Fact-checked by Grok 2 weeks ago

Viscosity solution

In mathematics, particularly the analysis of partial differential equations (PDEs), a viscosity solution is a generalized notion of solution introduced by Michael G. Crandall and Pierre-Louis Lions in 1983 for scalar nonlinear first-order Hamilton–Jacobi equations of the form F(x, u, Du) = 0. This framework was subsequently extended to fully nonlinear second-order elliptic and parabolic PDEs, such as F(x, u, Du, D^2u) = 0, where D and D^2 denote the gradient and Hessian, respectively. Unlike classical C^1 or C^2 solutions, viscosity solutions require only continuity of the function u and are defined through inequalities involving smooth test functions that touch u from above or below at local maxima or minima, thereby circumventing issues like characteristic crossing and nonexistence of smooth solutions. The concept relies on the notions of viscosity subsolutions and supersolutions. A continuous function u is a viscosity subsolution if, for every smooth test function \phi such that u - \phi has a local maximum at some interior point x_0, the PDE evaluated at (x_0, u(x_0), D\phi(x_0), D^2\phi(x_0)) satisfies F \leq 0; conversely, u is a viscosity supersolution if u - \phi has a local minimum at x_0 implies F \geq 0. A viscosity solution is then a function that is both a subsolution and a supersolution. This definition ensures consistency with classical solutions: if u is C^2 and satisfies the PDE pointwise, it is a viscosity solution. Viscosity solutions possess several key properties that make them powerful for global analysis. Under structural conditions on F—such as uniform continuity, ellipticity (for second-order cases), and monotonicity in the Hessian variable—they satisfy a comparison principle, implying uniqueness for Dirichlet boundary value problems. They are stable under uniform limits, allowing passage to limits in approximation schemes like the vanishing viscosity method, where solutions to regularized equations converge to the viscosity solution. Existence is often established via the Perron method or probabilistic representations. The theory has broad applications across pure and applied mathematics. In optimal control and dynamic programming, the value function of a control problem is the unique viscosity solution to the associated Hamilton–Jacobi–Bellman equation. It also arises in differential games, front propagation models (e.g., via level-set methods), mean curvature flow, and stochastic processes in finance and engineering.

Definition and Formulation

Formal Definition

Viscosity solutions provide a notion of weak solutions for fully nonlinear second-order partial differential equations (PDEs) of the form F(x, u, Du, D^2 u) = 0 in an open domain \Omega \subset \mathbb{R}^N, where F: \mathbb{R}^N \times \mathbb{R} \times \mathbb{R}^N \times \mathcal{S}^N \to \mathbb{R} is continuous, Du denotes the spatial gradient of u, D^2 u is the Hessian matrix, and \mathcal{S}^N is the space of N \times N symmetric matrices. Here, F is assumed to be degenerate elliptic, meaning that F(x, r, p, X) \geq F(x, r, p, Y) whenever Y - X is positive semidefinite (i.e., Y \geq X in the matrix sense). This condition ensures a form of monotonicity in the second-order term that aligns with the maximum principle for elliptic operators, though degeneracy allows coefficients to vanish. A function u: \Omega \to \mathbb{R} is a viscosity subsolution if it is upper semicontinuous and, for every test function \phi \in C^2(\Omega), at any point x_0 \in \Omega where u - \phi attains a local maximum (so \phi(x_0) = u(x_0) and \phi \geq u in a neighborhood of x_0), the inequality F\bigl(x_0, u(x_0), D\phi(x_0), D^2 \phi(x_0)\bigr) \leq 0 holds. Test functions \phi are smooth C^2 functions that "touch" u from above at x_0, approximating the behavior of u locally via its derivatives. This definition captures the PDE constraint at points of nondifferentiability by substituting the smooth \phi's derivatives for those of u. Dually, u is a viscosity supersolution if it is lower semicontinuous and, for every \phi \in C^2(\Omega), at any point x_0 \in \Omega where u - \phi attains a local minimum (so \phi(x_0) = u(x_0) and \phi \leq u in a neighborhood of x_0), F\bigl(x_0, u(x_0), D\phi(x_0), D^2 \phi(x_0)\bigr) \geq 0. Here, \phi touches u from below, enforcing the PDE in the opposite directional sense. A viscosity solution is a continuous function u: \Omega \to \mathbb{R} that is both a viscosity subsolution and a viscosity supersolution; the continuity follows directly from the upper and lower semicontinuity requirements. Unlike classical solutions, which require u \in C^2(\Omega) to satisfy the PDE pointwise, viscosity solutions relax differentiability while preserving key analytical properties through these test function conditions.

Sub- and Supersolutions

Viscosity subsolutions to the equation F(x, u, Du, D^2 u) = 0 are defined as upper semicontinuous functions u, satisfying \limsup_{y \to x} u(y) \leq u(x) for all x in the domain. This semicontinuity ensures that the relevant second-order jets are well-defined at points of local maxima. In contrast, viscosity supersolutions are lower semicontinuous functions v, with \liminf_{y \to x} v(y) \geq v(x), allowing evaluation at local minima. The core of the definition lies in the touching conditions using smooth test functions. For a subsolution u, if \phi \in C^2 is such that u - \phi attains a local maximum at some point x_0 in the domain (i.e., \phi touches u from above), then F(x_0, u(x_0), D \phi(x_0), D^2 \phi(x_0)) \leq 0. This inequality must hold for all such test functions and points. For a supersolution v, if \phi \in C^2 touches v from below, meaning v - \phi has a local minimum at x_0, the condition is F(x_0, v(x_0), D \phi(x_0), D^2 \phi(x_0)) \geq 0. These conditions capture the PDE inequality in a weak sense, applicable even where u or v is not differentiable. The nonlinearity F is required to satisfy degenerate ellipticity: for symmetric matrices X, Y, F(x, r, p, X) \geq F(x, r, p, Y) whenever X \leq Y (i.e., Y - X is positive semidefinite). This property implies that F is non-increasing in the second derivative argument, a relaxation of uniform ellipticity that accommodates fully nonlinear operators. Degenerate ellipticity underpins structural results, such as maximum principles for subsolutions, by ensuring that the operator preserves ordering in the matrix variable and supports comparison arguments without requiring positive definiteness. Strict subsolutions satisfy the touching inequality with F \leq -\varepsilon < 0 for some \varepsilon > 0, while strict supersolutions have F \geq \varepsilon > 0. These strict versions facilitate approximation schemes, with viscosity solutions emerging as limits in the case \varepsilon \to 0^+, often through regularization or vanishing viscosity methods.

Motivation and Examples

Need for Viscosity Solutions

Classical solutions to partial differential equations (PDEs) typically require the solution to be sufficiently smooth, such as twice continuously differentiable (C²), to allow pointwise substitution into the equation. However, in many nonlinear PDEs, particularly Hamilton-Jacobi equations, this smoothness assumption fails due to the formation of shocks or non-smooth fronts, where characteristics cross and derivatives become undefined or discontinuous. For instance, the evolution of interfaces or wavefronts often leads to singularities in finite time, rendering classical solutions nonexistent beyond initial smooth data. Weak solutions, defined through integral formulations or in a distributional sense, address some smoothness issues by allowing less regular functions that satisfy the PDE almost everywhere. Yet, for nonlinear PDEs such as conservation laws, these weak solutions suffer from non-uniqueness, as multiple functions can satisfy the integral condition without a clear physical selection criterion. This ambiguity arises because the nonlinearity prevents standard techniques like integration by parts from enforcing stability or selecting the appropriate solution among infinitely many candidates. Viscosity solutions emerged to overcome these limitations by drawing inspiration from physical regularization processes, where an artificial viscosity term, such as εΔu, is added to the PDE to produce smooth, unique solutions that converge to a well-defined limit as ε approaches zero. This vanishing viscosity method serves as a heuristic for understanding the behavior of solutions in the absence of regularization, ensuring the resulting viscosity solution captures the physically relevant dynamics. A key advantage of viscosity solutions is their ability to handle both first- and second-order nonlinear PDEs without imposing smoothness assumptions on the solution, while still preserving essential properties like the maximum principle. This framework provides stability under perturbations and applies broadly to problems where classical or weak approaches falter, such as in the eikonal equation modeling geometric optics or wavefront propagation.

Eikonal Equation Example

A prominent example illustrating the need for viscosity solutions arises in the one-dimensional eikonal equation, given by |u'(x)| = 1 for x \in (-1,1), subject to the boundary conditions u(-1) = u(1) = 0. The function u(x) = 1 - |x| serves as the viscosity solution to this problem. This solution is continuous on [-1,1] but not differentiable at x = 0, where it exhibits a kink. No classical C^1 solution exists for this equation under the given boundary conditions, as any differentiable function satisfying |u'(x)| = 1 almost everywhere would imply a linear growth incompatible with the zero boundary values at both endpoints, except for the trivial zero function which fails the equation. To verify that u(x) = 1 - |x| is a viscosity solution, consider test functions \phi \in C^\infty(\mathbb{R}). The function u is a viscosity subsolution if, whenever u - \phi attains a local maximum at x_0 \in (-1,1), then |\phi'(x_0)| \leq 1; it is a viscosity supersolution if, whenever u - \phi attains a local minimum at x_0, then |\phi'(x_0)| \geq 1. At points x_0 \neq 0, where u is smooth and |u'(x_0)| = 1, the classical solution condition directly implies the viscosity inequalities hold for any test function touching at x_0, as the derivative matches the equation. At the kink x_0 = 0, suppose u - \phi has a local maximum at 0. Then \phi must lie below u nearby, and the slope of \phi at 0 cannot exceed 1 in absolute value, satisfying |\phi'(0)| \leq 1; similarly for the minimum case, ensuring |\phi'(0)| \geq 1. For instance, a quadratic test function \phi(x) = 1 - \frac{1}{2} x^2 touches u at 0 from below, with |\phi'(0)| = 0 \leq 1, confirming the subsolution property, while adjusted linear tests verify the supersolution.

Core Properties

Comparison Principle

The comparison principle is a cornerstone property of viscosity solutions, asserting that if u is a viscosity subsolution and v is a viscosity supersolution to the equation F(x, u, Du, D^2 u) = 0 in a bounded open domain \Omega \subset \mathbb{R}^n, with u \leq v on \partial \Omega, then u \leq v throughout \Omega, provided u and v satisfy appropriate continuity and growth conditions. This principle builds on the definitions of sub- and supersolutions, where subsolutions satisfy F \leq 0 at points of local maximum for test functions touching from above, and supersolutions satisfy F \geq 0 at points of local minimum for test functions touching from below. Key conditions for the principle include the continuity of F, its degenerate ellipticity—meaning F(x, r, p, X) \leq F(x, r, p, Y) whenever Y \geq X in the sense of symmetric matrices—and monotonicity in the solution value, such as \gamma(r - s) \leq F(x, r, p, X) - F(x, s, p, X) for r \geq s and some \gamma > 0. Additionally, uniform continuity of F in its spatial arguments is often required, along with structural assumptions like uniform ellipticity for the second-order terms, expressed as \lambda |\xi|^2 \leq \trace(A(x) \xi \xi^T) \leq \Lambda |\xi|^2 for \xi \in \mathbb{R}^n, where A(x) is the diffusion matrix and \lambda, \Lambda > 0. These ensure the principle holds for a broad class of fully nonlinear second-order partial differential equations. A proof sketch relies on the doubling variables method: assume for contradiction that \sup_\Omega (u - v) > 0, and consider the auxiliary function \Phi(x, y) = u(x) - v(y) - \frac{\alpha}{2} |x - y|^2 for small \alpha > 0. Let (x_0, y_0) achieve the maximum of \Phi in \overline{\Omega} \times \overline{\Omega}; if x_0 = y_0 interior, then x_0 is a contact point where test functions apply, yielding F(x_0, u(x_0), Du(x_0), D^2 u(x_0)) \leq 0 and F(x_0, v(x_0), Dv(x_0), D^2 v(x_0)) \geq 0. Ellipticity and monotonicity then imply a contradiction via maximum principle arguments, as the difference u - v cannot exceed its boundary values without violating the equation's structure; regularization via sup-convolutions may be used to handle semiconcavity. The comparison principle implies uniqueness of viscosity solutions, as any solution w serves as both a subsolution and supersolution, ensuring it is the unique function satisfying u \leq w \leq v for any sub- and supersolution pair bounding it on the boundary. For instance, in the eikonal equation |Du| = 1, the principle guarantees that the viscosity solution coincides with the distance function, uniquely determined between sub- and supersolutions.

Existence and Stability

One primary method for establishing the existence of viscosity solutions to Hamilton-Jacobi equations of the form F(x, u, Du) = 0 involves the vanishing viscosity approach. In this technique, one considers the regularized equation F(x, u^\epsilon, Du^\epsilon) + \epsilon \Delta u^\epsilon = 0 for \epsilon > 0, where smooth solutions u^\epsilon exist under suitable conditions on F, such as continuity and growth bounds. As \epsilon \to 0, the solutions u^\epsilon converge uniformly on compact sets to a limit u that satisfies the original equation in the viscosity sense, provided the family \{u^\epsilon\} is locally uniformly bounded and the nonlinearity F converges appropriately. This method leverages the smoothing effect of the Laplacian term to approximate discontinuous or nonsmooth solutions. Another approach to existence, particularly for elliptic and parabolic equations, employs Perron's method, which constructs a solution as the supremum of all subsolutions or the infimum of all supersolutions bounded between suitable barriers. Under the comparison principle, this envelope function u = \sup \{ v \mid v \text{ is a subsolution}, \, \underline{u} \leq v \leq \overline{u} \} turns out to be the unique viscosity solution, assuming the sub- and supersolutions \underline{u}, \overline{u} satisfy the equation at test points and the domain is bounded or coercive conditions hold. This method avoids explicit regularization and relies on the monotonicity and stability properties inherent to the viscosity framework. Viscosity solutions exhibit strong stability properties with respect to uniform convergence. Specifically, if a sequence of viscosity subsolutions (respectively, supersolutions) converges uniformly on compact sets to a limit function u, then u is itself a viscosity subsolution (respectively, supersolution) to the limiting equation. Consequently, uniform limits of viscosity solutions are viscosity solutions, ensuring consistency under approximations like vanishing viscosity or numerical schemes. This stability holds in the sup norm on bounded domains and extends to unbounded domains under additional control on the growth of solutions. These existence and stability results typically require structural conditions on the nonlinearity F, such as coercivity—meaning F(x, r, p) \to +\infty as |p| \to \infty uniformly in bounded sets for x and r—to guarantee boundedness and compactness of approximating sequences. Linear growth bounds, like |F(x, r, p)| \leq C(1 + |x| + |r| + |p|), further ensure the applicability on whole spaces by preventing explosive behavior.

Boundary Value Problems

Handling Boundaries

In the context of viscosity solutions for partial differential equations (PDEs) posed on bounded domains, boundaries are handled through a generalized notion that extends the interior definition without requiring classical traces or continuity up to the boundary in the strong sense. This approach accommodates solutions that may not attain prescribed boundary values continuously but still satisfy them in a weak, viscosity-compatible manner. Consider a bounded open set \Omega \subset \mathbb{R}^n with a smooth boundary \partial \Omega. The PDE is of the form F(x, u, Du, D^2 u) = 0 in \Omega, where F is continuous, and Dirichlet data u = g is prescribed on \partial \Omega with g continuous. Viscosity solutions in this setting are defined using upper semicontinuous (USC) subsolutions and lower semicontinuous (LSC) supersolutions that incorporate boundary conditions via test functions approaching from within \Omega. A function u \in USC(\overline{\Omega}) is a boundary subsolution if it satisfies the interior viscosity subsolution condition—that is, for every test function \phi \in C^2(\Omega) such that u - \phi attains a local maximum at an interior point x \in \Omega, F(x, u(x), D\phi(x), D^2 \phi(x)) \leq 0—and additionally, u \leq g on \partial \Omega. At boundary points x \in \partial \Omega, the condition strengthens to: if \phi \in C^2(\Omega) touches u from inside \Omega (i.e., u - \phi attains a local maximum at x relative to \Omega), then F(x, u(x), D\phi(x), D^2 \phi(x)) \leq 0. This ensures the PDE holds in the viscosity sense at the boundary while enforcing the data inequality. Conversely, a function v \in LSC(\overline{\Omega}) is a boundary supersolution if it satisfies the interior viscosity supersolution condition—for every \phi \in C^2(\Omega) such that v - \phi attains a local minimum at x \in \Omega, F(x, v(x), D\phi(x), D^2 \phi(x)) \geq 0—and v \geq g on \partial \Omega. At x \in \partial \Omega, if \phi \in C^2(\Omega) touches v from inside (i.e., v - \phi attains a local minimum relative to \Omega), then F(x, v(x), D\phi(x), D^2 \phi(x)) \geq 0. A viscosity solution is then a continuous function that is both a subsolution and supersolution. To formalize these at the boundary, one employs generalized jet spaces: the second-order superjet J^{2,+} u(x) and subjet J^{2,-} v(x) for x \in \partial \Omega, defined using test functions \phi that touch from inside \Omega and account for the domain's geometry (e.g., via inward normals). Specifically, (p, X) \in J^{2,+} u(x) if there exists \phi \in C^2(\Omega) such that u(y) \leq u(x) + \phi(y) - \phi(x) + o(|y - x|^2) for y \in \Omega near x, with p = D\phi(x) and X = D^2 \phi(x); the boundary condition requires F(x, u(x), p, X) \leq 0 for all such jets. The interior jets coincide with the standard definition when restricted away from \partial \Omega. This framework ensures stability and comparison principles hold for bounded domains under suitable structural conditions on F.

Dirichlet Problems

The Dirichlet problem for a fully nonlinear second-order partial differential equation (PDE) seeks a viscosity solution u satisfying F(x, u, Du, D^2 u) = 0 \quad \text{in } \Omega, \quad u = g \quad \text{on } \partial \Omega, where \Omega \subset \mathbb{R}^n is a bounded open domain, F: \mathbb{R}^n \times \mathbb{R} \times \mathbb{R}^n \times \mathbb{S}^n \to \mathbb{R} (with \mathbb{S}^n denoting the space of n \times n symmetric matrices), and g: \partial \Omega \to \mathbb{R} is continuous. This formulation extends classical Dirichlet problems to nonlinear and possibly degenerate elliptic equations, where traditional smooth solutions may fail to exist. Uniqueness of viscosity solutions follows from a comparison principle that incorporates boundary data, building on the interior comparison by ensuring that any subsolution \underline{u} and supersolution \overline{u} satisfy \underline{u} \leq \overline{u} in \overline{\Omega}, provided F is uniformly elliptic (i.e., there exist \lambda, \Lambda > 0 such that \lambda I \leq X - Y \leq \Lambda I implies \lambda \leq F(x,r,p,X) - F(x,r,p,Y) \leq \Lambda for all relevant arguments) and proper (Lipschitz continuous in u, uniformly with respect to other variables). The boundary extension relies on the continuity of g and domain regularity (e.g., uniform exterior cone condition), preventing subsolutions from exceeding supersolutions at \partial \Omega. For Hamilton-Jacobi equations (first-order case, F(x,u,Du) = 0), a similar boundary-inclusive comparison holds under uniform continuity of F in its arguments and coercivity in Du. Existence is established via the Perron method adapted to boundaries, constructing the solution as the supremum of subsolutions bounded above by a boundary envelope function that matches g on \partial \Omega, or through barrier constructions that approximate the problem with solvable equations. These approaches require uniform ellipticity of F, Lipschitz continuity in u, and continuity of coefficients and g, yielding a unique continuous viscosity solution in C(\overline{\Omega}) for domains satisfying the uniform interior and exterior cone conditions. In the first-order Hamilton-Jacobi setting, existence can also be obtained via the vanishing viscosity method, perturbing the equation with a small Laplacian term and passing to the limit. Without uniform ellipticity, uniqueness fails, as demonstrated by constructions of multiple viscosity solutions to the Dirichlet problem for linear equations with measurable coefficients in dimensions n \geq 3. For instance, Nadirashvili exhibited distinct positive and negative solutions to \Delta u = f(x) u in the unit ball with zero boundary data, where f is chosen to violate ellipticity bounds. Safonov provided related counterexamples highlighting the necessity of structural assumptions for well-posedness.

Historical Development

Origins

The concept of viscosity solutions emerged from earlier efforts to address the limitations of classical solutions for nonlinear partial differential equations (PDEs), particularly in the context of hyperbolic conservation laws. In the 1950s and 1960s, researchers developed the vanishing viscosity method to construct generalized solutions for these equations, where small viscous terms are added and then taken to zero in the limit. A foundational contribution came from Olga Oleinik, whose 1957 work on discontinuous solutions of nonlinear differential equations introduced entropy conditions that ensure uniqueness for scalar conservation laws by characterizing admissible shocks. This approach paralleled physical interpretations, as the added viscosity term mimics dissipative effects in fluid dynamics, allowing the limit to select physically relevant solutions amid discontinuities. The modern notion of viscosity solutions was formally introduced in 1983 by Michael G. Crandall and Pierre-Louis Lions in their paper "Viscosity solutions of Hamilton-Jacobi equations," published in the Transactions of the American Mathematical Society. This work targeted first-order Hamilton-Jacobi PDEs, which frequently lack classical smooth solutions due to the formation of shocks or corners. The authors defined viscosity solutions using sub- and super-solution inequalities tested against smooth functions, providing a weak formulation that captures the essential behavior without requiring differentiability everywhere. The primary motivation stemmed from optimal control theory, where Hamilton-Jacobi equations describe value functions representing minimal costs or maximal rewards, often exhibiting non-differentiable profiles that render classical methods inadequate. Viscosity solutions addressed this by ensuring uniqueness through a comparison principle, even for discontinuous data, thus enabling rigorous analysis of control problems. The eikonal equation, |∇u| = 1, served as a key early test case, illustrating how the framework resolves singularities in geometric optics and front propagation. In the same year, Pierre-Louis Lions extended the theory to stochastic settings in his paper "Optimal control of diffusion processes and Hamilton-Jacobi-Bellman equations. Part 2: Viscosity solutions and uniqueness," applying it to fully nonlinear second-order Bellman equations arising in stochastic control. This adaptation proved crucial for handling value functions in diffusion processes with random perturbations, maintaining uniqueness despite the added complexity of second-order terms.

Key Contributions and Extensions

One of the pivotal extensions of the viscosity solution framework involved adapting it to second-order fully nonlinear elliptic partial differential equations (PDEs). In 1988, Robert Jensen established a maximum principle for viscosity solutions of such equations of the form F(D^2 u, Du, u, x) = 0, proving uniqueness under degenerate ellipticity and Lipschitz continuity assumptions on F, which extended the theory beyond first-order Hamilton-Jacobi equations. Complementing this, the theory was extended to parabolic cases in the late 1980s by Crandall, Lions, and collaborators, introducing viscosity solutions for equations like u_t + F(D^2 u, Du, u, x, t) = 0, with comparison principles that ensured well-posedness for time-dependent problems. Advancements in handling boundaries and initial-boundary value problems (IBVPs) further solidified the framework during the 1980s. Hiroshi Ishii contributed key results on comparison principles incorporating boundary conditions, demonstrating that viscosity subsolutions and supersolutions satisfy strict inequalities at the boundary for fully nonlinear elliptic PDEs, thus enabling uniqueness for Dirichlet and Neumann problems in bounded domains. Recent developments from 2020 to 2025 have broadened the applicability of viscosity solutions to more abstract settings. For path-dependent PDEs, Ekren and collaborators in 2020 refined the notion of viscosity solutions to accommodate randomized time horizons in path-dependent Hamilton-Jacobi-Bellman equations, establishing comparison and stability via probabilistic representations. In 2025, extensions to L^p-viscosity solutions emerged, where solutions are tested against L^p test functions rather than smooth ones, providing existence via Perron's method for fully nonlinear elliptic equations with p \geq n in \mathbb{R}^n. Comparison principles in the Wasserstein space were also advanced in 2025, yielding uniqueness for second-order parabolic PDEs on probability measures equipped with the W_2 metric, crucial for mean-field control problems. Additionally, relations to distribution solutions were clarified in 2025 through probabilistic methods, showing equivalence for semi-linear Neumann-type PDEs where viscosity solutions coincide with those in the sense of distributions under mild regularity conditions. Extensions to infinite-dimensional spaces and integro-PDEs have addressed stochastic processes and nonlocal operators. In infinite dimensions, viscosity solutions apply to Hamilton-Jacobi-Bellman equations driven by controlled stochastic differential equations in Hilbert spaces, with existence and uniqueness via dynamic programming principles for optimal control of diffusion processes. For integro-PDEs involving nonlocal operators, such as Lévy processes, the theory incorporates integro-differential terms like \int (u(x+y) - u(x) - Du(x) \cdot y) \nu(dy), with viscosity solutions defined via sub- and super-solution tests that yield comparison for second-order elliptic equations with jumps.

Applications

Hamilton-Jacobi-Bellman Equations

Viscosity solutions play a crucial role in solving Hamilton–Jacobi–Bellman (HJB) equations, which characterize the value function in optimal control problems. These equations typically lack classical solutions due to the nonlinearity introduced by the optimization over controls and potential discontinuities in the data. The viscosity solution framework, developed by Crandall and Lions, addresses this by providing a weak notion of solution that ensures existence, uniqueness, and stability for both deterministic and stochastic cases. The standard HJB equation for a stochastic optimal control problem is given by \sup_{a \in A} \left[ -\partial_t u(t,x) - H(x, Du(t,x)) + \frac{1}{2} \trace\left( \sigma(x,a) \sigma(x,a)^T D^2 u(t,x) \right) \right] = 0, for (t,x) \in [0,T) \times \mathbb{R}^n, with terminal condition u(T,x) = g(x). Here, H(x,p) is the Hamiltonian incorporating the drift and running cost, \sigma(x,a) is the diffusion coefficient depending on the control a, and A is the control set. In the deterministic case, \sigma \equiv 0, reducing to a first-order Hamilton–Jacobi equation. Viscosity solutions are defined using test functions, where subsolutions satisfy the inequality in the viscosity sense at maxima and supersolutions at minima, allowing handling of nonconvex Hamiltonians arising from the supremum over controls. The value function u(t,x) represents the infimum of the expected cost functional over admissible controls, starting from state x at time t: u(t,x) = \inf_{\alpha} \mathbb{E} \left[ \int_t^T l(s,X_s^\alpha, \alpha_s) ds + g(X_T^\alpha) \right], where X^\alpha is the controlled state process. Dynamic programming principles show that u satisfies the HJB equation in the viscosity sense, and the comparison principle for viscosity solutions guarantees its uniqueness among bounded continuous functions. This connection bridges optimal control theory with PDE analysis, ensuring the value function is the unique viscosity solution without requiring differentiability. A classic example is the eikonal equation |\nabla u(x)| = 1/c(x) in \mathbb{R}^n, a deterministic HJB form for the shortest path problem in a medium with speed $c(x)$; its viscosity solution is the minimal travel time from a source, coinciding with the value function of the corresponding control problem. Another application arises in inventory control, where the HJB equation models optimal ordering policies under stochastic demand, with the viscosity solution providing the minimal expected inventory cost as a function of current stock level.

Path-Dependent and Infinite-Dimensional PDEs

Viscosity solutions have been extended to path-dependent partial differential equations (PDEs), which arise in non-Markovian settings such as path-dependent option pricing or stochastic control with memory effects. These equations are formulated on the Skorokhod space of càdlàg paths, where the solution u(t, \omega) depends on the entire path \omega up to time t. A prototypical form is F(t, \omega, u, \delta u, \delta^2 u) = 0, with \delta denoting functional derivatives: the vertical derivative \partial_\omega u captures path variations at the current time, while the horizontal derivative D u accounts for time shifts along the path. This framework builds on the functional Itô calculus introduced by Dupire in 2009, which provides the necessary tools for differentiation in path space. The notion of viscosity solutions for these path-dependent PDEs (PPDEs) adapts the classical definition by employing horizontal and vertical test functions to define upper and lower semicontinuous envelopes. A function u is a viscosity subsolution if, at points of local maximum, the equation holds with appropriate inequalities involving these test functions; supersolutions are defined dually at local minima. This approach ensures comparison principles and stability, as established in seminal work where probabilistic representations via backward stochastic differential equations (BSDEs) confirm uniqueness under suitable growth conditions. Extensions from 2020 to 2025 have refined this theory, incorporating randomized stopping times to handle irregular paths and proving convergence of approximations for Hamilton-Jacobi-Bellman-type PPDEs. Approximate viscosity solutions have also been introduced to address vertical differentiability issues, enabling analysis of nonlinear PPDEs with non-smooth drivers. In infinite-dimensional settings, such as stochastic PDEs (SPDEs) or optimal control in Hilbert spaces, viscosity solutions require further adaptations to handle the lack of finite-dimensional structure. Solutions are often sought in spaces of continuous functions on Hilbert-valued paths, with well-posedness established via path-dependent equations in infinite dimensions, including comparison principles under B-continuity assumptions—a notion of regularity weaker than classical continuity but sufficient for test function approximations. For semilinear PDEs, recent developments provide Feynman-Kac-type representations linking viscosity solutions to expectations of controlled SPDEs in Hilbert spaces, ensuring existence and uniqueness for equations like \partial_t u + \mathcal{A} u + f(t, x, u) = 0, where \mathcal{A} is an unbounded generator. These results, from 2024, extend to mild solutions and highlight B-continuity's role in overcoming regularity barriers in infinite dimensions. Recent advances as of 2025 include comparison principles for viscosity solutions of path-dependent semilinear PDEs, proven via minimax formulations that equate viscosity notions to game-theoretic values, ensuring uniqueness without strict monotonicity assumptions on the nonlinearity. Key challenges in these extensions stem from non-local dependencies in path space or infinite dimensions, necessitating adapted notions of jets: horizontal/vertical derivatives replace standard partials, and B-continuity ensures test functions touch solutions without excessive oscillation. These adaptations preserve the robustness of viscosity theory while accommodating the functional or Hilbert-valued nature of the problems.

Other Applications

Viscosity solutions also find applications in front propagation models, such as those using level-set methods to track evolving interfaces, where the solution represents the arrival time of a front in heterogeneous media. In mean curvature flow, viscosity solutions provide a framework for handling singularities and proving convergence of approximations to the flow of convex hypersurfaces.

References

  1. [1]
  2. [2]
    [PDF] USER'S GUIDE TO VISCOSITY SOLUTIONS OF SECOND ORDER ...
    The notion of viscosity solutions of scalar fully nonlinear partial differential equations of second order provides a framework in which startling comparison ...
  3. [3]
    [PDF] Lecture notes on viscosity solutions
    These notes are concerned with viscosity solutions for fully nonlinear equa- tions. A majority of the notes are concerned with Hamilton-Jacobi equations.
  4. [4]
    [PDF] an introduction to viscosity solutions for fully nonlinear pde ... - arXiv
    Nov 11, 2014 · Today, though, it comprises an independent theory of “weak” solutions which applies to fully nonlinear elliptic and parabolic PDE and in most ...
  5. [5]
    [PDF] Introduction to Viscosity Solutions for Nonlinear PDEs
    1.2 Classical and weak solutions. . . . . . . . . . . . . . . . . . . . 5. 2 ... Hence the necessity of weaker notions of solution. Thinking of the ...
  6. [6]
    [PDF] viscosity solutions of the eikonal equations - UChicago Math
    Viscosity solutions are a theory for 'weak' solutions of nonlinear PDEs. For eikonal equations, a function is a viscosity solution if it is both a sub- and ...
  7. [7]
    [PDF] notes for math 6140 viscosity solutions
    Mar 18, 2015 · A suitable definition of viscosity solution can be obtained by reversing the inequal- ities above, and in the process defining the lower ...
  8. [8]
  9. [9]
    [PDF] Viscosity Solutions of Hamilton-Jacobi Equations at the Boundary.
    Jul 9, 1984 · Crandall, rvane and Lions [3] provides a simpler introduction to the subject while the book Lions 181 and the review paper Crandall and ...
  10. [10]
    Viscosity solutions of fully nonlinear second-order elliptic partial ...
    We investigate comparison and existence results for viscosity solutions of fully nonlinear, second-order, elliptic, possibly degenerate equations.
  11. [11]
    [PDF] existence results for boundary problems for uniformly elliptic and ...
    Jul 1, 1999 · [8] M.G. Crandall, H. Ishii and P.L. Lions, User's guide to viscosity solutions of second order partial differential equations, Bull. Amer ...
  12. [12]
    Viscosity Solutions of Hamilton-Jacobi Equations - jstor
    Volume 277, Number 1, May 1983. VISCOSITY SOLUTIONS OF HAMILTON-JACOBI EQUATIONS'. BY. MICHAEL G. CRANDALL AND PIERRE-LOUIS LIONS. ABSTRACT. Problems involving ...Missing: original | Show results with:original
  13. [13]
    Nonuniqueness for Second-Order Elliptic Equations with ... - SIAM.org
    A recent sensational result of Nikolai Nadirashvili states that there is no uniqueness of "weak" solutions to this problem if d ≥ 3 . He constructed two ...
  14. [14]
    Nonuniqueness for Second-Order Elliptic Equations with ...
    A recent sensational result of Nikolai Nadirashvili states that there is no uniqueness of "weak" solutions to this problem if d ≥ 3 d\ge 3 . ... Nadirashvili, ...
  15. [15]
    The maximum principle for viscosity solutions of fully nonlinear ...
    The maximum principle for viscosity solutions of fully nonlinear second order partial differential equations. Published: March 1988. Volume 101, pages 1–27, ( ...
  16. [16]
    Viscosity Solutions of Path-Dependent PDEs with Randomized Time
    We introduce a new definition of viscosity solution to path-dependent partial differential equations, which is a slight modification of the definition ...
  17. [17]
    Comparison of viscosity solutions for a class of second-order PDEs ...
    We prove a comparison result for viscosity solutions of second-order parabolic partial differential equations in the Wasserstein space.
  18. [18]
    Second-order elliptic integro-differential equations: viscosity ...
    The aim of this work is to revisit viscosity solutions' theory for second-order elliptic integro-differential equations and to provide a general framework.
  19. [19]
    [1109.5971] On viscosity solutions of path dependent PDEs - arXiv
    Sep 27, 2011 · The key ingredient of our approach is a functional Itô calculus recently introduced by Dupire [Functional Itô calculus (2009) Preprint].
  20. [20]
    On viscosity solutions of path dependent PDEs - Project Euclid
    Jan 9, 2014 · The key ingredient of our approach is a functional Itô calculus recently introduced by Dupire [Functional Itô calculus (2009) Preprint].
  21. [21]
    Approximate viscosity solutions of path-dependent PDEs and ...
    Dec 13, 2023 · We introduce a notion of approximate viscosity solutions for a class of nonlinear path-dependent PDEs (PPDEs), including the Hamilton–Jacobi–Bellman-type ...Missing: distribution | Show results with:distribution
  22. [22]
    Path-dependent equations and viscosity solutions in infinite dimension
    Feb 19, 2015 · In this paper, motivated by the study of models driven by path-dependent stochastic PDEs, we give a first well-posedness result for viscosity solutions of ...Missing: SPDEs | Show results with:SPDEs
  23. [23]
    [2303.10038] Semilinear Feynman-Kac Formulae for $B - arXiv
    Mar 17, 2023 · We prove the existence of a B-continuous viscosity solution for a class of infinite dimensional semilinear partial differential equations (PDEs) ...
  24. [24]
    Semilinear Feynman–Kac formulae for B-continuous viscosity ...
    Abstract. We prove the existence of a B-continuous viscosity solution for a class of infinite dimensional semilinear partial differential equations (PDEs) ...
  25. [25]
    Viscosity and minimax solutions for path-dependent Hamilton ... - arXiv
    Sep 19, 2025 · We introduce a new notion of a viscosity solution for this problem and show the equivalence of this notion to the notion of a minimax solution, ...<|separator|>
  26. [26]
    On the Relationship Between Viscosity and Distribution Solutions for ...
    Jan 18, 2025 · Based on probabilistic methods, we discuss the relationship between viscosity and distribution solutions for semi-linear partial ...
  27. [27]
    Path-dependent equations and viscosity solutions in infinite dimension
    In this paper, motivated by the study of models driven by path-dependent stochastic PDEs, we give a first well-posedness result for viscosity solutions of PPDEs ...Missing: SPDEs | Show results with:SPDEs