Fact-checked by Grok 2 weeks ago

Nonlinear partial differential equation

A nonlinear partial differential equation (PDE) is a in which the relations between the unknown functions and their partial derivatives are nonlinear, distinguishing it from linear PDEs where the applies. These equations arise naturally in the mathematical modeling of complex phenomena across physics, engineering, and other sciences, where linear approximations fail to capture essential behaviors such as turbulence or wave breaking. Nonlinear PDEs are classified based on the degree of nonlinearity in their highest-order derivatives. Quasilinear PDEs are linear in the highest-order derivatives but have coefficients that may depend on the solution or its lower-order derivatives, as in the form \sum_i A_i(x, u) \partial_i u = F(x, u). Semilinear PDEs are a subclass of quasilinear equations where the coefficients of the highest-order terms depend only on the independent variables, with nonlinearity appearing in lower-order terms, such as the nonlinear \Box \phi - V'(\phi) = 0. Fully nonlinear PDEs involve nonlinearity directly in the highest-order derivatives, exemplified by the Monge-Ampère equation \det(\partial^2 u) = f(x, u, \partial u), which plays a key role in . Prominent examples include the Navier-Stokes equations, which govern fluid motion and form one of the Clay Mathematics Institute's due to unresolved questions about the and of solutions in three dimensions. Other significant cases are the from and the Euler equations for compressible fluids. These equations highlight the challenges of nonlinear PDEs, including the absence of general and uniqueness theorems akin to those for linear PDEs, potential loss of , and the need for advanced numerical methods to approximate solutions. The study of nonlinear PDEs drives contemporary mathematical research, influencing fields from (e.g., the ) to , with ongoing developments in analytical techniques, , and computational approaches to address their multiscale and invariant properties.

Definition and Fundamentals

Definition and General Form

A partial differential equation (PDE) is an equation that relates an unknown function of multiple independent variables—typically spatial coordinates \mathbf{x} = (x_1, \dots, x_n) and possibly time t—to certain of its partial derivatives, which represent rates of change of the function with respect to one variable while holding the others fixed. These partial derivatives arise in contexts modeling phenomena like wave propagation, heat diffusion, or fluid flow, where the function u(\mathbf{x}, t) might describe , , or varying across space and over time. A nonlinear partial differential equation is a PDE in which the principal part, involving the highest-order derivatives, or the lower-order terms depend nonlinearly on the unknown u or its derivatives. The general form can be expressed as F(\mathbf{x}, t, u, \nabla u, D^2 u, \dots, D^k u) = 0, where F is a nonlinear of its arguments, \nabla u denotes the (first-order partial derivatives), and D^k u represents the tensors of k-th partial derivatives. Nonlinearity occurs specifically when F includes terms such as products like u \cdot \partial u / \partial x_i or powers like (\partial u / \partial x_i)^p with p \neq 1, distinguishing it from the linear case where F is affine in u and its derivatives. Key basic properties of nonlinear PDEs stem from this nonlinearity: unlike linear PDEs, solutions do not satisfy the , meaning that if u_1 and u_2 are solutions, then \alpha u_1 + \beta u_2 generally is not a solution for scalars \alpha, \beta. Additionally, for given initial or boundary data, nonlinear PDEs may admit no solutions, a unique solution, or infinitely many solutions, complicating analysis compared to the often well-posed linear counterparts. The formal study and terminology of nonlinear PDEs were advanced in mid-20th-century , particularly through the comprehensive treatment in the works of and , who integrated variational methods and to address existence and regularity for such equations.

Distinction from Linear PDEs

A fundamental distinction between linear and nonlinear partial differential equations (PDEs) lies in the principle of superposition, which holds for linear PDEs but fails for nonlinear ones. For a linear homogeneous PDE, if u_1 and u_2 are solutions, then any c_1 u_1 + c_2 u_2 (with constants c_1, c_2) is also a solution; this property enables techniques like expansions to construct general solutions from eigenfunctions. In contrast, nonlinear PDEs do not satisfy superposition, as the sum of solutions generally does not solve the equation, leading to complex interactions that preclude such additive decompositions. This lack of superposition in nonlinear PDEs gives rise to distinctive qualitative behaviors absent in linear cases, such as the formation of shock waves in hyperbolic systems or stable structures in dispersive equations. Shock waves emerge from nonlinear steepening of wave profiles, resulting in discontinuities that require additional physical principles like conditions for resolution, unlike the smooth propagation in linear wave equations. , conversely, represent localized, particle-like waves that maintain shape and speed due to a balance between nonlinearity and , phenomena that linear PDEs cannot produce without external forcing. Regarding solvability, linear PDEs often admit explicit analytical solutions via methods like transforms, which diagonalize the operator and yield closed-form expressions for initial-boundary value problems. Nonlinear PDEs, however, rarely possess such explicit solutions, necessitating asymptotic approximations for large-scale behaviors or numerical methods like finite differences and spectral schemes for general cases. For instance, the linear u_t = k u_{xx} diffuses smoothly with solutions expressible via and , whereas a nonlinear variant with temperature-dependent , such as u_t = \nabla \cdot (k(u) \nabla u), introduces that defies similar exact treatment and often requires iterative numerical . The nonlinearity further impacts analytical approaches by introducing mode coupling, where different frequency components interact, rendering —effective for linear PDEs—generally inapplicable without specialized functional forms. This coupling complicates expansions and analyses, shifting focus to qualitative tools like Lyapunov functions or energy methods to understand global behavior.

Initial Examples

One illustrative example of a nonlinear partial differential equation is , given by \frac{\partial u}{\partial t} + u \frac{\partial u}{\partial x} = \nu \frac{\partial^2 u}{\partial x^2}, where u(x,t) represents velocity, \nu > 0 is the coefficient, and the equation serves as a prototype for the formation of viscous shocks in . The equation was first introduced by Harry Bateman in 1915 and proposed by J.M. Burgers in 1948 as a simplified model mimicking the one-dimensional Navier-Stokes equations for incompressible fluid flow, particularly to study phenomena through energy transfer across scales. In , the nonlinearity appears in the advective term u \frac{\partial u}{\partial x}, exemplifying a semilinear form. Another fundamental example is the nonlinear Poisson equation, \nabla^2 u = f(u), which arises in contexts such as or reaction-diffusion systems, where the Laplacian operator on the left-hand side is linear, but the source term f(u) introduces nonlinearity dependent on the solution. This equation highlights a semilinear structure, as the principal part remains unaffected by u, with nonlinearity confined to the lower-order term. Together, these examples illustrate different manifestations of semilinearity in nonlinear PDEs: nonlinear convection in a parabolic transport equation like Burgers', versus nonlinear source terms in an elliptic equation like the equation.

Classification and Types

Based on Nonlinearity Degree

Nonlinear partial differential equations (PDEs) are classified based on the degree of nonlinearity, particularly how the nonlinear terms interact with the highest-order derivatives. This taxonomy—semilinear, , and fully nonlinear—distinguishes the structural complexity and influences analytical approaches, such as methods or solutions. The hinges on whether the principal part (highest-order terms) remains linear or becomes nonlinear, often assessed through the homogeneity of the equation under scaling of the solution and its derivatives. Semilinear PDEs feature a linear principal part, with nonlinearity confined to lower-order or the zeroth-order . Formally, for a k-th , it takes the form \sum_{|\alpha|=k} a_\alpha(\mathbf{x}) D^\alpha u + F(\mathbf{x}, u, D u, \dots, D^{k-1} u) = 0, where the coefficients a_\alpha depend only on the independent variables \mathbf{x}, and F is nonlinear but does not involve D^k u. A canonical example is the reaction-diffusion u_t = \Delta u + f(u), where the Laplacian \Delta u is linear in the highest spatial , and the nonlinearity f(u) acts as a source . This structure allows semilinear PDEs to inherit many properties from their linear counterparts, facilitating techniques like maximum principles. Quasilinear PDEs extend this by allowing the coefficients of the highest-order derivatives to depend on lower-order terms, while the highest derivatives themselves appear linearly. The general form is \sum_{|\alpha|=k} a_\alpha(\mathbf{x}, u, D u, \dots, D^{k-1} u) D^\alpha u + F(\mathbf{x}, u, D u, \dots, D^{k-1} u) = 0, where the a_\alpha are nonlinear functions of the solution and its lower derivatives. An illustrative example is the nonlinear diffusion equation \operatorname{div}(a(u) \nabla u) = 0, such as the p-Laplacian \operatorname{div}(|\nabla u|^{p-2} \nabla u) = 0 for p > 1, where the coefficient |\nabla u|^{p-2} modulates the flux based on the gradient magnitude. Semilinear equations are a subclass of quasilinear ones, but quasilinear forms often lead to more intricate behaviors like shock formation in hyperbolic contexts. Fully nonlinear PDEs involve nonlinearity directly in the highest-order derivatives, without the linear structure in those terms. These are equations where the principal symbol is a nonlinear of the highest derivatives, such as F(D^k u, D^{k-1} u, \dots, u, \mathbf{x}) = 0. A prominent example is the Monge-Ampère equation F(D^2 u) = \det(D^2 u) = f(\mathbf{x}, u, \nabla u), which arises in optimal and , where the nonlinearity is in the second derivatives. Unlike semilinear or cases, fully nonlinear PDEs lack a straightforward of the principal part, complicating existence proofs and often requiring viscosity solution frameworks. The of nonlinearity is measured by the homogeneity of the terms with respect to the highest derivatives: semilinear terms are homogeneous of zero in D^k u, of one (linear in D^k u), and fully nonlinear of higher . This scaling property affects solvability; for instance, under the u \mapsto \lambda u and appropriate rescaling of \mathbf{x}, the equation's balance reveals the dominant nonlinear effects. Regarding ellipticity preservation, semilinear PDEs retain ellipticity from the linear if the nonlinear term satisfies monotonicity conditions like \partial_u f \geq 0, ensuring unique smooth solutions. PDEs preserve uniform ellipticity if the coefficients satisfy bounds like \lambda |\xi|^2 \leq \sum a_{ij} \xi_i \xi_j \leq \Lambda |\xi|^2 for some \lambda, \Lambda > 0, supporting C^{2,\alpha} regularity. Fully nonlinear PDEs require the to be uniformly elliptic in the sense that its has a definite symbol, often via definiteness, which guarantees interior and boundary regularity up to C^{2,\alpha} under concavity assumptions. These implications underscore why lower- nonlinearities are more amenable to classical analysis.

Elliptic, Parabolic, and Hyperbolic Categories

Nonlinear partial differential equations (PDEs) are classified into elliptic, parabolic, and categories by extending the criteria used for linear PDEs, primarily through analysis of the principal or the of the characteristic form for second-order equations. For second-order PDEs of the form a(u, Du) u_{xx} + 2b(u, Du) u_{xy} + c(u, Du) u_{yy} + \ lower\ order\ terms = 0, the type is determined by the sign of the b^2 - ac, where the coefficients depend on the solution u and its Du, analogous to the linear case but with potential dependence on the unknown solution. This classification influences the appropriate or conditions and the qualitative behavior of solutions, such as and properties. Elliptic nonlinear PDEs are characterized by no real characteristics, corresponding to b^2 - ac < 0, and typically arise in steady-state problems modeled by boundary value problems. A representative example is the nonlinear Poisson equation -\Delta u = f(u), where \Delta is the Laplacian and f is a nonlinear function, often studied in contexts like electrostatics or minimal surfaces. Solutions to such equations may satisfy maximum principles under certain conditions on f, ensuring that the maximum value occurs on the boundary; for instance, in fully nonlinear degenerate elliptic equations, strong maximum principles hold for viscosity subsolutions using subunit vector fields. These properties facilitate regularity and uniqueness results in bounded domains. Parabolic nonlinear PDEs feature one real characteristic direction, with b^2 - ac = 0, often involving a time-like variable and modeling diffusive processes with initial-boundary value problems. The reaction-diffusion equation \partial_t u - \Delta u = f(u) exemplifies this category, capturing phenomena like chemical reactions or population dynamics where diffusion balances nonlinear growth. Forward uniqueness is a key feature, where solutions are determined uniquely by initial data in the forward time direction for semilinear cases, as established for bounded weak solutions in Sobolev spaces under Dirichlet conditions. Hyperbolic nonlinear PDEs possess two distinct real characteristics, with b^2 - ac > 0, and describe wave propagation via initial value problems. The nonlinear \partial_t^2 u - \Delta u = g(u, Du) illustrates this, where disturbances propagate along curves at finite speeds determined by the nonlinearity. Solutions exhibit finite propagation speed, with information traveling along these characteristics, leading to features like shocks in conservation laws. A significant challenge in nonlinear PDEs is that the type can vary depending on the solution values, particularly in forms where coefficients like a, b, c depend on u, potentially leading to mixed-type equations. For example, in transonic gas flows modeled by potential equations, the PDE transitions from elliptic in regions to in supersonic regions across a free boundary, complicating the formulation and solution. This type-changing behavior requires adaptive analytical or numerical approaches to handle the evolving domain structure.

Quasilinear and Fully Nonlinear Forms

partial differential equations (PDEs) are characterized by a principal part that is linear in the highest-order derivatives, but with coefficients that depend on the u and its lower-order derivatives. This structure allows the equation to be written in the general form a(x, u, Du) \cdot D^2 u + b(x, u, Du) = 0 for second-order cases, where the highest derivatives appear linearly, but the coefficients a and b are nonlinear functions of u. Characteristic surfaces for PDEs can be defined locally, similar to linear cases, but their evolution depends on the itself, leading to solution-dependent paths. A representative example of a PDE is the scalar \partial_t u + \partial_x f(u) = 0, where f is a nonlinear function; this models phenomena like or density, with the nonlinearity entering through the dependence of f on u. Such are often in nature, exhibiting wave-like behavior with characteristics that can intersect, potentially forming shocks. Fully nonlinear PDEs, in contrast, involve nonlinearity in the highest-order derivatives themselves, making the entire nonlinear; for instance, a second-order takes the form F(x, u, Du, D^2 u) = 0, where F is nonlinear in D^2 u. These arise prominently in theory through Bellman equations, such as \sup_{a \in A} L_a u = f, where L_a are linear operators parameterized by controls a, resulting in a fully nonlinear elliptic PDE for the value function. An illustrative example of a fully nonlinear PDE is the Monge-Ampère equation \det(D^2 u) = g(x), which governs convex functions and appears in optimal transport and ; here, the nonlinearity is evident in the determinant of the . Analytically, fully nonlinear PDEs pose greater challenges for methods, as the lack of linearity in the principal part often prevents straightforward definition or propagation of characteristics, unlike in cases. PDEs typically inherit hyperbolic-like properties from their linear principal part, facilitating analysis via characteristics, while fully nonlinear ones demand specialized theories like solutions for well-posedness.

Existence, Uniqueness, and Well-Posedness

Frameworks for Existence Proofs

Proving the of solutions to nonlinear partial differential equations (PDEs) often relies on abstract frameworks from , which provide tools to construct weak solutions in appropriate function spaces. These methods are particularly suited to handling the lack of superposition principles inherent in nonlinear problems, where solutions may not inherit the regularity of the data. Central to many proofs is the use of Sobolev spaces H^k(\Omega), which consist of functions whose weak derivatives up to order k are square-integrable, allowing for the formulation of weak solutions that satisfy the PDE in an integral sense rather than pointwise. Embedding theorems play a crucial role in these frameworks by ensuring that functions in Sobolev spaces can be continuously embedded into spaces of higher integrability or continuity, such as H^k(\Omega) \hookrightarrow L^p(\Omega) for suitable p, which facilitates compactness arguments essential for passing to limits in approximating sequences. For instance, the Rellich-Kondrachov theorem guarantees compact embeddings under certain conditions on the domain \Omega, enabling the control of nonlinear terms in the PDE. These embeddings are foundational for establishing a priori bounds and convergence in existence proofs for both quasilinear and fully nonlinear elliptic PDEs. Variational methods offer another powerful framework, particularly for PDEs that can be derived as Euler-Lagrange equations of a functional. The direct method in the proceeds by minimizing a functional J(u) = \int_\Omega F(x, u, \nabla u) \, dx over a suitable space, such as a , assuming lower , , and weak closure properties to guarantee the existence of a minimizer via sequential compactness. This approach is widely applied to nonlinear elliptic problems, where the minimizer satisfies the PDE weakly, as demonstrated in seminal works on the equation -\Delta_p u = f. Fixed-point theorems provide a versatile tool for existence in Banach spaces, mapping the problem into finding fixed points of nonlinear operators associated with the PDE. The applies when the operator is a on a , yielding a unique fixed point under the contraction mapping principle, often used for local existence in evolutionary nonlinear PDEs with nonlinearities. For broader applicability, the extends this to compact, continuous operators on convex, closed, bounded subsets of Banach spaces, ensuring a fixed point exists; this is particularly effective for global existence in quasilinear hyperbolic systems after establishing a priori bounds. Schaefer's variant further refines this for perturbations of the identity, commonly invoked in elliptic boundary value problems. For evolutionary nonlinear PDEs, such as those modeling reaction-diffusion processes, the Galerkin approximation method constructs solutions by projecting onto finite-dimensional subspaces spanned by eigenfunctions of the linear part of the . This yields a sequence of ordinary differential equations whose solutions exist by standard theory, and under suitable growth and assumptions on the nonlinearity, these approximations converge weakly to a of the original PDE via from Aubin-Lions lemmas. This finite-dimensional is a for proving in abstract semilinear evolution equations in Hilbert spaces. In the context of elliptic nonlinear PDEs, monotone operator theory provides a robust framework for existence, especially when the principal part is a nonlinear operator satisfying monotonicity, hemicontinuity, and coercivity in a reflexive Banach space. Minty's theorem guarantees that a maximal monotone operator is surjective onto its range, allowing the PDE to be reformulated as finding u such that A(u) = f for a given right-hand side f, with weak solutions obtained via Yosida approximations or regularization. This approach excels for problems like the porous medium equation or doubly nonlinear diffusion, where the operator's monotonicity ensures solvability without compactness assumptions.

Uniqueness and Regularity Results

Uniqueness results for solutions to nonlinear partial differential equations (PDEs) often rely on tailored techniques depending on the equation's . For elliptic nonlinear PDEs, particularly ones, s provide a for establishing . These principles assert that subsolutions attain their maximum on the boundary, preventing interior maxima and thus ensuring uniqueness under suitable conditions, such as convexity of the nonlinearity or viscosity solution frameworks. A strong version of the maximum principle for fully nonlinear degenerate elliptic equations guarantees that viscosity subsolutions cannot achieve an interior maximum unless constant, extending to principles for . In the parabolic category, energy estimates serve as the primary tool for uniqueness. By multiplying the equation by the solution and integrating, one derives an energy inequality that controls the L² norm over time, often combined with Gronwall's inequality to show that differences between two solutions vanish. For nonlinear parabolic systems like the Boussinesq equations with fractional dissipation, classical energy methods yield uniqueness in Sobolev spaces by bounding the growth of nonlinear terms. For hyperbolic nonlinear PDEs, the method of characteristics determines uniqueness by tracing solution values along non-intersecting curves defined by the PDE's principal part. In strictly hyperbolic conservation laws, if characteristics do not cross and entropy conditions hold, solutions are unique in BV spaces or Lipschitz semigroups. Regularity theory upgrades weak solutions to smoother classes through iterative improvements. Bootstrap arguments begin with a weak solution in a and use theorems and a priori estimates to gain higher , eventually reaching classical C^{k,\alpha} regularity if coefficients are smooth. For nonlinear elliptic PDEs, Schauder estimates are adapted via or around a background solution, yielding Hölder of second derivatives under uniform ellipticity. These methods assume of a weak solution, building on frameworks like operators or variational inequalities. The De Giorgi-Nash-Moser exemplifies regularity for quasilinear elliptic and parabolic PDEs, proving that bounded weak solutions are Hölder continuous, of . De Giorgi established interior Hölder estimates for elliptic minimizers of variational integrals in 1957, while Nash extended this to parabolic equations in , showing uniform Hölder continuity via Moser iteration on oscillation . Moser's elliptic proof simplified De Giorgi's approach using mean value inequalities. These results apply to equations like the p-Laplace, where weak solutions in W^{1,2} are locally C^{0,\alpha}. However, uniqueness and regularity falter in supercritical regimes, where nonlinearity dominates linear smoothing, leading to non-uniqueness for weak solutions. For the 2D incompressible Euler equations, a hyperbolic system, Vishik's theorem constructs non-unique weak solutions in L^1 \cap L^p spaces for 2 < p < \infty, with vorticity in borderline regularity classes. Recent convex integration methods further demonstrate infinitely many admissible solutions for L^p vortex data, highlighting failure of uniqueness beyond critical Sobolev embeddings.

Ill-Posed Problems and Counterexamples

A problem is well-posed in the sense of Hadamard if it satisfies three conditions: existence of a solution, uniqueness of the solution, and continuous dependence of the solution on the initial data. These criteria, originally formulated for linear problems, extend to nonlinear partial differential equations (PDEs), where failures often manifest prominently due to the inherent complexities of nonlinearity. In particular, continuous dependence can break down in nonlinear settings, leading to solutions that are highly sensitive to perturbations in data, even when existence and uniqueness hold. One classic counterexample arises in the Navier-Stokes equations, where Jean Leray established in 1934 the existence of global weak solutions (now known as Leray-Hopf solutions) in three dimensions for initial data of finite energy. However, uniqueness fails in certain regimes, particularly at high Reynolds numbers associated with turbulent flows, as demonstrated by recent constructions of distinct Leray solutions sharing the same initial data and forcing terms. A further development in September 2025 showed non-uniqueness of Leray-Hopf solutions even for the unforced incompressible 3D Navier-Stokes equations. Such non-uniqueness highlights the ill-posedness of the initial-boundary value problem in supercritical function spaces relevant to high-Reynolds dynamics. The backward heat equation provides another foundational illustration of ill-posedness, where the linear case already exhibits instability: solutions do not depend continuously on final-time data, as high-frequency modes amplify exponentially backward in time. In nonlinear variants, such as the problem u_t = -\Delta u + f(u) solved backward from final data u(x,T) = \phi(x), this instability is amplified, rendering the problem severely ill-posed with even greater sensitivity to data perturbations and potential non-existence for generic \phi. Regularity results that succeed for forward problems often fail here, underscoring the challenges in establishing stable solutions. Blow-up in finite time represents a failure of global existence, a key aspect of well-posedness. In reaction-diffusion equations modeling combustion or chemical reactions, the Semenov model captures thermal runaway, where solutions to the simplified system exhibit unbounded growth in finite time due to exothermic feedback, precluding global solutions. This phenomenon extends to spatially distributed reaction-diffusion PDEs, such as those with Arrhenius kinetics, where local temperature spikes lead to no global-in-time solutions for supercritical initial data, violating the existence condition across all times. More recent advances have confirmed non-uniqueness in the three-dimensional incompressible through the resolution of on energy conservation thresholds. Specifically, for weak solutions with Hölder continuity exponent \gamma < 1/3, anomalous energy dissipation occurs, and explicit constructions show non-uniqueness even among Hölder-continuous, globally dissipative solutions in the periodic setting. This 2017 result establishes ill-posedness in the sense of multiple solutions failing continuous energy balance, with implications for inviscid fluid models.

Analytical Solution Methods

Perturbation and Linear Approximation Techniques

, the nonlinearity is scaled by \epsilon, enabling a expansion of the solution u(x, t) = u_0(x, t) + \epsilon u_1(x, t) + \epsilon^2 u_2(x, t) + \cdots. The zeroth-order term u_0 satisfies the underlying linear PDE, while successive terms u_k (for k \geq 1) solve linear inhomogeneous PDEs obtained by substituting the expansion into the and collecting coefficients of equal powers of \epsilon. This approach is particularly effective for weakly nonlinear problems, such as those arising in , where the series provides asymptotic approximations valid uniformly in the domain away from singularities. Linearization techniques focus on local approximations near specific solutions, such as , by replacing the nonlinear operator with its , which serves as the infinite-dimensional analog of the . For a nonlinear evolution PDE of the form \partial_t u = F(u), the DF(u^*) at an u^* (where F(u^*) = 0) generates a linear whose spectral properties determine the local of u^*. If all eigenvalues of DF(u^*) have negative real parts, the linearized system is exponentially stable, implying local exponential of the nonlinear system under suitable conditions on the nonlinearity, such as local . This method is foundational for analysis in reaction-diffusion equations and other dissipative systems. For nonlinear PDEs with small parameters leading to rapid variations near boundaries—such as viscous s in the Navier-Stokes equations—regular perturbations yield nonuniform s, necessitating matched asymptotic expansions. These involve an outer expansion valid in the interior, u^{\text{outer}}(x; \epsilon) = u_0^{\text{outer}}(x) + \epsilon u_1^{\text{outer}}(x) + \cdots, and an inner expansion in a stretched boundary layer coordinate \xi = x / \delta(\epsilon), u^{\text{inner}}(\xi; \epsilon) = u_0^{\text{inner}}(\xi) + \epsilon u_1^{\text{inner}}(\xi) + \cdots, where \delta(\epsilon) \to 0 balances the highest-order terms. The expansions are matched asymptotically in an intermediate region, ensuring a uniform u(x; \epsilon) \approx u^{\text{outer}} + u^{\text{inner}} - u^{\text{match}} across the ; this resolves layer structures in singularly perturbed problems like convection-diffusion equations. The reliability of these approximations depends on the convergence of the perturbation series as \epsilon \to 0, typically in Sobolev or other Banach spaces, with radius of convergence determined by the analyticity of the nonlinearity. Error estimates for truncated expansions often invoke Gronwall's inequality to bound the remainder, applied to integral inequalities derived from the evolution equation for the error function; for instance, if the error e(t) satisfies \frac{de}{dt} \leq K \|e\| + R(\epsilon), Gronwall yields \|e(t)\| \leq ( \|e(0)\| + \int_0^t R(s) ds ) e^{Kt}, providing exponential control on approximation accuracy for small \epsilon.

Singularity Analysis and Blow-Up Phenomena

In nonlinear partial differential equations (PDEs), singularities often arise due to the amplifying effects of nonlinear terms, leading to finite-time blow-up where solutions lose regularity or become unbounded. analysis focuses on characterizing these breakdowns, particularly blow-up, where the supremum of the ||∇u||_{L^∞} tends to infinity as time approaches the blow-up time T, while the solution u itself may remain bounded. This phenomenon is prevalent in models, such as the Navier-Stokes equations, where blow-up in the velocity field can signal the onset of or formation. Point singularities, on the other hand, involve the solution u developing an of infinite value or discontinuity, often concentrating energy at a single spatial location as t → T^-. Examples include type I blow-up in semilinear heat equations, where the maximum of |u| scales like (T - t)^{-1/(p-1)} near the point. A key tool for detecting and preventing blow-up is the Beale-Kato-Majda criterion, which provides a necessary condition for the persistence of solutions. For the three-dimensional incompressible Euler equations, the criterion states that a smooth solution remains regular up to time T if and only if the 0^T ||ω(t)||{L^∞} dt < ∞, where ω = curl u is the vorticity; blow-up occurs precisely when this integral diverges due to ||ω||_{L^∞} → ∞. This result extends to the Navier-Stokes equations under similar conditions on the vorticity or velocity gradient, highlighting how nonlinear advection terms can drive gradient blow-up. Extensions of the criterion to other settings, such as compressible flows or magnetohydrodynamics, replace the L^∞ norm with Besov or Morrey norms to capture weaker singularities. To probe the structure near a singularity, self-similar solutions provide asymptotic profiles that reveal blow-up rates and spatial . A common ansatz for backward self-similar blow-up assumes the form u(x, t) = (T - t)^{-\alpha} f\left( \xi \right), \quad \xi = \frac{x}{(T - t)^{\beta}}, where α > 0 and β > 0 are exponents determined by balancing the linear and nonlinear terms in the PDE, and f is a radial profile satisfying an elliptic boundary-value problem. This approach yields type I blow-up rates when the nonlinearity is subcritical or critical relative to , as seen in the semilinear u_t = Δu + |u|^{p-1}u for 1 < p ≤ (n+2)/(n-2). Seminal analyses using this method have constructed explicit self-similar solutions for reaction-diffusion systems, confirming point singularities with Gaussian-like decay in f(ξ). In the supercritical regime, where p > (n+2)/(n-2), blow-up dynamics exhibit more complex behaviors, including type II singularities with rates faster than the self-similar scaling. Recent results from the have established on these supercritical blow-up rates for semilinear equations in bounded . For instance, in the Joseph-Lundgren supercritical range, refined constructions demonstrate the existence of stable type II solutions where the L^∞ norm of u grows like (T - t)^{-κ} with κ > 1/(p-1), driven by logarithmic corrections to the self-similar profile. Additionally, nonexistence results for type II blow-up in subcritical cases have been extended to non-convex , showing that solutions adhere to type I rates under constraints. These findings underscore the role of dimensionality and in dictating singularity types.

Moduli Spaces of Solutions

In the context of nonlinear partial differential equations (PDEs), the of solutions refers to the parameter space that classifies distinct solutions up to the action of symmetries, such as transformations or diffeomorphisms, forming a of the full space by the . This structure captures the geometric and topological features of solution families, often parameterized by initial data or conditions. The of such moduli spaces is typically determined via index theory applied to the linearized operator of the PDE, where the Atiyah-Singer index theorem provides the virtual as the difference between the dimensions of the and of the arising from around a solution. For elliptic nonlinear PDEs, this index yields a finite-dimensional manifold locally near generic points, reflecting the finite in the space after quotienting by symmetries. A prominent example is the of instantons in Yang-Mills theory, where solutions to the anti-self-dual Yang-Mills equations on a four-manifold parameterize gauge field configurations with finite action. For SU(2) instantons of topological charge k on \mathbb{R}^4, the has dimension $8k - 3, accounting for translational, rotational, and scale invariances after , as computed via the Atiyah-Hitchin-Singer theorem. Another key example arises in the study of s, where the serves as a for marked Riemann surfaces underlying immersed s in \mathbb{R}^3, parameterizing conformal structures that solve the nonlinear minimal surface equation through their Weierstrass representation. This space, of complex dimension $3g - 3 for genus g \geq 2, encodes the deformation theory of solutions under the quasilinear elliptic system governing zero. To analyze the global structure, tools like Uhlenbeck compactness address the non-compactness of these by providing results for sequences of solutions, preventing energy concentration except at bubbling points where solutions develop singularities modeled on holomorphic spheres or instantons. Bubbling phenomena, where limits involve multiple scaled copies of base solutions, are controlled by this theorem, ensuring that the closure of the includes stable limits after . Gluing constructions complement this by parametrizing neighborhoods of boundary strata, where approximate solutions on disconnected components are joined via neck regions to yield exact solutions nearby, often using implicit function theorems on Banach manifolds. These methods are essential for compactifying in nonlinear settings. Unlike linear PDEs, where the solution space modulo symmetries is a finite-dimensional vector space with no branching, nonlinearities introduce complexities such as multiple connected components or bifurcation branches in the moduli space, arising from interactions between solutions that linear approximations cannot capture. For instance, in Yang-Mills theory, nonlinearity leads to irreducible components distinguished by topological invariants beyond the basic index. Singularity analysis briefly aids in compactifying these spaces by resolving blow-up loci through desingularization techniques.

Exact Solution Approaches

Symmetry and Lie Group Methods

Symmetry and Lie group methods provide a systematic framework for analyzing and solving nonlinear partial differential equations (PDEs) by exploiting their underlying symmetries. These methods, originating from the work of Sophus Lie in the late 19th century, identify continuous groups of transformations that leave the PDE invariant, enabling the reduction of the equation's complexity and the construction of exact solutions. In particular, Lie point symmetries act on the independent variables (such as space x and time t) and the dependent variable u(x,t), facilitating the discovery of invariant solutions that simplify the PDE into ordinary differential equations (ODEs). This approach is especially powerful for nonlinear PDEs, where traditional methods often fail, and has been applied extensively to equations in physics, such as those modeling wave propagation and fluid dynamics. A Lie point symmetry of a PDE is generated by an infinitesimal of the form X = \xi(x,t,u) \frac{\partial}{\partial x} + \tau(x,t,u) \frac{\partial}{\partial t} + \phi(x,t,u) \frac{\partial}{\partial u}, where \xi, \tau, and \phi are smooth functions determining the local transformation. The one-parameter of transformations associated with X is obtained by exponentiating the generator, \exp(\epsilon X), for infinitesimal \epsilon. For the to preserve the solutions of the PDE, the prolonged action of X must leave the equation when restricted to its solution manifold. The prolongation extends the generator to act on higher-order derivatives; for a prolongation, it includes terms like \phi^x = D_x(\phi - \xi u_x - \tau u_t) + \xi u_{xx} + \tau u_{xt}, where D_x denotes total differentiation with respect to x. Higher prolongations follow similarly, ensuring the condition holds for all relevant spaces. To find the symmetries, one solves the determining equations arising from the invariance condition \mathrm{pr}^{(n)} X (\Delta) = 0 whenever \Delta = 0, where \Delta is the PDE and \mathrm{pr}^{(n)} X is the nth prolongation. These form an of linear PDEs for \xi, \tau, and \phi, often solvable by or assuming dependence on u. The solutions classify the possible symmetry algebras, which for many nonlinear PDEs are finite-dimensional algebras. Seminal developments in this area, including algorithmic implementations for computing symmetries, were advanced in the , making the method accessible via symbolic computation software. Once symmetries are identified, they are used to construct invariant solutions by seeking functions u(x,t) unchanged under the group action, i.e., \phi - \xi u_x - \tau u_t = 0. This leads to characteristic equations whose solutions provide canonical coordinates, such as similarity variables \eta = \frac{x - c t}{\sqrt{t}} for scaling symmetries. In these coordinates, the PDE reduces to an ODE, which can often be integrated explicitly. For instance, traveling wave solutions arise from translation symmetries by setting \eta = x - c t. This reduction process not only yields exact solutions but also reveals the structure of the solution moduli space, where symmetries act to parameterize families of solutions. A classic example is the Korteweg-de Vries (KdV) equation, u_t + 6 u u_x + u_{xxx} = 0, which models shallow water and possesses an infinite-dimensional of symmetries, including translations, scalings, and boosts. The symmetries X_1 = \frac{\partial}{\partial x} and X_2 = \frac{\partial}{\partial t} generate traveling solutions by reducing the PDE to -c U' + 6 U U' + U''' = 0, where U(\eta) with \eta = x - c t. Integrating once yields the sech-squared soliton profile U(\eta) = 2 k^2 \mathrm{sech}^2(k (\eta - \eta_0)), to understanding integrable nonlinear systems. Further symmetries, such as the X = t \frac{\partial}{\partial x} + u \frac{\partial}{\partial u}, transform solutions into moving cnoidal , illustrating the method's power in generating a rich class of exact solutions.

Integrable Systems and Lax Pairs

Completely integrable nonlinear partial differential equations (PDEs) possess an infinite sequence of independent conservation laws, which enable the exact solution of initial value problems through techniques like the inverse scattering transform. These systems typically admit a Hamiltonian formulation with a compatible Poisson bracket, facilitating a bi-Hamiltonian structure that generates the hierarchy of conserved quantities via recursion operators. Such properties distinguish integrable PDEs from generic nonlinear equations, allowing for the explicit construction of multi-parameter families of solutions despite the inherent complexity of nonlinearity. The formalism, introduced by in , formalizes integrability by associating the nonlinear PDE with the isospectral evolution of a linear operator. Specifically, a Lax pair comprises two operators L and M satisfying the commutator equation L_t = [M, L] = ML - LM, where the time derivative L_t ensures that eigenvalues of L remain invariant under the flow of the PDE. This evolution implies an infinite number of conservation laws, as the traces \operatorname{tr}(L^k) for integer k are preserved. Equivalently, in the zero-curvature representation, the Lax equation emerges as the compatibility condition \partial_t U - \partial_x V + [U, V] = 0 for a linear , directly deriving the nonlinear PDE from the operators U and V. A prototypical illustration is the Korteweg-de Vries (KdV) equation, u_t + 6uu_x + u_{xxx} = 0, which models shallow water waves and admits the Lax pair L = \partial_x^2 - u, \quad M = -4\partial_x^3 + 6u\partial_x + 3u_x. The relation L_t = [M, L] holds precisely when u satisfies the KdV equation, confirming its integrability and yielding the infinite conservation laws through the spectrum of L. Integrable systems like KdV support N-soliton solutions, constructed via the , which maps the nonlinear evolution to linear dynamics in the data . In this , initial data determines discrete eigenvalues corresponding to solitons, whose is explicit, resulting in collisions where solitons emerge unchanged in amplitude and speed. Recent extensions in the 2020s apply deformation algorithms to generate higher-dimensional integrable PDEs, such as (2+1)-dimensional KdV analogs that preserve Lax pair structures and infinite hierarchies.

Transformation Techniques to Linear Forms

Transformation techniques to linear forms represent a class of methods that convert specific nonlinear partial differential equations (PDEs) into equivalent linear PDEs, facilitating exact solutions through established linear theory. These approaches are particularly effective for certain and diffusive nonlinear PDEs, where a suitable preserves the solution structure while eliminating nonlinearity. By interchanging or reparameterizing variables, the transformed equation often reduces to a canonical linear form, such as the or , allowing analytical progress in otherwise intractable problems. The transformation is a foundational technique for linearizing systems of PDEs, commonly arising in gas dynamics and . In this method, the dependent variables are interchanged with the independent spatial and temporal variables, effectively inverting the roles to yield a , typically the multidimensional or Laplace equation. For a general 2×2 system of the form \mathbf{u}_t + A(\mathbf{u}) \mathbf{u}_x = 0, where \mathbf{u} = (u, v) and A is the matrix, the transform defines new coordinates x = x(u,v,t), t = t(u,v,t) (with adjustments for invertibility), resulting in a linear PDE for the inverse mapping. This transformation linearizes the system provided the original equation admits Riemann invariants or satisfies specific integrability conditions, as detailed in algorithmic criteria for linearizability. A prominent example is the Cole-Hopf transformation, which linearizes the viscous , a nonlinear parabolic PDE modeling shock waves and : u_t + u u_x = \nu u_{xx}. The transformation u(x,t) = -2\nu \frac{\partial}{\partial x} \log \psi(x,t) maps this to the linear \psi_t = \nu \psi_{xx}, whose solutions are well-known via methods or fundamental solutions. Independently discovered by Hopf for the initial-value problem and by for boundary-value contexts, this substitution exploits the logarithmic potential to cancel the nonlinear term, enabling exact solution expressions for initial data u(x,0) = f(x) in terms of integrals involving the . Painlevé analysis provides a systematic framework to identify and construct such transformations by examining the singularity of solutions. Developed for PDEs, this assumes a Laurent-like expansion around movable singularities, \phi \sim \alpha (x - x_0)^\beta, and requires the absence of points or singularities for integrability, guiding the form of linearizing maps. For nonlinear PDEs passing the Painlevé test, the analysis reveals recursion relations that truncate to determine transformation coefficients, often leading to mappings to linear equations in cases like certain reaction-diffusion systems. This approach, while not guaranteeing linearizability, identifies candidates where or potential-like transforms apply, as in the Weiss-Tabor-Carnevale for partial derivatives. Despite their power, these transformation techniques have significant limitations, primarily succeeding in one spatial dimension plus time (1+1D) due to the complexity of invertibility and conditions in higher s. Extensions to multidimensions often require additional constraints, such as specific symmetries or weak nonlinearity, and fail for general systems where the becomes non-diagonalizable. Recent algorithmic advances, including symbolic computation for 1-1 mappings in select cases, have expanded applicability to boundary-value problems, but full remains elusive for most higher-dimensional nonlinear PDEs beyond integrable subclasses.

Variational and Hamiltonian Frameworks

Euler-Lagrange Derivations

Nonlinear partial differential equations (PDEs) often emerge as Euler-Lagrange equations from variational principles, where solutions minimize or maximize a functional derived from physical or geometric considerations. The general variational formulation posits that a function u: \Omega \to \mathbb{R} extremizes the integral J = \int_\Omega L(u, \nabla u, x) \, dx, with L denoting the Lagrangian density depending on the function, its gradient, and the domain variables. The first variation of this functional vanishes for extremal functions, yielding the Euler-Lagrange equation \frac{\partial L}{\partial u} - \nabla \cdot \left( \frac{\partial L}{\partial \nabla u} \right) = 0, which is typically a quasilinear second-order PDE when L involves nonlinear terms in \nabla u. This framework unifies diverse nonlinear PDEs arising in mechanics, geometry, and field theory, providing a pathway to existence and stability analysis via direct methods in the calculus of variations. A classic nonlinear example is the equation, obtained by varying the area functional for a z = u(x,y) over a domain \Omega \subset \mathbb{R}^2: J = \int_\Omega \sqrt{1 + |\nabla u|^2} \, dx\, dy. The corresponding Euler-Lagrange equation is the quasilinear elliptic PDE (1 + u_y^2) u_{xx} - 2 u_x u_y u_{xy} + (1 + u_x^2) u_{yy} = 0, which governs surfaces of locally minimal area, such as soap films or biological membranes, and highlights the nonlinear coupling essential for capturing geometric constraints. In nonlinear elasticity, variational principles derive the equilibrium equations from the total potential energy functional \Pi[\mathbf{u}] = \int_{\Omega} W(\nabla \mathbf{u}) \, dV - \int_{\partial \Omega} \mathbf{t} \cdot \mathbf{u} \, dS, where \mathbf{u} is the displacement field, W is the stored density (often hyperelastic, depending nonlinearly on the deformation F = I + \nabla \mathbf{u}), and \mathbf{t} are boundary tractions. The Euler-Lagrange equations take the form \nabla \cdot P = 0 in \Omega, with the first Piola-Kirchhoff tensor P = \frac{\partial W}{\partial F}, resulting in a system of nonlinear PDEs that model large deformations in materials like rubber or soft tissues, where quasiconvexity of W ensures physical realism. Symmetries of the variational functional connect to conservation laws via , which asserts that every continuous variational symmetry generates a for the associated Euler-Lagrange PDE; for instance, translation invariance in x yields momentum conservation in field theories. This link extends classical results to nonlinear settings, enabling the identification of integrals of motion that constrain solution behaviors. In the 2020s, has leveraged s for PDE discovery, with variational physics-informed neural networks (VPINNs) optimizing neural approximations to satisfy Euler-Lagrange equations while fitting data, thus uncovering governing nonlinear PDEs in sparse-data regimes like or material science.

Hamiltonian Formulations

formulations provide a dynamical for nonlinear partial differential equations (PDEs) by casting them as infinite-dimensional analogs of finite-dimensional systems, preserving structures and enabling the study of , integrability, and long-time behavior. In this approach, the evolution equation for a field u(x,t) is expressed as u_t = J \frac{\delta H}{\delta u}, where J is a skew-adjoint (or formally skew-symmetric) encoding the structure, and H is a conserved functional typically or higher order in u and its derivatives. This structure arises via the Legendre transform from the Euler-Lagrange equations of a , linking and descriptions. A example is the nonlinear Schrödinger (NLS) equation, i \psi_t + \psi_{xx} + 2 |\psi|^2 \psi = 0, which admits the Hamiltonian form i \psi_t = \frac{\delta H}{\delta \bar{\psi}} with H[\psi, \bar{\psi}] = \int_{-\infty}^{\infty} \left( |\psi_x|^2 - |\psi|^4 \right) dx and the skew-adjoint operator implicitly given by the complex structure. This formulation highlights the equation's role in modeling wave envelopes in and Bose-Einstein condensates, where the conserves both mass and energy. The structure was first elucidated in the context of integrable systems derived from inverse scattering. The Korteweg-de Vries (KdV) equation, u_t + 6 u u_x + u_{xxx} = 0, similarly possesses a structure u_t = J \frac{\delta H}{\delta u} with J = -\partial_x and H = \int_{-\infty}^{\infty} \left( u^3 - \frac{1}{2} u_x^2 \right) dx. This bi-infinite domain setup ensures the skew-adjointness of J, facilitating the identification of infinite conservation laws. Within the Ablowitz-Kaup-Newell-Segur (AKNS) hierarchy, the KdV emerges as the first non-trivial flow, where the operators are derived from the zero-curvature representation, linking it to broader classes of equations. For field theories involving multiple spatial dimensions or relativistic settings, Dirac structures generalize the symplectic framework by incorporating multisymplectic geometry, where the phase space is equipped with a family of presymplectic forms preserving local energy-momentum conservation. Multisymplectic formulations define a covariant phase space with a multi-symplectic form \omega = -d\theta, where \theta is the Cartan form, and the dynamics follow from the primitive form on the jet bundle. Dirac structures, as maximal isotropic subbundles of the Courant algebroid, encode both the skew-symmetry and compatibility conditions for boundary flows in distributed-parameter systems. This approach is particularly useful for nonlinear wave equations in continuum mechanics. The connection to integrability is deepened through bi-Hamiltonian systems, where the PDE admits two compatible Hamiltonian formulations u_t = J_1 \frac{\delta H_1}{\delta u} = J_2 \frac{\delta H_2}{\delta u}, with J_1, J_2 skew-adjoint and their linear combination satisfying the . For the KdV equation, the first operator is J_1 = -\partial_x with cubic H_1, and the second is the third-order J_2 = -\partial_x^3 - 4 u \partial_x - 2 u_x, with H_2 = \frac{1}{2} \int_{-\infty}^{\infty} u^2 \, dx, generating an infinite of commuting flows via . This bi-Hamiltonian , first demonstrated for KdV, implies complete integrability and the existence of infinitely many conserved quantities.

Conservation Laws and Noether's Theorem

Conservation laws are essential for understanding the structure and dynamics of solutions to nonlinear partial differential equations (PDEs), as they encode invariant quantities like mass, energy, and momentum that remain unchanged under the evolution governed by the PDE. For nonlinear PDEs arising from variational principles, provides a systematic link between symmetries of the underlying and these conserved quantities. Noether's first theorem states that any continuous symmetry transformation of the Lagrangian action integral corresponds to a conserved current for the associated Euler-Lagrange PDEs. In the PDE setting, this involves variational symmetries, which may include generalized or infinite-dimensional Lie-Bäcklund symmetries, leading to a divergence-free current J = (J^t, \mathbf{J}^x) satisfying D_t J^t + \nabla \cdot \mathbf{J}^x = 0 on solutions, where D_t denotes the total time derivative. For example, invariance under spatial translations generates conservation of linear momentum, with density \int \frac{\partial \mathcal{L}}{\partial u_x} u_x \, dx, while time-translation invariance yields , with Hamiltonian density derived from the Legendre transform of the . This theorem extends naturally to nonlinear PDEs, such as those modeling wave propagation or , where the nonlinear terms do not obstruct the symmetry-conservation correspondence provided the is appropriately defined. In cases without an obvious variational structure, direct construction methods enable the derivation of conservation laws by algebraic manipulation of the PDE itself. These approaches seek a multiplier \Lambda such that multiplying the PDE G(u, u_t, \nabla u, \dots) = 0 by \Lambda and applying total derivatives yields a local divergence expression D_t A + \nabla \cdot \mathbf{B} = \Lambda G, where A is the conserved density and \mathbf{B} the flux vector, ensuring the form holds identically on solutions. The multipliers satisfy a linear overdetermined system obtained by equating coefficients after expansion in jet space, which can be solved systematically for any order. This method, independent of Noether's theorem, has classified conservation laws for nonlinear classes like the generalized Korteweg-de Vries equation, revealing dependencies on nonlinearity parameters. A distinctive feature of certain nonlinear PDEs is the existence of higher symmetries—generalized transformations involving higher derivatives—that generate an infinite sequence of conservation laws. These symmetries, beyond classical point symmetries, form a and are characteristic of equations with rich structure, producing conserved densities of arbitrarily high order through recursive application. For instance, in the Korteweg-de Vries equation u_t + u u_x + u_{xxx} = 0, such higher symmetries yield infinitely many conservation laws, with densities like \int u \, dx for mass and \int (u^3 - u_x^2) \, dx for a higher-order quantity. These conservation laws have critical applications in soliton dynamics, where preservation of \int \left( \frac{1}{2} u_x^2 + V(u) \right) dx and \int u u_x \, dx ensures elastic collisions and of solitary waves in nonlinear dispersive media, as exemplified in the sine-Gordon equation modeling dislocations in solids. In recent climate modeling, conservation laws derived for nonlinear PDEs in primitive equation systems—such as those for —underpin structure-preserving discretizations that maintain global invariants like total and , reducing numerical errors in long-term forecasts of phenomena like jet streams and ocean eddies. structures in these PDEs further aid in preserving such conservations through integrators.

Numerical and Computational Methods

Discretization Techniques

Discretization techniques for nonlinear partial differential equations (PDEs) involve approximating the continuous problem on a or , transforming it into a solvable of nonlinear algebraic equations. These methods are essential for handling the complexities introduced by nonlinearity, such as steep gradients, shocks, and singularities, while preserving key physical like conservation laws where applicable. Common approaches include finite differences, finite elements, and finite volumes, each tailored to specific classes of nonlinear PDEs, with adaptivity enhancing efficiency near solution irregularities. Finite difference methods approximate derivatives using Taylor expansions on structured grids, proving particularly effective for nonlinear hyperbolic conservation laws. Upwind schemes bias the stencil in the direction of wave propagation to stabilize solutions against oscillations, as developed in early work on systems of nonlinear hyperbolic equations. For instance, in scalar conservation laws like the , upwind differencing resolves nonlinear by evaluating fluxes based on the sign of the speed. Nonlinear in these schemes is ensured by the Courant-Friedrichs-Lewy (CFL) , which limits the time step \Delta t such that \Delta t \leq \frac{\Delta x}{\max |f'(u)|}, preventing information from overtaking the numerical domain. The employs the Galerkin projection, where test and trial functions from a finite-dimensional —typically polynomials—weakly enforce the PDE, making it suitable for nonlinear PDEs with variational structure, such as those arising from energy minimization. In this framework, the weak form integrates the nonlinear operator against test functions over elements, yielding a nonlinear algebraic after . Solving this often relies on , which iteratively linearizes the residual around the current approximation using the , converging quadratically near solutions for well-posed problems. Finite volume methods discretize the domain into control volumes and enforce integral conservation laws by balancing fluxes across faces, ideal for nonlinear PDEs exhibiting shocks. The Godunov scheme, a cornerstone for hyperbolic systems, solves local Riemann problems at interfaces to compute exact or approximate fluxes, enabling robust capture of discontinuities without artificial smoothing. Originally formulated for equations, it applies broadly to quasilinear forms like u_t + \nabla \cdot f(u) = 0, where shocks form due to nonlinearity. Adaptive strategies, such as h-p refinement, dynamically adjust the mesh resolution (h) and polynomial order (p) to target singularities or high-gradient regions in nonlinear PDE solutions, achieving exponential convergence rates. In h-refinement, mesh size decreases near irregularities, while p-refinement increases polynomial degree for smoother parts; combining both optimizes computational cost for problems like those with corner singularities in elliptic nonlinearities. These techniques rely on error estimators to guide refinement, ensuring efficient resolution without global over-refinement.

Error Analysis and Stability

In the numerical solution of nonlinear partial differential equations (PDEs), error analysis and stability assessment are crucial for ensuring the reliability of approximations derived from discretization techniques. Consistency measures how well the discrete scheme approximates the continuous PDE as the grid size approaches zero, while stability ensures that small perturbations in initial data or rounding errors do not amplify uncontrollably over time. Convergence, the ultimate goal, follows from these properties under certain conditions, but nonlinearities complicate the classical frameworks applicable to linear PDEs. The Lax equivalence theorem, which states that for linear PDEs with a well-posed , a consistent scheme is convergent it is , does not generally hold for nonlinear PDEs due to the lack of superposition principles and the potential for nonlinear instabilities. Instead, nonlinear settings require adapted analyses, such as nonlinear stability, where the scheme is linearized around a background solution by freezing nonlinear coefficients at specific points, allowing mode amplification factors to be assessed for bounded growth. This approach reveals that stability in nonlinear cases often demands stricter conditions than in linear ones, as interactions between modes can lead to unpredicted error growth. Error estimates quantify the difference between the exact u and the numerical u_h, typically yielding bounds of the form \|u - u_h\|_{L^2} \leq C h^k for spatial step size h and order k, where C is a constant independent of h under suitable regularity assumptions. A priori bounds, derived via methods, exploit conserved or dissipated quantities in the PDE—such as functionals—to establish these estimates without solving the ; for instance, in semilinear , multiplying the by a test and integrating by parts yields estimates bounding the in Sobolev norms. These methods are particularly effective for nonlinear problems preserving structure, like , where the norm provides a natural control on errors. Stability challenges in nonlinear PDEs often arise from stiff nonlinearities, which introduce rapid variations or multi-scale behaviors that explicit schemes struggle to handle without violating time-step restrictions. Implicit schemes address this by solving algebraic systems at each step, offering unconditional for certain dissipative nonlinearities, though at higher computational . For nonlinear PDEs, the Courant-Friedrichs-Lewy (CFL) condition remains relevant for explicit methods, requiring the time step \Delta t to satisfy \Delta t \leq C \frac{h}{\lambda_{\max}}, where \lambda_{\max} is the maximum characteristic speed, which can vary nonlinearly and tighten near shocks or fronts. Nonlinear extensions of analysis confirm that violating such conditions leads to instability, as seen in discretizations where explicit upwind schemes require CFL numbers below 1 to prevent oscillations. Recent advancements in the have integrated to enhance error control in solvers for nonlinear PDEs, particularly through adaptive strategies that dynamically adjust resolutions based on residual or posterior indicators. For example, neural network-augmented solvers correct coarse-grid in , achieving adaptive refinement that reduces function evaluations while maintaining accuracy in stiff regimes, as demonstrated in frameworks learning surrogate models for time-stepping. These ML-enhanced methods, often combined with physics-informed networks, provide a posteriori estimates via , enabling reliable convergence for complex nonlinear dynamics like those in fluid turbulence.

High-Performance Computing Applications

High-performance computing (HPC) plays a crucial role in simulating nonlinear partial differential equations (PDEs), where the inherent nonlinearity amplifies computational demands, requiring scalable algorithms to handle large-scale, high-dimensional problems in fields like and . Traditional serial solvers often fail to meet these needs due to the exponential growth in , prompting the adoption of ization strategies that distribute workloads across thousands of processors. , such as the finite element tearing and interconnecting (FETI) and balancing domain decomposition by constraints (BDDC) approaches, enable efficient solution of nonlinear elliptic and parabolic PDEs by partitioning the spatial into subdomains solved concurrently, with consistency enforced through coarse-grid corrections. These methods have demonstrated up to 100,000 cores for nonlinear problems, reducing solution times from days to hours on supercomputers. For spectral methods, which excel in resolving smooth solutions of nonlinear PDEs like the Navier-Stokes equations, GPU acceleration has emerged as a key enabler for high-throughput computations. By leveraging the parallel architecture of graphics processing units (GPUs), pseudo-spectral integrations can achieve speedups of over 100x compared to CPU-based implementations for stochastic nonlinear PDEs, such as those modeling turbulent flows. Tools like cuPSS facilitate this by implementing fast transforms and nonlinear term evaluations directly on GPUs, allowing efficient processing of high-resolution simulations. Error analysis from techniques informs these choices, ensuring that parallel implementations maintain without excessive overhead. Multigrid methods adapted for nonlinear PDEs, particularly the full approximation scheme (FAS), address iterative convergence challenges by performing corrections across multiple grid levels, incorporating the full nonlinear residual for accurate prolongation and restriction. FAS-based solvers, such as those for the or heterogeneous diffusion problems, converge in a mesh-independent number of cycles, often 10-20 for complex nonlinearities, making them ideal for large-scale HPC applications where traditional methods diverge. Recent extensions, including aggregation-based nonlinear multigrid, further enhance robustness for variational inequalities arising in nonlinear elasticity. Uncertainty quantification (UQ) in nonlinear PDEs, which account for parametric variabilities like random coefficients in porous media flow, relies on methods parallelized via HPC to sample high-dimensional probability spaces efficiently. Multilevel variants reduce variance through hierarchical sampling, achieving mean-square error convergence at costs near-optimal for expectations of quantities of interest, with parallel efficiency scaling to petascale systems for problems like elliptic PDEs with log-normal conductivities. These approaches have been applied to quantify risks in engineering designs, estimating failure probabilities with 10^6-10^8 samples in feasible wall-clock times. As of 2025, AI-driven surrogate models, including (PINNs) and neural operators, are revolutionizing real-time simulations of nonlinear PDEs in engineering contexts such as and structural optimization. PINNs embed PDE constraints directly into training, enabling surrogate predictions with relative errors below 2% compared to finite element solutions for nonlinear elliptic problems in , and offering computation 4x faster than classical PINNs once trained. Latent neural operators extend this by learning low-dimensional representations for parametric families, facilitating real-time predictions for multiscale physical systems significantly faster than traditional numerical solvers.

Applications and Recent Developments

Physical Sciences Applications

Nonlinear partial differential equations (PDEs) play a central role in modeling complex phenomena in , where the Navier-Stokes equations capture the evolution of viscous, incompressible flows and are essential for understanding . These equations describe how and fields interact through nonlinear terms, leading to chaotic, multiscale behaviors observed in turbulent regimes, such as those in atmospheric flows or engineering applications like design. The nonlinearity introduces challenges like energy cascades from large to small scales, as highlighted in studies of triad interactions that reveal the dynamic interplay driving turbulent dissipation. In inviscid , the Euler equations govern the formation of shock waves, where nonlinear characteristics propagate discontinuities across the flow field. These equations model compressible flows in scenarios like supersonic or astrophysical jets, where abrupt changes in and occur, necessitating shock-capturing techniques to resolve the singularities. Seminal analyses have shown that multidimensional shocks in the Euler framework can be formulated as free boundary problems, providing insights into stability and in high-speed flows. The of exemplify nonlinear hyperbolic PDEs in gravitational physics, relating to the distribution of mass and energy through a of ten coupled, equations. Their nature allows for well-posed initial value problems, enabling the study of phenomena like mergers or cosmological expansion, where nonlinearities amplify small perturbations into global structures. Reviews of hyperbolic formulations emphasize how these equations ensure causal propagation of , aligning with observational data from events such as those detected by . In , the models the propagation of solitary waves, or solitons, in optical fibers, where balances to maintain pulse integrity over long distances. This enables high-bit-rate by mitigating signal , with solitons serving as stable carriers for data transmission in fiber networks. Seminal work demonstrated that these solitons arise from the integrable structure of the equation, allowing exact solutions that predict stable propagation in anomalous regimes. Recent developments in the 2020s have advanced the application of nonlinear PDEs in plasma physics for fusion modeling, particularly in tokamak devices where magnetohydrodynamic (MHD) equations simulate edge turbulence and transport barriers critical to sustaining high-temperature plasmas. These models incorporate nonlinear wave interactions, such as lower hybrid waves in collisional environments, to predict confinement efficiency and mitigate instabilities like edge-localized modes. Data-driven approaches have further refined reduced-order PDE models from full kinetic simulations, accelerating predictions for ITER-scale fusion experiments. Numerical methods have enabled large-scale simulations of these systems, facilitating real-time control strategies.

Biological and Engineering Contexts

In biological systems, nonlinear partial differential equations (PDEs) play a crucial role in modeling and spatial dynamics. Reaction-diffusion equations, pioneered by , describe how chemical morphogens interact through and autocatalytic reactions to generate self-organizing patterns such as spots and stripes observed in animal coats and . These models demonstrate instability in uniform states leading to Turing patterns when rates differ between activator and inhibitor species, providing a foundational mechanism for in embryos. Another key application in is the Fisher-Kolmogorov-Petrovsky-Piskunov (Fisher-KPP) , which models the propagation of population fronts in ecological and evolutionary contexts. This nonlinear PDE captures traveling wave solutions representing the invasion of a into unoccupied , with the wave speed determined by the growth rate and diffusion coefficient, influencing predictions of epidemic spread and gene propagation. In engineering, nonlinear PDEs address challenges in and control systems. The von Kármán equations model large deflections of thin elastic plates under transverse loading, incorporating geometric nonlinearity from membrane stresses that couple in-plane and out-of-plane deformations, essential for designing components and lightweight structures. In , the Hamilton-Jacobi-Bellman (HJB) equation provides the value function for optimal control of nonlinear dynamical systems, deriving feedback policies that minimize cost functionals in applications like and guidance. For , phase-field models simulate microstructure evolution during phase transformations and solidification. These diffuse-interface approaches use order parameters to track interfaces without explicit tracking, capturing nonlinear interactions in alloy processing and predicting or driven by minimization. Post-2020 developments have extended nonlinear PDEs to modeling with nonlocal interactions, where terms account for long-range , improving forecasts of outbreaks beyond local assumptions. In bioengineering, advances as of 2025 incorporate nonlinear PDEs into models of tissue growth, such as simulations that couple vascularization with oxygen dynamics to guide scaffold design for .

Emerging Techniques (Post-2020 Advances)

Since 2020, (PINNs) have emerged as a powerful data-driven approach for discovering and solving nonlinear partial differential equations (PDEs) from sparse observational data, integrating physical laws directly into training to enforce PDE constraints. This method addresses the challenge of problems, where underlying PDEs are inferred from data without prior knowledge of the exact form, achieving high accuracy in scenarios with limited measurements. For instance, extensions like conservative PINNs incorporate conservation laws to improve stability in simulations, reducing training errors by up to 50% compared to traditional least-squares methods. Recent advancements, such as PIKANs (physics-informed Kolmogorov-Arnold networks), further enhance expressivity by using learnable activation functions, enabling better handling of high-dimensional nonlinearities with convergence rates improved by factors of 2-5 in benchmark tests. These techniques have been applied to discover governing equations in complex systems, demonstrating robustness in noisy data environments typical of real-world measurements. Nonlocal operators in fractional nonlinear PDEs have gained traction post-2020 for modeling processes, where standard diffusion fails to capture long-range interactions or memory effects in heterogeneous media. Fractional derivatives, such as Caputo or Riemann-Liouville types, introduce nonlocality that better describes subdiffusive or superdiffusive behaviors observed in porous media and biological tissues. Numerical schemes like methods combined with L1 approximations have been developed to solve these equations efficiently, achieving second-order accuracy in space and first-order in time for nonlinear advection-diffusion models. Recent work on regularized general fractional derivatives with Sonin kernels extends this to multiterm , providing analytical solutions for stability and uniqueness under mild regularity assumptions, which has implications for phase transitions and viscoelastic materials. These advances highlight the role of fractional nonlinear PDEs in simulating irregular transport, such as in tumor growth models where nonlocal effects motivate biological applications. Machine learning-enhanced reduced-order models (ROMs) have addressed the computational bottleneck of high-dimensional nonlinear PDEs since 2020, projecting solutions onto low-dimensional manifolds while preserving dynamics through data-driven techniques. (POD) integrated with deep neural networks enables non-intrusive ROMs that capture nonlinear interactions, reducing simulation times from hours to minutes for parametrized systems like Navier-Stokes equations. Koopman , randomized for efficiency, has led to twin models that not only approximate but also explain nonlinear behaviors via orthogonal decompositions, with error reductions of 30-40% over classical POD-Galerkin methods in turbulent flow predictions. These ROMs excel in real-time applications, such as in high-dimensional parameter spaces, by leveraging autoencoders for robust . Theoretical progress from 2023 to 2025 has focused on resolution in supercritical nonlinear PDEs, where classical regularity theory breaks down due to strong nonlinearities. solutions for fully nonlinear elliptic PDEs satisfying Jacobi inequalities exhibit removability of half-line singularities, implying that isolated singularities can be extended to classical solutions under geometric constraints, as proven via barrier methods and maximum principles. In stochastic settings, supercritical SQG equations demonstrate well-posedness in Besov spaces, resolving potential blow-up through estimates that bound formation. AI-guided ansatze, using large language models to propose trial solutions or transformations, have facilitated linearization of select nonlinear PDEs by automating for change-of-variable substitutions, achieving exact reductions in toy models like with minimal human intervention. These hybrid approaches combine rigorous analysis with computational discovery to tackle longstanding open problems in management.

Notable Equations

Fluid Dynamics Equations

In fluid dynamics, nonlinear partial differential equations (PDEs) play a central role in modeling the motion of viscous and inviscid fluids, capturing phenomena such as , shock waves, and wave propagation. These equations arise from principles of , , and , often leading to systems that are challenging to solve analytically due to their nonlinearity. Prominent examples include the Navier-Stokes equations for viscous flows, the Euler equations for inviscid flows, and the for approximating long-wave dynamics in shallow fluids. The Navier-Stokes equations describe the motion of Newtonian fluids, balancing inertial forces, gradients, viscous diffusion, and external forces. For an incompressible fluid with constant density \rho and kinematic \nu, the momentum equation is given by \frac{\partial \mathbf{u}}{\partial t} + (\mathbf{u} \cdot \nabla) \mathbf{u} = -\frac{\nabla p}{\rho} + \nu \Delta \mathbf{u} + \mathbf{f}, where \mathbf{u} is the velocity field, p is the , and \mathbf{f} represents body forces per unit mass, coupled with the incompressibility constraint \nabla \cdot \mathbf{u} = 0. This system, originally derived by in 1822 and refined by George Gabriel Stokes in 1845, exhibits nonlinear terms that can lead to complex behaviors like boundary layers and . The Euler equations represent the inviscid limit of the Navier-Stokes equations, obtained by setting \nu = 0, which simplifies the equation to \frac{\partial \mathbf{u}}{\partial t} + (\mathbf{u} \cdot \nabla) \mathbf{u} = -\frac{\nabla p}{\rho} + \mathbf{f}, again with \nabla \cdot \mathbf{u} = 0 for incompressible flows. These equations, formulated by Leonhard Euler in 1757, focus on inviscid, irrotational, or rotational flows and are particularly useful for high-Reynolds-number regimes where is negligible. In form, taking the of the equation yields the equation \frac{D \boldsymbol{\omega}}{Dt} = (\boldsymbol{\omega} \cdot \nabla) \mathbf{u} in three dimensions, where \boldsymbol{\omega} = \nabla \times \mathbf{u} is the , highlighting vortex stretching and mechanisms absent in two dimensions. The model the dynamics of free-surface flows where the horizontal length scale greatly exceeds the vertical depth, such as tsunamis or tidal waves, approximating the vertically integrated Navier-Stokes equations. For a of constant under g, the system consists of the \frac{\partial h}{\partial t} + \nabla \cdot (h \mathbf{u}) = 0 and the momentum equation \frac{\partial (h \mathbf{u})}{\partial t} + \nabla \cdot (h \mathbf{u} \otimes \mathbf{u} + \frac{1}{2} g h^2 \mathbf{I}) = -g h \nabla b, where h is the water depth, \mathbf{u} is the depth-averaged horizontal velocity, and b is the bottom ; this forms a nonlinear system supporting wave propagation and shocks. Derived in the late and widely used in , these equations capture nonlinear effects like wave steepening and breaking. A major open challenge in nonlinear PDEs from is the and of solutions to the three-dimensional incompressible Navier-Stokes equations. Posed as one of the Clay Mathematics Institute's in 2000, the question asks whether smooth, globally defined solutions exist for all smooth initial data and forces in \mathbb{R}^3 or periodic domains, or if finite-time blow-up occurs; no resolution has been achieved despite partial results on weak solutions and regularity criteria.

Wave and Quantum Equations

Nonlinear partial differential equations (PDEs) modeling wave propagation often incorporate nonlinear terms to capture phenomena such as wave steepening, shock formation, and soliton stability, which are absent in linear models. A prototypical example is the nonlinear , given by \frac{\partial^2 u}{\partial t^2} - c^2 \Delta u = f(u), where c > 0 is the wave speed, \Delta is the , and f(u) represents nonlinear forcing, such as f(u) = |u|^{p-1}u for power-law nonlinearity. This equation arises in contexts like and relativistic field theories, where solutions can develop finite-time blow-up or behaviors depending on the exponent p and spatial dimension. Seminal analyses of its well-posedness and global existence rely on energy methods and Strichartz estimates, as detailed in foundational treatments of semilinear wave equations. The Korteweg-de Vries (KdV) equation exemplifies nonlinear dispersive waves, particularly in shallow-water dynamics: \frac{\partial u}{\partial t} + u \frac{\partial u}{\partial x} + \frac{\partial^3 u}{\partial x^3} = 0. Derived in the late to resolve paradoxes in solitary wave observations, it balances nonlinear steepening with , yielding stable solutions that maintain shape during propagation. These solitons, first experimentally noted by John Scott Russell in 1834, were theoretically modeled by KdV to describe long waves in rectangular canals with small amplitude relative to water depth. In water wave applications, the equation approximates surface waves where nonlinearity arises from the advective term u u_x, enabling predictions of wave interactions without dissipation. The original derivation expanded the Euler equations for irrotational flow, incorporating higher-order corrections beyond linear shallow-water theory. In , the nonlinear Schrödinger (NLS) equation governs envelope in coherent wave packets, such as i \frac{\partial \psi}{\partial t} + \Delta \psi + |\psi|^2 \psi = 0, where \psi is the complex and the cubic nonlinearity |\psi|^2 \psi accounts for self-interaction. This focusing form models phenomena like modulation instability and bright solitons, while the defocusing variant (-|\psi|^2 \psi) stabilizes dark solitons. In the context of Bose-Einstein condensates (BECs), the equation takes the form of the Gross-Pitaevskii equation, describing the macroscopic of a dilute quantum gas at near-absolute zero, where interatomic repulsion or attraction drives the nonlinearity. First proposed for superfluid vortices, it accurately predicts BEC ground states and following experimental realizations in 1995, with applications to vortex lattices and superfluid flow in trapped atomic gases. The equation emerges from of the second-quantized many-body , treating quantum fluctuations via the Bogoliubov approximation.

Reaction-Diffusion Systems

Reaction-diffusion systems are a class of nonlinear partial differential equations that model the interplay between diffusive transport and nonlinear terms, often arising in chemical, biological, and ecological contexts. These equations typically take the form of \partial_t \mathbf{u} = D \Delta \mathbf{u} + \mathbf{f}(\mathbf{u}), where \mathbf{u} represents concentrations or densities, D is a , \Delta is the Laplacian, and \mathbf{f} captures local . Such systems are renowned for producing spatiotemporal patterns and propagating fronts due to the nonlinearity in \mathbf{f}, distinguishing them from linear equations. The -Kolmogorov-Petrovsky-Piskunov (Fisher-KPP) equation is a prototypical one-component equation given by \partial_t u = D \Delta u + r u (1 - u), where u(\mathbf{x}, t) denotes the density of a or substance, D > 0 is the diffusion coefficient, and r > 0 is the growth rate. This equation admits traveling wave solutions that propagate at a minimum speed of $2\sqrt{rD}, representing the advance of a stable state (e.g., u=1) invading an unstable one (e.g., u=0). The model was independently introduced by to describe the spatial spread of advantageous genes in populations and by Kolmogorov, Petrovsky, and Piskunov in the context of with growth. The Gray-Scott model extends reaction- to two components, simulating an autocatalytic where U and V interact via \partial_t U = D_U \Delta U - U V^2 + F(1 - U), \partial_t V = D_V \Delta V + U V^2 - (F + k) V, with D_U, D_V as diffusion rates, F as an external feed rate, and k as a rate. This system is capable of generating complex Turing patterns, such as spots and stripes, through diffusion-driven instability when D_V \ll D_U. The model draws from autocatalytic reactions studied in and gained prominence through numerical explorations of . The Nagumo equation, a one-dimensional reaction-diffusion model for excitable media, is expressed as \partial_t u = D \partial_{xx} u + f(u), where f(u) is typically a cubic nonlinearity like f(u) = u(1 - u)(u - a) with $0 < a < 0.5, enabling bistable dynamics. It supports propagating fronts that model impulses along nerve axons, transitioning between resting and excited states. This equation originates from simplifications of neuronal and has been analyzed for wave propagation properties. These equations find applications in , where reaction-diffusion mechanisms underpin in developing organisms, as theorized by Turing; in , modeling flame propagation fronts; and recently in , where Fisher-KPP describes the invasion speed of species like . In biological , such systems briefly reference contexts like markings.

References

  1. [1]
    [PDF] A review of numerical methods for nonlinear partial differential ...
    Jul 20, 2012 · A PDE is said to be nonlinear if the relations between the unknown functions and their partial derivatives involved in the equation are ...
  2. [2]
    [PDF] Introduction to Partial Differential Equations - UCSB Math
    May 21, 2003 · Nonlinear partial differential equations include the Einstein field equations from general relativity and the Navier-Stokes equations which ...
  3. [3]
    [PDF] partial differential equations - Math (Princeton)
    Fully nonlinear equations appear however in connection to the theory of characteristics of linear PDE, which we discuss at length below, or in geometry. 1.
  4. [4]
    [PDF] Partial Differential Equations - Berkeley Math
    A partial differential equation. (PDE) is an equation involving an unknown func- tion u of more than one variable and certain of its partial derivatives. The ...<|separator|>
  5. [5]
    [PDF] introduction to partial differential equations - UNCW
    Jun 19, 2015 · PDEs. Let u = u(x, y). Then, the general form of a linear second order partial differential equation is given by a(x, y)uxx +2b(x, y)uxy +c ...
  6. [6]
    [PDF] Partial Differential Equation: Penn State Math 412 Lecture Notes
    These equations are representatives of the three classes of second order partial differential equations. Definition 1.57 (Classes of Second Order PDE's [Far93]) ...
  7. [7]
    Methods of Mathematical Physics, vol. II. Partial Differential Equations
    Methods of Mathematical Physics, vol. II. Partial Differential Equations Available ... This content is only available via PDF. Open the PDF for in another window.
  8. [8]
    [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 ...
  9. [9]
    [PDF] classification and principle of superposition for
    Linear Partial Differential Equations. A partial differential equation (PDE) ... to (7) and (12) are linear PDES while equations (8) to (11) are nonlinear PDEs.
  10. [10]
    [PDF] Solitons and Dispersive Shock Waves - LSEC
    Nonlinear shock waves. ▷ Classical or viscous shock waves (VSWs). ▻ Characterized by localized steep gradient across a shock front and dissipation of energy.
  11. [11]
    [PDF] 10 Partial Differential Equations and Fourier methods
    10.2 Solving PDEs with Fourier methods. The Fourier transform is one example of an integral transform: a general technique for solving differential equations ...<|control11|><|separator|>
  12. [12]
    [PDF] Nonlinear heat transfer - arXiv
    This paper considered heat transfer with temperature-dependent thermal conductivity. The one-dimensional unsteady heat conduction equation was solved ...
  13. [13]
    [PDF] Separation of Variables and Exact Solutions to Nonlinear PDEs
    Exact solutions to nonlinear equations facilitate a better understanding of the mechanisms of complex nonlinear effects such as the spatial localization of ...
  14. [14]
    [PDF] 6 The Burgers equation
    Burgers proposed equation (143) as a made-up, toy model for turbulence. It mimics the Navier-Stokes equations of fluid motion through its fluid-like ex-.
  15. [15]
    Burgers, J.M. (1948) A Mathematical Model Illustrating the Theory of ...
    In this paper we study inviscid and viscid Burgers equations with initial conditions in the half plane.
  16. [16]
    [PDF] 1 Introduction
    A k-th order partial differential equation is fully nonlinear if the highest-order deriva- tives of u appear nonlinearly in the equation. Example 7. • u2 x + u2.
  17. [17]
    [PDF] Nonlinear Partial Differential Equations - Uni Ulm
    N = {1,2,...}, N0 = {0,1,...} is the natural numbers without and with 0, respectively;. Z, Q, R, C the set of integers, rational numbers, real numbers, ...
  18. [18]
  19. [19]
    [PDF] Notes on Partial Differential Equations John K. Hunter - UC Davis Math
    Abstract. These are notes from a two-quarter class on PDEs that are heavily based on the book Partial Differential Equations by L. C. Evans, together.
  20. [20]
    [PDF] Nonlinear Elliptic Equations - Michael Taylor
    Nonlinear elliptic equations arise from calculus of variations in geometry and continuum mechanics. Semilinear equations are the simplest form, with a linear ...
  21. [21]
    Classification of PDEs
    ### Summary of Classification of Nonlinear PDEs
  22. [22]
    New strong maximum and comparison principles for fully nonlinear ...
    Dec 22, 2018 · We introduce a notion of subunit vector field for fully nonlinear degenerate elliptic equations. We prove that an interior maximum of a viscosity subsolution ...
  23. [23]
    Existence and Uniqueness of Bounded Weak Solutions of a ...
    Oct 30, 2009 · In part I, the existence and uniqueness are established for Sobolev solutions of a class of semilinear parabolic partial differential equations.
  24. [24]
    [PDF] nonlinear wave equations - UChicago Math
    Aug 29, 2016 · This means that solutions of hyperbolic PDE have the properties of solutions to wave equations. For instance, finite speed of propagation of ...
  25. [25]
    [PDF] Nonlinear PDEs of mixed type arising in mechanics and geometry
    Mar 10, 2008 · The mixed type equations are corresponding to the transonic flows, which means the speed is above sound speed at some place and below sound ...
  26. [26]
    [PDF] The method of characteristics applied to quasi-linear PDEs
    Oct 26, 2005 · The PDE (5) is called quasi-linear because it is linear in the derivatives of u. It is NOT linear in u(x, t), though, and this will lead to ...
  27. [27]
    [PDF] LECTURE 3 A conservation law is a first order PDE of the form ut + ...
    Conservation laws are examples of quasilinear equations, that is, equations that are linear in the highest order derivatives, with coefficients that depend ...
  28. [28]
    [PDF] Hyperbolic Partial Differential Equations and Conservation Laws
    Quasilinear if c = c(u) for example. Standard multidimensional form: Ut + X ... • Conservation laws or quasilinear hyperbolic equations ut + uux = 0.
  29. [29]
    [PDF] Existence, uniqueness and removable singularities for nonlinear ...
    Classically, a fully nonlinear partial differential equation for a smooth function u(x) on an open set X ⊂ Rn is written in the form f(x, u, Du, D2u) = 0.
  30. [30]
    [PDF] An Accelerated Method for Nonlinear Elliptic PDE
    Sep 29, 2015 · In this work, we are concerned with second order elliptic operators, which can be written in the form of H(u) = H(x,u, Du, D2u).
  31. [31]
    [PDF] The Monge-Amp`ere Equation - UCI Mathematics
    Structure of the Equation: The Monge-Ampere equation is. detD2u = f. In order to get a maximum principle, we restrict ourselves to u convex and f ≥ 0; if u and.
  32. [32]
    [PDF] an introduction to viscosity solutions for fully nonlinear pde ... - arXiv
    Nov 11, 2014 · The purpose of these notes is to give a quick and elementary, yet rigorous, presentation of the rudiments of the so-called theory of Viscosity ...
  33. [33]
    [PDF] Chapter 3: Sobolev spaces - UC Davis Math
    We can use the Sobolev inequality to prove various embedding theorems. In general, we say that a Banach space X is continuously embedded, or embedded for.
  34. [34]
    (PDF) Sobolev spaces and embedding theorems - ResearchGate
    The last part (nonlinear PDE) is subdivided into four chapters dealing with variational methods (first and second variation, existence of minimizers for ...
  35. [35]
    [PDF] The Calculus of Variations - College of Science and Engineering
    Jun 6, 2024 · The calculus of variations is a field of mathematics concerned with minimizing (or maximizing) functionals (that is, real-valued functions ...
  36. [36]
    [PDF] Partial Differential Equations 2 Variational Methods - mediaTUM
    Using its methods to prove existence of a minimizer is called the direct method of the calculus of variations. That such a minimizer solves the Euler equation ...
  37. [37]
    [PDF] Fixed Point Methods in Nonlinear Analysis - UChicago Math
    Aug 29, 2014 · When applicable, the Banach fixed point theorem is especially useful as it both guarantees the existence and uniqueness of a fixed point.
  38. [38]
    [PDF] Schauder fixed point theorem - Let X be a real Banach space.
    Now we will show how Schaefer's theorem applies to prove existence of solutions to some non-linear elliptic PDE'S. Consider the boundary value problem.
  39. [39]
    existence of solutions and galerkin approximations for nonlinear ...
    Introduction—Preliminaries. In this paper we are concerned with existence and approximation results for nonlinear functional evolution.
  40. [40]
    [PDF] Nonlinear Evolution Equations1 - UC Davis Math
    Partial differential equations (PDE's) can be regarded as evolution equations on an infinite dimensional state space. The solution u(x, t) belongs to a function ...
  41. [41]
    Monotone Operators in Banach Space and Nonlinear Partial ...
    This book is about monotone operators and nonlinear semigroup theory and their application to partial differential equations. The primary readership is advanced ...<|control11|><|separator|>
  42. [42]
    Global well-posedness for the nonlinear generalized parabolic ...
    Sep 21, 2021 · Furthermore, we prove the uniqueness of the solution by using classical energy estimates.
  43. [43]
    A remark on the uniqueness of solutions to hyperbolic conservation ...
    May 26, 2023 · Given a strictly hyperbolic n\times n system of conservation laws, it is well known that there exists a unique Lipschitz semigroup of weak solutions.
  44. [44]
    [2507.01818] Schauder-type estimates and applications - arXiv
    Jul 2, 2025 · The Schauder estimates are among the oldest and most useful tools in the modern theory of elliptic partial differential equations (PDEs).
  45. [45]
    Instability and nonuniqueness for the 2d Euler equations in vorticity ...
    In this expository work, we present Vishik's theorem on non-unique weak solutions to the two-dimensional Euler equations on the whole space.
  46. [46]
    Non-uniqueness of admissible solutions for the 2D Euler equation ...
    Abstract page for arXiv paper 2304.09578: Non-uniqueness of admissible solutions for the 2D Euler equation with $L^p$ vortex data.
  47. [47]
    [PDF] Well-Posed Problems - UNL Math
    According to Hadamard, a problem is well-posed (or correctly-set) if a. it has a solution, b. the solution is unique, c. the solution depends continuously ...
  48. [48]
    [PDF] Lecture 1. Introduction to well- and ill-posed problems.
    Nov 19, 2009 · Definition 1.2.1. (Hadamard, 1932) The problem (1.1) is said to be well-posed in the sense of Hadamard if the following conditions ...
  49. [49]
    The “Well-Posedness” of Differential Equations - Cantor's Paradise
    Nov 21, 2020 · Hadamard is suggesting that the well posed problem will be the one which corresponds intuitively to a physical scenario! Hadamard's first ...
  50. [50]
    Non-uniqueness of Leray solutions of the forced Navier-Stokes ...
    Non-uniqueness ... In a seminal work, Leray (1934) demonstrated the existence of global weak solutions to the Navier-Stokes equations in three dimensions.
  51. [51]
    Non-uniqueness of Leray solutions of the forced Navier-Stokes ...
    Dec 6, 2021 · We exhibit two distinct Leray solutions with zero initial velocity and identical body force. Our approach is to construct a `background' solution which is ...Missing: high Reynolds
  52. [52]
    [PDF] 3. Backward heat equation ? - People
    In a word, the backward heat equation is ill-posed because all solutions are instantly swamped by high-frequency noise.Missing: nonlinear | Show results with:nonlinear
  53. [53]
    A nonlinearly backward heat problem: uniqueness, regularization ...
    Nov 28, 2010 · We consider the problem of finding, from the final data u(x,T) = ϕ(x), the temperature u satisfying. The problem is nonlinearly ill-posed.<|separator|>
  54. [54]
    A Mathematical Analysis of Blowup for Thermal Reactions—the ...
    The equations describing the induction period process for a super-critical, high-activation energy thermal explosion in a bounded domain are studied.
  55. [55]
    The Problem Of Blow-Up In Nonlinear Parabolic Equations
    Aug 7, 2025 · theory of blow-up, and, more generally, the study of singularities. Blow-up in reaction-diffusion equations: basic problems and ...
  56. [56]
    Perturbation methods in fluid mechanics : Van Dyke, Milton
    Jul 24, 2019 · Perturbation methods in fluid mechanics. by: Van Dyke, Milton. Publication date: 1964. Topics: Fluid dynamics, Perturbation (Mathematics).
  57. [57]
    [PDF] Linearized stability analysis of nonlinear partial differential equations
    Sep 18, 2015 · The key to using Lyapunov's Indirect Method is showing that the linearized generator corresponds to the generator of the Fréchet derivative of ...
  58. [58]
    [PDF] Chapter 4 - UC Davis Math
    These problems can be analyzed using the method of matched asymptotic ex- pansions (MMAE), in which we construct different asymptotic solutions inside and ...Missing: PDE | Show results with:PDE
  59. [59]
    Matched Asymptotic Expansion Solutions of Nonlinear Partial ...
    The solution is found for the case where the null space of B is one-dimensional by a method of matched asymptotic expansions. The form of the solution is shown ...
  60. [60]
    Type Ⅱ finite time blow-up for the energy critical heat equation in
    In this paper we construct type Ⅱ blow-up solutions of (3) for ... , Nonradial type Ⅱ blow up for the energy-supercritical semilinear heat equation, Anal.
  61. [61]
    Blow-up rate for the subcritical semilinear heat equation in non ...
    Oct 20, 2025 · Our main result shows the nonexistence of type II blow-up for possibly sign-changing solutions in the energy subcritical range (n-2)p<n+2.
  62. [62]
    [PDF] Representability in non-linear elliptic Fredholm analysis
    We summarize current work aimed at showing that moduli spaces of solutions to non-linear elliptic Fredholm partial differential equations are derived log smooth.
  63. [63]
    Applications of Lie Groups to Differential Equations - SpringerLink
    In stockThis book provides a solid introduction to those applications of Lie groups to differential equations which have proved to be useful in practice. The ...
  64. [64]
    Applications of Symmetries to Nonlinear Partial Differential Equations
    This review begins with the standard Lie symmetry theory for nonlinear PDEs and explores extensions of symmetry analysis. First, it introduces three key ...
  65. [65]
    [PDF] Symmetry and Explicit Solutions of Partial Differential Equations
    Lie point symmetries of large systems of differential equations, Comp. Phys ... [15] Olver, P.J., Symmetry Groups of Partial Differential Equations, Thesis, ...
  66. [66]
    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.Missing: inverse scattering
  67. [67]
    Integrals of nonlinear equations of evolution and solitary waves
    Lax,. Peter D. Lax. Search for more papers by this author. First published: September 1968. https://doi.org/10.1002/cpa.3160210503. Citations: 2,314. †. This ...
  68. [68]
    [PDF] Lax Pairs
    May 25, 2012 · The idea was originally published by Peter Lax in a seminal paper in 1968 [Lax-68]. A Lax pair consists of the Lax operator L (which is self- ...Missing: original | Show results with:original
  69. [69]
    Multidimensional Integrable Deformations of Integrable PDEs - arXiv
    May 15, 2023 · We present a novel example of a (2+1)-dimensional deformation of KdV equation obtained by generalized deformation.
  70. [70]
    Hodograph Transformations of Linearizable Partial Differential ...
    In this paper an algorithmic method is developed for transforming quasilinear partial differential equations of the form $u_t = g( u )u_{nx} + f( u, ...Missing: original | Show results with:original
  71. [71]
    [1210.3673] Conservation Laws, Hodograph Transformation and ...
    Oct 13, 2012 · For the hyperbolic system of quasilinear first-order partial differential equations, linearizable by hodograph transformation, the conservation ...
  72. [72]
    Minimal Surface Equation - an overview | ScienceDirect Topics
    Examples of Area-Minimizing Surfaces ... When A(f) is the area of the graph of f, the associated Euler–Lagrange equation is the minimal surface equation.<|separator|>
  73. [73]
    [PDF] Minimal surfaces for undergraduates - arXiv
    Jan 7, 2021 · In this section we derive Lagrange's equation of minimal graphs, which is one of the first examples in the calculus of variations for functions ...
  74. [74]
    A Hidden Convexity of Nonlinear Elasticity | Journal of Elasticity
    Jul 16, 2024 · A technique for developing convex dual variational principles for the governing PDE of nonlinear elastostatics and elastodynamics is presented.
  75. [75]
    Existence of a variational principle for PDEs with symmetries and ...
    It is well-known from Noether's theorem that symmetries of a variational functional lead to conservation laws of the corresponding Euler-Lagrange equation.
  76. [76]
    Solving PDEs by variational physics-informed neural networks
    Sep 19, 2022 · We consider the discretization of elliptic boundary-value problems by variational physics-informed neural networks (VPINNs), in which test ...
  77. [77]
    Physics-informed learning of governing equations from scarce data
    Oct 21, 2021 · This work introduces a novel approach called physics-informed neural network with sparse regression to discover governing partial differential equations from ...
  78. [78]
    [PDF] Emmy Noether: Symmetry and Conservation
    An infinite-dimensional variational symmetry group depending upon an arbitrary function corresponds to a nontrivial differential relation among its. Euler– ...<|control11|><|separator|>
  79. [79]
    [math-ph/0108023] Direct construction method for conservation laws ...
    Aug 27, 2001 · Anco, George Bluman. View a PDF of the paper titled Direct construction method for conservation laws of partial differential equations. Part ...
  80. [80]
    Direct construction method for conservation laws of partial ...
    Nov 28, 2002 · An effective algorithmic method is presented for finding the local conservation laws for partial differential equations with any number of ...
  81. [81]
    [PDF] Higher Order Symmetries of Underdetermined Systems of Partial ...
    A system of Euler-Lagrange equations is underdetermined if and only if it admits an infinite dimensional variational symmetry group depending on an arbitrary ...
  82. [82]
    [PDF] On the Partial Difference Equations of Mathematical Physics
    Problems involving the classical linear partial differential equations of mathematical physics can be reduced to algebraic ones of a very much simpler ...
  83. [83]
    A Finite Element Method for Nonlinear Elliptic Problems - SIAM.org
    We present a Galerkin method with piecewise polynomial continuous elements for fully nonlinear elliptic equations. A key tool is the discretization proposed ...Missing: influential papers
  84. [84]
    Truncated newton methods for nonlinear finite element analysis
    References. 1. M. Papadrakakis ... 7. B. Nour-Omid, B.N. Parlett, R.L. Taylor. A Newton-Lanczos method for solution of nonlinear finite element equations.Missing: key reference
  85. [85]
    Aspects of an adaptive hp-finite element method - ScienceDirect.com
    The main components needed for an adaptive hp-version finite element algorithm are discussed: an adaptive hp-refinement strategy, effective methods for ...Missing: seminal | Show results with:seminal
  86. [86]
    [PDF] Nonlinear Stability - HAL
    Jun 3, 2020 · Unfortunately, the Lax Equivalence Theorem no longer holds and we cannot use the same approach (which relies heavily on linearity) to prove ...
  87. [87]
    Stability analysis of a non-linear diffusion-type kinetic equation
    A diffusion-type partial differential equation with non-linear coefficients is analysed for stability in the von Neumann sense, and some numerical examples ...
  88. [88]
    What does the Von Neumann's stability analysis tell us about non ...
    Mar 6, 2016 · However, as the authors realize, this is only applicable to linear PDE's. So the authors work around this by "freezing" the non-linear term, ...Von Neumann analysis on a finite difference hyperbolic schemeFrozen coefficient method (von Neumann stability analysis)More results from mathoverflow.net
  89. [89]
    On Optimal Order Error Estimates for the Nonlinear Schrödinger ...
    This paper focuses on unconditionally optimal error analysis of an uncoupled and linearized Crank--Nicolson Galerkin finite element method for the time- ...
  90. [90]
    [PDF] Error analysis of kernel/GP methods for nonlinear and ... - Caltech
    Oct 10, 2024 · The error estimates demonstrate dimension-benign convergence rates if the solution space of the PDE is smooth enough. We illustrate these points ...
  91. [91]
    [PDF] A discontinuous Galerkin method and its error estimate for nonlinear ...
    We prove the energy conserving property of our scheme and its optimal error estimates in the L2-norm for the solution itself as well as for the auxiliary.
  92. [92]
    Linearly implicit methods for nonlinear PDEs with linear dispersion ...
    A method which removes the CFL stability constraint ... Two classes of implicit-explicit multistep methods for nonlinear stiff initial-value problems.
  93. [93]
    [PDF] 11. Finite Difference Methods for Partial Differential Equations
    May 18, 2008 · With this geometric formulation, the CFL criterion can be applied to both linear and nonlinear transport equations that have non-uniform wave.
  94. [94]
    Nonlinear stability and convergence of finite-difference methods for ...
    In this paper we study the nonlinear stability and the convergence of some simple finite-difference schemes for the numerical solution of problems involving the ...
  95. [95]
    [PDF] Learning differential equations that are easy to solve
    The NFE taken by an adaptive solver depends on how far it can extrapolate the trajectory forward without introducing too much error. For example, for a standard ...
  96. [96]
    ML-PDE: A Framework for a Machine Learning Enhanced PDE Solver
    ML-PDE is a hybrid machine learning technique using a deep neural network to correct errors in coarse-grid simulations of turbulent flows, while estimating ...Missing: adaptive control 2020s<|control11|><|separator|>
  97. [97]
    Nonlinear FETI-DP and BDDC Methods: A Unified Framework and ...
    By changing the order of the operations, new parallel communication can be constructed, where the loosely coupled local problems are nonlinear. We discuss ...Missing: parallelization | Show results with:parallelization
  98. [98]
    A simple GPU implementation of spectral-element methods ... - arXiv
    Sep 30, 2023 · We present in this paper a simple but extremely fast MATLAB implementation on a modern GPU, which can be easily reproduced, for solving 3D Poisson type ...
  99. [99]
    A Multigrid Algorithm for the p-Laplacian - SIAM.org
    We introduce a full approximation storage (FAS) multigrid algorithm to find the finite element solution for a class of nonlinear monotone elliptic problems.
  100. [100]
    $$hp$-Multilevel Monte Carlo Methods for Uncertainty Quantification ...
    We propose a novel $hp$-multilevel Monte Carlo method for the quantification of uncertainties in the compressible Navier--Stokes equations, ...<|separator|>
  101. [101]
    Multi-level physics informed deep learning for solving partial ...
    Nov 1, 2024 · Here we develop a multi-level physics-informed neural network framework where an aggregation model is developed by combining multiple neural networks.
  102. [102]
    Multidimensional Transonic Shock Waves and Free Boundary ...
    Sep 21, 2021 · The nonlinear PDEs under our analysis include the steady Euler equations for potential flow, the steady full Euler equations, the unsteady Euler ...
  103. [103]
    Hyperbolic Methods for Einstein's Equations | Living Reviews in ...
    I review evolutionary aspects of general relativity, in particular those related to the hyperbolic character of the field equations and to the applications.
  104. [104]
    Nonlinear lower hybrid wave equations in collisional tokamak plasmas
    Jul 26, 2024 · A new set of coupled integro-differential nonlinear lower hybrid (LH) wave equations is derived within the framework of a kinetic theory coupled to the Maxwell ...<|separator|>
  105. [105]
    Data-driven discovery of reduced plasma physics models from fully ...
    Sep 9, 2022 · Here we report the data-driven discovery of accurate reduced plasma models, in the form of partial differential equations, directly from first-principles ...
  106. [106]
    [PDF] The Chemical Basis of Morphogenesis AM Turing | Caltech
    May 5, 2007 · A system of reactions and diffusion on a sphere is also con- sidered. Such a system appears to account for gastrulation. Another reaction system ...Missing: seminal | Show results with:seminal
  107. [107]
    Turing patterns, 70 years later | Nature Computational Science
    Aug 14, 2022 · On the 70th anniversary of Alan Turing's seminal paper on morphogenesis, we look back at the history of the paper and its many applications.
  108. [108]
    Revisiting the Fisher–Kolmogorov–Petrovsky–Piskunov equation to ...
    Sep 4, 2019 · A key feature of the Fisher–KPP model is that any positive initial condition with compact support will always evolve into a travelling wave that ...Abstract · Introduction · Results and discussion · Conclusion
  109. [109]
    The Föppl–von Kármán equations of elastic plates with initial stress
    May 18, 2022 · The Föppl–von Kármán (FvK) equations are a set of nonlinear partial differential equations describing the large deflection of linear elastic ...
  110. [110]
    Jacobi--Bellman Equation of Nonlinear Filtering - SIAM.org
    A Max-Plus-Based Algorithm for a Hamilton--Jacobi--Bellman Equation of Nonlinear Filtering | SIAM Journal on Control and Optimization.
  111. [111]
    Phase-Field Models for Microstructure Evolution - Annual Reviews
    Aug 1, 2002 · The phase-field method has recently emerged as a powerful computational approach to modeling and predicting mesoscale morphological and ...
  112. [112]
    [2104.09915] Phase-field modeling of microstructure evolution - arXiv
    Apr 20, 2021 · We briefly review the state-of-the-art in phase-field modeling of microstructure evolution. The focus is placed on recent applications of phase-field ...
  113. [113]
    Modeling nonlocal behavior in epidemics via a reaction–diffusion ...
    Sep 15, 2022 · In this work, we attempt to combine these modeling paradigms via the introduction of a network structure within a reaction–diffusion PDE system.Missing: post- | Show results with:post-
  114. [114]
    Analytical and numerical properties of an extended angiogenesis ...
    Oct 22, 2025 · This paper presents an extended mathematical model for tumor angiogenesis incorporating oxygen dynamics as a main regulator.
  115. [115]
    Anatomy-DT: A Cross-Diffusion Digital Twin for Anatomical Evolution
    Sep 29, 2025 · In contrast, mechanistic models of tumor growth and tissue dynamics, often formulated as reaction–diffusion partial differential equations (PDEs) ...<|control11|><|separator|>
  116. [116]
    A comprehensive review of advances in physics-informed neural ...
    Oct 2, 2024 · This paper discusses the latest advancements in PINNs and their potential applications in complex fluid dynamics, including turbulence, multiphase flows, multi ...Missing: post- | Show results with:post-
  117. [117]
    Iterative Investigation of the Nonlinear Fractional Cahn–Allen and ...
    This work provides significant insights into the modeling of anomalous diffusion and complex dynamic processes in fields such as phase transitions, biological ...
  118. [118]
    Robust reduced-order machine learning modeling of high ...
    This work develops a novel machine-learning-based reduced-order modeling method by integrating SpectralDense layers into autoencoders and incorporating them ...
  119. [119]
    From Jean Leray to the millennium problem: the Navier–Stokes ...
    Nov 9, 2020 · In this article, we will introduce the Navier–Stokes equations, describe their main mathematical problems, discuss several of the most important results.
  120. [120]
    [PDF] The First Five Births of the Navier-Stokes Equation
    As a preliminary step toward a more realistic theory of elasticity, in 1821 Navier announced the general equations of equilibrium and motion for an (isotropic, ...Missing: formulation seminal
  121. [121]
    Navier-Stokes Equation - Clay Mathematics Institute
    This is the equation which governs the flow of fluids such as water and air. However, there is no proof for the most basic questions one can ask.
  122. [122]
    [PDF] An Introduction to the Incompressible Euler Equations - UC Davis Math
    Sep 25, 2006 · The left-hand side of this equation is the material time-derivative of the vorticity and describes the advection of vorticity by the fluid flow.
  123. [123]
    The shallow water equations - Clawpack
    The shallow water equations are: ht+(hu)x=0, (hu)t+(hu2+12gh2)x=0, where h is depth, u is velocity, and g is gravity. These are depth averaged equations.