Fact-checked by Grok 2 weeks ago

Inverse scattering transform

The inverse scattering transform (IST) is a nonlinear analogue of the , employed to solve the initial-value problems of certain integrable nonlinear evolution equations, such as the Korteweg–de Vries (KdV) equation, by linearizing them through a process. It consists of three main steps: a direct transform that maps the initial potential to data (including discrete eigenvalues and a ), a linear of that data, and an inverse transform that reconstructs the solution at later times using techniques like the Marchenko or Riemann-Hilbert problems. Originally discovered in 1967 by Gardner, Greene, Kruskal, and Miura to exactly solve the KdV equation u_t + 6uu_x + u_{xxx} = 0, the method revealed the nature of its solutions and established a one-to-one correspondence between rapidly decaying potentials and their data in the context of the associated Schrödinger operator. The term "inverse scattering transform" was coined in 1974 by Ablowitz, Kaup, Newell, and Segur, who generalized it to a broader class of equations, including the , emphasizing its Fourier-like structure for nonlinear problems. IST has become a cornerstone of soliton theory and integrable systems, enabling the exact construction of multi-soliton solutions that emerge unchanged from nonlinear interactions, as well as periodic and other exact solutions for equations modeling phenomena in water waves, optics, and plasma physics. Its formulation relies on the existence of a Lax pair—compatible linear operators whose compatibility condition yields the nonlinear equation—ensuring integrability and conservation laws. Beyond one-dimensional cases, extensions to higher dimensions and discrete systems have been developed, with applications in inverse problems for quantum mechanics and signal processing. The method's power lies in transforming complex nonlinear dynamics into tractable linear algebra, profoundly influencing mathematical physics since its inception.

Historical Development

Origins in Soliton Studies

In August 1834, Scottish engineer John Scott Russell observed a remarkable solitary wave while conducting experiments on canal boat designs along the Union Canal near . As a boat suddenly halted upon hitting an obstruction, a large, rounded swell detached from the bow and propagated forward at a constant speed of about 8-9 miles per hour, maintaining its shape—a 30-foot-long, 1- to 1.5-foot-high hump—over a considerable distance without dispersion or alteration. Russell termed this a "wave of translation," distinguishing it from typical oscillatory waves, and documented his findings through subsequent experiments in a wave tank, confirming the wave's stability and dependence on water depth and amplitude. He reported these observations formally in 1844 to the British Association for the Advancement of Science, sparking initial interest in such nonlinear wave behaviors. The study of water waves in the had roots in earlier linear theories, but Russell's discovery highlighted the role of nonlinear phenomena in producing stable, localized waves. Pioneering work by in the 1780s derived linearized equations for small-amplitude surface waves, yielding the for wave speed \sqrt{gh} in shallow water, where g is and h is depth. However, observations of real-world waves, including wind-driven and tidal motions, revealed deviations from linearity, such as steepening and breaking, prompting investigations into nonlinear extensions. François Gerstner's 1802 exact solution for finite-amplitude trochoidal waves marked an early nonlinear advancement, balancing centrifugal and gravitational forces without assuming small perturbations. These developments underscored how nonlinearity could interact with dispersion to sustain coherent structures, contrasting with the dissipative tendencies of linear models. Building on Russell's empirical insights, Joseph Boussinesq provided the first mathematical for solitary waves in 1872. In his theory of weakly nonlinear, weakly dispersive waves, Boussinesq derived equations capturing the balance between nonlinear steepening and dispersive spreading, yielding a solitary wave profile of the form \eta = a \sech^2 \left[ \sqrt{\frac{3a}{d^3}} (x - ct) \right], where a is , d is depth, and speed c = \sqrt{gd} \left(1 + \frac{a}{2d}\right). This , developed for shallow-water , explained Russell's observed without full of the primitive Euler equations. Other contemporaries, like George Green and Philip Kelland, attempted similar models in the 1830s-1840s, but Boussinesq's work stood out for its rigor in addressing undular bores and periodic waves akin to solitons. These 19th-century observations and approximations profoundly motivated the pursuit of exact solutions to nonlinear partial differential equations (PDEs) governing wave propagation. Russell's solitary wave, defying expectations of wave dispersion, demonstrated that integrable nonlinear systems could support persistent, particle-like structures, challenging the dominance of linear approximations in hydrodynamics. This empirical drive, coupled with Boussinesq's foundational models, encouraged deeper theoretical exploration of balance mechanisms in nonlinear PDEs, setting the stage for later exact formulations like the 1895 model for shallow water waves.

Key Formulations and Contributors

Interest in the Korteweg–de Vries (KdV) equation, proposed in 1895 to model solitary waves, waned after the early until the , when numerical studies revived attention to phenomena. In 1965, Norman J. Zabusky and Martin D. Kruskal conducted computer simulations of a discretized KdV-like equation in the context of the Fermi–Pasta–Ulam problem on nonlinear lattice dynamics. Their results revealed stable, non-dispersive wave packets—dubbed "solitons" by analogy to particle interactions—that passed through each other with only phase shifts, demonstrating recurrence and integrability-like behavior without energy equipartition. This numerical insight prompted further analytical investigation into exact solutions for the KdV equation. The inverse scattering transform emerged as a powerful analytical tool in the mid-1960s, building on earlier observations of soliton phenomena in nonlinear wave equations. In 1967, Clifford S. Gardner, John M. Greene, Martin D. Kruskal, and Robert M. Miura published a seminal paper introducing the method specifically for solving the initial-value problem of the Korteweg-de Vries (KdV) equation. Their approach drew an explicit analogy to quantum mechanical theory, where the nonlinear evolution is linearized through a direct scattering transform that decomposes the potential into spectral data, followed by of that data and an inverse reconstruction step. This formulation not only yielded exact multi- solutions but also revealed an infinite number of conserved quantities, highlighting the integrability of the KdV equation. A pivotal advancement came in 1968 from Peter D. , who formalized the concept of a Lax pair—consisting of two linear operators whose compatibility condition generates the nonlinear evolution —to rigorously prove integrability for systems like KdV. Lax's framework provided a general criterion: if a nonlinear partial differential can be expressed as the zero-curvature condition of such an operator pair, it admits infinitely many integrals of motion, ensuring solvability via inverse scattering techniques. This operator-theoretic perspective extended the GGKM method beyond KdV, influencing subsequent developments in integrable . In the early 1970s, Mark J. Ablowitz, David J. Kaup, Alan C. Newell, and Harvey Segur significantly broadened the scope of the inverse scattering transform. Their 1973 paper identified a class of nonlinear evolution equations, including the , that are solvable by the method through a unified spectral framework. Building on this, their detailed 1974 study elaborated the underlying the transform, demonstrating its application to equations modeling optical and water wave phenomena, and establishing the Ablowitz-Kaup-Newell-Segur (AKNS) system as a cornerstone for further extensions. The 1970s saw rapid generalizations of the inverse scattering transform to multi-dimensional problems and alternative operator structures. In 1971–1972, and Anatoly B. Shabat introduced a novel non-self-adjoint operator pair, now known as the Zakharov-Shabat system, which enabled exact solutions for the focusing and laid the groundwork for two-dimensional integrable models like the Kadomtsev-Petviashvili equation. Their scheme for integrating nonlinear equations via inverse scattering, detailed across two papers, facilitated applications to self-focusing in and modulated waves. These contributions marked a timeline of expansion, with multi-dimensional extensions appearing by mid-decade, solidifying the transform's role in modern .

Mathematical Framework

Integrable Nonlinear Evolution Equations

The integrable nonlinear evolution equations amenable to solution via the inverse scattering transform are partial differential equations (PDEs) of the general form u_t + N(u, u_x, \dots, u_{xxx}, \dots) = 0, where u = u(x,t) is a real- or complex-valued depending on a single spatial variable x and time t, and N denotes a nonlinear in u and its spatial derivatives. These 1+1-dimensional equations arise in modeling phenomena such as shallow-water waves, optical solitons, and certain quantum theories, where the nonlinearity balances to permit stable wave propagation. Integrability of such PDEs is characterized by the existence of infinitely many conserved quantities (integrals of motion) that are functionally independent and commute under the , allowing the system to be fully diagonalized and solved exactly. This complete integrability manifests in the formation of solutions—localized, stable waves that interact elastically, preserving their individual identities and speeds after collisions, with only shifts occurring. Unlike generic nonlinear PDEs, integrable ones avoid chaotic behavior or formation in finite time due to this rich conservation structure. Faddeev and Takhtajan established that integrability for 1+1-dimensional nonlinear equations can be framed within a formalism, where the equations belong to an infinite hierarchy generated by commuting . A prominent feature is the bi-Hamiltonian structure, involving two compatible (nondegenerate and Poisson-commuting) operators J_0 and J_1, such that the evolution can be expressed as u_t = J_0 \frac{\delta H_1}{\delta u} = J_1 \frac{\delta H_0}{\delta u} for suitable functionals H_0 and H_1. By Magri's theorem, this bi-Hamiltonian setup recursively generates an infinite sequence of conserved quantities in involution, ensuring complete integrability and compatibility with the inverse scattering method. In contrast, non-integrable PDEs, such as the inviscid u_t + u u_x = 0, possess only a finite number of conserved quantities and develop discontinuities (shocks) in finite time due to steepening of nonlinear waves without balancing dispersion. Solutions to such equations typically require perturbative or numerical approximations, like adding small to regularize shocks, and cannot be addressed exactly by scattering. The role of pairs in confirming integrability for the above class of equations is explored in later sections.

Lax Pair Operators

The Lax pair consists of a spatial operator L and a temporal operator M, both depending on the spatial variable x, time t, and the potential u(x,t) of the underlying nonlinear evolution equation. These operators satisfy the Lax equation L_t = [M, L], where [M, L] = M L - L M denotes the , ensuring that the time evolution preserves the spectrum of L. This formulation, introduced by in 1968, provides a linear algebraic condition for the integrability of nonlinear partial differential equations (PDEs). The Lax equation arises as the compatibility condition of an overdetermined linear system for an eigenfunction \psi(x,t): \psi_x = L \psi, \quad \psi_t = M \psi. Differentiating the first equation with respect to t and the second with respect to x yields \psi_{xt} = L_t \psi + L \psi_t = L_t \psi + L M \psi and \psi_{tx} = M_x \psi + M \psi_x = M_x \psi + M L \psi. Equating these for consistency gives L_t + L M = M_x + M L, or equivalently, L_t = M_x + [M, L]. In many one-dimensional cases on infinite domains, the M_x term is negligible or absorbed, reducing to the commutator form L_t = [M, L]. Substituting the explicit forms of L and M (which depend on u and its derivatives) into this condition derives the nonlinear evolution equation u_t + N(u) = 0, where N(u) is a nonlinear differential operator specific to the system. For one-dimensional integrable systems, the operators in the Lax pair are often s. A canonical example from the Ablowitz–Kaup–Newell–Segur (AKNS) framework is the spatial operator L = i \partial_x + q(x,t) \sigma, where \sigma represents (e.g., \sigma_3 = \begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix} for the diagonal part, with off-diagonal terms involving the potential q), and the full form incorporates the spectral parameter. The temporal operator M is similarly a matrix , ensuring the Lax equation holds and generating equations like the . For the , L takes the form of a third-order L = \partial_x^2 + u(x,t), paired with a higher-order M. The Lax pair linearizes the original nonlinear problem by enforcing an isospectral evolution: the eigenvalues of L remain constant in time due to the structure, transforming the nonlinear PDE into a linear problem whose data evolves simply (often multiplicatively). This isospectral flow allows the inverse scattering transform to reconstruct solutions by tracking conserved spectral invariants.

Conditions for Integrability

The inverse scattering transform requires that the underlying (PDE) possesses a of commuting flows, which manifests as an infinite sequence of evolution s sharing the same conserved quantities and evolving independently without . This ensures the of infinitely many integrals of motion, a hallmark of complete integrability, as demonstrated in the KdV where higher-order flows commute with the primary . Additionally, the PDE must admit asymptotic states, where solutions decay sufficiently at spatial infinities to define a well-posed linear problem, enabling the mapping of nonlinear dynamics to linear evolution of data. A key test for such integrability is the Painlevé property, which posits that the general solution of the PDE, when expanded around movable singularities, contains no branch points or essential singularities other than poles. Developed for PDEs by Weiss, Tabor, and Carnevale, this criterion applies to both continuous and discrete nonlinear equations; for continuous cases, it involves checking that all symmetry reductions of the PDE yield ordinary differential equations (ODEs) satisfying the property, while discrete variants extend it to lattice equations via similar singularity analysis. If the Painlevé property holds, it signals potential solvability by inverse scattering, as seen in equations like the where the test confirms the absence of non-pole singularities in Laurent expansions. Symmetry reductions play a crucial role in verifying integrability by transforming the PDE into lower-dimensional ODEs that must individually satisfy the Painlevé property, thereby confirming the equation's global solvability. Complementarily, Bäcklund transformations provide a constructive confirmation by generating new solutions from known ones through a nonlinear , often linking the PDE to a linearizable form or revealing infinite-dimensional algebras indicative of integrability. These transformations, such as the Miura map connecting the KdV and modified KdV equations, not only produce solutions but also underpin the existence of conserved quantities essential for the inverse scattering framework. In higher dimensions, the inverse scattering transform faces significant limitations, as the standard 1+1-dimensional formulation relies on one-dimensional , and most PDEs in 2+1 or greater dimensions lack the necessary commuting hierarchies or decay properties for direct application. To address integrable cases in two dimensions, additional structures like d-bar problems are required, which reformulate the as a nonlinear over the to handle multidimensional data. These extensions, while successful for specific systems like the Kadomtsev-Petviashvili , highlight the rarity of full integrability beyond one spatial dimension without specialized adaptations. The connection to Lax pairs serves primarily for verification, ensuring the pair's yields the original PDE under integrability conditions.

Core Components of the Method

Direct Scattering Transform

The direct scattering transform constitutes the initial phase of the inverse scattering method, wherein the nonlinear potential q(x), representing the initial condition of an integrable nonlinear evolution equation, is mapped to a set of linearizable scattering data through the solution of a linear eigenvalue problem. This process linearizes the inherently nonlinear problem by associating the potential with spectral properties of an associated linear operator, analogous to a Fourier transform but adapted for nonlinear dynamics. Specifically, for systems like the Korteweg–de Vries (KdV) equation, the Lax operator L is the time-independent Schrödinger operator defined as L = -\frac{d^2}{dx^2} + q(x), and the eigenvalue problem L \psi = \lambda \psi is solved, where \lambda = -k^2 parameterizes the spectrum. The solutions to this eigenvalue problem, known as Jost solutions, are constructed based on their asymptotic behavior as |x| \to \infty, assuming the potential q(x) \to 0 sufficiently rapidly to ensure well-defined limits. The Jost solution from the left, \psi_l(x, k), satisfies \psi_l(x, k) e^{-ikx} \to 1 as x \to -\infty, while the Jost solution from the right, \psi_r(x, k), satisfies \psi_r(x, k) e^{ikx} \to 1 as x \to +\infty. These solutions enable the definition of the scattering matrix, which relates the incoming and outgoing waves. For real wavenumbers k, the t(k) and reflection coefficients r_l(k) (from the left) and r_r(k) (from the right) are determined via the Wronskian relations between the Jost solutions, such as t(k) = W(\psi_l, \psi_r) / (2ik), where W denotes the . In symmetric potentials, r_l(k) = -r_r(k) = r(k), and the r(k) encodes the continuous spectrum of the scattering data. The scattering data further comprises discrete eigenvalues corresponding to bound states, which occur for complex k in the appropriate half-plane of analyticity. For the KdV case, these eigenvalues lie on the imaginary axis, k_n = i \kappa_n with \kappa_n > 0, manifesting as poles of the t(k) in the upper half-plane. The Jost functions possess analytic properties in the complex k-plane: \psi_l(x, k) is analytic in the upper half-plane, and \psi_r(x, k) in the lower half-plane, facilitating the identification of bound states through residue analysis. The norming constant \beta_n associated with each discrete eigenvalue k_n is defined as the residue \beta_n = \operatorname{Res}_{k=k_n} \left[ \frac{1}{a(k)} \right], or equivalently \beta_n = \lim_{k \to k_n} (k - k_n) \frac{\psi_l(x, k)}{\psi_r(x, k)}, which quantifies the residue at the pole and directly influences the amplitude of the corresponding in the reconstructed solution. In more general frameworks, such as the Ablowitz–Kaup–Newell–Segur (AKNS) system, the direct scattering problem extends to matrix-valued operators for vector potentials, yielding analogous Jost matrices and scattering data, including off-diagonal reflection coefficients for non-self-adjoint cases like the . The completeness of the scattering data—combining the discrete (eigenvalues and norming constants) with the continuous (reflection coefficient r(k) for real k)—ensures a correspondence with the original potential, up to the integrable class of rapidly decaying functions. This transform preserves the infinite number of conservation laws inherent to the , as the scattering data evolve simply under the nonlinear flow.

Time Evolution of Scattering Data

The time evolution of the scattering data in the inverse scattering transform is a key feature that linearizes the nonlinear dynamics of the underlying (PDE), allowing for explicit computation of the solution at later times. For integrable nonlinear equations admitting a Lax pair , the scattering data—comprising discrete eigenvalues, norming constants, and the —evolves according to a set of linear equations (ODEs). This preserves the structure of the data while transforming the complex nonlinear PDE into a simpler linear problem in the spectral domain. The discrete eigenvalues \kappa_n, associated with bound states or solitons, remain constant in time due to the isospectral nature of the induced by the pair. This constancy arises from the commutativity of the spatial and temporal operators in the Lax formulation, ensuring that the of the spatial operator L is under the PDE's . For the continuous , the reflection coefficient r(k, t) evolves as r(k, t) = r(k, 0) \exp\left(-i \int_0^t \omega(k, s) \, ds\right), where \omega(k) is the specific to the PDE. In the case of the Korteweg–de Vries (KdV) equation, this takes the form \omega(k) = 8k^3, leading to r(k, t) = r(k, 0) \exp(-8i k^3 t). The norming constants, which encode amplitude and phase information for solitons, also undergo simple phase shifts proportional to the . For soliton solutions, the positions and phases of individual solitons evolve deterministically from the discrete scattering data. Specifically, the position parameter \xi_n(t) for the n-th soliton satisfies \xi_n(t) = \xi_n(0) + 4 \kappa_n^2 t, reflecting the soliton velocity $4 \kappa_n^2 derived from the eigenvalue \kappa_n. This linear motion in the spectral domain corresponds to the nonlinear interaction-free propagation of solitons in physical space, with higher-speed solitons (larger \kappa_n) overtaking slower ones without altering their forms. Multi-soliton interactions are thus captured exactly through the time evolution of the full discrete spectrum. The analyticity and properties of the initial scattering data are preserved throughout the evolution, maintaining the data's suitability for the inverse transform. For real-valued potentials in the Schrödinger , symmetries such as r(-k, t) = \overline{r(k, t)} hold at all times, ensuring the remains consistent with the physical reality of the potential. This preservation stems from the unitary evolution in the spectral space governed by the linear ODEs. Furthermore, the time independence of the eigenvalues links directly to the number of conserved quantities in the ; these are given by the traces of even powers of the L, \operatorname{Tr}(L^{2m}), which remain and provide integrals of motion for the nonlinear PDE.

Inverse Scattering Reconstruction

The inverse scattering reconstruction constitutes the final phase of the inverse scattering transform (IST), wherein the potential u(x,t) of the underlying integrable nonlinear evolution is recovered from the time-evolved obtained in prior steps. This maps the spectral information—comprising reflection coefficients, transmission coefficients, and discrete eigenvalues—back to the spatial domain, enabling the explicit solution of the for equations such as the Korteweg–de Vries (KdV) or nonlinear Schrödinger (NLS) . The process ensures that the reconstructed potential satisfies the original nonlinear partial differential , leveraging the integrability conditions to avoid ill-posedness. A foundational approach to this is the Marchenko integral equation, which formulates the as a of the second kind for an auxiliary kernel K(x,y). The equation is given by K(x,y) + F(x+y) + \int_x^\infty K(x,z) F(z+y) \, dz = 0, \quad y \geq x, where F(x+y) is constructed from the and the discrete spectrum (norming constants associated with bound states) via the of r(k) \exp(-8 i k^3 t) plus sum terms \sum \beta_n \exp(- \kappa_n (x+y) + i 8 \kappa_n^3 t). This equation is solved iteratively or numerically for K(x,y), with the solution existing and unique under the condition that the reflection coefficient is square-integrable to ensure convergence. The Marchenko method originates from the one-dimensional inverse scattering theory for the Schrödinger operator and extends to more general Lax operators in IST frameworks. The potential u(x,t) is then directly obtained from the kernel via the relation u(x,t) = -2 \partial_x K(x,x,t), where the derivative is taken with respect to the first argument evaluated on the diagonal. This extraction formula arises from the asymptotic behavior of the Jost solutions and ensures that u(x,t) incorporates both the discrete contributions (manifesting as soliton components) and the continuous (corresponding to dispersive ). For instance, in the KdV , discrete eigenvalues yield sech-squared soliton profiles, while the continuous part reconstructs the scattering tail. The method's efficacy relies on the analytic properties of the scattering data, with stability proven for small perturbations in the data under suitable Sobolev norms. An alternative and often more versatile formulation of the reconstruction employs the Riemann-Hilbert (RH) problem, which reformulates the inverse as a on a suitable contour in the . The RH problem is posed for a -valued M(\zeta; x,t) analytic off the contour \Sigma (typically the real line for the continuous ), satisfying the jump condition \Psi_+(x,\zeta,t) = \Psi_-(x,\zeta,t) G(\zeta,t), \quad \zeta \in \Sigma, where \Psi_\pm denote the boundary values from the upper and lower half-planes, and the jump G(\zeta,t) encodes the evolved data: the off-diagonal entries involve the reflection coefficients for the continuous , while simple poles in G at discrete eigenvalues \zeta_n (in the upper half-plane for bound states) incorporate the norming constants and residue information. The solution M is normalized at infinity, M(\zeta) \to I as |\zeta| \to \infty, and the potential is recovered from the (1,2) entry of M via u(x,t) = 2i \lim_{\zeta \to \infty} \zeta [M_{12}(\zeta; x,t)]. This approach unifies the treatment of discrete and continuous spectra, with the continuous handled by the oscillatory jump on \Sigma and discrete by residue at the poles, facilitating and numerical implementations via steepest descent methods. The RH formulation was rigorously developed for IST in the context of operators and extends to non-self-adjoint cases like the AKNS hierarchy. In both the Marchenko and RH frameworks, the discrete corresponds to bound states, contributing localized structures to u(x,t), while the continuous captures extended, radiative features modulated by the . The reconstruction handles these components separably: discrete eigenvalues remain time-independent under the IST evolution (up to factors), allowing explicit superposition, whereas the continuous data evolves deterministically via exponential factors e^{-i \theta(k,t)}, with \theta the function from the pair. This ensures the reconstructed potential is uniquely determined and asymptotically stable, with error bounds scaling with the data's L^2 norm. For multi- cases, the discrete dominates for large |x|, yielding pure solutions when the continuous part vanishes.

Case Study: Korteweg–de Vries Equation

Defining Lax Operators

In the case study of the Korteweg–de Vries (KdV) equation within the inverse scattering transform, the Lax operators are specifically formulated to linearize the nonlinear evolution, building on the general concept of a Lax pair consisting of spatial and temporal operators whose compatibility yields the target equation. The KdV equation, which models weakly nonlinear dispersive waves such as shallow water solitons, is expressed as u_t + 6 u u_x + u_{xxx} = 0, where u(x,t) represents the wave profile, u_t = \partial u / \partial t, u_x = \partial u / \partial x, and u_{xxx} = \partial^3 u / \partial x^3. The spatial Lax operator L is the one-dimensional time-dependent Schrödinger L = -\partial_x^2 + u(x,t), which acts on eigenfunctions \psi(x,t) via the spectral problem L \psi = \lambda \psi, where \lambda is the spectral parameter. The temporal Lax operator M, which governs the of the eigenfunctions, is the third-order M = -4 \partial_x^3 - 6 u \partial_x - 3 u_x. This ensures that the time dependence of the eigenfunctions satisfies \psi_t = M \psi. The integrability of the KdV equation is confirmed by the zero-curvature condition, or Lax equation, L_t = [M, L] = M L - L M, where the commutator [M, L] is computed explicitly as a acting on test functions. Direct expansion yields [M, L] = -6 u u_x - u_{xxx}. Since L_t corresponds to multiplication by u_t, this matches the KdV evolution u_t + 6 u u_x + u_{xxx} = 0, thus verifying the compatibility of the pair.

Applying Direct Scattering

In the application of the direct scattering transform to the Korteweg–de Vries , the spatial component of the pair serves as the Schrödinger operator L = -\frac{d^2}{dx^2} + u(x,0), where u(x,0) is the initial potential satisfying suitable decay conditions such as \int_{-\infty}^{\infty} (1 + |x|) |u(x,0)| \, dx < \infty. The direct scattering problem centers on solving the associated eigenvalue problem -\psi''(x) + u(x,0) \psi(x) = k^2 \psi(x), for the spectral parameter k \in \mathbb{C}, yielding the continuous and discrete spectral data from which the time evolution can later be inferred. The eigenfunctions are constructed via Jost functions, which are unique solutions to the differential equation with prescribed exponential asymptotics at spatial infinity; the right Jost function, behaving as e^{ikx} for x \to +\infty, admits the Volterra integral representation f(k,x) = e^{ikx} \left( 1 + \int_x^\infty A(x,y) e^{2iky} \, dy \right), where the kernel A(x,y) is uniquely determined by u(x,0) through an auxiliary integral equation ensuring the asymptotic condition. From the Jost functions and their left counterpart (asymptotic to e^{-ikx} as x \to -\infty), the scattering coefficients a(k) and b(k) are obtained via the Wronskian determinant a(k) = \frac{1}{2ik} W(f_l, f_r), with b(k) from the coefficient of the incoming wave in the left asymptotic expansion. The reflection coefficient is then r(k) = b(k)/a(k) for real k, capturing the continuous spectrum and satisfying unitarity |r(k)|^2 + |1/a(k)|^2 = 1. Bound states in the discrete spectrum arise as poles of r(k) in the upper half-plane, specifically at purely imaginary values k = ik_n (\kappa_n > 0) where a(ik_n) = 0, with the number of such states finite and interlacing the zeros of the initial potential's . Each is supplemented by a norming constant \gamma_n = \log |c_n / a'(ik_n)|, where c_n encodes the relative of the Jost functions at the eigenvalue, essential for the in the .

Evolving Scattering Data

In the inverse scattering transform applied to the Korteweg–de Vries (KdV) equation, the of the scattering data occurs linearly, decoupling the nonlinear dynamics into simpler components that correspond to and . This evolution is determined by the compatibility condition of the Lax pair, ensuring that the scattering data at time t can be obtained explicitly from the initial data at t = 0. The discrete part of the , associated with bound states, governs the soliton contributions, while the continuous handles the dispersive . The eigenvalues k_n of the discrete spectrum remain constant in time, preserving the number and characteristics of the solitons. The phases linked to these eigenvalues evolve according to \exp(-8 i k_n^3 t), which modulates the position and phase of each soliton without altering their intrinsic properties. For the continuous spectrum, the reflection coefficient r(k, t) undergoes a simple phase shift: r(k, t) = r(k, 0) e^{-8 i k^3 t}. This evolution arises directly from the time-dependent part of the pair and ensures that the radiation component disperses predictably. The linear evolution manifests in the soliton parameters as follows: each soliton's position shifts with time as x_n(t) = x_n(0) - 4 k_n^2 t, reflecting their uniform motion at speed -4 k_n^2, while the remains fixed at $2 k_n^2. These shifts are determined by the imaginary parts of the eigenvalues and the phase factors, leading to non-interacting propagation despite their nonlinear origins. The dispersive , stemming from the continuous , arises from the oscillating , producing an extended wave train that spreads and decays due to the k^3 inherent in the KdV . This separation highlights the power of the inverse scattering method in resolving the KdV dynamics into asymptotically stable and transient components.

Performing Inverse Scattering

The reconstruction of the solution to the Korteweg–de Vries (KdV) equation from its evolved data is performed using the Marchenko , which solves an to recover the potential underlying the Schrödinger operator in the Lax pair. The evolved data, briefly, consists of the time-dependent r(k,t) and the eigenvalues \{k_n\} with associated norming constants \{\gamma_n\}, which remain constant except for phase factors under KdV evolution. The Marchenko kernel is formed from this data as F(x) = \sum_n \gamma_n e^{2 i k_n x} + \frac{1}{2\pi} \int r(k) e^{2 i k x}\, dk, where the sum runs over the bound states (discrete spectrum) and the integral over the continuous spectrum; the time dependence enters through the evolved r(k,t) and phase-adjusted \gamma_n(t). The function K(x,y,t) is then obtained by solving the Fredholm integral equation K(x,y,t) + F(x+y,t) + \int_x^\infty K(x,z,t) F(z+y,t)\, dz = 0 for y \geq x, which is well-posed under suitable decay conditions on the scattering data. The KdV potential, which is the solution u(x,t), is given by u(x,t) = -2 K_x(x,x,t), where K_x denotes the derivative with respect to the first spatial argument evaluated on the diagonal. This yields the full nonlinear solution incorporating both discrete and continuous contributions. In the pure soliton case, where r(k) = 0 (reflectionless ), the kernel simplifies to a finite sum, and the solution reduces to an N- wave given explicitly by u(x,t) = -2 \partial_x^2 \log \det(I + A), with A an N \times N matrix constructed from the discrete data \{k_n, \gamma_n\} and their evolved phases. Here, I is the , and the determinant form captures the exact nonlinear interactions among the solitons without approximation. The pure soliton scenario describes stable, particle-like waves that interact elastically and separate asymptotically, corresponding to initial conditions producing no in the direct scattering step. In contrast, nonzero r(k) introduces a component, representing dispersive that decay and interact nonlinearly with the solitons via the full Marchenko kernel, leading to more complex dynamics including potential instabilities or long-time asymptotics dominated by solitons.

Broader Applications and Examples

Nonlinear Schrödinger Equation

The (NLSE) models the evolution of slowly varying wave envelopes in nonlinear dispersive media, such as optical fibers or Bose-Einstein condensates. In its standard focusing form, it is expressed as i q_t + q_{xx} + 2 |q|^2 q = 0, where q(x,t) is a complex-valued function representing the wave amplitude, with subscripts denoting partial derivatives with respect to time t and space x. This equation differs fundamentally from the Korteweg–de Vries (KdV) equation, which governs real-valued surface waves and admits a scalar second-order Lax operator; the NLSE requires a first-order matrix formulation to capture its complex nonlinearity and support phenomena like modulation instability. The adaptation of the inverse scattering transform (IST) to the NLSE employs the Zakharov-Shabat (ZS) Lax pair, which linearizes the nonlinear evolution. The spatial operator is L = i \partial_x + \begin{pmatrix} 0 & q \\ -q^* & 0 \end{pmatrix}, where q^* denotes the of q, and the time-evolution operator M ensures compatibility with the NLSE via the zero-curvature condition L_t - M_x + [L, M] = 0. Unlike the Sturm-Liouville operator in the KdV case, this non- Dirac-type system accommodates complex potentials, leading to on the real axis and bound states in the . The direct problem involves solving the eigenvalue L \psi = \lambda \psi for the Jost solutions, yielding the data. The scattering data for the NLSE consists of the \rho(k) = b(k)/a(k) for real wavenumbers k, where a(k) and b(k) are the transmission and reflection coefficients from the Jost functions, and discrete eigenvalues \lambda_j located in the upper half-plane (\operatorname{Im} \lambda_j > 0), each associated with a norming constant. These eigenvalues correspond to solitons, contrasting with the KdV's real-line bound states. The time evolution of the scattering data is simple: \rho(k,t) = \rho(k,0) \exp(-4i k^2 t) for the continuous part, while discrete eigenvalues remain constant, and norming constants acquire phases \exp(2i \lambda_j^2 t). Reconstruction of q(x,t) from the evolved scattering data proceeds via a Riemann-Hilbert problem, formulated in terms of analytic functions in the that match the jump condition across the real axis determined by the scattering data. This yields exact solutions, including multi-soliton forms; for the focusing NLSE, the discrete spectrum produces bright solitons with sech-shaped envelopes that propagate without distortion, while the defocusing variant (-2 |q|^2 q) supports dark solitons featuring intensity dips on a continuous background.

Sine-Gordon Equation

The sine-Gordon equation, a prototypical integrable relativistic field theory model, is expressed in light-cone coordinates x = (X + T)/\sqrt{2} and t = (X - T)/\sqrt{2} as \phi_{xt} = \sin \phi, where \phi(x, t) is a real and subscripts denote partial derivatives. This form underscores the equation's Lorentz invariance, as the light-cone structure preserves , distinguishing it from non-relativistic integrable systems like the . The inverse scattering transform solves the \phi(x, 0) = \phi_0(x) and \phi_t(x, 0) = \psi_0(x) exactly, yielding multi-soliton solutions that maintain topological and dynamical properties under Lorentz transformations. Within the Ablowitz–Kaup–Newell–Segur (AKNS) formalism, the direct scattering problem is posed using a pair in light-cone coordinates. The spatial component of the Lax operator is U = \begin{pmatrix} i\lambda & \sin(\phi/2) e^{-i\theta} \\ -\sin(\phi/2) e^{i\theta} & -i\lambda \end{pmatrix}, where \lambda is the complex spectral parameter and \theta(x, t) is a dynamical phase satisfying \theta_{xt} = 0, often chosen as \theta = \int^x \phi_s(s, t) \, ds to ensure . The associated is \Psi_x = U \Psi, with \Psi a vector . The is governed by a V, and the zero-curvature condition U_t - V_x + [U, V] = 0 reproduces the sine-Gordon equation. This setup maps the nonlinear problem to a linear Zakharov–Shabat scattering problem, where the potential is encoded in the off-diagonal terms involving \phi. The scattering data consist of the r(\lambda) from the continuous spectrum and discrete eigenvalues from bound states. Solitons appear as single kinks or antikinks, corresponding to simple poles on the imaginary axis in the \lambda-plane; a single kink is \phi = 4 \arctan(e^{\gamma (x - v t)}), with velocity v and Lorentz factor \gamma = (1 - v^2)^{-1/2}, reflecting the relativistic nature. solutions, oscillatory bound states of a kink-antikink pair, emerge from eigenvalue pairs off the real axis, exemplified by \phi = 4 \arctan\left( \frac{\sin(\omega t)}{\cosh(\kappa x)} \right) with internal frequency \omega and width \kappa. Time evolution of the scattering data is trivial for discrete eigenvalues (phase factors only) and simple for the , preserving integrability and Lorentz symmetry. Reconstruction proceeds via the inverse scattering step, adapting the Gelfand–Levitan–Marchenko s to the 1+1-dimensional Lorentz-invariant framework. The kernel K(x, y) satisfies an involving the data transformed by the \exp(i (\lambda x - \mu t)), solved to recover the s. The field \phi is then extracted from the of the components, \phi_x = -2i \partial_x \ln(\psi_1 / \psi_2), yielding the full solution. This method constructs general N-kink and configurations while respecting the equation's topological and boost invariance.

Additional Integrable Models

The Toda lattice serves as a discrete analog to the , modeling a chain of particles interacting via exponential springs, and is completely integrable through the inverse scattering transform applied to a tridiagonal matrix. The scattering data for this system typically involves reflectionless potentials, enabling the construction of solutions that preserve the integrable structure under . This approach, originally developed for the finite lattice, extends to infinite chains, highlighting the lattice's role in bridging continuous and discrete integrable models. The modified Korteweg–de Vries (mKdV) equation and the Harry Dym equation emerge as reductions within the KdV hierarchy, both solvable by the inverse scattering transform using spectral problems akin to the Zakharov–Shabat operator. For the mKdV equation, which governs odd solutions in the hierarchy, the scattering data evolves simply under the flow, yielding N-soliton interactions without radiation. The Harry Dym equation, related via a transformation to the KdV, admits an inverse spectral transform that linearizes its evolution, facilitating exact solutions for initial data in appropriate function spaces. The vector nonlinear Schrödinger equation and the Manakov system extend the scalar case to multi-component waves, maintaining integrability through a formulation of the inverse scattering transform. In the Manakov system, which describes polarization dynamics in , the scattering problem involves a 2×2 AKNS-type , with conserved quantities ensuring the of eigenvalues and norming constants. This framework captures vector solitons and their interactions, applicable to birefringent fibers. Two-dimensional extensions, such as the Kadomtsev–Petviashvili (KP) equation, employ the \overline{\partial}-method as a generalization of the inverse scattering transform for (2+1)-dimensional systems. The \overline{\partial}-dressing approach constructs solutions from analytic scattering data on the complex plane, accommodating lump and line-soliton configurations in the KP hierarchy. This method underscores the adaptability of inverse scattering techniques to higher dimensions while preserving complete integrability.

Modern Extensions and Implementations

Physical and Engineering Applications

In , the inverse scattering transform (IST) plays a crucial role in analyzing propagation within optical fibers, where the governs the dynamics of pulses in Kerr . By decomposing the initial pulse into and via direct scattering, IST enables the exact prediction of nonlinear effects such as and interactions over long distances, which is essential for designing dispersion-managed fiber systems that support stable, high-bit-rate data transmission in telecommunication networks. This approach, originally developed by Zakharov and Shabat for the , has been instrumental in mitigating and achieving error-free transmission in submarine cables and metropolitan networks. In plasma physics, IST applied to the Korteweg-de Vries (KdV) equation provides a framework for understanding the formation and stability of ion-acoustic solitons and Langmuir waves in collisionless plasmas. These solitons, which maintain their shape during propagation and collisions, model electrostatic waves where ion inertia balances electron pressure, with IST revealing the discrete spectrum of bound states corresponding to soliton amplitudes. The Gross-Pitaevskii equation, a variant of the NLS equation incorporating trapping potentials, describes the mean-field dynamics of Bose-Einstein condensates (BECs), and IST facilitates the exact solution for matter-wave in these ultracold quantum gases. In one- and multi-component BECs, IST identifies higher-order pole solutions that represent complex soliton structures, such as dark and bright , whose interactions preserve particle number and coherence, enabling precise control of atomic wave packets in optical lattices. This has implications for simulating quantum fluids and observing topological defects in experiments with or sodium atoms cooled to nanokelvin temperatures. Recent applications of IST extend to quantum information processing through integrable spin chains, where the quantum inverse scattering method constructs exact eigenstates for models like the Heisenberg XXX chain, facilitating the simulation of and many-body dynamics on quantum hardware. Post-2020 developments leverage IST in designing integrable quantum circuits with Yang-Baxter gates, enabling efficient computation of correlation functions and error-corrected qubits in noisy intermediate-scale quantum devices.

Numerical Methods and Computational Tools

Numerical methods for the inverse scattering transform (IST) primarily focus on discretizing the equations that arise in the reconstruction phase, such as the Marchenko equations, which relate the scattering data to the underlying potential. These equations are typically solved using approximations or boundary methods to handle the Volterra-type integrals efficiently on a grid. For instance, potential splitting techniques combined with inverse Fourier transforms have been employed to approximate solutions, enabling stable computations for the . Structured matrix algorithms further exploit the Toeplitz-like structure of the discretized Marchenko kernels to achieve fast inversion via Levinson-Durbin recursion, reducing complexity from O(N^3) to O(N^2) for grid size N. Riemann-Hilbert (RH) problems provide an alternative formulation for the inverse step, particularly advantageous for analytic continuations and in IST. Numerical solvers for RH problems often rely on orthogonal methods, where the solution is represented via a sequence of monic polynomials satisfying a derived from the jump conditions across contours. The Fokas-Its-Kitaev approach integrates this with RH representations to yield algorithms for computing partition functions in random matrix theory, adaptable to IST for integrable hierarchies. Iterative techniques, such as the nonlinear steepest descent method, enhance convergence by deforming contours to avoid oscillatory regions, allowing high-accuracy solutions for large-time evolutions in dynamics. Software tools for IST computations have matured, with open-source libraries facilitating direct and inverse transforms across languages. The FNFT library, implemented in C with bindings for and , computes nonlinear Fourier transforms for equations like the nonlinear Schrödinger and Korteweg-de Vries, supporting efficient evaluation of scattering coefficients and Marchenko kernels via layer-peeling algorithms. In , packages like SIE.jl implement IST components for equilibrium measure computations in , leveraging the language's speed for spectral problems. Recent advancements include GPU-accelerated solvers in broader numerical frameworks, though specific IST implementations remain CPU-dominant; for example, 2024 updates to 's SciML ecosystem enable parallelized ODE integrations tied to scattering evolutions, achieving up to 10x speedups on GPUs for related nonlinear simulations. Challenges persist in extending IST to higher dimensions, where complete analytic frameworks are lacking, necessitating approximate methods like multidimensional Riemann-Hilbert reductions or asymptotic expansions to manage increased computational demands. For non-integrable perturbations, such as those in the perturbed KdV equation, IST approximations involve adiabatic invariants or multiple-scale analyses to track soliton evolution, with ensured via modified Marchenko solvers that incorporate damping terms. The core in IST, whether via Marchenko integrals or RH factorization, thus underpins these extensions but requires careful to preserve integrability properties.

References

  1. [1]
    The Inverse Scattering Transform‐Fourier Analysis for Nonlinear ...
    A comprehensive presentation of the inverse scattering method is given and general features of the solution are discussed. The relationship of the scattering ...Missing: key | Show results with:key
  2. [2]
    [PDF] What is... the Inverse Scattering Transform
    In the inverse-scattering transform theory, the reflection coefficient R(λ) is a nonlinear analogue of the Fourier transform of ψ(x). Definition 1 (Direct ...
  3. [3]
    Method for Solving the Korteweg-deVries Equation | Phys. Rev. Lett.
    Method for Solving the Korteweg-deVries Equation. Clifford S. Gardner, John M. Greene, Martin D. Kruskal, and Robert M. Miura. Plasma Physics Laboratory ...Missing: inverse scattering
  4. [4]
  5. [5]
    Inverse Scattering Transform - an overview | ScienceDirect Topics
    The inverse scattering transform is defined as a mathematical procedure that reconstructs a potential from given scattering data, establishing a one-to-one ...
  6. [6]
    [PDF] The Korteweg-de Vries equation: a historical essay
    Boussinesq's work on weakly nonlinear, weakly dispersive wave propagation is developed in three papers (1871 a, b, 1872) and in his 1877 monograph. The first ...Missing: source | Show results with:source
  7. [7]
    [PDF] advancement of science
    ON WAVES. 311. Report on Waves. By J. ScorT RUSSELL, Esq., M.A., F.R.S. Edin., made to the Meetings in 1842 and 1843.... Members of Committee { Sir JOHN ROBISON ...
  8. [8]
    Report on Waves: Made to the Meetings of the British Association in ...
    Aug 8, 2009 · Report on Waves: Made to the Meetings of the British Association in 1842-43. by: John Scott Russell. Publication date: 1845. Collection ...Missing: translation | Show results with:translation
  9. [9]
    [PDF] THE ORIGINS OF WATER WAVE THEORY
    Joseph-. Louis Lagrange (1781, 1786), perhaps independently, derived the linearized gov- erning equations for small-amplitude waves, and obtained the solution ...
  10. [10]
    [PDF] 1 Brief History and Overview of Nonlinear Water Waves - Elsevier
    In this way we hope to enhance our understanding of nonlinear water wave dynamics. 1.2 The Nineteenth Century. It is safe to say that the systematic study of ...
  11. [11]
    Joseph Boussinesq's legacy in fluid mechanics - ScienceDirect.com
    Solitary waves. The most important of Boussinesq's early works in hydrodynamics concerned wave propagation on the surface of a liquid. For plane sine waves of ...
  12. [12]
    Integrals of nonlinear equations of evolution and solitary waves
    10 Peter, D. Lax, Integrals of nonlinear equations of evolution and solitary waves, New York Univ., Courant Inst. Math. Sciences, Report NYO-1480-87, Jan. 1968.
  13. [13]
    Nonlinear-Evolution Equations of Physical Significance
    Jul 9, 1973 · Nonlinear-Evolution Equations of Physical Significance. Mark J. Ablowitz, David J. Kaup, Alan C. Newell, and Harvey Segur.
  14. [14]
    The Inverse scattering transform fourier analysis for nonlinear ...
    Dec 1, 1974 · The Inverse scattering transform fourier analysis for nonlinear problems · M. Ablowitz, D. Kaup, +1 author. H. Segur · Published 1 December 1974 ...
  15. [15]
    [PDF] The Inverse Scattering Transform and Integrability of Nonlinear ...
    In this expostion, we will give a semi-historical account of some of this work, specifically concerning the inverse scattering transform, which among other ...
  16. [16]
    [PDF] INTEGRABLE SYSTEMS - DAMTP
    May 10, 2012 · These soliton solutions will be constructed by means of an inverse problem: recovering a potential from the scattering data. 2.1 History of two ...
  17. [17]
    Hamiltonian Methods in the Theory of Solitons
    **Summary of Conditions for Integrability, Bi-Hamiltonian Structure, General Form, and Role in Inverse Scattering (Faddeev and Takhtajan)**
  18. [18]
    [PDF] BiHamiltonian systems
    In 1978 Magri, [7], proved a remarkable theorem that, subject to certain technical hypotheses, any such biHamiltonian system is completely integrable in the ...
  19. [19]
    [PDF] Inverse Scattering Transform and the Theory of Solitons
    Glossary. AKNS method A method introduced by Ablowitz, Kaup,. Newell, and Segur in 1973 that identifies the nonlinear partial differential equation (NPDE) ...
  20. [20]
    [PDF] KdV equation, direct and inverse scattering
    Dec 29, 2024 · The main goal of these notes is to study integrable sytems with finite and infinite degree of freedoms. We will first study the inverse ...<|control11|><|separator|>
  21. [21]
  22. [22]
    The Riemann–Hilbert Problem and Inverse Scattering
    This provides an effective way of treating the inverse scattering problem for first-order systems.
  23. [23]
  24. [24]
    Solitons and the Inverse Scattering Transform
    Using these ideas Ablowitz, Kaup, Newell and Segur (19736) and (1974) developed a method to find a rather wide class of nonlinear evolution equations ...Missing: paper | Show results with:paper
  25. [25]
    [PDF] Solitons and the Inverse Scattering Transform - DTIC
    The KdV equation governs the evolution on a slow time-scale of a small amplitude wave that satisfied the linear wave equation on a fast time-scale. To stop at 0 ...
  26. [26]
    [PDF] EXPLICIT SOLUTIONS TO THE KORTEWEG-DE VRIES EQUATION ...
    A pure N-soliton solution to the KdV equation can be written explicitly as [4,5,7,8] u(x, t) = −2. ∂2 log (det Γ(x;t)). ∂x2. ,. (1.2) where Γ(x;t) is the N ...Missing: \partial_x^ | Show results with:\partial_x^
  27. [27]
    Exact Theory of Two-dimensional Self-focusing and One-dimensional Self-modulation of Waves in Nonlinear Media
    - **Citation**: Zakharov, V. E., & Shabat, A. B. (1972). Exact Theory of Two-dimensional Self-focusing and One-dimensional Self-modulation of Waves in Nonlinear Media. *Journal of Experimental and Theoretical Physics*, 34, 62.
  28. [28]
    Integrals of the Toda lattice | Phys. Rev. B
    Feb 15, 1974 · The exponential lattice introduced by Toda is shown to be an integrable dynamical system. An explicit set of n integrals is given for a lattice of n particles.
  29. [29]
  30. [30]
    A numerical solution to the Zakharov-Shabat inverse scattering ...
    A numerical solution to the Zakharov-Shabat inverse scattering problem with application to solitary wave propagation in nonlinear optical fibers☆. Author ...
  31. [31]
    The Inverse Scattering Problem for the Zakharov–Shabat System
    'The Inverse Scattering Problem for the Zakharov–Shabat System' published in 'Integrable Hamiltonian Hierarchies'
  32. [32]
    Nonlinear spectral analysis of ion acoustic solitons arising from a ...
    Nov 17, 2022 · The homogeneous KdV equation, even though it is nonlinear, supports an exact spectra found through the inverse scattering transform (IST).21,22 ...
  33. [33]
    Ion-acoustic solitons in a relativistic Fermi plasma at finite temperature
    Nov 6, 2024 · In the former, the Korteweg–de Vries (KdV) equation is derived to describe the evolution of one-dimensional solitons. However, the KdV equation ...
  34. [34]
    Multiple higher-order poles solutions in spinor Bose-Einstein ... - arXiv
    Feb 13, 2024 · In this study, we explore multiple higher-order pole solutions in spinor Bose--Einstein condensates. These solutions are associated with different pairs of ...
  35. [35]
    Inverse Scattering Transform for Nonlinear Schrödinger Systems on ...
    Jun 12, 2023 · Both the direct and the inverse problems make use of the first operator in the Lax pair, the “scattering problem”, while the time evolution is ...
  36. [36]
    [PDF] arXiv:2406.12695v3 [quant-ph] 6 Oct 2024
    Oct 6, 2024 · In this paper, we address the problem of Yang-Baxter integrability of doubled quantum circuit of qubits (spins 1/2) with open boundary conditions where the ...
  37. [37]
    [PDF] arXiv:2307.16781v3 [quant-ph] 8 Feb 2024
    Feb 8, 2024 · The Leningrad school of Faddeev constructed a systematic method to solve integrable quantum systems, called the quan- tum inverse scattering method, where the ...
  38. [38]
    Full wave 3D inverse scattering transmission ultrasound tomography ...
    Nov 19, 2020 · We present here a quantitative ultrasound tomographic method yielding a sub-mm resolution, quantitative 3D representation of tissue characteristics in the ...
  39. [39]
    Whole-Body Imaging Using Low Frequency Transmission Ultrasound
    However, here we show high-resolution transmission ultrasound images resulting from clinical data (i.e., not simulation) using an inverse scattering algorithm.
  40. [40]
    [PDF] Potential splitting and numerical solution of the inverse scattering ...
    In Section 2 we describe an explicit method that involves solving the Marchenko equation ... mated by using a standard inverse discrete Fourier transform, ...
  41. [41]
    Structured matrix algorithms for inverse scattering on the line
    In this article the Marchenko integral equations leading to the solution of the inverse scattering problem for the 1-D Schrödinger equation on the lin.
  42. [42]
    Riemann-Hilbert Problems, Their Numerical Solution and the ...
    In this dissertation, the applied theory of Riemann–Hilbert problems, using both Hölder and. Lebesgue spaces, is reviewed. The numerical solution of RHPs is ...
  43. [43]
    [PDF] arXiv:2302.12930v4 [math.NA] 17 Jan 2024
    Jan 17, 2024 · Our approach exploits the Fokas–Its–Kitaev Riemann–Hilbert representation of the orthogonal polynomials to produce an O(N) method to compute the ...
  44. [44]
    [PDF] Nonlinear steepest descent and the numerical solution of Riemann ...
    May 8, 2012 · Abstract. The effective and efficient numerical solution of Riemann–Hilbert problems has been demonstrated in recent work.
  45. [45]
    FastNFT/FNFT: Fast numerical computation of (inverse) nonlinear ...
    FNFT is a software library for the numerical computation of (inverse) nonlinear Fourier transforms, which are also known as (inverse) scattering transforms.
  46. [46]
    Full length article: Computation of equilibrium measures
    Aug 9, 2025 · We present a method to compute the inverse scattering transform (IST) ... The Julia software package SIE.jl implements our method with a ...
  47. [47]
    GPU Acceleration of Julia's SciML: ODEs, Optimization, and more
    Feb 12, 2025 · GPU Acceleration of Julia's SciML: ODEs, Optimization, and more | Smith, Smith | JuliaCon 2024 · Comments.
  48. [48]
    THE INVERSE SCATTERING TRANSFORM, NONLINEAR WAVES ...
    Oct 1, 2017 · Lax formalism, neither a higher dimensional soliton (local) nor a fully satisfactory inverse scattering theory has been worked out.
  49. [49]
    [PDF] Generalised Fourier Transform and Perturbations to Soliton Equations
    Jul 12, 2009 · Then we present the perturbation theory for the Camassa-Holm hierarchy. 2 Basic facts for the inverse scattering method for the KdV hierarchy.