˇ𝑓𝑛subscriptsuperscriptitalic-ϕ′′𝑛𝑥f^{\prime\prime}(x)=\sum_{n=-\infty}^{\infty}\check{f}_{n}\phi^{\prime\prime}_% {n}(x).italic_f start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_n = - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_ϕ start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) .

The goal is to express this function itself in terms of MT functions, that is

f′′(x)=n=fˇnϕn′′(x)=n=fˇn′′ϕn(x),superscript𝑓′′𝑥superscriptsubscript𝑛subscriptˇ𝑓𝑛subscriptsuperscriptitalic-ϕ′′𝑛𝑥superscriptsubscript𝑛subscriptsuperscriptˇ𝑓′′𝑛subscriptitalic-ϕ𝑛𝑥f^{\prime\prime}(x)=\sum_{n=-\infty}^{\infty}\check{f}_{n}\phi^{\prime\prime}_% {n}(x)=\sum_{n=-\infty}^{\infty}\check{f}^{\prime\prime}_{n}\phi_{n}(x),italic_f start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_n = - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_ϕ start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_n = - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT overroman_ˇ start_ARG italic_f end_ARG start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) ,

where fˇn′′subscriptsuperscriptˇ𝑓′′𝑛\check{f}^{\prime\prime}_{n}overroman_ˇ start_ARG italic_f end_ARG start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT denote the second derivative coefficients, which we intend to find. Differentiating the first derivative relation in (6) yields the five-term second derivative relation

d2dx2ϕn(x)superscript𝑑2𝑑superscript𝑥2subscriptitalic-ϕ𝑛𝑥\displaystyle\frac{d^{2}}{dx^{2}}\phi_{n}(x)divide start_ARG italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_d italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) =nddxϕn1(x)+i(2n+1)ddxϕn(x)+(n+1)ddxϕn+1(x)absent𝑛𝑑𝑑𝑥subscriptitalic-ϕ𝑛1𝑥𝑖2𝑛1𝑑𝑑𝑥subscriptitalic-ϕ𝑛𝑥𝑛1𝑑𝑑𝑥subscriptitalic-ϕ𝑛1𝑥\displaystyle=-n\frac{d}{dx}\phi_{n-1}(x)+i(2n+1)\frac{d}{dx}\phi_{n}(x)+(n+1)% \frac{d}{dx}\phi_{n+1}(x)= - italic_n divide start_ARG italic_d end_ARG start_ARG italic_d italic_x end_ARG italic_ϕ start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT ( italic_x ) + italic_i ( 2 italic_n + 1 ) divide start_ARG italic_d end_ARG start_ARG italic_d italic_x end_ARG italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) + ( italic_n + 1 ) divide start_ARG italic_d end_ARG start_ARG italic_d italic_x end_ARG italic_ϕ start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT ( italic_x ) (17)
=n(n1)ϕn2(x)4in2ϕn1(x)(2+6n(n+1))ϕn(x)+4i(n+1)2ϕn+1(x)+(n+2)(n+1)ϕn+2(x).absent𝑛𝑛1subscriptitalic-ϕ𝑛2𝑥4𝑖superscript𝑛2subscriptitalic-ϕ𝑛1𝑥26𝑛𝑛1subscriptitalic-ϕ𝑛𝑥4𝑖superscript𝑛12subscriptitalic-ϕ𝑛1𝑥𝑛2𝑛1subscriptitalic-ϕ𝑛2𝑥\displaystyle=n(n-1)\phi_{n-2}(x)-4in^{2}\phi_{n-1}(x)-(2+6n(n+1))\phi_{n}(x)+% 4i(n+1)^{2}\phi_{n+1}(x)+(n+2)(n+1)\phi_{n+2}(x).= italic_n ( italic_n - 1 ) italic_ϕ start_POSTSUBSCRIPT italic_n - 2 end_POSTSUBSCRIPT ( italic_x ) - 4 italic_i italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT ( italic_x ) - ( 2 + 6 italic_n ( italic_n + 1 ) ) italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) + 4 italic_i ( italic_n + 1 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT ( italic_x ) + ( italic_n + 2 ) ( italic_n + 1 ) italic_ϕ start_POSTSUBSCRIPT italic_n + 2 end_POSTSUBSCRIPT ( italic_x ) .

Note that this involves five nodal values, {n2,n1,n,n+1,n+2}𝑛2𝑛1𝑛𝑛1𝑛2\{n-2,n-1,n,n+1,n+2\}{ italic_n - 2 , italic_n - 1 , italic_n , italic_n + 1 , italic_n + 2 }. Similar to first-order derivative, the second derivative coefficients are obtained by exploiting the orthogonality of the functions and shifting the indices of the series and thus the MT coefficients of the second derivative can be represented as

n=fˇn′′ϕn(x)=n=[n(n1)fˇn2+4in2fˇn1(2+6n(n+1))fˇn4i(n+1)2fˇn+1+(n+2)(n+1)fˇn+2]ϕn(x).superscriptsubscript𝑛subscriptsuperscriptˇ𝑓′′𝑛subscriptitalic-ϕ𝑛𝑥superscriptsubscript𝑛delimited-[]𝑛𝑛1subscriptˇ𝑓𝑛24𝑖superscript𝑛2subscriptˇ𝑓𝑛126𝑛𝑛1subscriptˇ𝑓𝑛4𝑖superscript𝑛12subscriptˇ𝑓𝑛1𝑛2𝑛1subscriptˇ𝑓𝑛2subscriptitalic-ϕ𝑛𝑥\sum_{n=-\infty}^{\infty}\check{f}^{\prime\prime}_{n}\phi_{n}(x)=\sum_{n=-% \infty}^{\infty}\left[n(n-1)\check{f}_{n-2}+4in^{2}\check{f}_{n-1}-(2+6n(n+1))% \check{f}_{n}-4i(n+1)^{2}\check{f}_{n+1}+(n+2)(n+1)\check{f}_{n+2}\right]\phi_% {n}(x).∑ start_POSTSUBSCRIPT italic_n = - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT overroman_ˇ start_ARG italic_f end_ARG start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) = ∑ start_POSTSUBSCRIPT italic_n = - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT [ italic_n ( italic_n - 1 ) overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_n - 2 end_POSTSUBSCRIPT + 4 italic_i italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT - ( 2 + 6 italic_n ( italic_n + 1 ) ) overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - 4 italic_i ( italic_n + 1 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT + ( italic_n + 2 ) ( italic_n + 1 ) overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_n + 2 end_POSTSUBSCRIPT ] italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) . (18)

Using this information, the MT coefficients of the second derivative can be computed via

fˇ′′=(,fˇ1′′,fˇ0′′,fˇ1′′,)T,fˇ′′=𝔻2fˇ,fˇ=(,fˇ1,fˇ0,fˇ1,)T,formulae-sequencesuperscriptˇf′′superscriptsubscriptsuperscriptˇ𝑓′′1subscriptsuperscriptˇ𝑓′′0subscriptsuperscriptˇ𝑓′′1𝑇formulae-sequencesuperscriptˇf′′subscript𝔻2ˇfˇfsuperscriptsubscriptˇ𝑓1subscriptˇ𝑓0subscriptˇ𝑓1𝑇\check{\textbf{f}}^{\prime\prime}=\left(\dots,\check{f}^{\prime\prime}_{-1},% \check{f}^{\prime\prime}_{0},\check{f}^{\prime\prime}_{1},\dots\right)^{T},~{}% ~{}~{}~{}\check{\textbf{f}}^{\prime\prime}=\mathbb{D}_{2}\check{\textbf{f}},~{% }~{}~{}~{}\check{\textbf{f}}=\left(\dots,\check{f}_{-1},\check{f}_{0},\check{f% }_{1},\dots\right)^{T},overroman_ˇ start_ARG f end_ARG start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT = ( … , overroman_ˇ start_ARG italic_f end_ARG start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT , overroman_ˇ start_ARG italic_f end_ARG start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , overroman_ˇ start_ARG italic_f end_ARG start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT , overroman_ˇ start_ARG f end_ARG start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT = blackboard_D start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT overroman_ˇ start_ARG f end_ARG , overroman_ˇ start_ARG f end_ARG = ( … , overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT , overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT , (19)

where 𝔻2subscript𝔻2\mathbb{D}_{2}blackboard_D start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT is the banded matrix

𝔻2=[30100i12264i122064i7436i61236i3816i2616i144i024i2000024i204i1416i6216i3836i12636i7464i201264i122100i30].subscript𝔻2delimited-[]missing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpression30100𝑖12264𝑖12missing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpression2064𝑖7436𝑖6missing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpression1236𝑖3816𝑖2missing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpression616𝑖144𝑖0missing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpression24𝑖200missing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpression0024𝑖2missing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpression04𝑖1416𝑖6missing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpression216𝑖3836𝑖12missing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpression636𝑖7464𝑖20missing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpression1264𝑖122100𝑖30missing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpressionmissing-subexpression\mathbb{D}_{2}=\left[\begin{array}[]{*{16}c}\ddots&\ddots&\ddots&&&&&&&&&&&&&% \\ \ddots&\ddots&\ddots&\ddots&&&&&&&&&&&&\\ \ddots&\ddots&\ddots&\ddots&\ddots&&&&&&&&&&&\\ &30&100i&-122&64i&12&&&&&&&&&&\\ &&20&64i&-74&36i&6&&&&&&&&&\\ &&&12&36i&-38&16i&2&&&&&&&&\\ &&&&6&16i&-14&-4i&0&&&&&&&\\ &&&&&2&4i&-2&0&0&&&&&&\\ &&&&&&0&0&-2&-4i&2&&&&&\\ &&&&&&&0&4i&-14&-16i&6&&&&\\ &&&&&&&&2&16i&-38&-36i&12&&&\\ &&&&&&&&&6&36i&-74&-64i&20&&\\ &&&&&&&&&&12&64i&-122&-100i&30&\\ &&&&&&&&&&&\ddots&\ddots&\ddots&\ddots&\ddots\\ &&&&&&&&&&&&\ddots&\ddots&\ddots&\ddots\\ &&&&&&&&&&&&&\ddots&\ddots&\ddots\end{array}\right].blackboard_D start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = [ start_ARRAY start_ROW start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL 30 end_CELL start_CELL 100 italic_i end_CELL start_CELL - 122 end_CELL start_CELL 64 italic_i end_CELL start_CELL 12 end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL 20 end_CELL start_CELL 64 italic_i end_CELL start_CELL - 74 end_CELL start_CELL 36 italic_i end_CELL start_CELL 6 end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL 12 end_CELL start_CELL 36 italic_i end_CELL start_CELL - 38 end_CELL start_CELL 16 italic_i end_CELL start_CELL 2 end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL 6 end_CELL start_CELL 16 italic_i end_CELL start_CELL - 14 end_CELL start_CELL - 4 italic_i end_CELL start_CELL 0 end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL 2 end_CELL start_CELL 4 italic_i end_CELL start_CELL - 2 end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL - 2 end_CELL start_CELL - 4 italic_i end_CELL start_CELL 2 end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL 0 end_CELL start_CELL 4 italic_i end_CELL start_CELL - 14 end_CELL start_CELL - 16 italic_i end_CELL start_CELL 6 end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL 2 end_CELL start_CELL 16 italic_i end_CELL start_CELL - 38 end_CELL start_CELL - 36 italic_i end_CELL start_CELL 12 end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL 6 end_CELL start_CELL 36 italic_i end_CELL start_CELL - 74 end_CELL start_CELL - 64 italic_i end_CELL start_CELL 20 end_CELL start_CELL end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL 12 end_CELL start_CELL 64 italic_i end_CELL start_CELL - 122 end_CELL start_CELL - 100 italic_i end_CELL start_CELL 30 end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL end_CELL start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL start_CELL ⋱ end_CELL end_ROW end_ARRAY ] .

Notice that this matrix is pentadiagonal and Hermitian (𝔻2=𝔻2superscriptsubscript𝔻2subscript𝔻2\mathbb{D}_{2}^{\dagger}=\mathbb{D}_{2}blackboard_D start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT = blackboard_D start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT). Similar to 𝔻1subscript𝔻1\mathbb{D}_{1}blackboard_D start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, this differentiation matrix is banded and fairly sparse; this is a general theme. Based on recurrence relation (6), increasing the order of the derivative increases the bandwidth of the differentiation matrix by one super and one sub diagonal. To implement this in practice, the system in (19) is truncated to some sufficiently large set of modes and the truncated version of 𝔻2subscript𝔻2\mathbb{D}_{2}blackboard_D start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT is applied (see A).

A natural question to ask is whether to approximate x2superscriptsubscript𝑥2\partial_{x}^{2}∂ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT by 𝔻12=𝔻1𝔻1superscriptsubscript𝔻12subscript𝔻1subscript𝔻1\mathbb{D}_{1}^{2}=\mathbb{D}_{1}\mathbb{D}_{1}blackboard_D start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = blackboard_D start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT blackboard_D start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT (two applications of the first derivative matrix (16)), rather than 𝔻2subscript𝔻2\mathbb{D}_{2}blackboard_D start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. When the system is infinite (no truncation), 𝔻2=𝔻12subscript𝔻2superscriptsubscript𝔻12\mathbb{D}_{2}=\mathbb{D}_{1}^{2}blackboard_D start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = blackboard_D start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. However, upon truncation the matrices differ at the first and last modes due to different boundary conditions. In our experience, approximation by 𝔻2subscript𝔻2\mathbb{D}_{2}blackboard_D start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT is slightly more accurate (by about a factor of 2), but one can use 𝔻12superscriptsubscript𝔻12\mathbb{D}_{1}^{2}blackboard_D start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and still expect spectral convergence rates.

Refer to caption
Figure 3: Infinity norm error of the (a) first and (b) second MT derivative approximations applied to the rational function f(x)=(1+x2)1𝑓𝑥superscript1superscript𝑥21f(x)=(1+x^{2})^{-1}italic_f ( italic_x ) = ( 1 + italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT as a function of the number of MT modes, N𝑁Nitalic_N. The error is observed to converge exponentially fast. All derivatives were computed using

A typical set of convergence results is highlighted in Fig. 3. For the rational function f(x)=1/(1+x2)𝑓𝑥11superscript𝑥2f(x)=1/(1+x^{2})italic_f ( italic_x ) = 1 / ( 1 + italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ), the MT approximation of derivatives is remarkably effective. The error converges exponentially fast to zero until we reach round-off error effects at around N=64𝑁64N=64italic_N = 64 modes. We notice with each order of derivative, the error increases by about one order of magnitude. These rapid convergence results are expected based on the exponentially decay of the MT coefficients shown in Fig. 2. Given that we are using fast modified Fourier transforms in (12) and (13), these calculations are performed quickly.

4 Numerical approximation of PDEs

In this section we develop a method for numerically approximating rational solutions of nonlinear PDEs. An explicit split-step time-stepping approach is implemented using the differentiation matrices described in Sec. 3. These techniques will be used to approximate rational solutions of a well-known PDE, the nonlinear Schrödinger (NLS) equation.

To begin, consider the general nonlinear initial boundary value problem

ut=u+[u],u0(x)=u(x,t0),(x,t)×[t0,T],formulae-sequencesubscript𝑢𝑡𝑢delimited-[]𝑢formulae-sequencesubscript𝑢0𝑥𝑢𝑥subscript𝑡0𝑥𝑡subscript𝑡0𝑇u_{t}=\mathcal{L}u+\mathcal{M}[u],~{}~{}~{}~{}~{}u_{0}(x)=u(x,t_{0}),~{}~{}~{}% ~{}~{}(x,t)\in\mathbb{R}\times[t_{0},T],italic_u start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = caligraphic_L italic_u + caligraphic_M [ italic_u ] , italic_u start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x ) = italic_u ( italic_x , italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) , ( italic_x , italic_t ) ∈ blackboard_R × [ italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_T ] , (20)

where the linear and nonlinear operators are denoted by \mathcal{L}caligraphic_L and \mathcal{M}caligraphic_M, respectively. The linear operator is assumed to be a constant coefficient differential operator with the form

=c0+c1x+c22x2+,subscript𝑐0subscript𝑐1𝑥subscript𝑐2superscript2superscript𝑥2\mathcal{L}=c_{0}+c_{1}\frac{\partial}{\partial x}+c_{2}\frac{\partial^{2}}{% \partial x^{2}}+\dots~{},caligraphic_L = italic_c start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT divide start_ARG ∂ end_ARG start_ARG ∂ italic_x end_ARG + italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ∂ italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + … ,

for scalar coefficients c0,c1,subscript𝑐0subscript𝑐1c_{0},c_{1},\dotsitalic_c start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … The nonlinear operator is assumed to be local and depend on some product of u𝑢uitalic_u and it’s derivatives. If both operators in (20) are linear ([u]=Mudelimited-[]𝑢𝑀𝑢\mathcal{M}[u]=Mucaligraphic_M [ italic_u ] = italic_M italic_u) and constant coefficient, then the solution is

u(x,t)=exp[(tt0)(+)]u0(x).𝑢𝑥𝑡𝑡subscript𝑡0subscript𝑢0𝑥u(x,t)=\exp\left[(t-t_{0})(\mathcal{L}+\mathcal{M})\right]u_{0}(x).italic_u ( italic_x , italic_t ) = roman_exp [ ( italic_t - italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ( caligraphic_L + caligraphic_M ) ] italic_u start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x ) . (21)

We consider solutions that approach a constant value at the boundaries, i.e. u(x,t)Ub𝑢𝑥𝑡subscript𝑈𝑏u(x,t)\rightarrow U_{b}italic_u ( italic_x , italic_t ) → italic_U start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT as x±𝑥plus-or-minusx\rightarrow\pm\inftyitalic_x → ± ∞. Ultimately we have in mind rational solutions where the degree of the denominator polynomial is greater than or equal to the degree of the numerator polynomial.

4.1 MT-SS4 approximation of the NLS equation

We now introduce a fourth-order Malmquist-Takenaka split-step (MT-SS4) method for integrating nonlinear PDEs. The idea here is to split the linear and nonlinear parts of (20) into two parts that are easier to solve separately. This method has the benefit of being quite accurate and straightforward to implement while also possessing a relatively large stability region. A thorough derivation of the split-step and its properties can be found in [20].

Consider splitting (20) into two equations

𝐰t=𝐰,𝐯t=[𝐯].formulae-sequence𝐰𝑡𝐰𝐯𝑡delimited-[]𝐯\frac{\partial{\bf w}}{\partial t}=\mathcal{L}{\bf w},~{}~{}~{}~{}~{}~{}~{}~{}% \frac{\partial{\bf v}}{\partial t}=\mathcal{M}[{\bf v}].divide start_ARG ∂ bold_w end_ARG start_ARG ∂ italic_t end_ARG = caligraphic_L bold_w , divide start_ARG ∂ bold_v end_ARG start_ARG ∂ italic_t end_ARG = caligraphic_M [ bold_v ] . (22)

The idea behind the method is to solve the equations in (22) separately and then re-combine them in a specific order to approximate the solution. For example, the first-order approximation of (21) for t0=0subscript𝑡00t_{0}=0italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0 is exp(t(+))exp(t)exp(t).𝑡𝑡𝑡\exp(t(\mathcal{L}+\mathcal{M}))\approx\exp(t\mathcal{L})\cdot\exp(t\mathcal{M% }).roman_exp ( italic_t ( caligraphic_L + caligraphic_M ) ) ≈ roman_exp ( italic_t caligraphic_L ) ⋅ roman_exp ( italic_t caligraphic_M ) . Only when \mathcal{L}caligraphic_L and \mathcal{M}caligraphic_M commute does this become an equality. That is, solve one equation (22) and then use its solutions as initial condition for the other equation. The nonlinear version operates in a similar manner [20].

The fourth-order split-step method for one time-step ΔtΔ𝑡\Delta troman_Δ italic_t is given by [20, 42]

𝐮(t+Δt)exp[α4Δt]exp[β3Δt]exp[α3Δt]exp[β2Δt]exp[α2Δt]exp[β1Δt]exp[α1Δt]𝐮(t),𝐮𝑡Δ𝑡subscript𝛼4Δ𝑡subscript𝛽3Δ𝑡subscript𝛼3Δ𝑡subscript𝛽2Δ𝑡subscript𝛼2Δ𝑡subscript𝛽1Δ𝑡subscript𝛼1Δ𝑡𝐮𝑡{\bf u}(t+\Delta t)\approx\exp\left[\alpha_{4}\Delta t\mathcal{M}\right]\exp% \left[\beta_{3}\Delta t\mathcal{L}\right]\exp\left[\alpha_{3}\Delta t\mathcal{% M}\right]\exp\left[\beta_{2}\Delta t\mathcal{L}\right]\exp\left[\alpha_{2}% \Delta t\mathcal{M}\right]\exp\left[\beta_{1}\Delta t\mathcal{L}\right]\exp% \left[\alpha_{1}\Delta t\mathcal{M}\right]{\bf u}(t),bold_u ( italic_t + roman_Δ italic_t ) ≈ roman_exp [ italic_α start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT roman_Δ italic_t caligraphic_M ] roman_exp [ italic_β start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT roman_Δ italic_t caligraphic_L ] roman_exp [ italic_α start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT roman_Δ italic_t caligraphic_M ] roman_exp [ italic_β start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_Δ italic_t caligraphic_L ] roman_exp [ italic_α start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_Δ italic_t caligraphic_M ] roman_exp [ italic_β start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT roman_Δ italic_t caligraphic_L ] roman_exp [ italic_α start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT roman_Δ italic_t caligraphic_M ] bold_u ( italic_t ) , (23)

in terms of the coefficients

α1=12c,α2=12(1c),α3=α2,α4=α1,β1=c,β2=12c,β3=β1,formulae-sequencesubscript𝛼112𝑐formulae-sequencesubscript𝛼2121𝑐formulae-sequencesubscript𝛼3subscript𝛼2formulae-sequencesubscript𝛼4subscript𝛼1formulae-sequencesubscript𝛽1𝑐formulae-sequencesubscript𝛽212𝑐subscript𝛽3subscript𝛽1\alpha_{1}=\frac{1}{2}c,\hskip 3.61371pt\alpha_{2}=\frac{1}{2}(1-c),\hskip 3.6% 1371pt\alpha_{3}=\alpha_{2},\hskip 3.61371pt\alpha_{4}=\alpha_{1},\hskip 3.613% 71pt\beta_{1}=c,\hskip 3.61371pt\beta_{2}=1-2c,\hskip 3.61371pt\beta_{3}=\beta% _{1},italic_α start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_c , italic_α start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( 1 - italic_c ) , italic_α start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = italic_α start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_α start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT = italic_α start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_β start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_c , italic_β start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 1 - 2 italic_c , italic_β start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = italic_β start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ,

where c=1221/3𝑐12superscript213c=\frac{1}{2-2^{1/3}}italic_c = divide start_ARG 1 end_ARG start_ARG 2 - 2 start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT end_ARG. More specifically, starting from the solution at time t𝑡titalic_t, 𝐮(t)𝐮𝑡{\bf u}(t)bold_u ( italic_t ), we first integrate the \mathcal{M}caligraphic_M-operator equation (22) by a step α1Δtsubscript𝛼1Δ𝑡\alpha_{1}\Delta titalic_α start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT roman_Δ italic_t to get an intermediate solution 𝐯1subscript𝐯1{\bf v}_{1}bold_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. Then we integrate the \mathcal{L}caligraphic_L-operator equation (22) by a step β1Δtsubscript𝛽1Δ𝑡\beta_{1}\Delta titalic_β start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT roman_Δ italic_t, using 𝐯1subscript𝐯1{\bf v}_{1}bold_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT as the initial condition to get another intermediate solution 𝐰1subscript𝐰1{\bf w}_{1}bold_w start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. This process repeats itself until the final step in (23).

Note that the overall accuracy of the method does not depend on the choice of which operator is first and second. That is, the fourth-order method in (23) starts and ends with \mathcal{M}caligraphic_M, but it is equally valid to begin and end with \mathcal{L}caligraphic_L, and the order of the method will not be affected. Due to its simplicity, for the NLS equation below we choose to solve the nonlinear part of four times (αnsubscript𝛼𝑛\alpha_{n}italic_α start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT coefficients) and the linear part three times (βnsubscript𝛽𝑛\beta_{n}italic_β start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT coefficients). Whenever possible the exponential operators in (23) should be pre-computed.

From a practical point-of-view, this is an effective technique only when the equations in (22) are solved efficiently and accurately. Typically, the linear part of equation in (22) can be rapidly solved using some spectral approximation, e.g. Fourier or MT. The crucial step is the nonlinear one. For an equation like NLS, this is a nearly trivial solve (see next section) [38]. However, for other equations, like those in the Korteweg-de Vries family, solving this is not so straightforward and this method is not as attractive [39].

Refer to caption
Figure 4: A top (left) and side (right) view of the modulus using MT-SS4 method for the NLS equation (24) and Peregrine initial condition (26) with t0=2subscript𝑡02t_{0}=-2italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = - 2. These results were obtained using N=256𝑁256N=256italic_N = 256 MT modes and a time-step of Δt=0.0001Δ𝑡0.0001\Delta t=0.0001roman_Δ italic_t = 0.0001.

Consider the nonlinear Schrödinger equation

iut+uxx+2(|u|21)u=0,u0(x)=u(x,t0),(x,t)×[t0,T],formulae-sequence𝑖subscript𝑢𝑡subscript𝑢𝑥𝑥2superscript𝑢21𝑢0formulae-sequencesubscript𝑢0𝑥𝑢𝑥subscript𝑡0𝑥𝑡subscript𝑡0𝑇iu_{t}+u_{xx}+2(|u|^{2}-1)u=0,~{}~{}~{}~{}~{}u_{0}(x)=u(x,t_{0}),~{}~{}~{}~{}(% x,t)\in\mathbb{R}\times[t_{0},T],italic_i italic_u start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + italic_u start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT + 2 ( | italic_u | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1 ) italic_u = 0 , italic_u start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_x ) = italic_u ( italic_x , italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) , ( italic_x , italic_t ) ∈ blackboard_R × [ italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_T ] , (24)

with boundary conditions u1𝑢1u\to 1italic_u → 1 as x±𝑥plus-or-minusx\to\pm\inftyitalic_x → ± ∞. Note that through the phase transformation u(x,t)=ψ(x,t)e2it𝑢𝑥𝑡𝜓𝑥𝑡superscript𝑒2𝑖𝑡u(x,t)=\psi(x,t)e^{-2it}italic_u ( italic_x , italic_t ) = italic_ψ ( italic_x , italic_t ) italic_e start_POSTSUPERSCRIPT - 2 italic_i italic_t end_POSTSUPERSCRIPT this equation can be recast in the more standard form

iψt+ψxx+2|ψ|2ψ=0.𝑖subscript𝜓𝑡subscript𝜓𝑥𝑥2superscript𝜓2𝜓0i\psi_{t}+\psi_{xx}+2|\psi|^{2}\psi=0.italic_i italic_ψ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + italic_ψ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT + 2 | italic_ψ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ψ = 0 .

The form in (24) is more convenient for our purposes. Two well-known solutions of the NLS equaiton are the plane wave solution

uc(x,t)=1,subscript𝑢𝑐𝑥𝑡1u_{c}(x,t)=1,italic_u start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_x , italic_t ) = 1 , (25)

and Peregrine soliton solution

up(x,t)=4x216it+16t234x2+16t2+1,subscript𝑢𝑝𝑥𝑡4superscript𝑥216𝑖𝑡16superscript𝑡234superscript𝑥216superscript𝑡21u_{p}(x,t)=\frac{4x^{2}-16it+16t^{2}-3}{4x^{2}+16t^{2}+1},italic_u start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ( italic_x , italic_t ) = divide start_ARG 4 italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 16 italic_i italic_t + 16 italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 3 end_ARG start_ARG 4 italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 16 italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 1 end_ARG , (26)

both of which are rational functions with unity boundary conditions. Physically, (25) corresponds to a nonlinear periodic wave train. The Peregrine solution in (26) is a model for rogue waves; a depiction of it’s dynamics is shown in Fig. 4. Unlike most solitons that maintain their profile during propagation, the Peregrine soliton is localized in both space and time.

Consider solving the NLS equation (24) solution via split-step method (23) where =ix2𝑖superscriptsubscript𝑥2\mathcal{L}=i\partial_{x}^{2}caligraphic_L = italic_i ∂ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and [u]=2i(|u|21)delimited-[]𝑢2𝑖superscript𝑢21\mathcal{M}[u]=2i\left(|u|^{2}-1\right)caligraphic_M [ italic_u ] = 2 italic_i ( | italic_u | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1 ) in (20). Explicitly, the split-step equations in (22) are

wt=i2wx2,vt=2i(|v|21)v.formulae-sequence𝑤𝑡𝑖superscript2𝑤superscript𝑥2𝑣𝑡2𝑖superscript𝑣21𝑣\frac{\partial w}{\partial t}=i\frac{\partial^{2}w}{\partial x^{2}},~{}~{}~{}~% {}~{}~{}~{}~{}\frac{\partial v}{\partial t}=2i\left(|v|^{2}-1\right)v.divide start_ARG ∂ italic_w end_ARG start_ARG ∂ italic_t end_ARG = italic_i divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_w end_ARG start_ARG ∂ italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , divide start_ARG ∂ italic_v end_ARG start_ARG ∂ italic_t end_ARG = 2 italic_i ( | italic_v | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1 ) italic_v . (27)

The second equation in (27) is clearly solved by

v(x,t0+Δt)=e2iΔt(|v(x,t0)|21)v(x,t0).𝑣𝑥subscript𝑡0Δ𝑡superscript𝑒2𝑖Δ𝑡superscript𝑣𝑥subscript𝑡021𝑣𝑥subscript𝑡0v(x,t_{0}+\Delta t)=e^{2i\Delta t\left(|v(x,t_{0})|^{2}-1\right)}v(x,t_{0}).italic_v ( italic_x , italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + roman_Δ italic_t ) = italic_e start_POSTSUPERSCRIPT 2 italic_i roman_Δ italic_t ( | italic_v ( italic_x , italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1 ) end_POSTSUPERSCRIPT italic_v ( italic_x , italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) . (28)

To numerically solve the first equation in (27) we wish to approximate spatial derivative using an MT spectral approach. Normally the first step is to expand

Refer to caption
Figure 5: Top (left) and side (right) views of the solution modulus using MT-SS4 method. All parameters are the same as Fig. 4 except the time-step Δt=0.1Δ𝑡0.1\Delta t=0.1roman_Δ italic_t = 0.1. The large time-step reveals an instability in the approximation for t0𝑡0t\geq 0italic_t ≥ 0.

the solution w(x,t)𝑤𝑥𝑡w(x,t)italic_w ( italic_x , italic_t ) in basis of MT functions (see (8)). However, the solution w(x,t)1𝑤𝑥𝑡1w(x,t)\rightarrow 1italic_w ( italic_x , italic_t ) → 1 as |x|𝑥|x|\rightarrow\infty| italic_x | → ∞ and is not in L2()superscript𝐿2L^{2}(\mathbb{R})italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( blackboard_R ); an expansion in terms of the MT functions requires functions decay to zero at infinity. As a result, the functions in (25) and (26) cannot directly be expressed in terms of a MT expansion.

To get around this barrier, we first observe the trivial calculus fact

dfdx=ddx(f1),𝑑𝑓𝑑𝑥𝑑𝑑𝑥𝑓1\frac{df}{dx}=\frac{d}{dx}(f-1),divide start_ARG italic_d italic_f end_ARG start_ARG italic_d italic_x end_ARG = divide start_ARG italic_d end_ARG start_ARG italic_d italic_x end_ARG ( italic_f - 1 ) ,

for some differential function f(x)𝑓𝑥f(x)italic_f ( italic_x ) such that f(x)1𝑓𝑥1f(x)\rightarrow 1italic_f ( italic_x ) → 1 as |x|𝑥|x|\rightarrow\infty| italic_x | → ∞. Crucially, even though both sides of the equation have the same derivatives, the function on the right-hand side has zero boundary conditions and can be directly expanded in a basis of MT functions. That is, we make the simple change of variables w~(x,t)=w(x,t)1~𝑤𝑥𝑡𝑤𝑥𝑡1\widetilde{w}(x,t)=w(x,t)-1over~ start_ARG italic_w end_ARG ( italic_x , italic_t ) = italic_w ( italic_x , italic_t ) - 1 to the linear (first) equation in (27), solve, and then return the unit boundary conditions. In this way we can approximate derivatives of functions u(x)𝑢𝑥u(x)italic_u ( italic_x ) with constant and nonzero boundary conditions, without resorting to a reformulation of the problem.

After shifting by one, we expand w~(x,t)~𝑤𝑥𝑡\widetilde{w}(x,t)over~ start_ARG italic_w end_ARG ( italic_x , italic_t ) as (8), and apply the differentiation matrix given in (19). The system is truncated to N𝑁Nitalic_N MT modes and the linear equation is reformulated as

d𝐰ˇdt=i𝔻2𝐰ˇ,𝑑ˇ𝐰𝑑𝑡𝑖subscript𝔻2ˇ𝐰\frac{d\check{\bf{w}}}{dt}=i\mathbb{D}_{2}{\check{\bf w}},divide start_ARG italic_d overroman_ˇ start_ARG bold_w end_ARG end_ARG start_ARG italic_d italic_t end_ARG = italic_i blackboard_D start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT overroman_ˇ start_ARG bold_w end_ARG ,

where 𝔻2subscript𝔻2\mathbb{D}_{2}blackboard_D start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT is the truncated N×N𝑁𝑁N\times Nitalic_N × italic_N differentiation matrix given in (19) and 𝐰ˇˇ𝐰\check{{\bf w}}overroman_ˇ start_ARG bold_w end_ARG consists of the MT coefficients associated with the function w~(x,t)~𝑤𝑥𝑡\widetilde{w}(x,t)over~ start_ARG italic_w end_ARG ( italic_x , italic_t ). This is a finite system of ODEs and can be solved exactly by

𝐰ˇ(t0+Δt)=eiΔt𝔻2𝐰ˇ(t0).ˇ𝐰subscript𝑡0Δ𝑡superscript𝑒𝑖Δ𝑡subscript𝔻2ˇ𝐰subscript𝑡0\check{{\bf w}}(t_{0}+\Delta t)=e^{i\Delta t\mathbb{D}_{2}}\check{{\bf w}}(t_{% 0}).overroman_ˇ start_ARG bold_w end_ARG ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + roman_Δ italic_t ) = italic_e start_POSTSUPERSCRIPT italic_i roman_Δ italic_t blackboard_D start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT overroman_ˇ start_ARG bold_w end_ARG ( italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) . (29)

Afterward, the solution in physical space is recovered by computing a truncated sum from (8) (see discrete version in (33)).

Refer to caption
Figure 6: (a) Convergence of the MT-SS4 approximation for the Peregrine soliton (26) solved on the interval 2t22𝑡2-2\leq t\leq 2- 2 ≤ italic_t ≤ 2. The error at time t=2𝑡2t=2italic_t = 2 for N=256𝑁256N=256italic_N = 256 MT nodes is shown. The numerical approximation apparently converges at a fourth-order rate when time-step ΔtΔ𝑡\Delta troman_Δ italic_t is decreased, until other errors become significant. (b) Average CPU runtime as a function of ΔtΔ𝑡\Delta troman_Δ italic_t and fixed N𝑁Nitalic_N. The runtime clearly increases linearly with ΔtΔ𝑡\Delta troman_Δ italic_t.

4.2 A convergence study

Typical results obtained by solving the NLS equation (24) with the MT-SS4 method for the Pergrine soliton solution (26) are discussed next. The evolution of a typical Pergrine soliton obtained using this method is shown in Fig. 4 over the time interval [2,2]22[-2,2][ - 2 , 2 ]. Notice that the soliton peaks at the origin (x,t)=(0,0)𝑥𝑡00(x,t)=(0,0)( italic_x , italic_t ) = ( 0 , 0 ).

A benefit of the split-step method is the relatively large stability region. That is, the method yields stable results for quite large values of ΔtΔ𝑡\Delta troman_Δ italic_t.

A numerically induced stability is shown in Fig. 5. This is the result of taking an apparently too large time-step. The derivation of a rigorous stability bound for the MT-SS4 method is an open problem.

Refer to caption
Figure 7: Final time snapshot of the real (a) and imaginary (b) parts of the MT-SS4 approximation for the Peregrine soliton shown in Fig. 4; computational parameters are the same. (c) Absolute difference between MT-SS4 approximation and exact solution at t=2𝑡2t=2italic_t = 2.

Next, the rate of convergence and runtime as a function of ΔtΔ𝑡\Delta troman_Δ italic_t is discussed. In Fig. 6(a) the convergence rate of the MT-SS4 method applied to the NLS equation (24) for the Peregrine solution (26) is shown. A clear fourth-order convergence rate is observed until other errors become significant. We have observed some apparent aliasing errors for certain ΔtΔ𝑡\Delta troman_Δ italic_t values. These anomalies and a padding approach to mitigate them is discussed in Sec. 4.3.

The corresponding runtimes are shown in Fig. 6(b). This timer encapsulates only the split-step solver portion of the code. We observe a roughly linear dependence for this range of computational parameters. For all runs considered in Fig. 6, the total CPU time ranged from 0.5-10 seconds. All runs were performed on a HP Laptop personal computer with an 1.60 GHz 4 Core Intel processor with 8.00 GB of Ram and 15.7 GB of memory. The data values are the average of over 20 runs to mitigate various CPU-related fluctuations in the runtime.

A comparison between the exact and numerical values is highlighted in Fig. 7 at the final time t=2𝑡2t=2italic_t = 2. For a time-step size of Δt=104Δ𝑡superscript104\Delta t=10^{-4}roman_Δ italic_t = 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT, the difference between the exact and numerical approximation is 𝒪(109)𝒪superscript109\mathcal{O}(10^{-9})caligraphic_O ( 10 start_POSTSUPERSCRIPT - 9 end_POSTSUPERSCRIPT ) or better over this time interval. The largest source of the error is observed to occur near the soliton peak located near the origin. We point out that under transformation (9), the spatial points cluster near the origin, even though the points in θ𝜃\thetaitalic_θ are evenly spaced. In practice, we observe it essential to cluster points near nontrivial portions of the solution, e.g. rogue wave peak, to achieve maximal accuracy.

Lastly, we do not include any results for the plane wave solution (25) as it is solved exactly by the MT-SS4 method. Examining (28) we see that the nonlinear integrator simply returns the initial plane wave solution, i.e. v(x,t+Δt)=v(x,t)𝑣𝑥𝑡Δ𝑡𝑣𝑥𝑡v(x,t+\Delta t)=v(x,t)italic_v ( italic_x , italic_t + roman_Δ italic_t ) = italic_v ( italic_x , italic_t ). For the linear integrator (4.1), we note that ucsubscript𝑢𝑐u_{c}italic_u start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT has boundary condition Ub=1subscript𝑈𝑏1U_{b}=1italic_U start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT = 1 and so w~(x,t)=w(x,t)1=0~𝑤𝑥𝑡𝑤𝑥𝑡10\widetilde{w}(x,t)=w(x,t)-1=0over~ start_ARG italic_w end_ARG ( italic_x , italic_t ) = italic_w ( italic_x , italic_t ) - 1 = 0. As a result, all MT coefficients are zero, 𝐰ˇ=𝟎ˇ𝐰0\check{\bf w}={\bf 0}overroman_ˇ start_ARG bold_w end_ARG = bold_0, and the (trivial) solution to (4.1) is zero. Upon returning the boundary condition Ub=1subscript𝑈𝑏1U_{b}=1italic_U start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT = 1, we recover the exact constant solution, and so on.

4.3 Aliasing errors

In this section we address an aliasing-type error that can occur with this method. Aliasing effects are a well-known consequence in certain spectral method approximations of nonlinear PDEs [18]. In our case, sometimes when implementing the MT-SS4 algorithm for a given time-step ΔtΔ𝑡\Delta troman_Δ italic_t, we observe anomalously large errors relative to similar time-step sizes. This is due to unexpectedly large spectral coefficient values corresponding to high-frequency MT modes. We propose a spectral padding approach to mitigate these effects.

The error convergence rate for the MT-SS4 method applied to the NLS equation (24) is shown in Fig. 8 for the Peregrine soliton. The only difference between this graph and that of Fig. 6(a) is that we include more values of ΔtΔ𝑡\Delta troman_Δ italic_t. Specifically, in Fig. 6(a) we decrease ΔtΔ𝑡\Delta troman_Δ italic_t by a factor of 10/7 each time, whereas in Fig. 8(A) we multiply the value of ΔtΔ𝑡\Delta troman_Δ italic_t by 0.990.990.990.99 each time (equivalently, decrease by a factor of 100/9910099100/99100 / 99). The overall convergence rate appears fourth-order except for a few problematic ΔtΔ𝑡\Delta troman_Δ italic_t values. The most egregious example occurs at (b): Δt=0.001(0.99)60Δ𝑡0.001superscript0.9960\Delta t=0.001(0.99)^{60}roman_Δ italic_t = 0.001 ( 0.99 ) start_POSTSUPERSCRIPT 60 end_POSTSUPERSCRIPT where the error is 𝒪(106)𝒪superscript106{\mathcal{O}}(10^{-6})caligraphic_O ( 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT ), meanwhile a slightly larger value of (c): Δt=0.001(0.99)59Δ𝑡0.001superscript0.9959\Delta t=0.001(0.99)^{59}roman_Δ italic_t = 0.001 ( 0.99 ) start_POSTSUPERSCRIPT 59 end_POSTSUPERSCRIPT or slightly smaller value of (a): Δt=0.001(0.99)61Δ𝑡0.001superscript0.9961\Delta t=0.001(0.99)^{61}roman_Δ italic_t = 0.001 ( 0.99 ) start_POSTSUPERSCRIPT 61 end_POSTSUPERSCRIPT have errors that are 𝒪(107)𝒪superscript107{\mathcal{O}}(10^{-7})caligraphic_O ( 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT ). This order of magnitude increase of error is concerning and requires further investigation.

To understand the source of this unexpectedly large error, we examine the evolution of the MT coefficients of the function u(x,t)1𝑢𝑥𝑡1u(x,t)-1italic_u ( italic_x , italic_t ) - 1. In Fig. 9(a) the modulus of the MT-coefficients at the initial condition u(x,2)1𝑢𝑥21u(x,-2)-1italic_u ( italic_x , - 2 ) - 1 for the solution given in (26) is shown. In the other panels we display the MT coefficients at t=2𝑡2t=2italic_t = 2 (final time) for different choices of ΔtΔ𝑡\Delta troman_Δ italic_t. In Fig. 9(c) the MT coefficients for the problematic case are observed to be unexpectedly large at large modal values, i.e. at large values of n𝑛nitalic_n in (1). The MT functions ϕn(x)subscriptitalic-ϕ𝑛𝑥\phi_{n}(x)italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) at large values of |n|𝑛|n|| italic_n | correspond to highly oscillatory functions (see Fig. 1). In contrast, the MT coefficient profiles in Figs. 9(b) and 9(d) appear to remain relatively small at large values of n𝑛nitalic_n.

The discrepancy in the MT coefficients for Δt=0.001(0.99)60Δ𝑡0.001superscript0.9960\Delta t=0.001(0.99)^{60}roman_Δ italic_t = 0.001 ( 0.99 ) start_POSTSUPERSCRIPT 60 end_POSTSUPERSCRIPT compared to the others appears to be an aliasing-type error. For certain values of ΔtΔ𝑡\Delta troman_Δ italic_t, high MT modes (large n𝑛nitalic_n) are unexpectedly large. This is a result of the nonlinearity in the NLS equation. For Fourier spectral methods, there are a myriad of techniques for de-aliasing, see [18]. One of the simplest techniques, which we will employ, is that of padding. Spectral padding involves setting a certain number of high MT modes (large n𝑛nitalic_n) to zero every time-step. Explicitly, we apply the following filter once at each linear solve in the MT-SS4

u~n={uˇn,|n|<N2P0,otherwise,subscript~𝑢𝑛casessubscriptˇ𝑢𝑛𝑛𝑁2𝑃0otherwise\tilde{u}_{n}=\begin{cases}\check{u}_{n},&|n|<\frac{N}{2}-P\\ 0,&\text{otherwise}\\ \end{cases},over~ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = { start_ROW start_CELL overroman_ˇ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , end_CELL start_CELL | italic_n | < divide start_ARG italic_N end_ARG start_ARG 2 end_ARG - italic_P end_CELL end_ROW start_ROW start_CELL 0 , end_CELL start_CELL otherwise end_CELL end_ROW , (30)

for P0𝑃0P\geq 0italic_P ≥ 0 whereby all MT modes with corresponding modenumber |n|𝑛|n|| italic_n | greater than or equal to N/2P𝑁2𝑃N/2-Pitalic_N / 2 - italic_P (N𝑁Nitalic_N is even) are effectively truncated.

Refer to caption
Figure 8: Convergence of the MT-SS4 approximation for the Peregrine soliton (26). Other than the time-step ΔtΔ𝑡\Delta troman_Δ italic_t, all computational parameters are the same as Fig. 6. Overall, a fourth-order convergence rate is observed for both cases. (A) The numerical approximation converges as expected except for a few problematic ΔtΔ𝑡\Delta troman_Δ italic_t values. Specific points are indicated: (a) Δt=0.001(0.99)61Δ𝑡0.001superscript0.9961\Delta t=0.001(0.99)^{61}roman_Δ italic_t = 0.001 ( 0.99 ) start_POSTSUPERSCRIPT 61 end_POSTSUPERSCRIPT, (b) Δt=0.001(0.99)60Δ𝑡0.001superscript0.9960\Delta t=0.001(0.99)^{60}roman_Δ italic_t = 0.001 ( 0.99 ) start_POSTSUPERSCRIPT 60 end_POSTSUPERSCRIPT, and (c) Δt=0.001(0.99)59Δ𝑡0.001superscript0.9959\Delta t=0.001(0.99)^{59}roman_Δ italic_t = 0.001 ( 0.99 ) start_POSTSUPERSCRIPT 59 end_POSTSUPERSCRIPT. (B) Convergence rate of the padded MT-SS4 method for the Peregrine soliton (26). The spectral padding defined in (30) with P=1𝑃1P=1italic_P = 1 has been applied; otherwise, this is the same as (A).
Refer to caption
Figure 9: Modulus of MT coefficients without padding for the exact solution given in (26) with N=256𝑁256N=256italic_N = 256 MT modes. (a) MT coefficients at the initial condition u(x,2)1𝑢𝑥21u(x,-2)-1italic_u ( italic_x , - 2 ) - 1. Numerical and exact MT coefficients at final time t=2𝑡2t=2italic_t = 2 for (b) Δt=0.001(0.99)59Δ𝑡0.001superscript0.9959\Delta t=0.001(0.99)^{59}roman_Δ italic_t = 0.001 ( 0.99 ) start_POSTSUPERSCRIPT 59 end_POSTSUPERSCRIPT, (c) Δt=0.001(0.99)60Δ𝑡0.001superscript0.9960\Delta t=0.001(0.99)^{60}roman_Δ italic_t = 0.001 ( 0.99 ) start_POSTSUPERSCRIPT 60 end_POSTSUPERSCRIPT, and (d) Δt=0.001(0.99)61Δ𝑡0.001superscript0.9961\Delta t=0.001(0.99)^{61}roman_Δ italic_t = 0.001 ( 0.99 ) start_POSTSUPERSCRIPT 61 end_POSTSUPERSCRIPT.

The convergence rate for the padded MT-SS4 method is shown in Fig. 8(B) applied to the NLS equation. Here we take a minimal padding (P=1𝑃1P=1italic_P = 1) which zeros out the n=N/2,N/2+1,N/21𝑛𝑁2𝑁21𝑁21n=-N/2,-N/2+1,N/2-1italic_n = - italic_N / 2 , - italic_N / 2 + 1 , italic_N / 2 - 1 modes. This approach is found to prevent the growing tails observed in Fig. 9(c). The convergence rate is fourth-order and should be compared with the convergence rate in Fig. 8(A). Unlike the unpadded case, we do not observe any sharp jumps in error for the ΔtΔ𝑡\Delta troman_Δ italic_t values considered before. In the event it does appear, the size of pad can be extended by increasing P𝑃Pitalic_P in (30), thereby truncated more high modenumber MT coefficients.

5 Instability and rogue waves in the NLS equation

Having developed the MT-SS4 method in the previous sections, we now look to apply it on a worthy problem. In this section we explore two models for rogue waves: the Peregrine soliton (26) and instability. The Peregrine soliton is found to be unstable. This will limit its ability as a robust rogue wave mechanism. Next, we examine a perturbation of the constant background. A variety of localized perturbations are found to generate unstable rogue wave, Peregine-like modes.

The instability of the Peregrine soliton

To begin, we examine the error introduced by the MT-SS4 method when approximating the Peregrine soliton. The infinity-norm error as a function of time is shown in Fig. 10. Importantly, the error is observed to grow exponentially fast. For linearly stable problems, often one can bound the error as a function of term that grows algebraically in t𝑡titalic_t [43]. This is a serious error growth. Decreasing the time-step size is found to postpone, but not eliminate this error growth. We have included a set of fits in Fig. 10 that illuminate that the error growth rate, before and after the peak, is on the order of e2tsuperscript𝑒2𝑡e^{2t}italic_e start_POSTSUPERSCRIPT 2 italic_t end_POSTSUPERSCRIPT. Countless simulations reveal that, regardless of the number of MT nodes N𝑁Nitalic_N, time-step ΔtΔ𝑡\Delta troman_Δ italic_t, or initial time t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, the growth rate of the error is eventually 𝒪(e2t)𝒪superscript𝑒2𝑡\mathcal{O}(e^{2t})caligraphic_O ( italic_e start_POSTSUPERSCRIPT 2 italic_t end_POSTSUPERSCRIPT ).

So then, what is the nature of the significant error? Here we attribute it not necessarily to a deficiency in our method per se, but rather to the intrinsically unstable nature of the Peregrine soliton and the NLS equation with nonzero boundary conditions. Several recent works [14, 16, 44] have highlighted the unstable nature of the Peregrine solution. Moreover, in [16, 46] the authors precisely predicted a linear instability growth rate of e2tsuperscript𝑒2𝑡e^{2t}italic_e start_POSTSUPERSCRIPT 2 italic_t end_POSTSUPERSCRIPT. Hence, the MT-SS4 method has revealed this expected instability growth rate.

Refer to caption
Figure 10: Error over time interval [4,6]46[-4,6][ - 4 , 6 ] using the MT-SS4 for the NLS equation (24) and initial condition u(x,t0)𝑢𝑥subscript𝑡0u(x,t_{0})italic_u ( italic_x , italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) where t0=4subscript𝑡04t_{0}=-4italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = - 4 for the solution given in (26). The numerical method uses N=512𝑁512N=512italic_N = 512 MT modes and Δt=0.0001Δ𝑡0.0001\Delta t=0.0001roman_Δ italic_t = 0.0001.

This instability will most likely always be a challenge in simulation of the Peregrine solution. The reason is that numerics always makes some sort of approximation, e.g. truncation, round-off error. Hence, the moment we try to simulate this solution, regardless of method, we introduce some error. Decreasing the time-step or increasing the number of modes can mitigate these errors, but eventually they will become significant.

In a physical context, we conjecture this instability will make it difficult for Peregrine soliton to form naturally. In a controlled environment, like a wave flume, these waves have been realized [45]. However, in the open ocean a wide variety of disturbances such as wind or collision with any object will certainly introduce perturbations that will affect the formation of the genuine solution in (26). A more likely scenario is a disturbance of a periodic wave train, or the constant solution (25). We explore this next and find that it generates Peregrine-like structures.

Rogue waves generated by instability

The constant background of the NLS equation is unstable to perturbations and so next we examine the rogue wave-like states that form from this instability. Recall that the Peregrine soliton (26) itself approaches the constant solution at large times when t±𝑡plus-or-minust\rightarrow\pm\inftyitalic_t → ± ∞. In this section we consider several examples of localized perturbations of the constant solution. This would be akin to an ocean wave that is perturbed by a gust of wind or another similar stimuli.

Starting at an initial time of t0=2subscript𝑡02t_{0}=-2italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = - 2 we perturb the constant solution, (25), with initial condition

uj(x,2)=1+εfj(x),j=1,2,3,4,formulae-sequencesubscript𝑢𝑗𝑥21𝜀subscript𝑓𝑗𝑥𝑗1234u_{j}(x,-2)=1+\varepsilon f_{j}(x),~{}~{}~{}~{}j=1,2,3,4,italic_u start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_x , - 2 ) = 1 + italic_ε italic_f start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_x ) , italic_j = 1 , 2 , 3 , 4 , (31)

where 0<ε10𝜀much-less-than10<\varepsilon\ll 10 < italic_ε ≪ 1 and

f1=11+x2,f2=11+x4,f3=ex2,f4=sech(x).formulae-sequencesubscript𝑓111superscript𝑥2formulae-sequencesubscript𝑓211superscript𝑥4formulae-sequencesubscript𝑓3superscript𝑒superscript𝑥2subscript𝑓4sech𝑥f_{1}=\frac{1}{1+x^{2}},~{}~{}~{}~{}~{}~{}f_{2}=\frac{1}{1+x^{4}},~{}~{}~{}~{}% ~{}~{}f_{3}=e^{-x^{2}},~{}~{}~{}~{}~{}~{}f_{4}=\text{sech}(x).italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 1 + italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 1 + italic_x start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT end_ARG , italic_f start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = italic_e start_POSTSUPERSCRIPT - italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT , italic_f start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT = sech ( italic_x ) .

and then evolve the equation using the MT-SS4 method. In each case considered we choose ε=0.1.𝜀0.1\varepsilon=0.1.italic_ε = 0.1 . For each case considered, we observe single-hump Peregrine-like structures peak at approximately t=0𝑡0t=0italic_t = 0, with a peak height of around 2.6-2.8. A typical evolution is shown in Fig. 11. Notice the peak amplitude is a factor of 26-28 larger than the initial perturbation amplitude (ε=0.1𝜀0.1\varepsilon=0.1italic_ε = 0.1). The structure that forms around t=0𝑡0t=0italic_t = 0 has an uncanny resemblance to that of the Pergreine soliton (see Fig. 4). Namely, a localized peak of approximate amplitude 2.7 appears and then suddenly disappears. Moreover, near the peak, we observe a set of adjacent troughs. We do not dwell on them here, but near the final time (t1𝑡1t\approx 1italic_t ≈ 1), a structure consisting of two-humps forms. This is reminiscent of modulational instability on periodic finite domains [47].

To further establish the Peregrine (rational) nature of this peak, in Fig. 12 we show some solution profiles at t=0𝑡0t=0italic_t = 0 (near the peak time) generated by each initial condition in (31). In each case considered, the overall decay rate of |u(x,0)1|𝑢𝑥01|u(x,0)-1|| italic_u ( italic_x , 0 ) - 1 | is found to decay rationally. Notice that we have subtracted off the background to isolate the decay rate. Furthermore, when we apply a variety of rational fits, we consistently find the tails to decay like 𝒪(1x2)𝒪1superscript𝑥2\mathcal{O}\left(\frac{1}{x^{2}}\right)caligraphic_O ( divide start_ARG 1 end_ARG start_ARG italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) as they approach the constant background. This matches the rational structure of (26) at t=0𝑡0t=0italic_t = 0.

These results suggest a likely mechanism for generating rogues wave is this instability. We have established the first peak is nearly the Peregrine soliton at t=0𝑡0t=0italic_t = 0. All four of initial perturbations considered in (31) appear to be generating roughly the same rational structure (see Fig. 11(left)). These results were obtained by applying the MT-SS4 method developed in the first part of this paper.

Refer to caption
Figure 11: Typical results obtained by applying the MT-SS4 using the perturbed initial condition (31) with f1(x)=11+x2subscript𝑓1𝑥11superscript𝑥2f_{1}(x)=\frac{1}{1+x^{2}}italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_x ) = divide start_ARG 1 end_ARG start_ARG 1 + italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG. Top (left) and side (right) views of the solution modulus. The computational parameters used are N=512𝑁512N=512italic_N = 512,Δt=0.0001Δ𝑡0.0001\Delta t=0.0001roman_Δ italic_t = 0.0001, and with a pad of P=1𝑃1P=1italic_P = 1.
Refer to caption
Figure 12: Plots of ln|u(x,0)1|𝑢𝑥01\ln|u(x,0)-1|roman_ln | italic_u ( italic_x , 0 ) - 1 | obtained by solving the NLS via the MT-SS4 method with the perturbed initial conditions (31). The computational parameters are the same as Fig. 11. These results are obtained from initial perturbations (a) f1(x)=(1+x2)1subscript𝑓1𝑥superscript1superscript𝑥21f_{1}(x)=(1+x^{2})^{-1}italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_x ) = ( 1 + italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, (b) f2(x)=(1+x4)1subscript𝑓2𝑥superscript1superscript𝑥41f_{2}(x)=(1+x^{4})^{-1}italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_x ) = ( 1 + italic_x start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, (c) f3(x)=ex2subscript𝑓3𝑥superscript𝑒superscript𝑥2f_{3}(x)=e^{-x^{2}}italic_f start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( italic_x ) = italic_e start_POSTSUPERSCRIPT - italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT, (d) f4(x)=sech(x)subscript𝑓4𝑥sech𝑥f_{4}(x)=\text{sech}(x)italic_f start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ( italic_x ) = sech ( italic_x ). The rational function ln|(1+x2)1|superscript1superscript𝑥21\ln|(1+x^{2})^{-1}|roman_ln | ( 1 + italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT | is indicated by curve (e); it indicates that the decay rate of all the perturbed approximations is rational and decays like 1x21superscript𝑥2\frac{1}{x^{2}}divide start_ARG 1 end_ARG start_ARG italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG as x±𝑥plus-or-minusx\rightarrow\pm\inftyitalic_x → ± ∞.

6 Conclusion

In this work we have developed a numerical method for approximating rogue wave models of certain partial differential equations. These rogue wave models are rational in form, so we expand them in a basis of Malmquist-Takenaka rational functions. The spectral coefficients can decay exponentially fast and can be numerically computed in terms of (fast) Fourier methods. A fourth-order split-step method for approximation the nonlinear Schrödinger equation was introduced. The accuracy of the method for rational solutions was observed to be spectral in space and fourth-order in time.

This method was applied to approximating rogue wave models. The Peregrine soliton is linearly unstable and the instability manifests itself in the simulations. Any slight perturbation will excite this instability making the genuine Peregrine soliton solution challenging to see in a deep water, open ocean setting. On the other hand, when a constant background (traveling plane wave) is perturbed, a Peregrine-like rogue wave is found to emerge.

Appendix A Malmquist-Takenaka approximation of derivatives

In this appendix we give the general algorithm for approximating spatial derivatives using the MT spectral approximation. The semi-discrete motivations are shown in (12) and (13). The discrete modified Fourier transforms are computed using standard FFT algorithms. We abbreviate the modified FFT by MFFT and the modified inverse Fourier transform by MIFFT. Their precise definitions are given in B.

Input:N(even),function f(x),differentiation matrix 𝔻mInput:𝑁(even)function 𝑓𝑥differentiation matrix subscript𝔻𝑚\displaystyle\text{Input:}~{}~{}~{}N\text{(even)},\hskip 10.84006pt\text{% function }f(x),\hskip 14.45377pt\text{differentiation matrix }\mathbb{D}_{m}Input: italic_N (even) , function italic_f ( italic_x ) , differentiation matrix blackboard_D start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT
xj+N2+1=12tan(πjN),j=N2,N2+1,,N21formulae-sequencesubscript𝑥𝑗𝑁2112𝜋𝑗𝑁𝑗𝑁2𝑁21𝑁21\displaystyle x_{j+\frac{N}{2}+1}=\frac{1}{2}\tan\left(\frac{\pi j}{N}\right),% ~{}~{}~{}~{}j=-\frac{N}{2},-\frac{N}{2}+1,...,\frac{N}{2}-1italic_x start_POSTSUBSCRIPT italic_j + divide start_ARG italic_N end_ARG start_ARG 2 end_ARG + 1 end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_tan ( divide start_ARG italic_π italic_j end_ARG start_ARG italic_N end_ARG ) , italic_j = - divide start_ARG italic_N end_ARG start_ARG 2 end_ARG , - divide start_ARG italic_N end_ARG start_ARG 2 end_ARG + 1 , … , divide start_ARG italic_N end_ARG start_ARG 2 end_ARG - 1
x=(x1,x2,,xN)Txsuperscriptsubscript𝑥1subscript𝑥2subscript𝑥𝑁𝑇\displaystyle\textbf{x}=(x_{1},x_{2},...,x_{N})^{T}x = ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_x start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT
f=(f1,f2.,,fN)T,fjf(xj)\displaystyle\textbf{f}=(f_{1},f_{2}.,...,f_{N})^{T},\hskip 7.22743ptf_{j}% \approx f(x_{j})f = ( italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT . , … , italic_f start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT , italic_f start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ≈ italic_f ( italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT )
fˇ=MFFT[f]ˇfMFFTdelimited-[]f\displaystyle\check{\textbf{f}}=\text{MFFT}[\textbf{f}]overroman_ˇ start_ARG f end_ARG = MFFT [ f ]
fˇ(m)=𝔻mfˇsuperscriptˇf𝑚subscript𝔻𝑚ˇf\displaystyle\check{\textbf{f}}^{(m)}=\mathbb{D}_{m}\check{\textbf{f}}overroman_ˇ start_ARG f end_ARG start_POSTSUPERSCRIPT ( italic_m ) end_POSTSUPERSCRIPT = blackboard_D start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT overroman_ˇ start_ARG f end_ARG
f(m)=MIFFT[fˇ(m)]superscriptf𝑚MIFFTdelimited-[]superscriptˇf𝑚\displaystyle\textbf{f}^{(m)}=\text{MIFFT}\left[\check{\textbf{f}}^{(m)}\right]f start_POSTSUPERSCRIPT ( italic_m ) end_POSTSUPERSCRIPT = MIFFT [ overroman_ˇ start_ARG f end_ARG start_POSTSUPERSCRIPT ( italic_m ) end_POSTSUPERSCRIPT ]
Output:f(m)=(f1(m),f2(m),,fN(m))TOutput:superscriptf𝑚superscriptsuperscriptsubscript𝑓1𝑚superscriptsubscript𝑓2𝑚superscriptsubscript𝑓𝑁𝑚𝑇\displaystyle\text{Output:}~{}~{}~{}\textbf{f}^{(m)}=(f_{1}^{(m)},f_{2}^{(m)},% ...,f_{N}^{(m)})^{T}Output: f start_POSTSUPERSCRIPT ( italic_m ) end_POSTSUPERSCRIPT = ( italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_m ) end_POSTSUPERSCRIPT , italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_m ) end_POSTSUPERSCRIPT , … , italic_f start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_m ) end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT

This algorithm gives derivative approximations at the equally spaced angular points θj=2πjNsubscript𝜃𝑗2𝜋𝑗𝑁\theta_{j}=\frac{2\pi j}{N}italic_θ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = divide start_ARG 2 italic_π italic_j end_ARG start_ARG italic_N end_ARG for j=N/2,N/2+1,,N/21𝑗𝑁2𝑁21𝑁21j=-N/2,-N/2+1,\dots,N/2-1italic_j = - italic_N / 2 , - italic_N / 2 + 1 , … , italic_N / 2 - 1. Differentiation matrices for the first two derivatives are shown in (16) and (19). The output is the approximation of the mthsuperscript𝑚thm^{\rm th}italic_m start_POSTSUPERSCRIPT roman_th end_POSTSUPERSCRIPT-order derivative evaluated at the grid points, i.e. fj(m)f(m)(xj)superscriptsubscript𝑓𝑗𝑚superscript𝑓𝑚subscript𝑥𝑗f_{j}^{(m)}\approx f^{(m)}(x_{j})italic_f start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_m ) end_POSTSUPERSCRIPT ≈ italic_f start_POSTSUPERSCRIPT ( italic_m ) end_POSTSUPERSCRIPT ( italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ). Once the MT coefficients are obtained, one can approximate f(x)𝑓𝑥f(x)italic_f ( italic_x ) and its derivatives at arbitrary values of x𝑥xitalic_x (not just the spatial grid points) by summing the truncated MT series f(m)(x)Σn=N/2N/21fˇn(m)ϕn(x),m=0,1,2,formulae-sequencesuperscript𝑓𝑚𝑥𝑁21𝑛𝑁2Σsubscriptsuperscriptˇ𝑓𝑚𝑛subscriptitalic-ϕ𝑛𝑥𝑚012f^{(m)}(x)\approx\overset{N/2-1}{\underset{n=-N/2}{\Sigma}}\check{f}^{(m)}_{n}% \phi_{n}(x),~{}m=0,1,2,italic_f start_POSTSUPERSCRIPT ( italic_m ) end_POSTSUPERSCRIPT ( italic_x ) ≈ start_OVERACCENT italic_N / 2 - 1 end_OVERACCENT start_ARG start_UNDERACCENT italic_n = - italic_N / 2 end_UNDERACCENT start_ARG roman_Σ end_ARG end_ARG overroman_ˇ start_ARG italic_f end_ARG start_POSTSUPERSCRIPT ( italic_m ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) , italic_m = 0 , 1 , 2 , for the MT functions given in (1).

Appendix B Discrete modified Fourier and inverse Fourier transforms

In this appendix we discuss the practical implementation of (12) and (13) to compute the MT coefficients and sum them via the discrete Fourier transform. We refer to these as the modified fast Fourier transform (MFFT) and modified inverse fast Fourier transform (MIFFT), respectively. First we discretize θ[π,π]𝜃𝜋𝜋\theta\in[-\pi,\pi]italic_θ ∈ [ - italic_π , italic_π ] by θj=jhsubscript𝜃𝑗𝑗\theta_{j}=jhitalic_θ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_j italic_h, for j=N2,N2+1,,N21𝑗𝑁2𝑁21𝑁21j=-\frac{N}{2},-\frac{N}{2}+1,\dots,\frac{N}{2}-1italic_j = - divide start_ARG italic_N end_ARG start_ARG 2 end_ARG , - divide start_ARG italic_N end_ARG start_ARG 2 end_ARG + 1 , … , divide start_ARG italic_N end_ARG start_ARG 2 end_ARG - 1, where N𝑁Nitalic_N is an even integer with angular spacing h=2πN2𝜋𝑁h=\frac{2\pi}{N}italic_h = divide start_ARG 2 italic_π end_ARG start_ARG italic_N end_ARG. There is no point at j=N/2𝑗𝑁2j=N/2italic_j = italic_N / 2 due to the periodic boundary conditions. The discrete MT (Fourier) coefficients are given by

fˇn=(i)nπ21Nj=N/2N/21f(12tan(θj2))(1itan(θj2))einθj,n=N2,N2+1,,N21.formulae-sequencesubscriptˇ𝑓𝑛superscript𝑖𝑛𝜋21𝑁superscriptsubscript𝑗𝑁2𝑁21𝑓12subscript𝜃𝑗21𝑖subscript𝜃𝑗2superscript𝑒𝑖𝑛subscript𝜃𝑗𝑛𝑁2𝑁21𝑁21\check{f}_{n}=(-i)^{n}\sqrt{\frac{\pi}{2}}\frac{1}{N}\sum_{j=-N/2}^{N/2-1}f% \left(\frac{1}{2}\tan\left(\frac{\theta_{j}}{2}\right)\right)\left(1-i\tan% \left(\frac{\theta_{j}}{2}\right)\right)e^{-in\theta_{j}},~{}~{}~{}~{}n=-\frac% {N}{2},-\frac{N}{2}+1,\dots,\frac{N}{2}-1.overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = ( - italic_i ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT square-root start_ARG divide start_ARG italic_π end_ARG start_ARG 2 end_ARG end_ARG divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_j = - italic_N / 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N / 2 - 1 end_POSTSUPERSCRIPT italic_f ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_tan ( divide start_ARG italic_θ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG ) ) ( 1 - italic_i roman_tan ( divide start_ARG italic_θ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG ) ) italic_e start_POSTSUPERSCRIPT - italic_i italic_n italic_θ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT , italic_n = - divide start_ARG italic_N end_ARG start_ARG 2 end_ARG , - divide start_ARG italic_N end_ARG start_ARG 2 end_ARG + 1 , … , divide start_ARG italic_N end_ARG start_ARG 2 end_ARG - 1 . (32)

This is the discrete approximation of (11). Summing the MT coefficients yields the corresponding function

fj=2π11itan(θj2)n=N/2N/21(i)nfˇneinθj,j=N2,,N21,formulae-sequencesubscript𝑓𝑗2𝜋11𝑖subscript𝜃𝑗2superscriptsubscript𝑛𝑁2𝑁21superscript𝑖𝑛subscriptˇ𝑓𝑛superscript𝑒𝑖𝑛subscript𝜃𝑗𝑗𝑁2𝑁21f_{j}=\sqrt{\frac{2}{\pi}}\frac{1}{1-i\tan\left(\frac{\theta_{j}}{2}\right)}% \sum_{n=-N/2}^{N/2-1}(i)^{n}\check{f}_{n}e^{in\theta_{j}},~{}~{}~{}~{}j=-\frac% {N}{2},\cdots,\frac{N}{2}-1,italic_f start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = square-root start_ARG divide start_ARG 2 end_ARG start_ARG italic_π end_ARG end_ARG divide start_ARG 1 end_ARG start_ARG 1 - italic_i roman_tan ( divide start_ARG italic_θ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG ) end_ARG ∑ start_POSTSUBSCRIPT italic_n = - italic_N / 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N / 2 - 1 end_POSTSUPERSCRIPT ( italic_i ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT overroman_ˇ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_n italic_θ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT , italic_j = - divide start_ARG italic_N end_ARG start_ARG 2 end_ARG , ⋯ , divide start_ARG italic_N end_ARG start_ARG 2 end_ARG - 1 , (33)

where fj=f(12tan(θj2))subscript𝑓𝑗𝑓12subscript𝜃𝑗2f_{j}=f\left(\frac{1}{2}\tan\left(\frac{\theta_{j}}{2}\right)\right)italic_f start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_f ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_tan ( divide start_ARG italic_θ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG ) ). This is the discrete approximation of the sum in (8). Once the MT coefficients are known, the corresponding derivative coefficients are found using N×N𝑁𝑁N\times Nitalic_N × italic_N truncated versions of the differentiation matrices derived in Sec. 3.

Appendix C Fourth-order split-step integration

All PDEs in this work are solved using the explicit Malmquist-Takenaka split-step (MT-SS4) method described in Sec. 4.1. These approximations tends to be spectrally accurate in space and fourth-order accurate in time. The algorithm pseduo-code is given below.

The temporal interval [t0,T]subscript𝑡0𝑇[t_{0},T][ italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_T ] is discretized into S+1𝑆1S+1italic_S + 1 equally spaced time levels tl=t0+lΔtsubscript𝑡𝑙subscript𝑡0𝑙Δ𝑡t_{l}=t_{0}+l\Delta titalic_t start_POSTSUBSCRIPT italic_l end_POSTSUBSCRIPT = italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_l roman_Δ italic_t for l=0,1,,S𝑙01𝑆l=0,1,\dots,Sitalic_l = 0 , 1 , … , italic_S with time-steps of Δt=Tt0SΔ𝑡𝑇subscript𝑡0𝑆\Delta t=\frac{T-t_{0}}{S}roman_Δ italic_t = divide start_ARG italic_T - italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_S end_ARG. The real line (,)(-\infty,\infty)( - ∞ , ∞ ) is discretized by xj+N2+1=12tan(πjN),subscript𝑥𝑗𝑁2112𝜋𝑗𝑁x_{j+\frac{N}{2}+1}=\frac{1}{2}\tan\left(\frac{\pi j}{N}\right),italic_x start_POSTSUBSCRIPT italic_j + divide start_ARG italic_N end_ARG start_ARG 2 end_ARG + 1 end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_tan ( divide start_ARG italic_π italic_j end_ARG start_ARG italic_N end_ARG ) , for j=N2,N2+1,,N21𝑗𝑁2𝑁21𝑁21j=-\frac{N}{2},-\frac{N}{2}+1,...,\frac{N}{2}-1italic_j = - divide start_ARG italic_N end_ARG start_ARG 2 end_ARG , - divide start_ARG italic_N end_ARG start_ARG 2 end_ARG + 1 , … , divide start_ARG italic_N end_ARG start_ARG 2 end_ARG - 1 and 𝐱=(x1,x2,,xN)T𝐱superscriptsubscript𝑥1subscript𝑥2subscript𝑥𝑁𝑇{\bf x}=(x_{1},x_{2},...,x_{N})^{T}bold_x = ( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_x start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT for even N𝑁Nitalic_N. Denote the numerical approximation of the solution by 𝐮(t)=(u1,u2,,uN)T𝐮𝑡superscriptsubscript𝑢1subscript𝑢2subscript𝑢𝑁𝑇{\bf u}(t)=(u_{1},u_{2},...,u_{N})^{T}bold_u ( italic_t ) = ( italic_u start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_u start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_u start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT where uju(xj,t)subscript𝑢𝑗𝑢subscript𝑥𝑗𝑡u_{j}\approx u(x_{j},t)italic_u start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ≈ italic_u ( italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , italic_t ). The N×N𝑁𝑁N\times Nitalic_N × italic_N second differentiation matrix 𝔻2subscript𝔻2\mathbb{D}_{2}blackboard_D start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT is defined in (19). The initial condition is given by 𝐮0=u(𝐱,t0)subscript𝐮0𝑢𝐱subscript𝑡0{\bf u}_{0}=u({\bf x},t_{0})bold_u start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_u ( bold_x , italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ).

Consider the NLS equation (24) and splitting (27). The solution of the nonlinear equation was given in (28); notice this is solved in the physical domain. To solve the linear equation, first the solution is shifted, 𝐮1𝐮1{\bf u}-1bold_u - 1, to ensure zero boundary conditions. Next, the MFFT transform (32) is taken to find the coefficients and the linear system is solved by (29). Afterward, the function is reconstructed by MIFFT transform (33) and the unity boundary conditions are restored. The split-step method given in (23) gives the appropriate combination of steps to achieve fourth-order accuracy.

One MT-SS4 time-step for the NLS equation

𝐯1=exp[2iα1Δt(|𝐮j|21)]𝐮jsubscript𝐯12𝑖subscript𝛼1Δ𝑡superscriptsubscript𝐮𝑗21subscript𝐮𝑗\displaystyle{\bf v}_{1}=\exp\left[2i\alpha_{1}\Delta t\left(|{\bf u}_{j}|^{2}% -1\right)\right]{\bf u}_{j}bold_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = roman_exp [ 2 italic_i italic_α start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT roman_Δ italic_t ( | bold_u start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1 ) ] bold_u start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT
𝐯ˇ1=MFFT[𝐯11]subscriptˇ𝐯1MFFTdelimited-[]subscript𝐯11\displaystyle\check{{\bf v}}_{1}=\text{MFFT}[{\bf v}_{1}-1]overroman_ˇ start_ARG bold_v end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = MFFT [ bold_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - 1 ]
𝐰ˇ1=exp[iβ1Δt𝔻2]𝐯ˇ1subscriptˇ𝐰1𝑖subscript𝛽1Δ𝑡subscript𝔻2subscriptˇ𝐯1\displaystyle\check{{\bf w}}_{1}=\exp\left[i\beta_{1}\Delta t\mathbb{D}_{2}% \right]\check{{\bf v}}_{1}overroman_ˇ start_ARG bold_w end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = roman_exp [ italic_i italic_β start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT roman_Δ italic_t blackboard_D start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ] overroman_ˇ start_ARG bold_v end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT
𝐲1=MIFFT[𝐰ˇ1]+1subscript𝐲1MIFFTdelimited-[]subscriptˇ𝐰11\displaystyle{\bf y}_{1}=\text{MIFFT}\left[\check{{\bf w}}_{1}\right]+1bold_y start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = MIFFT [ overroman_ˇ start_ARG bold_w end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ] + 1
𝐯2=exp[2iα2Δt(|𝐲1|21)]𝐲1subscript𝐯22𝑖subscript𝛼2Δ𝑡superscriptsubscript𝐲121subscript𝐲1\displaystyle{\bf v}_{2}=\exp\left[2i\alpha_{2}\Delta t\left(|{\bf y}_{1}|^{2}% -1\right)\right]{\bf y}_{1}bold_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = roman_exp [ 2 italic_i italic_α start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_Δ italic_t ( | bold_y start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1 ) ] bold_y start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT
𝐯ˇ2=MFFT[𝐯21]subscriptˇ𝐯2MFFTdelimited-[]subscript𝐯21\displaystyle\check{{\bf v}}_{2}=\text{MFFT}[{\bf v}_{2}-1]overroman_ˇ start_ARG bold_v end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = MFFT [ bold_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - 1 ]
𝐰ˇ2=exp[iβ2Δt𝔻2]𝐯ˇ2subscriptˇ𝐰2𝑖subscript𝛽2Δ𝑡subscript𝔻2subscriptˇ𝐯2\displaystyle\check{{\bf w}}_{2}=\exp\left[i\beta_{2}\Delta t\mathbb{D}_{2}% \right]\check{{\bf v}}_{2}overroman_ˇ start_ARG bold_w end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = roman_exp [ italic_i italic_β start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT roman_Δ italic_t blackboard_D start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ] overroman_ˇ start_ARG bold_v end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT
𝐲2=MIFFT[𝐰ˇ2]+1subscript𝐲2MIFFTdelimited-[]subscriptˇ𝐰21\displaystyle{\bf y}_{2}=\text{MIFFT}\left[\check{{\bf w}}_{2}\right]+1bold_y start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = MIFFT [ overroman_ˇ start_ARG bold_w end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ] + 1
𝐯3=exp[2iα3Δt(|𝐲2|21)]𝐲2subscript𝐯32𝑖subscript𝛼3Δ𝑡superscriptsubscript𝐲221subscript𝐲2\displaystyle{\bf v}_{3}=\exp\left[2i\alpha_{3}\Delta t\left(|{\bf y}_{2}|^{2}% -1\right)\right]{\bf y}_{2}bold_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = roman_exp [ 2 italic_i italic_α start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT roman_Δ italic_t ( | bold_y start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1 ) ] bold_y start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT
𝐯ˇ3=MFFT[𝐯31]subscriptˇ𝐯3MFFTdelimited-[]subscript𝐯31\displaystyle\check{{\bf v}}_{3}=\text{MFFT}[{\bf v}_{3}-1]overroman_ˇ start_ARG bold_v end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = MFFT [ bold_v start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT - 1 ]
𝐰ˇ3=exp[iβ3Δt𝔻2]𝐯ˇ3subscriptˇ𝐰3𝑖subscript𝛽3Δ𝑡subscript𝔻2subscriptˇ𝐯3\displaystyle\check{{\bf w}}_{3}=\exp\left[i\beta_{3}\Delta t\mathbb{D}_{2}% \right]\check{{\bf v}}_{3}overroman_ˇ start_ARG bold_w end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = roman_exp [ italic_i italic_β start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT roman_Δ italic_t blackboard_D start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ] overroman_ˇ start_ARG bold_v end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT
𝐲3=MIFFT[𝐰ˇ3]+1subscript𝐲3MIFFTdelimited-[]subscriptˇ𝐰31\displaystyle{\bf y}_{3}=\text{MIFFT}\left[\check{{\bf w}}_{3}\right]+1bold_y start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = MIFFT [ overroman_ˇ start_ARG bold_w end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ] + 1
𝐮j+1=exp[2iα4Δt(|𝐲3|21)]𝐲3subscript𝐮𝑗12𝑖subscript𝛼4Δ𝑡superscriptsubscript𝐲321subscript𝐲3\displaystyle{\bf u}_{j+1}=\exp\left[2i\alpha_{4}\Delta t\left(|{\bf y}_{3}|^{% 2}-1\right)\right]{\bf y}_{3}bold_u start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT = roman_exp [ 2 italic_i italic_α start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT roman_Δ italic_t ( | bold_y start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1 ) ] bold_y start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT
tj+1=tj+Δtsubscript𝑡𝑗1subscript𝑡𝑗Δ𝑡\displaystyle t_{j+1}=t_{j}+\Delta titalic_t start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT = italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + roman_Δ italic_t

References

  • [1] O. E. Hansteen, H. P. Jostad, and T. I. Tjelta, Observed platform response to a “monster” wave, Field Measurements in Geomechanics 13738 pp. 15-18 (2003).
  • [2] C. Kharif and E. Pelinovsky, Physical mechanisms of the rogue wave phenomenon, European. J. Mechanics-B/Fluids 22 pp. 603-634 (2003).
  • [3] G. Lawton, Monsters of the deep (The perfect wave), New Scientist 170 pp. 28-32 (2001).
  • [4] I. V. Lavrenov, The wave energy concentration at the Agulhas current off South Africa, Natural Hazards 17 pp. 117-127 (1998).
  • [5] V. E. Zhakharov, Stability of periodic waves of finite amplitude on the surface of deep water, J. Appl. Mech. Tech. Phys. 9 pp. 190-194 (1968).
  • [6] G. Agrawal, Nonlinear fiber optics, Academic Press, Elsevier (2013).
  • [7] V. E. Zhakharov, Collapse of Langmuir waves, Soviet Physics JETP. 35 pp. 908-914 (1972).
  • [8] M. J. Ablowitz, Nonlinear dispersive waves, Cambridge University Press (2011).
  • [9] Y. Ohta and J. Yang, General high-order rogue waves and their dynamics in the nonlinear Schrödinger equation, Proc. Royal Soc. A 468, pp. 1716-1740 (2012).
  • [10] B. Yang and J. Yang, Rogue wave patterns in the nonlinear Schrödinger equation, Physica D 419, 132850 (2021).
  • [11] D. H. Peregrine, Water waves, nonlinear Schrödinger equations and their solutions, J. Austral. Math. Soc. Ser. B 25, pp. 16-43 (1983).
  • [12] B. T. Benjamin and J. E. Feir, The disintegration of wave trains on deep water. Part 1. Theory, J. Fluid Mech. 27 pp. 417-430 (1967).
  • [13] B. T. Benjamin, Instability of periodic wavetrains in nonlinear dispersive systems, Proc. Royal Soc. London A 299 pp. 59-76 (1967).
  • [14] J. Cuevas-Maraver, P. G. Kevrekidis, D. J. Frantzeskakis, N. I. Karachalios, M. Haragus, and G. James, Floquet analysis of Kuznetsov-Ma breathers: A path towards spectral stability of rogue waves, Phys. Rev. E 96 012202 (2017).
  • [15] C. Klein and M. Haragus, Numerical study of the stability of the Peregrine solution, Annals Math. Sci. and Appl. 2 pp. 217-239 (2017).
  • [16] M. J. Ablowitz and J. T. Cole, Transverse instability of rogue waves, Phys. Rev. Lett. 127 104101 (2021).
  • [17] B. Fornberg, A practical guide to pseudospectral methods, Cambridge University Press (1996).
  • [18] C. Canuto, M. Y. Hussaini, A. Quarteroni, and T. A. Zang, Spectral Methods: Fundamentals in Single Domains, Springer-Verlag, Berlin (2006).
  • [19] B. Fornberg and G. B. Whitham, A numerical and theoretical study of certain nonlinear wave phenomena, Philos. Trans. Royal Soc. A 289 pp. 373-404 (1978).
  • [20] J. Yang, Nonlinear Waves in Integrable and Nonintegrable Systems, SIAM, Philadelphia (2010).
  • [21] J. P. Boyd, The rate of convergence of Hermite function series, Math. of Comp. 35, pp. 1309-1316 (1980).
  • [22] B. Guo, J. Shen, and C. Xu, Spectral and pseudospectral approximations using Hermite functions: application to the Dirac equation, Adv. in Comp. Math. 19, pp. 35-55 (2003).
  • [23] C. E. Grosch and S. A. Orszag, Numerical solution of problems in unbounded regions: coordinate transforms, J. Comp. Phys. 25, pp. 273-296 (1977).
  • [24] A. Islas and C. M. Schober, Numerical investigation of the stability of the rational solutions of the nonlinear Schrödinger equation, Appl. Math. Comp.305, pp. 17-26 (2017).
  • [25] J. P. Boyd, Chebyshev and Fourier Spectral methods 2nd ed., Dover, New York, (2001).
  • [26] A. Islas and C. M. Schober, Numerical investigation of the stability of the rational solutions of the nonlinear Schrödinger equation, Applied Mathematics and Computation 305 pp. 17-26 (2017).
  • [27] J. P. Boyd, Spectral methods using rational basis functions on an infinite interval, J. Comp. Phys. 69, pp. 112-142 (1987).
  • [28] F. Malmquist, Sur la détermination d’une classe de fonctions analytiques par leurs valeurs dans un ensemble donné de points, Comptes Rendus du Sixieme Congres. 26, pp. 253-259 (1925).
  • [29] S. Takenaka, On the orthogonal functions and a new formula of interpolation, Japan J. Maths 2, pp. 129-145 (1926).
  • [30] C. Christov, A complete orthonormal system of functions in l2(,)superscript𝑙2l^{2}(-\infty,\infty)italic_l start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( - ∞ , ∞ ) space, SIAM J. Appl. Math. 42, pp. 1337-1344 (1982).
  • [31] J. P. Boyd, The orthogonal rational functions of Higgins and Christov and algebraically mapped Chebyshev polynomials, J. Approx. Theory 61, pp. 98-105 (1990).
  • [32] M. J. Ablowitz and A. S. Fokas, Complex variables: introduction and applications, Cambridge University Press (2003).
  • [33] J. A. C. Weideman, Computing the Hilbert transform on the real line, Math. of Comp. 64, pp. 745-762 (1995).
  • [34] A. Iserles and M. Webb, A family of orthogonal rational functions and other orthogonal systems with a skew-Hermitian differentiation matrix, J. Fourier Anal. and Appl. 26, 19 (2020).
  • [35] R. E. A. C. Paley and N. Wiener, Fourier transforms in the complex domain, Amer. Math. Soc. Colloq. Publ., 19 Providence, (1934).
  • [36] J. P. Boyd, The optimization of convergence for Chebyshev polynomial methods in an unbounded domain, J. Comp. Phys. 45 pp. 43–79 (1982).
  • [37] S. Shindin, N. Parumasur, and O. Aluko, Analysis of Malmquist-Takenaka-Christov rational approximations with applications to the nonlinear Benjamin equation, Commun. Nonlinear Sci. Numer. Simulat. 94, 105571 (2021).
  • [38] T. R. Taha and M. J. Ablowitz Analytical and numerical aspects of certain nonlinear evolution equations. II. Numerical, nonlinear Schrödinger equation, J. Comp. Phys. 55 pp. 203-230 (1984).
  • [39] T. R. Taha and M. J. Ablowitz Analytical and numerical aspects of certain nonlinear evolution equations. III. Numerical, Korteweg-de Vries equation, J. Comp. Phys. 55 pp. 231-253 (1984).
  • [40] L. N. Trefethen, Spectral Methods in MATLAB, SIAM, Philadelphia, (2000).
  • [41] S. Roudenko, Z. Wang, and K. Yang, Dynamics of solutions in the generalized Benjamin-Ono equation: A numerical study, J. Comp. Physics 445 110570 (2021).
  • [42] H. Yoshida, Construction of higher order symplectic integrators, Phys. Lett. A 150 pp. 262-268 (1990).
  • [43] J. T. Cole and T. I. Johnson, Stability and error analysis of a Malmquist-Takenaka time-stepping method, In progress (2024).
  • [44] A. Calini, C. M. Schober, and M. Strawn Linear instability of the Peregrine breather: Numerical and analytical investigations, Appl. Num. Math. 141 pp. 36-43 (2019).
  • [45] A. Chabchoub, N. P Hoffmann, and N. Akhmediev, Rogue wave observation in a water wave tank, Phys. Rev. Lett. 106 204502 (2011).
  • [46] A. Calini, C. M. Schober, and M.Strawn Linear instability of the Peregrine breather: Numerical and analytical investigations, Applied Numerical Mathematics 141 pp. 36-43 (2019).
  • [47] H. C. Yuen and W. E. Ferguson, Relationship between Benjamin-Feir instability and recurrence in the nonlinear Schrödinger equation, Phys. Fluids 21 pp. 1275-1278 (1978).