Fact-checked by Grok 2 weeks ago

Variational inequality

A variational inequality is a mathematical formulation that seeks to find an element u^* in a closed subset K of a such that \langle Au^*, v - u^* \rangle \geq 0 for all v \in K, where A is a (often nonlinear) from the space to its , and \langle \cdot, \cdot \rangle denotes the duality pairing. This generalizes classical variational principles and conditions by incorporating constraints directly into the formulation, allowing solutions to satisfy directional inequalities rather than equalities. The concept was introduced by Philip Hartman and Guido Stampacchia in their seminal 1966 paper, where it was developed as a tool to address nonlinear elliptic differential-functional equations arising in boundary value problems with obstacles or unilateral constraints. Building on earlier work in and functional equations, variational inequalities provide a unified framework for studying a broad class of problems, including those where traditional equations fail due to discontinuities or non-smoothness. Key existence results, such as the Hartman-Stampacchia theorem, guarantee solutions under conditions like monotonicity (where \langle A(u) - A(v), u - v \rangle \geq 0), , and of A, ensuring the problem is well-posed even in infinite-dimensional Hilbert or Banach spaces. In finite dimensions, the variational inequality problem VI(F, K)—finding x^* \in K \subseteq \mathbb{R}^n such that F(x^*)^T (x - x^*) \geq 0 for all x \in K, with F continuous and K closed convex—extends to applications in optimization, where it characterizes the optimality conditions for convex programs, and in , modeling Nash equilibria in noncooperative games. Uniqueness often follows from strict monotonicity of F. Beyond , variational inequalities have profound impacts in applied fields: in , they model contact problems like Signorini's problem for bodies; in , they describe spatial price equilibria and oligopolistic markets; and in engineering, they arise in traffic networks (via Wardrop equilibria), porous media flow, and of option pricing under constraints. Numerical methods, including projection algorithms and fixed-point iterations, are well-developed for solving these problems efficiently, with extensions to and multivalued variants addressing and set-valued mappings.

History

Origins and Early Concepts

The origins of variational inequalities can be traced to the classical , where the Euler-Lagrange equations served as key precursors by providing necessary conditions for extremal functions in optimization problems. Developed by Leonhard Euler in the 1740s through works such as his 1744 publication Methodus inveniendi lineas curvas maximi minimive proprietate gaudentes sive solutio problematis isoperimetrici latissimo sensu accepti, these equations addressed the minimization of functionals subject to constraints, revealing the need for modified conditions when inequalities restricted the admissible functions—such as in scenarios involving barriers or unilateral restrictions that prevent certain variations from being negative. In particular, problems like finding minimal surfaces above an obstacle highlighted how equality-based constraints in the Euler-Lagrange framework naturally extended to inequality forms to ensure the functional's variation remained non-negative on the feasible set. Building on Euler's foundations, advanced the theory in 1879 by refining necessary conditions for extrema in variational problems under constraints, emphasizing sufficient conditions and the role of second variations to distinguish maxima, minima, and saddle points. His improvements addressed shortcomings in earlier approaches, particularly for constrained extrema where inequality barriers required careful of the Legendre condition and corner , laying groundwork for handling unilateral constraints without assuming smooth equality restrictions. David Hilbert's 1904 work on the Dirichlet principle further extended these ideas by rigorously justifying the existence of minimizers for variational integrals associated with elliptic partial differential equations. In his "Über das Dirichletsche Prinzip," Hilbert used arguments in suitable spaces to ensure the integral attains its minimum. Early 20th-century contributions, notably Richard Courant's analysis of elliptic partial differential equations with constraints, applied variational methods in the context of physical problems like membrane vibrations and equilibrium states subject to rigid supports. Courant's , derived for solutions bounded by obstacle-like conditions on the , provided bounds on eigenvalues and functionals, illustrating how variational methods could incorporate unilateral constraints to model real-world barriers in elliptic value problems. These developments bridged classical variational calculus with more general formulations, paving the way for mid-20th-century formalizations.

Key Developments in the Mid-20th Century

The mid-20th century marked the formal emergence of variational inequalities as a unified mathematical framework, building on prior work in and partial differential equations. Independently, Gaetano Fichera around 1963–1964 developed variational inequality formulations for unilateral constraints in elasticity, notably addressing the Signorini problem of contact with frictionless rigid obstacles, establishing existence principles for such problems. In 1962, George J. Minty provided a pivotal generalization of Felix Browder's to operators in Hilbert spaces, establishing the Minty-Browder formulation that ensured the surjectivity of the sum of the identity and a maximal operator under appropriate conditions. This result laid essential groundwork for solving nonlinear problems via resolvent operators, influencing subsequent developments in . A landmark contribution came in 1964 from Guido Stampacchia, who introduced the classical variational inequality in the context of elliptic boundary value problems with discontinuous coefficients. In his seminal paper, Stampacchia extended the to coercive bilinear forms on closed subsets of Hilbert spaces, proving existence and uniqueness for solutions to the problem: find u \in K such that a(u, v - u) \geq 0 for all v \in K, where K is a closed subset, a(\cdot, \cdot) is a continuous coercive on the space, and the associated operator A satisfies a(u, v) = \langle Au, v \rangle. This formulation, often denoted as \langle Au, v - u \rangle \geq 0 for all v \in K, provided a powerful tool for handling inequalities arising in and . In 1966, Philip Hartman and Guido Stampacchia published a seminal paper on nonlinear elliptic differential-functional equations, further developing variational inequalities as a tool for boundary value problems with obstacles or unilateral constraints. By 1969, Jacques-Louis Lions significantly advanced the theory by extending variational inequalities to time-dependent evolution equations and nonlinear partial differential equations. In his , Lions developed abstract frameworks for parabolic and hyperbolic problems, incorporating variational inequalities to address noncoercive cases and obstacle-type constraints in dynamic settings. These extensions enabled the treatment of complex phenomena such as flow through porous media and , solidifying variational inequalities as a versatile method for nonlinear analysis.

Formal Definition

Finite-Dimensional Formulation

In finite-dimensional \mathbb{R}^n, the standard inner product is defined as \langle \mathbf{x}, \mathbf{y} \rangle = \mathbf{x}^T \mathbf{y} for \mathbf{x}, \mathbf{y} \in \mathbb{R}^n, inducing the norm \|\mathbf{x}\| = \sqrt{\langle \mathbf{x}, \mathbf{x} \rangle}. A nonempty K \subseteq \mathbb{R}^n is if, for all \mathbf{x}, \mathbf{y} \in K and \lambda \in [0, 1], the \lambda \mathbf{x} + (1 - \lambda) \mathbf{y} \in K. The finite-dimensional variational inequality problem, denoted \mathrm{VI}(F, K), is to find a vector \mathbf{x}^* \in K such that \langle F(\mathbf{x}^*), \mathbf{y} - \mathbf{x}^* \rangle \geq 0 \quad \forall \, \mathbf{y} \in K, where K \subseteq \mathbb{R}^n is a nonempty and F: \mathbb{R}^n \to \mathbb{R}^n is a single-valued . This formulation, often assuming F is continuous, provides an accessible framework for problems in optimization and analysis. For set-valued mappings F: \mathbb{R}^n \rightrightarrows \mathbb{R}^n, the problem generalizes to finding \mathbf{x}^* \in K and \mathbf{w} \in F(\mathbf{x}^*) such that \langle \mathbf{w}, \mathbf{y} - \mathbf{x}^* \rangle \geq 0 \quad \forall \, \mathbf{y} \in K, with of the selection or of F typically imposed for theoretical analysis. A solution \mathbf{x}^* to \mathrm{VI}(F, K) equivalently satisfies the fixed-point equation \mathbf{x}^* = \proj_K \bigl( \mathbf{x}^* - F(\mathbf{x}^*) \bigr), where \proj_K: \mathbb{R}^n \to K is the orthogonal onto the closed K, defined as \proj_K(\mathbf{z}) = \arg\min_{\mathbf{u} \in K} \|\mathbf{z} - \mathbf{u}\|^2.

General Formulation in Banach Spaces

The general formulation of a variational inequality in a Banach space setting extends the finite-dimensional case to infinite-dimensional spaces, leveraging tools from functional analysis such as duality pairings and weak topologies. Let V be a reflexive Banach space with dual space V^*, and let K \subseteq V be a nonempty, convex, and closed subset. Given a nonlinear operator A: V \to V^*, the variational inequality problem seeks an element u \in K satisfying \langle A(u), v - u \rangle \geq 0 \quad \forall v \in K, where \langle \cdot, \cdot \rangle denotes the duality pairing between V and V^*. This formulation captures equilibrium conditions in abstract spaces, where the operator A represents forces or gradients, and the constraint K models admissible configurations. The dual space V^* plays a crucial role, as it allows the inequality to be expressed in terms of linear functionals, facilitating analysis in spaces without inner products, unlike the Hilbert space case. Key assumptions underpin this setup to ensure well-posedness in the . The reflexivity of V guarantees the weak compactness of closed bounded sets via the Banach-Alaoglu theorem, which is essential for handling infinite-dimensional phenomena like . The set K must be convex and closed to preserve the variational structure under weak limits. The operator A is typically assumed to be continuous, or more generally hemicontinuous—meaning that the map t \mapsto \langle A((1-t)u + tv), w \rangle is continuous on [0,1] for all u, v, w \in V—to align with weak convergence properties. These conditions enable the extension of finite-dimensional intuition to broader function spaces without relying on coordinate representations. A related, relaxed variant is the Minty variational inequality, which seeks u \in K such that \langle A(v), v - u \rangle \geq 0 \quad \forall v \in K. Originally introduced for operators in Hilbert spaces, this form extends naturally to reflexive Banach spaces and serves as an auxiliary problem. Under the assumptions of —i.e., \langle A(u) - A(v), u - v \rangle \geq 0 for all u, v \in V—and of A, solutions to the Minty inequality coincide with those of the standard (Stampacchia) formulation. This equivalence simplifies proofs involving surjectivity or maximality of operators. In applications to nonlinear problems, the operator A is often the Gâteaux derivative of a functional J: V \to \mathbb{R}, defined as A(u) = J'(u) where the directional derivative satisfies \lim_{t \to 0^+} \frac{J(u + t h) - J(u)}{t} = \langle J'(u), h \rangle for h \in V. This perspective links variational inequalities to subdifferential inclusions for functionals, with the V^* hosting the subgradients. The Gâteaux differentiability assumption accommodates nonlinearities beyond linear s, such as those arising in elasticity or , while maintaining the duality framework.

Theoretical Properties

Monotonicity Conditions

In the context of variational inequalities, monotonicity conditions on the operator A play a fundamental role in establishing the existence of solutions. A single-valued A: X \to X^*, where X is a reflexive with dual X^*, is said to be if for all u, v \in X, \langle A(u) - A(v), u - v \rangle \geq 0, where \langle \cdot, \cdot \rangle denotes the duality pairing between X and X^*. This condition generalizes the increasing nature of functions to nonlinear operators in infinite-dimensional spaces. If the inequality is strict for u \neq v, then A is strictly monotone, which strengthens the separation properties between points. Furthermore, A is strongly monotone if there exists \mu > 0 such that \langle A(u) - A(v), u - v \rangle \geq \mu \|u - v\|^2 for all u, v \in X, providing a quantitative measure of how much A "expands" differences, often leading to unique solutions in associated problems. Hemicontinuity is another key regularity condition that complements monotonicity. An operator A is hemicontinuous if it is locally bounded and continuous with respect to the weak topology along every line segment in its domain, meaning that for any u, v \in X with v \in D(A), the mapping t \mapsto \langle A((1-t)u + t v), w \rangle is continuous in t \in [0,1] for every w \in X. This property ensures that monotone operators behave sufficiently regularly to apply topological arguments, such as those in fixed-point theorems, without requiring full continuity. Maximal monotonicity extends the notion of monotonicity to set-valued operators A: X \rightrightarrows X^*, where the graph of A, defined as \{(u, u^*) \in X \times X^* \mid u^* \in A(u)\}, cannot be properly contained in the graph of any other monotone operator. Minty's characterization provides an equivalent condition for maximality: a monotone operator A is maximal if and only if, for every \lambda > 0, the range of I + \lambda A covers the entire X^*, where I is the operator. This surjectivity criterion is particularly useful in Hilbert spaces, where maximal monotone operators are densely defined and single-valued on the interior of their . These properties underpin results for variational inequalities by ensuring that the operator cannot be extended further while preserving monotonicity, thus capturing the "full" behavior of the mapping.

Existence and Uniqueness Results

The existence of solutions to variational inequalities in reflexive s is guaranteed under conditions of monotonicity, , and for the involved. Specifically, the Browder-Minty theorem establishes that if A: V \to V^* is a and hemicontinuous on a reflexive V, and coercive (i.e., \lim_{\|u\| \to \infty, u \in K} \frac{\langle A(u), u \rangle}{\|u\|} = +\infty), with K \subset V a nonempty, closed, and , then the variational inequality \langle A(u), v - u \rangle \geq 0 for all v \in K admits at least one solution u \in K. This result extends Minty's earlier work in Hilbert spaces to the more general Banach setting, relying on the surjectivity of the I + A for maximal monotone extensions. Uniqueness of solutions follows from stricter monotonicity assumptions. If A is strictly monotone, meaning \langle A(u) - A(v), u - v \rangle > 0 for all u \neq v in K, then the variational inequality has at most one solution, as any two solutions would contradict the strict inequality. Furthermore, if A is strongly monotone with constant \mu > 0, i.e., \langle A(u) - A(v), u - v \rangle \geq \mu \|u - v\|^2, and Lipschitz continuous with constant L, then under the condition \mu > L/2, the proximal mapping associated with A acts as a contraction, ensuring a unique solution via fixed-point arguments. A key assumption for boundedness of solution sets is of the operator A, defined by \lim_{\|u\| \to \infty, u \in K} \frac{\langle A(u), u \rangle}{\|u\|} = +\infty. This condition ensures that the level sets \{u \in K : \langle A(u), u \rangle \leq C\} are bounded for any C > 0, preventing solutions from escaping to infinity and facilitating arguments in proofs. For pseudomonotone operators, which generalize monotonicity: whenever u_n \to u strongly in V and \liminf_{n \to \infty} \langle A(u_n), u_n - v \rangle \geq 0 for all v \in V, then \langle A(u), u - v \rangle \geq 0 for all v \in V, the Lions-Stampacchia theorem provides existence in reflexive Banach spaces when A is continuous, pseudomonotone, and coercive, with K closed and convex. In the context of partial differential equations, Gårding's inequality ensures coercivity for bilinear forms associated with elliptic operators, stating that for a strongly elliptic operator of even order, there exists c > 0 such that \operatorname{Re} a(u, u) \geq c \|u\|_{H^s}^2 - C \|u\|_{H^{s-1}}^2 for u in a Sobolev space, which underpins existence for variational inequalities modeling obstacle problems or unilateral constraints. Proofs of these existence results often employ the , approximating the variational inequality on finite-dimensional subspaces of V (e.g., spanned by basis functions) to obtain a of solutions that converges weakly to a of the original problem via monotonicity and in reflexive spaces, as detailed in the Lions-Stampacchia . Alternatively, the proximal point method iteratively solves regularized inclusions u^{k+1} \in (I + \lambda_k A)^{-1}(u^k) for \lambda_k > 0, leveraging the maximal monotonicity of A to ensure to a , with weak in Hilbert spaces under the assumption.

Examples

Relation to Optimization Problems

Variational inequalities provide a unifying for classical optimization problems, particularly in the finite-dimensional setting. Specifically, for a continuously differentiable f: \mathbb{R}^n \to \mathbb{R} and a nonempty closed K \subseteq \mathbb{R}^n, a point x^* \in K minimizes f over K it solves the variational inequality \langle \nabla f(x^*), y - x^* \rangle \geq 0 for all y \in K. This equivalence establishes the variational inequality as the first-order necessary and sufficient optimality for minimization subject to convex constraints, extending the unconstrained case where \nabla f(x^*) = 0. A canonical example is the projection problem onto a . Consider minimizing \frac{1}{2} \|x - b\|^2 over K, where b \in \mathbb{R}^n. The unique minimizer x^* = \mathrm{proj}_K(b) satisfies the variational inequality \langle x^* - b, y - x^* \rangle \geq 0 for all y \in K, which corresponds to the operator F(x) = x - b. This formulation is central in proximal algorithms and , highlighting how variational inequalities capture minimization under constraints. In , variational inequalities model Nash equilibria in noncooperative games with strategy sets. For an n-player game where player i chooses x_i \in K_i (convex compact) to maximize concave payoff f_i(x_1, \dots, x_n), a Nash equilibrium x^* = (x_1^*, \dots, x_n^*) \in \prod_{i=1}^n K_i satisfies \langle \nabla_{x_i} f_i(x^*), y_i - x_i^* \rangle \geq 0 for all y_i \in K_i and each i. This is precisely the variational inequality over the product set with operator F(x) whose i-th block is -\nabla_{x_i} f_i(x). Furthermore, convex optimization problems with inequality constraints can be reformulated as variational inequalities via the Karush-Kuhn-Tucker (KKT) conditions. For \min f(x) subject to g_j(x) \leq 0 (j=1,\dots,m), assuming convexity and differentiability, the KKT system—\nabla f(x^*) + \sum \lambda_j^* \nabla g_j(x^*) = 0, \lambda^* \geq 0, \lambda_j^* g_j(x^*) = 0, g(x^*) \leq 0—reduces to a variational inequality over the feasible set K = \{x : g(x) \leq 0\} with F(x) = \nabla f(x), or equivalently, a mixed variational inequality incorporating the multipliers. This reformulation enables unified treatment of constrained convex programs.

Complementarity and Obstacle Problems

The (LCP) provides a fundamental discrete example of a variational inequality arising in optimization and . Given a q \in \mathbb{R}^n and a M \in \mathbb{R}^{n \times n}, the LCP seeks vectors w, z \in \mathbb{R}^n satisfying w \geq 0, z \geq 0, w^T z = 0, and z = q + M w. This system enforces complementary slackness, where each pair (w_i, z_i) has at most one positive component. The LCP can be equivalently expressed as a variational inequality over the nonnegative : find x \in \mathbb{R}_+^n such that (q + M x)^T (y - x) \geq 0 \quad \forall y \in \mathbb{R}_+^n. This formulation highlights the LCP as a special case of the finite-dimensional variational inequality with the operator F(x) = q + M x and K = \mathbb{R}_+^n. The LCP was originally introduced by Lemke in the of bimatrix equilibrium points and . Its connection to variational inequalities was further elucidated in subsequent surveys. In continuum mechanics, the Signorini problem exemplifies variational inequalities modeling physical constraints due to unilateral contact between an elastic body and a rigid obstacle. Consider a linearly elastic body in equilibrium under given forces, with a portion of its boundary subject to nonpenetration against a rigid support defined by \psi. The key conditions on this contact boundary \Gamma are the nonpenetration u \geq \psi, the nonadhesion stress \sigma_n(u) \geq 0 (normal component of the stress tensor), and the complementarity (u - \psi) \sigma_n(u) = 0, ensuring no interpenetration or tensile forces across the contact set. These conditions lead to a variational inequality formulation in the appropriate Sobolev space, where the solution u satisfies the weak equilibrium equations interior to the domain and the boundary inequality on \Gamma. The problem was posed by Signorini in 1959 to describe ambiguous boundary conditions in elasticity, and its existence and uniqueness were established using variational methods by Fichera in 1964. The obstacle problem offers another canonical illustration of variational inequalities in the context of constrained minimization for elliptic partial differential equations. It models the deflection of an elastic membrane fixed at the boundary of a \Omega \subset \mathbb{R}^d and resting above a given obstacle function \psi, minimizing the subject to u \geq \psi in \Omega. Equivalently, the solution u solves the variational inequality: find u \in K = \{ v \in H^1_0(\Omega) : v \geq \psi \ \text{a.e. in } \Omega \} such that a(u, v - u) \geq 0 \quad \forall v \in K, where a(u,v) = \int_\Omega \nabla u \cdot \nabla v \, dx is the standard associated with the Laplacian. A key property is that the solution u coincides with the least superharmonic majorant of \psi, meaning u is the smallest superharmonic function dominating \psi. This interpretation links the problem to classical while capturing the free where u > \psi. The variational formulation of the obstacle problem was developed by Stampacchia in the mid-1960s as part of early work on elliptic variational inequalities.

Applications

In Mechanics and PDEs

Variational inequalities play a central role in modeling elastoplasticity, particularly in capturing the stress-strain relations under yield conditions. In the Prandtl-Reuss model for perfect plasticity, the evolution of plastic strain is governed by a monotone variational inequality that enforces the flow rule and the yield criterion, such as the von Mises condition. This formulation ensures that the stress tensor remains within the elastic domain while allowing irreversible deformation when the yield surface is reached, leading to a quasistatic rate-independent system in three dimensions. Unilateral contact problems in , such as the between an elastic body and a rigid , are naturally expressed as variational inequalities in Sobolev spaces like H^1(\Omega). For frictionless contact, the field u satisfies a variational inequality that incorporates the non-penetration condition \langle \sigma n, v - u \rangle \geq 0 on the potential contact boundary, where \sigma is the stress tensor and n the outward normal, ensuring unilateral constraints without interpenetration. This setup arises in boundary value problems for and has been foundational since the development of mixed formulations for such inequalities. Evolutionary variational inequalities extend these ideas to time-dependent phenomena, including phase transitions modeled by the Stefan problem. In the classical two-phase Stefan problem, the temperature evolution across a moving interface is reformulated as a parabolic variational inequality, where the phase change is captured by an obstacle-like condition on the or field, enforcing the release. This approach handles the free boundary dynamics through monotonicity and provides existence results for weak solutions in multidimensional settings. A specific application arises in the mixed formulation for past obstacles, where the velocity u and pressure p satisfy the variational inequality a(u, v) + b(v, p) - b(u, q) \geq (f, v) \quad \forall (v, q) \in V \times Q, with a(\cdot, \cdot) the viscous , b(\cdot, \cdot) the term, f the forcing, and the spaces V, Q incorporating no-slip and incompressibility, alongside obstacle constraints on u. This inequality models low-Reynolds-number flows with unilateral barriers, such as in or . Regularity theory for solutions of variational inequalities in mechanics provides optimal Hölder estimates for the free boundary in obstacle problems. In two dimensions, Caffarelli established that the solution exhibits C^{1,\alpha} regularity near regular points of the coincidence set, with explicit exponents depending on the dimension, which is crucial for analyzing contact interfaces and phase boundaries in PDE models.

In Economics and Game Theory

Variational inequalities provide a powerful framework for modeling states in systems and strategic interactions in , where agents optimize their objectives subject to constraints, leading to conditions that capture , supply-demand balances, and best-response behaviors. In , they extend classical concepts to scenarios involving transportation costs, congestion, and incomplete information, allowing for the analysis of decentralized decision-making without assuming full cooperation. In game theory, variational inequalities unify the formulation of Nash equilibria across continuous and discrete strategy spaces, facilitating the study of noncooperative outcomes where no player benefits from unilateral deviation. Recent applications include modeling equilibria in , such as in generative adversarial networks (GANs) and , where variational inequalities capture adversarial training dynamics and policy optimization under constraints as of 2024. Spatial price equilibrium problems, which determine prices and flows across regions accounting for supply, , and costs, can be formulated as a variational inequality. In this setting, the x includes supply quantities s, quantities d, and link flows f, with the feasible set K being the of nonnegative supply and sets and the feasible flow set. The F(x) is defined such that its components reflect effective s minus marginal costs for supplies, marginal costs minus delivered prices for demands, and costs minus differentials for flows, often expressed as F(x) = c(f) + \nabla p(s) - \pi(d) where c denotes costs, p supply prices, and \pi delivered prices, though variations like F(x) = c(x) - \nabla p appear in generalized forms. A x^* \in K satisfies \langle F(x^*), x - x^* \rangle \geq 0 for all x \in K, ensuring no opportunities exist. This formulation, introduced in seminal work on , allows for stability studies under parameter perturbations like cost changes. Traffic network equilibrium, based on Wardrop's first principle that all used paths between an origin-destination pair have equal and minimal travel times, is equivalently expressed as a variational inequality over feasible flow patterns. Here, the operator F assigns to each edge e the travel cost c_e(f_e), which may depend on the flow f_e on that edge due to congestion, and the convex set K comprises all nonnegative flows satisfying conservation of flow and demand constraints. An equilibrium flow f^* \in K solves \sum_e c_e(f^*_e) (f_e - f^*_e) \geq 0 for all f \in K, implying that no user can reduce their cost by switching routes. This VI structure, established in foundational traffic models, supports existence proofs under monotonicity assumptions on costs and enables extensions to asymmetric or multi-class user behaviors. In noncooperative games, equilibria correspond to solutions of variational inequalities where select strategies from convex compact sets to maximize individual payoffs. For a game with n , strategy profiles x = (x_1, \dots, x_n) \in K = K_1 \times \dots \times K_n, and payoff functions u_i(x_i, x_{-i}), the pseudogradient operator has components F_i(x) = \nabla_{x_i} (-u_i(x)), so the VI seeks x^* \in K such that \sum_i \langle F_i(x^*), x_i - x^*_i \rangle \geq 0 for all x \in K, equivalent to each player optimizing given others' strategies. This unification applies to both potential and general-sum s, with monotonicity of F ensuring uniqueness under strict conditions. Extensions of the Arrow-Debreu general model to variational inequalities in the 1980s accommodated by incorporating asset trading constraints and uncertainty, where prices clear both and futures markets without full spanning of states. These developments reformulated the excess into a VI operator over price and allocation spaces, allowing existence results via fixed-point mappings even when securities do not complete the . For finite bimatrix games, the variational inequality reduces to finding mixed strategies \sigma^* in the such that \sum_i (A \sigma^* - b)_i (\tau_i - \sigma^*_i) \geq 0 \quad \forall \tau \geq 0, \sum_i \tau_i = 1, where A is the payoff for one player, b is the of ones, and the ensures optimality against deviations; a symmetric VI holds for the opponent's . This formulation highlights how VI captures mixed-strategy Nash equilibria in zero-sum or general bimatrix settings.

Numerical Methods

Projection and Iterative Algorithms

Projection-based methods play a central role in iterative algorithms for solving variational inequalities (VIs) in finite-dimensional Hilbert spaces, where the feasible set K is a nonempty, closed, and convex subset. The orthogonal projection onto K, denoted \proj_K(z), is defined as the unique point in K that minimizes the Euclidean distance to a given point z: \proj_K(z) = \arg\min_{y \in K} \|y - z\|^2. This operator is firmly nonexpansive and thus nonexpansive, making it a foundational building block for ensuring convergence in projection-based iterations. Assuming the VI operator F is monotone (i.e., \langle F(x) - F(y), x - y \rangle \geq 0 for all x, y \in K), many algorithms leverage projections to handle the constraint K while approximating solutions to \mathrm{VI}(F, K). The extragradient , introduced by Korpelevich in , is a seminal projection-based for solving VIs where F is and continuous with constant L > 0. The iteration proceeds in two steps: first, compute an auxiliary point y^k = \proj_K(x^k - \alpha F(x^k)), where \alpha \in (0, 1/L); then, update x^{k+1} = \proj_K(x^k - \alpha F(y^k)). Under these assumptions on F, the converges to a solution of the VI starting from any initial x^0 \in K. Recent improvements have established an O(1/k) last-iterate rate for the extragradient in the monotone case. This rate measures the gap function or , providing non-ergodic guarantees that enhance practical reliability. For the case of strong monotonicity (i.e., \langle F(x) - F(y), x - y \rangle \geq \mu \|x - y\|^2 for some \mu > 0), the achieves linear rates. The proximal point algorithm, developed by Rockafellar in , extends the projection framework to handle maximal operators F, which are and have full with no proper enlargements. At each iteration, x^{k+1} solves the regularized : \mathrm{VI}(F + (1/\lambda) \mathrm{Id}, K), where \lambda > 0 is a stepsize and \mathrm{Id} is the identity operator; equivalently, x^{k+1} = \proj_K(x^k - \lambda F(x^{k+1})). For maximal F, the algorithm converges weakly to a solution of the original from any starting point, with the regularization ensuring solvability of the subproblems. This method is particularly robust for ill-conditioned problems, as the added proximal term promotes stability. Douglas-Rachford splitting addresses VIs reformulated as finding zeros of the sum of two maximal monotone operators, such as A + B where A = F (monotone) and B = N_K (normal cone to K). The iteration involves reflections: compute \mathrm{refl}_B = 2 \prox_{\lambda B} - \mathrm{Id} and \mathrm{refl}_A = 2 \prox_{\lambda A} - \mathrm{Id}, then update via z^{k+1} = \frac{1}{2} ( \mathrm{Id} + \mathrm{refl}_A \circ \mathrm{refl}_B ) (z^k), with the projected iterate x^{k+1} = \prox_{\lambda B}(z^{k+1}). Lions and Mercier proved weak convergence to a zero of A + B in 1979 for maximal monotone operators in Hilbert spaces. This splitting technique is versatile for structured VIs, enabling efficient computation when individual proximals are available.

Finite Element and Discretization Techniques

Discretization techniques for variational inequalities (VIs) often employ the Galerkin method to approximate solutions in finite-dimensional subspaces of the underlying function space V. In this approach, a VI defined over a convex set K \subset V is projected onto a discrete convex set K_h \subset V_h, where V_h is a finite element space, such as piecewise linear polynomials on a triangulation of the domain. For instance, in the obstacle problem, V_h consists of continuous functions that are linear on each element and satisfy the discrete obstacle constraint, leading to a finite-dimensional VI: find u_h \in K_h such that a(u_h, v_h - u_h) \geq \langle f, v_h - u_h \rangle for all v_h \in K_h, where a(\cdot, \cdot) is the bilinear form and f is the data functional. This formulation preserves the monotonicity and coercivity properties of the continuous problem when a is symmetric and positive definite, enabling efficient computation via standard finite element assembly. Error estimates for these Galerkin approximations rely on the quasi-best approximation property for monotone VIs. Under suitable regularity assumptions on the solution, the error \|u - u_h\|_V satisfies \|u - u_h\|_V \leq C \inf_{v_h \in K_h} \|u - v_h\|_V, where C is a constant independent of the mesh size h. For piecewise polynomial elements of degree r, this yields convergence rates of O(h^r) in the energy norm, with optimal rates achieved for linear elements (r=1) in problems like the obstacle or Signorini type, provided the exact solution is sufficiently smooth. These estimates have been rigorously established for both conforming and mixed formulations, highlighting the robustness of finite elements in handling the inequality constraints without loss of accuracy compared to elliptic PDEs. Solving the resulting nonlinear discrete VI requires specialized nonlinear solvers, such as the Newton method adapted to the variational structure. The method linearizes the discrete VI around an iterate by solving a linear saddle-point problem on the active set of constraints, with globalization achieved via or trust-region techniques to ensure descent in a merit function like the energy functional. This approach converges quadratically locally for strongly monotone operators and has been effectively combined with preconditioned conjugate gradient solvers for the linear steps, particularly in large-scale and simulations. For contact problems modeled as VIs, active set strategies provide an efficient alternative by iteratively identifying and updating the active (contact) and inactive sets based on Lagrange multipliers or gap functions. Starting from an initial guess, the algorithm solves a problem on the predicted active set, updates the sets according to Signorini conditions, and iterates until , often requiring only a few outer iterations for moderate . These methods are particularly suited to finite element discretizations in three dimensions, offering superlinear and integration with techniques for non-matching meshes in multi-body contact. Nonconforming finite elements are valuable for mixed formulations of VIs, such as those arising in Stokes-flow with obstacles or plate , where standard conforming spaces may fail . These elements relax inter-element , approximating the solution in a larger space while enforcing weak via additional terms in the . is ensured by extending Brezzi's inf-sup to the VI setting, requiring that the discrete spaces satisfy \inf_{q_h \neq 0} \sup_{v_h \neq 0} b(v_h, q_h) / (\|v_h\| \|q_h\|) \geq \beta > 0 on the kernel of the , which guarantees unique solvability and optimal error estimates akin to conforming methods. This extension, originally for saddle-point problems, has been adapted to VIs to handle incompressibility and without spurious modes.

References

  1. [1]
    [PDF] VARIATIONAL INEQUALITIES - HAL
    The variational inequalities generalize the theory of equations ; on the other hand a variational inequality can be reduced to an equation with the following.Missing: sources | Show results with:sources
  2. [2]
    [PDF] Variational Inequalities
    Variational inequality theory was introduced by Hart- man and Stampacchia (1966) as a tool for the study of partial differential equations with applications ...Missing: primary sources
  3. [3]
    On some non-linear elliptic differential-functional equations
    1966 On some non-linear elliptic differential-functional equations. Philip Hartman, Guido Stampacchia ... DOI: 10.1007/BF02392210. ABOUT; FIRST PAGE; CITED BY ...
  4. [4]
    An Introduction to Variational Inequalities and Their Applications
    Variational inequalities (equilibrium or evolution problems typically with convex constraints) are carefully explained in An Introduction to Variational ...Missing: primary sources
  5. [5]
    Leonhard Euler (1707 - 1783) - Biography - University of St Andrews
    It applies variational principles to determine the optimal ship design and first established the principles of hydrostatics ... Euler here also begins ...
  6. [6]
    [PDF] The Calculus of Variations - College of Science and Engineering
    Jan 7, 2022 · Some of these minimization problems played a key role in the historical development ... leading to a second order parametric variational problem, ...
  7. [7]
    Über das Dirichletsche Prinzip - EUDML
    Hilbert, David. "Über das Dirichletsche Prinzip." Mathematische Annalen 59 (1904): 161-186. ... Dirichlet principle, Hilbert space, harmonic functions ...<|control11|><|separator|>
  8. [8]
    [PDF] An appreciation of R. Courant's `variational methods for the solution ...
    Physically, rigid conditions correspond to rigid constraints of the system at the boundary C while natural conditions express equilib- rium of the system of C ...<|control11|><|separator|>
  9. [9]
    Monotone (nonlinear) operators in Hilbert space - Project Euclid
    Monotone (nonlinear) operators in Hilbert space. George J. Minty. DOWNLOAD PDF + SAVE TO MY LIBRARY. Duke Math. J. 29(3): 341-346 (September 1962).
  10. [10]
    Lions, J.-L. (1969) Quelques méthodes de résolution des problèmes ...
    Lions, J.-L. (1969) Quelques méthodes de résolution des problèmes aux limites non linéaires. Gauthier-Villars, Paris.
  11. [11]
    Finite-dimensional variational inequality and nonlinear ...
    Feb 17, 1989 · This paper provides a state-of-the-art review of these developments as well as a summary of some open research topics in this growing field.
  12. [12]
    Variational inequalities - Wiley Online Library
    Variational inequalities. J. L. Lions,. J. L. Lions. University of Paris. Search for more papers by this author · G. Stampacchia, ... Download PDF. back ...
  13. [13]
    Nonlinear monotone operators and convex sets in Banach spaces
    September 1965 Nonlinear monotone operators and convex sets in Banach spaces. Felix E. Browder · DOWNLOAD PDF + SAVE TO MY LIBRARY. Bull. Amer. Math. Soc.Missing: original | Show results with:original
  14. [14]
    [PDF] A Theoretical Perspective of Convex Optimization and Variational ...
    the relationship between convex optimization and variational inequality problems. ... The following result shows that the first-order condition for unconstrained ...
  15. [15]
    [PDF] Finite-Dimensional-Variational-Inequality-and-Nonlinear ...
    P.T. Harker, J.S. Pang Variational inequality and complementarity problems. An important special case of VI(X, F') is the nonlinear complementarity problem.
  16. [16]
    Variational Inequalities and the Signorini Problem for Nonlinear ...
    It is conceivable that the contact set could be especially cornplicated. Fichera (1964) was the first to study the existence and uniqueness of this problern ...
  17. [17]
    A Generalized Strange Term in Signorini\'s Type Problems
    The general question which will make the object of this paper is the homogenization of Signorini's type-like problems in perforated domains. Their classical ...
  18. [18]
    The Obstacle Problem and Best Superharmonic Approximation
    Least superharmonic majorants are fundamental in Potential Theory, where they are called réduites, and they occur in the the theory of elliptic PDEs as ...
  19. [19]
    A generalized Norton-Hoff model and the Prandtl-Reuss law of ...
    The aim of this article is to study the quasistatic evolution of a three-dimensional elastic-perfectly plastic solid which satisfies the Prandtl-Reuss law.
  20. [20]
    Variational methods in the stefan problem | SpringerLink
    Sep 12, 2006 · —The solution of a two-phase Stefan problem by a variational inequality, in: J.R. Ockendon, A.R. Hodgkins Eds., “Moving Boundary Problems in ...
  21. [21]
    [PDF] arXiv:2108.00046v3 [math.NA] 11 Oct 2022
    Oct 11, 2022 · We propose a mixed formulation of the Stokes variational inequality where a La- grange multiplier is used to enforce the contact conditions ...
  22. [22]
    Sensitivity Analysis for the General Spatial Economic Equilibrium ...
    Stella Dafermos, Anna Nagurney, (1984) Sensitivity Analysis for the General Spatial Economic Equilibrium Problem. Operations Research 32(5):1069-1086. https ...
  23. [23]
    Traffic Equilibrium and Variational Inequalities - PubsOnLine
    Exchange rates and multicommodity international trade: insights from spatial price equilibrium modeling with policy instruments via variational inequalities.
  24. [24]
    A penalty/Newton/conjugate gradient method for the solution of ...
    The numerical methodology combines penalty and Newton's method, the linearized problems being solved by a conjugate gradient algorithm.
  25. [25]
    A Primal-Dual Active Set Algorithm for Three-Dimensional Contact ...
    In this paper, efficient algorithms for contact problems with Tresca and Coulomb friction in three dimensions are presented and analyzed.
  26. [26]
    Error estimates for the finite element solution of variational inequalities
    We study the mixed finite element approximation of variational inequalities, taking as model problems the so called “obstacle problem” and “unilateral problem”.