Fact-checked by Grok 2 weeks ago

List of nonlinear partial differential equations

Nonlinear partial differential equations (PDEs) are mathematical equations that relate a function of multiple independent variables to its partial derivatives, where the highest-order derivatives appear nonlinearly, often modeling complex phenomena in physics, , and that linear PDEs cannot capture. A list of such equations serves as a reference compilation of prominent examples, highlighting their forms, applications, and analytical challenges, drawn from fields like , wave propagation, and reaction-diffusion systems. These equations are fundamental in describing real-world processes where interactions lead to emergent behaviors, such as or formation, and their study often involves advanced techniques like methods, numerical simulations, or laws due to the lack of general superposition principles. Unlike linear PDEs, nonlinear ones can exhibit phenomena like shock waves, blow-up solutions, or traveling waves, making them central to modern . Notable examples include (u_t + u u_x = \nu u_{xx}), a simplified model for viscous fluid flow that illustrates shock formation and diffusion; the Korteweg-de Vries (KdV) equation (u_t + u u_x + u_{xxx} = 0), which governs shallow water waves and supports soliton solutions; and (u_t = u_{xx} + u(1 - u)), used in population genetics to describe reaction-diffusion processes with traveling wavefronts. Other canonical cases are the (i \psi_t + \psi_{xx} + |\psi|^2 \psi = 0), modeling optical solitons and Bose-Einstein condensates, and the Navier-Stokes equations for incompressible fluids (\partial_t \mathbf{u} + (\mathbf{u} \cdot \nabla) \mathbf{u} = -\nabla p + \nu \Delta \mathbf{u}, \nabla \cdot \mathbf{u} = 0), which remain unsolved in three dimensions for general smooth solutions.

Introduction

Definition and Characteristics

A (PDE) involves an unknown and its partial derivatives that appear in a nonlinear fashion, such as through products of the with its derivatives, powers of the or derivatives, or within other nonlinear s. This contrasts with linear PDEs, where the equation takes the form of a of the and its derivatives with coefficients depending only on independent variables. Key characteristics of nonlinear PDEs include the absence of the , which holds for linear PDEs and allows linear combinations of solutions to yield new solutions. Instead, nonlinearity introduces phenomena such as shock waves from intersecting characteristics, localized solitons that preserve upon interaction, and blow-up solutions where the solution becomes unbounded in finite time. These features arise because wave speeds or propagation depend on the solution amplitude itself, leading to breakdowns in smooth solutions that do not occur in linear cases like the or equations. Nonlinear PDEs are categorized by the location of nonlinearity: semilinear PDEs are linear in highest-order derivatives but include nonlinear source terms in the function; quasilinear PDEs are linear in highest-order derivatives with coefficients depending on the function and lower-order derivatives; and fully nonlinear PDEs feature nonlinearity directly in the highest-order derivatives. This classification aids in analyzing solvability and stability, assuming prior knowledge of linear PDEs where superposition simplifies solution construction. Early recognition of nonlinear PDEs appeared in 18th- and 19th-century works by Euler and Lagrange on and variational principles, which generated such equations through Euler-Lagrange formulations. Modern study accelerated post-1950s with computational advances enabling numerical simulations and the exploration of integrable systems via symmetry methods pioneered by .

Importance and Applications

Nonlinear partial differential equations (PDEs) are crucial for modeling complex, real-world phenomena where linear approximations break down, such as in flows, pattern formation in biological systems, and relativistic effects in gravitational dynamics. These equations capture the intricate interactions that govern natural processes, providing essential insights into systems exhibiting emergent behaviors not predictable by simpler linear models. In physics, they describe wave propagation and shock formations; in , they model and growth; and in engineering, they inform designs for under nonlinear stresses. Key applications span multiple disciplines. In , nonlinear PDEs like the Navier-Stokes equations simulate shock waves and turbulent flows critical for and weather prediction. In , they model waves that represent stable particle-like structures in nonlinear media, aiding advancements in and . In biology, reaction-diffusion equations underpin , explaining in animal coats and embryonic development through mechanisms like Turing instability. These applications highlight how nonlinear PDEs enable quantitative predictions in scenarios where superposition principles fail. Solving nonlinear PDEs presents significant challenges, primarily due to the absence of general and theorems that apply universally, unlike their linear counterparts. While linear PDEs often yield explicit via or Fourier transforms, nonlinear ones lack such universality, leading to difficulties in proving , , or long-term behavior. Common approaches include methods for small nonlinearities, numerical simulations using finite element or schemes for approximate , and specialized techniques like inverse scattering for integrable cases such as the Korteweg-de Vries equation. These methods are indispensable but computationally intensive, often requiring for practical implementation. As of 2025, modern developments have integrated and to address these challenges, with (PINNs) emerging as a powerful tool for approximating solutions to nonlinear PDEs from data. PINNs embed PDE constraints directly into training, enabling efficient solving of forward and inverse problems in high dimensions and significant reductions in computational time compared to traditional methods. Additionally, links to data-driven discovery allow for uncovering hidden PDE structures from observational data, enhancing applications in emerging fields like climate modeling—where nonlinear PDEs simulate atmospheric turbulence and ocean currents—and quantum computing simulations, which leverage quantum algorithms to tackle intractable nonlinear dynamics in weather prediction and molecular interactions. These advancements, combined with quantum-inspired methods, promise to revolutionize the analysis of nonlinear PDEs in complex systems.

Alphabetical List

A–F

The Allen–Cahn equation models the dynamics of and domain coarsening in binary alloys and other materials with conserved order parameters. It is a nonlinear reaction-diffusion typically formulated in one spatial dimension plus time (1+1 dimensions) as u_t = \epsilon^2 u_{xx} + u - u^3, where u represents the concentration difference between two phases, \epsilon is a small controlling the interface width, and the nonlinear term u - u^3 arises from a driving phase transitions. This equation captures the motion of antiphase boundaries through curvature-driven flow and has been pivotal in phase-field modeling for microstructure evolution. The Bateman-Burgers equation extends the classical by incorporating a variable diffusion coefficient, often used in to study turbulent flows and shock waves with position-dependent . In 1+1 dimensions, it takes the form u_t + u u_x = \nu(x) u_{xx}, where \nu(x) is the diffusion term, allowing for more realistic modeling of inhomogeneous compared to constant-viscosity cases. This variant highlights the interplay between nonlinear advection and in propagating disturbances. The Benjamin–Bona–Mahony equation provides a regularized long-wave for unidirectional wave propagation in nonlinear dispersive media, such as shallow water or nonlinear lattices, avoiding the ill-posedness of certain systems. Defined in 1+1 dimensions as u_t + u_x + u u_x - u_{xxt} = 0, it features a mixed term that introduces , leading to stable solitary wave solutions. Its applications include modeling and ion-acoustic waves in plasmas. The Benjamin–Ono equation describes the evolution of internal waves in stratified fluids, particularly in deep water where the involves the . In 1+1 dimensions, it is expressed as u_t + H(u_{xx}) + u u_x = 0, with H denoting the , which captures nonlocal effects essential for wave stability. This equation is integrable and exhibits interactions relevant to and atmospheric dynamics. The Born–Infeld equation reformulates classical electrodynamics to impose a finite speed of propagation for electromagnetic signals, addressing singularities in . In 1+1 dimensions, it appears as a nonlinear for the E: \left(1 - E^2\right)^{3/2} E_{tt} - E_{xx} = 0, bounding the field strength to prevent infinite energies. Originally proposed for relativistic electron theory, it influences modern studies in nonlinear optics and string theory. The Boussinesq equation approximates the propagation of surface water waves in shallow channels, incorporating both nonlinearity and dispersion for improved accuracy over linear theories. A common 1+1 dimensional form is the "good" Boussinesq equation: u_{tt} - u_{xx} - u_{xxxx} - 3(u^2)_{xx} = 0, which supports cnoidal waves and solitons. It is fundamental in and hydraulic modeling. The Bretherton equation governs the stability and dynamics of thin liquid films flowing down inclined surfaces under , capturing the onset of fingering instabilities due to nonlinear effects. In 1+1 dimensions for film thickness h: h_t + (h^3 + 3 h^2 \delta h_x)_x = 0, approximately, where \delta relates to slip length. This model is key in and processes. Burgers' equation serves as a canonical one-dimensional model for the formation and dissipation of shock waves in compressible fluids, simplifying the Navier-Stokes equations while retaining essential nonlinear and diffusive features. In 1+1 dimensions: u_t + u u_x = \nu u_{xx}, with \nu > 0 as viscosity; in the inviscid limit (\nu \to 0), it develops discontinuities. It is widely used in turbulence theory and traffic flow simulations. The Cahn–Hilliard equation describes the conserved dynamics of phase separation in binary mixtures, emphasizing mass conservation through a fourth-order structure. In arbitrary spatial dimensions d, it is u_t = \nabla \cdot (M(u) \nabla \mu), \quad \mu = -\gamma \Delta u + f'(u), where M(u) is mobility, \gamma interfacial energy, and f(u) the free energy density (e.g., double-well). Applications span alloy solidification and polymer blends. The Camassa–Holm equation models shallow water waves with peaked solitons (peakons) that preserve particle-like properties under collisions, arising from reanalysis of the Green-Naghdi equations. In 1+1 dimensions: u_t - u_{xxt} + 3u u_x = 2 u_x u_{xx} + u u_{xxx}, or equivalently in nonlocal form involving a Helmholtz operator. It is integrable and applied to tsunamis and . The Chafee–Infante equation, a reaction-diffusion in bounded domains, illustrates bistable dynamics where stable uniform states compete, often studied in 1+1 dimensions as u_t = u_{xx} + u(1 - u)(u - a), with $0 < a < 0.5 for bistability. Though ODE-like in zero spatial dimensions for equilibria, its PDE form analyzes front propagation in combustion and ecology. The Degasperis–Procesi equation generalizes the for shallow water waves, featuring a higher nonlinearity parameter and also supporting peakons. In 1+1 dimensions: u_t - u_{xxt} + 4u u_x = 3u_x u_{xx} + u u_{xxx}, it is completely integrable via inverse scattering. Used in modeling breaking waves and orbital stability. The is an integrable nonlinear PDE exhibiting soliton solutions through its connection to the via reciprocal transformations. In 1+1 dimensions: u_t = u^3 u_{xxx}, it preserves positivity and is linked to random matrix theory. Applications include statistical mechanics of eigenvalues. The Eikonal equation arises as the high-frequency limit of wave equations in geometrical optics and as a Hamilton-Jacobi equation in control theory. In arbitrary dimensions: |\nabla u| = n(\mathbf{x}), where n(\mathbf{x}) is the refractive index, it determines ray paths via characteristics. Solved by fast marching methods in imaging and seismology. The Einstein field equations form the core of general relativity, coupling spacetime geometry to matter-energy distribution through nonlinear tensor relations. In 3+1 dimensions (or general d+1): R_{\mu\nu} - \frac{1}{2} R g_{\mu\nu} = \frac{8\pi G}{c^4} T_{\mu\nu}, where R_{\mu\nu} is the Ricci tensor, R the scalar curvature, g_{\mu\nu} the metric, G gravitational constant, c speed of light, and T_{\mu\nu} the stress-energy tensor. They predict black holes, gravitational waves, and cosmology. The inviscid describe the motion of ideal fluids without viscosity, balancing nonlinear convection with pressure gradients. In 1+3 dimensions for velocity \mathbf{u} and pressure p: \partial_t \mathbf{u} + (\mathbf{u} \cdot \nabla) \mathbf{u} = -\nabla p, \quad \nabla \cdot \mathbf{u} = 0, with incompressibility; they support vortex dynamics and shocks. Central to and astrophysical flows.

G–K

The is a fundamental model in superconductivity and superfluidity, describing the behavior of an order parameter ψ near critical points. In its time-dependent form for one time and three spatial dimensions, it reads: \partial_t \psi = \nabla^2 \psi + \psi (1 - |\psi|^2) This equation captures phase transitions and vortex dynamics in type-II superconductors. The Gross–Pitaevskii equation models the dynamics of Bose–Einstein condensates, representing the macroscopic wave function of weakly interacting bosons. For one time and n spatial dimensions, its standard form is: i \hbar \partial_t \psi = -\frac{\hbar^2}{2m} \nabla^2 \psi + V \psi + g |\psi|^2 \psi Here, V denotes an external potential, and g the interaction strength; solutions exhibit solitons and vortices in trapped condensates. The Hasegawa–Mima equation governs drift wave turbulence in plasmas, particularly ion acoustic and electrostatic modes in magnetized settings. In one time and three spatial dimensions, it takes the form: \partial_t (\nabla^2 \phi - \phi) + [\phi, \nabla^2 \phi] + V_{0x} \partial_y \phi = 0 where [·,·] denotes the Poisson bracket and V_{0x} a constant shear flow; it predicts zonal flows and anomalous transport in tokamaks. The Hunter–Saxton equation describes the dynamics of liquid crystal directors, focusing on orientation fields in nematic phases. For one time and one spatial dimension, it is given by: u_t + u u_x = \frac{1}{2} \int_0^x u_x^2 \, ds This integrable system exhibits wave breaking and singular solutions, relevant to defect formation in aligned molecules. The Kadomtsev–Petviashvili (KP) equation extends the KdV model to two spatial dimensions, analyzing the stability of long waves with weak transverse perturbations. In one time and two spatial dimensions, the standard form is: (u_t + 6 u u_x + u_{xxx})_x + 3 \sigma^2 u_{yy} = 0 with σ = ±1 distinguishing KP-I and KP-II hierarchies; it supports line solitons and resonant interactions in water waves and ion-acoustic plasmas. The Kaup–Kupershmidt equation is an integrable fifth-order KdV-type system, arising in the classification of nonlinear evolution equations with bi-Hamiltonian structure. For one time and one spatial dimension, it reads: u_t + u_{xxxxx} + 10 u u_{xxx} + 25 u_x u_{xx} + 20 u^2 u_x = 0 This equation admits multi-soliton solutions via inverse scattering and models higher-order dispersive waves. The Korteweg–de Vries (KdV) equation models shallow water solitons and unidirectional wave propagation with dispersion and nonlinearity balancing. In one time and one spatial dimension, the canonical form is: u_t + 6 u u_x + u_{xxx} = 0 A modified variant replaces the quadratic nonlinearity: v_t + 6 v^2 v_x + v_{xxx} = 0 These exhibit exact N-soliton solutions and infinite conservation laws, applicable to ion-acoustic waves and lattice vibrations. The Kuramoto–Sivashinsky equation captures chaotic pattern formation in combustion fronts and thin film flows, combining instability and stabilizing diffusion. For one time and n spatial dimensions, it is: u_t + \nabla^2 u + \nabla^4 u + \frac{1}{2} |\nabla u|^2 = 0 Solutions display spatiotemporal chaos and cellular structures, with low-dimensional attractors in bounded domains.

L–Q

The Landau–Lifshitz equation models the precessional and damping dynamics of magnetization in ferromagnetic materials under an effective magnetic field. In its standard form, the equation reads \partial_t \mathbf{m} = -\gamma \mathbf{m} \times \mathbf{H}_{\mathrm{eff}} + \alpha \mathbf{m} \times \partial_t \mathbf{m}, where \mathbf{m} is the unit magnetization vector, \gamma > 0 is the , \mathbf{H}_{\mathrm{eff}} incorporates , , and external fields, and \alpha \geq 0 is the Gilbert damping coefficient. This 1+3 dimensional PDE, first derived in 1935, captures micromagnetic phenomena such as spin waves and domain wall motion on nanosecond timescales. The Liouville equation arises in the study of conformal metrics on surfaces with prescribed constant . It takes the form \Delta u = -K e^{2u}, where u is a scalar potential function, \Delta is the Laplace–Beltrami operator, and K is a constant (often normalized to \pm 1). This nonlinear elliptic PDE, dating to 1853, enables explicit parametrization of surfaces like pseudospheres (K < 0) or catenoids, with solutions expressible via logarithms of Jacobi elliptic functions for bounded domains. Its any-dimensional variants appear in complex analysis for prescribing curvatures on Riemann surfaces. The Monge–Ampère equation is a fully nonlinear elliptic PDE central to convex analysis, Kähler geometry, and optimal transport problems. The canonical form is \det(D^2 u) = f(x), where D^2 u is the Hessian matrix of the convex function u: \mathbb{R}^n \to \mathbb{R}, and f > 0 is a given density function. Originating in for minimal surfaces, modern applications include solving the Monge–Kantorovich mass transport problem, where solutions yield optimal maps between probability measures via \nabla u. Regularity theory ensures C^{1,\alpha} solutions under strict convexity assumptions. The Navier–Stokes equations describe the motion of viscous, incompressible fluids in three spatial dimensions plus time. The momentum equation is \partial_t \mathbf{u} + (\mathbf{u} \cdot \nabla) \mathbf{u} = -\nabla p + \nu \nabla^2 \mathbf{u} + \mathbf{f}, coupled with the incompressibility condition \nabla \cdot \mathbf{u} = 0, where \mathbf{u} is the velocity field, p is , \nu > 0 is kinematic , and \mathbf{f} represents external forces. Formulated in the 1820s–1840s, these equations model phenomena like and boundary layers, but global existence and smoothness for all smooth initial data in 3D remain unsolved, earning a $1 million Millennium Prize. The cubic nonlinear Schrödinger equation governs envelope solitons in nonlinear wave propagation. In normalized 1+1 dimensions, it is i \partial_t \psi + \partial_{xx} \psi + 2 |\psi|^2 \psi = 0, where \psi is a complex amplitude. Derived in 1973 for anomalous in optical fibers, it predicts self-focusing and bright solitons stable under perturbations. In quantum many-body systems, this form underlies the for Bose–Einstein condensates at zero temperature, describing superfluid dynamics and vortex formation in dilute gases. The derivative nonlinear Schrödinger equation extends the cubic model with self-steepening terms relevant to femtosecond pulse propagation in birefringent fibers. A key variant is i q_t + q_{xx} + 2 i |q|^2 q_x + |q|^2 q = 0, where q is the complex envelope. Introduced in 1979 for plasma physics, it applies in nonlinear optics to circularly polarized waves, yielding chirped solitons and peakons via inverse scattering, with applications to high-intensity laser filamentation. The porous medium equation models slow diffusion processes where permeability depends on saturation, such as gas flow through soils. In n spatial dimensions plus time, it is u_t = \Delta (u^m), with m > 1 yielding degenerate nonlinearity (diffusivity vanishes at u=0). First analyzed mathematically in the 1950s, it exhibits finite-speed propagation and self-similar Barenblatt solutions u \sim t^{-k} (C - \xi^2)_+^{1/(m-1)}, applied to glaciology, combustion fronts, and population dynamics with Allee effects.

R–Z, α–ω

The is a that governs the evolution of a Riemannian g_{ij} on a manifold, given by \partial_t g_{ij} = -2 R_{ij}, where R_{ij} is the tensor. Introduced by Richard Hamilton in 1982, this equation deforms the metric in a manner analogous to the , smoothing out irregularities in the geometry while preserving volume when normalized. It has applications in , particularly for studying the topology of manifolds. The Ricci flow played a pivotal role in Grigori Perelman's proof of the in 2002–2003, where techniques were used to handle singularities, demonstrating that every simply connected, closed is homeomorphic to the . The Seiberg–Witten equations form a system of nonlinear partial differential equations in gauge theory, defined on a 4-manifold with a U(1) connection A and a spinor \phi as D_A \phi = 0, \quad F^+_A = \sigma(\phi), where D_A is the Dirac operator, F^+_A is the self-dual part of the curvature 2-form, and \sigma(\phi) is a quadratic map from the spinor bundle. Developed by Nathan Seiberg and Edward Witten in 1994, these equations arise from the low-energy effective action of \mathcal{N}=2 supersymmetric Yang–Mills theory and provide invariants for smooth 4-manifolds, simplifying earlier Donaldson invariants. They are elliptic and gauge-invariant, enabling the study of monopole moduli spaces and Donaldson-type theorems with reduced technical complexity. The sine–Gordon equation is a nonlinear wave equation in 1+1 dimensions, \phi_{tt} - \phi_{xx} + \sin \phi = 0, exhibiting soliton solutions such as and breathers due to its integrability via . First appearing in the Frenkel–Kontorova model for crystal dislocations in 1938–1939, it models topological defects and has applications in for bosonic solitons and in condensed matter for phenomena like Josephson junctions in superconductors, where the sine term arises from nonlinear periodic potentials. The Swift–Hohenberg equation describes in systems near the onset of , typically in any spatial dimension but often studied in 1D or 2D, \partial_t u = r u - (1 + \nabla^2)^2 u - g u^3, with r as a control parameter for and g > 0 for supercriticality. Derived by Jack Swift and Pierre Hohenberg in 1977 to model , it captures the emergence of striped or hexagonal patterns through linear at 1, stabilized by higher-order . This amplitude equation framework highlights universal behaviors in nonequilibrium systems, such as Turing patterns in reaction- models. The source-free Yang–Mills equations are a system of nonlinear partial differential equations for on in any dimension, D_\mu F^{\mu\nu} = 0, where F^{\mu\nu} is the curvature 2-form and D_\mu is the . Formulated by Chen Ning Yang and Robert Mills in 1954 as a non-Abelian generalization of Maxwell's equations, they describe the dynamics of gauge fields in quantum field theory, underlying the strong and electroweak interactions in the Standard Model. Instanton solutions and moduli spaces of these equations reveal topological structures, with applications in Yang–Mills mass gap problems and Donaldson–Uhlenbeck–Yau theorem for stable bundles. The Zakharov system models the interaction of high-frequency Langmuir waves with low-frequency ion acoustic waves in unmagnetized in 1+3 dimensions, i E_t + \nabla^2 E = n E, \quad n_{tt} - \nabla^2 n = \nabla^2 |E|^2, where E is the envelope and n the perturbation. Introduced by Vladimir Zakharov in 1972, it captures wave collapse and self-focusing phenomena, leading to singularities in finite time under certain initial conditions. This coupled system is integrable in 1+1 dimensions but exhibits dispersive turbulence in higher dimensions, with relevance to laser-plasma interactions and . The \phi^4 theory refers to the nonlinear for a in 1+1 or higher dimensions, \square \phi + \frac{\partial V}{\partial \phi} = 0, \quad V(\phi) = \frac{\lambda}{4} (\phi^2 - v^2)^2, with enabling and kink solitons. A cornerstone of since the 1950s, it serves as a prototype for renormalizable interacting theories, modeling analogs and phase transitions. In formulation, it provides a PDE framework for constructive QFT in , addressing triviality bounds and divergences through quantization. For completeness, the Broer–Kaup equations extend shallow-water models in 1+1 dimensions to bidirectional long waves, p_t - q_{xx} + p p_x = 0, \quad q_t + p_x = 0, preserving Hamiltonian structure and integrable via Hirota method. Developed by Leo Broer and David Kaup in the late 1970s as a higher-order Boussinesq-type system, they describe nonlinear dispersive waves without breaking, applicable to coastal engineering and tsunami propagation. The Dodd–Bullough equation is an integrable model in quantum field theory, featuring exponential nonlinearities in a relativistic scalar field. In 1+1 dimensions, it takes the Klein-Gordon form u_{tt} - u_{xx} = e^{u} + e^{-2u}, admitting soliton solutions through Bäcklund transformations. Introduced by Roger Dodd and Robin Bullough in 1970 as a variant of the sine-Gordon equation with asymmetric exponential terms, it arises in integrable hierarchies and affine Toda theories, with connections to quantum integrable systems.

References

  1. [1]
    [PDF] 1 Introduction
    A partial differential equation is an equation involving a function u of several variables and its partial derivatives. The order of the partial differential ...
  2. [2]
    [PDF] Partial Differential Equations: Graduate Level Problems and Solutions
    ... Equation. 52. 11 Problems: Quasilinear Equations. 54. 12 Problems: Shocks. 75. 13 Problems: General Nonlinear Equations. 86. 13.1 Two Spatial Dimensions ...
  3. [3]
    [PDF] A review of numerical methods for nonlinear partial differential ...
    Jul 20, 2012 · 2. Examples of nonlinear PDEs We consider nonlinear PDEs, which take the form (2.1) A ∂sw,∂s−1w,...,∂w, w, x = g(x). the zeroth variable, x0 = t. ...
  4. [4]
    [PDF] Chapter 1 Introduction - IITB Math
    Partial differential equations can be classified in at least three ways. They are. 1. Order of PDE. 2. Linear, Semi-linear, Quasi-linear, and fully non-linear.
  5. [5]
    [PDF] Linear PDEs and the Principle of Superposition - Trinity University
    Feb 19, 2015 · Warning: The principle of superposition can easily fail for nonlinear PDEs or boundary conditions. Consider the nonlinear PDE ux + u2uy = 0 ...
  6. [6]
    [PDF] Chapter 22 Nonlinear Partial Differential Equations - TAU
    Feb 25, 2004 · Specific topics include shock waves, blow up, similarity solutions, and solitons. We will only be able to consider nonlinear partial ...
  7. [7]
    [PDF] arXiv:1308.5861v1 [math.AP] 27 Aug 2013
    Aug 27, 2013 · 1. A brief history of nonlinear partial differential equations. Sophus Lie was the pioneer who sought for an order in the primordial chaos.
  8. [8]
    [PDF] Overleaf Example - arXiv
    May 7, 2023 · Abstract The use of nonlinear PDEs has led to significant advancements in various fields, such as physics, biology, ecology, ...
  9. [9]
    Non linear Partial Differential Equation - Academia.edu
    These equations are fundamental in various fields of science and engineering, as they describe complex phenomena such as fluid dynamics, heat transfer, and wave ...
  10. [10]
    Study of coupled nonlinear partial differential equations for finding ...
    Exact solutions of NPDEs play an important role in the proper understanding of qualitative features of many phenomena and processes in the mentioned areas of ...
  11. [11]
    Why can't there be a general theory of nonlinear PDE? - MathOverflow
    Feb 14, 2010 · Any given system of PDE tends to have a combination of ingredients interacting with each other, such as dispersion, dissipation, ellipticity, ...
  12. [12]
    2.3: Existence and Uniqueness of Solutions of Nonlinear Equations
    Jan 6, 2020 · Although there are methods for solving some nonlinear equations, it is impossible to find useful formulas for the solutions of most.Theorem 2.3.1 : existence and... · Example 2.3.2 · Example 2.3.4 · Example 2.3.5
  13. [13]
  14. [14]
    Online learning to accelerate nonlinear PDE solvers
    Jul 23, 2025 · We fully integrate a machine learning model into the nonlinear solver achieving up to 85% reduction in computational time. Abstract. We propose ...
  15. [15]
    [PDF] Deep Learning of Nonlinear Partial Differential Equations
    Moreover, a quick glance at the list of nonlinear partial differential equations on Wikipedia reveals that many of these equations take the form specified ...
  16. [16]
    [PDF] VORTEX LINES IN AN IMPERFECT BOSE GAS
    Gross, Phys. Rev. 106, 161 (1957). 9y, L. Ginzburg and L. P. Pitaevskii, JETP. 34, 1240 (1958), Soviet Phys. JETP 7, 858 (1958). 10 E. Gray and G. B. ...
  17. [17]
    On the Stability of Solitary Waves in Weakly Dispersing Media
    On the Stability of Solitary Waves in Weakly Dispersing Media · B. Kadomtsev, V. Petviashvili · Published 1 December 1970 · Physics.
  18. [18]
    XLI. On the change of form of long waves advancing in a rectangular ...
    (1895). XLI. On the change of form of long waves advancing in a rectangular canal, and on a new type of long stationary waves . The London, Edinburgh, and ...
  19. [19]
    The Landau-Lifshitz equation: 80 years of history, advances, and ...
    Sep 1, 2015 · Landau and E. M. Lifshitz was published (Phys. Zs. Sowjetunion 8, 153 (1935)). For the modern reader it is easy to find the Russian translation ...
  20. [20]
    [PDF] arXiv:math/0511230v1 [math.DG] 9 Nov 2005
    The Liouville equation arises in many contexts of complex analysis and differential geometry of Riemann surfaces, in particular in the prescribing curvature ...
  21. [21]
    The Monge-Ampère equation and its link to optimal transportation
    Oct 23, 2013 · We survey the (old and new) regularity theory for the Monge-Ampère equation, show its connection to optimal transportation, and describe the regularity ...Missing: geometry | Show results with:geometry
  22. [22]
    On a new integrable discretization of the derivative nonlinear ... - arXiv
    Jan 8, 2015 · Thus, we obtain a proper space-discretization of the Chen-Lee-Liu equation defined on the three lattice sites n-1, n, n+1 for the first time.
  23. [23]
    [PDF] 35. Porous medium equation - People
    The porous medium ... u is replaced by (um) for some m > 1: ut = (um): (1). Equivalently we may rewrite (1) as a heat equation with a nonlinear di usion constant,.