Fact-checked by Grok 2 weeks ago

Calculus of variations

The calculus of variations is a field of mathematical analysis concerned with optimizing functionals, which are real-valued mappings defined on spaces of functions, by determining the functions that achieve extrema such as minima or maxima. Unlike standard calculus, which optimizes finite-dimensional functions of variables, it addresses infinite-dimensional problems by considering variations—small perturbations—of candidate functions to derive necessary conditions for optimality. The origins of the calculus of variations trace back to ancient Greek geometry, with early problems like finding the curve of minimal length enclosing a given area, but systematic development began in the 17th century. Isaac Newton applied variational ideas to minimize resistance in fluid flow, while Johann Bernoulli posed the famous brachistochrone problem in 1696, seeking the curve of fastest descent under gravity, which spurred further advances. Leonhard Euler formalized the subject in the 1740s through his work on isoperimetric problems and introduced the Euler-Lagrange equation as the core tool for solving variational problems. Joseph-Louis Lagrange extended these ideas in the 1760s, providing a more general framework that integrated dynamics and optimization, influencing the naming of the field as "calculus of variations." By the 19th century, contributions from Carl Gustav Jacobi, William Rowan Hamilton, and Karl Weierstrass refined the theory, addressing sufficiency conditions and higher-order variations. Central to the calculus of variations is the Euler-Lagrange equation, derived by setting the first variation of a functional to zero, which yields a differential equation that extremal functions must satisfy. For a functional of the form I = \int_a^b F(x, y, y') \, dx, the equation is \frac{\partial F}{\partial y} - \frac{d}{dx} \left( \frac{\partial F}{\partial y'} \right) = 0, generalizing the derivative concept to function spaces. Extensions include problems with constraints, leading to Lagrange multipliers in functional form, and higher-dimensional cases for surfaces or fields. Modern developments incorporate partial differential equations and numerical methods for complex, non-classical problems. The calculus of variations has profound applications across physics, engineering, and optimization. In classical mechanics, it underpins Hamilton's principle of stationary action, deriving equations of motion from minimizing the action integral. It optimizes trajectories in aerospace engineering, such as minimal-fuel paths for spacecraft, and shapes in structural design, like beams of least weight under load. In field theory, it formulates the Euler-Lagrange equations for electromagnetism and general relativity, while in economics and control theory, it solves optimal control problems for dynamic systems. These applications highlight its role in bridging pure mathematics with real-world modeling and simulation.

Fundamentals

Basic Definitions and Functionals

In the calculus of variations, a functional is defined as a mapping from a space of functions to the real numbers, often described as a "function of a function." The primary focus is on integral functionals, which take the form J = \int_a^b L\bigl(x, y(x), y'(x)\bigr) \, dx, where L is a given integrand function (known as the Lagrangian), y(x) is the dependent variable function, y'(x) is its derivative with respect to the independent variable x, and the integration is over a fixed interval [a, b]. The domain of such functionals consists of admissible functions, which are typically required to possess sufficient smoothness to ensure the integral is well-defined and differentiable in the variational sense. In classical settings, this space comprises continuously differentiable functions, denoted C^1[a, b], satisfying y \in C[a, b] and y' \in C[a, b]. More generally, in modern treatments, admissible functions may belong to Sobolev spaces W^{1,p}(a, b) for $1 \leq p \leq \infty, which allow for functions with weaker regularity while still controlling the integrability of derivatives. A central concern in the calculus of variations is the classical variational problem: to determine a function y from the space of admissible functions that extremizes (either minimizes or maximizes) the given functional J. This optimization is usually subject to boundary conditions at the endpoints of the interval. For fixed (or Dirichlet) boundary conditions, the function must satisfy y(a) = A and y(b) = B for prescribed constants A and B. When one or both endpoints are free, natural boundary conditions are applied instead, which are conditions derived intrinsically from the requirement that y extremizes J. These concepts find motivation in physics, particularly through principles such as the principle of stationary action, where physical trajectories correspond to functions that extremize an associated integral functional representing the system's action.

Variations and Derivatives

In the calculus of variations, a variation of a function y(x) is defined as a small perturbation given by \delta y(x) = \epsilon \eta(x), where \epsilon is a small scalar parameter and \eta(x) is an arbitrary smooth function that typically vanishes at the endpoints if the boundary conditions are fixed. This form represents an infinitesimal change in the function while preserving the boundary constraints, allowing analysis of how functionals respond to such perturbations. The Gateaux derivative of a functional J at a function y in the direction of \eta extends the concept of directional derivatives to infinite-dimensional spaces and is defined as \langle DJ(y), \eta \rangle = \lim_{\epsilon \to 0} \frac{J[y + \epsilon \eta] - J}{\epsilon}, provided the limit exists. For a typical functional of the form J = \int_a^b L(x, y(x), y'(x)) \, dx, this derivative evaluates to \langle DJ(y), \eta \rangle = \int_a^b \left( \frac{\partial L}{\partial y} \eta + \frac{\partial L}{\partial y'} \eta' \right) dx. The first variation \delta J[y; \eta] is identified with this Gateaux derivative, and for y to be a stationary point (a candidate for an extremum) of the functional, the first variation must vanish for all admissible directions \eta, i.e., \delta J[y; \eta] = 0. This condition is necessary for local extrema and forms the basis for deriving optimality criteria in variational problems. In constrained variational problems, such as those with integral or pointwise restrictions, admissible variations are those perturbations \eta that satisfy the constraint conditions, ensuring the perturbed function remains feasible. These variations are crucial for applying the stationarity condition within the feasible set, often leading to the use of Lagrange multipliers to handle the constraints.

History

Early Origins

The earliest precursors to the calculus of variations trace back to ancient Greek geometry, where problems involving optimization of curves and areas were contemplated without formal methods. One of the most famous is the isoperimetric problem, which seeks the curve of fixed perimeter that encloses the maximum area; this was known to Greek mathematicians such as Zenodorus around the 2nd century BCE, who proved that the circle provides the optimal solution among plane figures. A legendary formulation appears in the myth of Queen Dido, founder of Carthage, who purportedly used a strip of oxhide to enclose the largest possible territory along the North African coast, intuitively leading to a semicircular boundary as the maximizing shape. In the 17th century, variational ideas gained momentum through problems in physics and optics, often solved via geometric intuition rather than systematic calculus. A key precursor was Pierre de Fermat's principle of least time, articulated in 1657, which posits that light rays follow paths minimizing travel time between two points, thereby deriving the laws of reflection and refraction; this optical minimization foreshadowed the functional optimization central to variational methods. The brachistochrone problem, posed by Johann Bernoulli in 1696 as a challenge in the Acta Eruditorum, asked for the curve allowing a particle to descend under gravity from one point to another in the shortest time, counterintuitively a cycloid rather than a straight line. This was independently solved by Isaac Newton (anonymously), Gottfried Wilhelm Leibniz, Jacob Bernoulli, and Louis de l'Hôpital, all employing geometric analogies to refraction in optics, highlighting the need for a general approach to such extremal problems. Leonhard Euler's early work marked the transition toward formalization, building on these 17th-century challenges. In his 1744 treatise Methodus inveniendi lineas curvas maximi minimive proprietate gaudentes, Euler developed methods to solve variational problems by treating them as infinite-dimensional analogs of ordinary extrema, applying differential techniques to derive equations for optimal curves; this text systematically addressed problems like the brachistochrone and isoperimetric cases, laying the groundwork for the field. Euler introduced the term "calculus of variations" (calculus variationum) in his 1766 work Elementa calculi variationum, reflecting the emerging discipline's focus on varying functions to extremize integrals. From Euler's methods emerged key results like the Euler-Lagrange equation for finding stationary paths.

Modern Developments

The formalization of the calculus of variations began with Joseph-Louis Lagrange's publication of Mécanique Analytique in 1788, where he introduced the δ notation to denote small variations in functions, providing a systematic framework for analyzing functionals in mechanics. In the 19th century, Karl Weierstrass advanced the theory during the 1870s by distinguishing between strong and weak extrema, emphasizing the need for rigorous conditions on function smoothness to ensure minimizers in variational problems. Carl Gustav Jacobi developed sufficient conditions for extrema, including the Jacobi condition, while William Rowan Hamilton applied variational methods to mechanics, introducing the Hamilton-Jacobi equation. Adrien-Marie Legendre contributed to conditions for maxima and minima via analysis of the second variation, introducing the Legendre condition (F_{y'y'} > 0 for minima). This period also saw the emergence of Beltrami's identity in 1868, a first integral simplifying cases where the integrand is independent of the independent variable. The 20th century brought greater rigor to existence questions. In 1904, David Hilbert proved Dirichlet's principle using integral methods that anticipated Sobolev spaces, establishing the existence of solutions to boundary value problems via minimization of quadratic functionals. Mikhail Lavrentiev's 1926 counterexample highlighted gaps in existence theory, demonstrating that minimizers in certain spaces may not exist or may differ across admissible function classes, known as the Lavrentiev phenomenon. In the 1930s, Constantin Carathéodory generalized the calculus to non-holonomic systems, incorporating constraints via differential forms and extending variational principles to broader mechanical contexts. Contemporary extensions integrate the calculus of variations with optimal control and computation. Lev Pontryagin's maximum principle in the 1950s linked variational methods to control theory, providing necessary conditions for optimal trajectories in dynamical systems through adjoint variables and Hamiltonian maximization. From the 1970s onward, finite element methods emerged as a computational tool for approximating solutions to variational problems, discretizing domains to solve partial differential equations arising from functionals in engineering and physics.

Core Theory

Extrema of Functionals

In the calculus of variations, extrema of functionals represent the maximum or minimum values attained by a functional J over a suitable class of functions y(x). A local extremum occurs at a function y_0(x) if there exists a neighborhood such that J \geq J[y_0] (or \leq) for all y in that neighborhood. The primary necessary condition for y_0 to be a local extremum is that it is a stationary point, meaning the first variation of the functional vanishes: \delta J[y_0; h] = 0 for all admissible variations h(x). This condition implies that small perturbations around y_0 do not change the functional to first order, analogous to the derivative being zero for finite-dimensional extrema. Extrema are classified as weak or strong based on the topology of the function space considered for nearby functions. A weak local minimum (or maximum) at y_0 requires that J \geq J[y_0] (or \leq) for all y sufficiently close to y_0 in the C^1 norm, meaning both y and y' converge uniformly to y_0 and y_0'. In contrast, a strong local minimum holds if the inequality is satisfied for functions close in the C^0 norm, allowing only uniform convergence of y to y_0 without restrictions on the derivatives, which permits more irregular perturbations. Weak extrema are easier to characterize analytically, as they align with smoothness assumptions common in classical variational problems. To establish sufficient conditions for local minima, the second variation \delta^2 J[y_0; h] plays a central role, serving as the quadratic form analogous to the second derivative test. If the second variation is positive definite—meaning \delta^2 J[y_0; h] > 0 for all nontrivial admissible variations h—then y_0 is a strict weak local minimum, provided the first variation vanishes. Positive definiteness ensures convexity in a neighborhood, guaranteeing that the functional increases quadratically for small perturbations. This condition relates to the Legendre condition on the Hessian of the integrand but focuses here on the global quadratic form. Variational problems often incorporate constraints that affect the nature of extrema. Fixed boundary conditions specify y(a) and y(b) at the endpoints, restricting variations to h(a) = h(b) = 0, while free boundaries allow natural conditions to emerge from the stationarity requirement. Simple integral constraints, known as isoperimetric constraints, impose equality conditions like \int_a^b K(x, y, y') \, dx = c for some constant c, preserving quantities such as arc length in geometric problems; these lead to augmented functionals or Lagrange multipliers for stationary points. The existence of extrema, particularly minima, is addressed through direct methods in the calculus of variations, which rely on topological arguments rather than solving differential equations. These methods typically assume the functional is coercive (unbounded below only at infinity), lower semicontinuous, and defined on a reflexive Banach space like a Sobolev space; compactness from embedding theorems then yields a minimizing sequence converging to a global minimizer. The vanishing first variation at this minimizer provides the necessary condition, often leading to the Euler–Lagrange equation as a characterization.

Euler–Lagrange Equation

The Euler–Lagrange equation provides the necessary condition for a function y(x) to extremize a functional of the form J = \int_a^b L(x, y(x), y'(x)) \, dx, where L is a smooth Lagrangian depending on x, y, and y'. To derive this equation, consider the first variation \delta J induced by a small variation \eta(x) in y(x), with \eta(a) = \eta(b) = 0 to respect fixed endpoints. The variation expands as \delta J = \int_a^b \left( \frac{\partial L}{\partial y} \eta + \frac{\partial L}{\partial y'} \eta' \right) dx = 0 for all admissible \eta. Integrating the second term by parts yields \int_a^b \frac{\partial L}{\partial y'} \eta' \, dx = \left[ \frac{\partial L}{\partial y'} \eta \right]_a^b - \int_a^b \frac{d}{dx} \left( \frac{\partial L}{\partial y'} \right) \eta \, dx. The boundary term vanishes due to the conditions on \eta, so \delta J = \int_a^b \left( \frac{\partial L}{\partial y} - \frac{d}{dx} \left( \frac{\partial L}{\partial y'} \right) \right) \eta \, dx = 0. By the fundamental lemma of the calculus of variations, which states that if the integral of a continuous function times arbitrary \eta is zero, then the function itself must be zero, it follows that \frac{\partial L}{\partial y} - \frac{d}{dx} \left( \frac{\partial L}{\partial y'} \right) = 0. This is the standard form of the Euler–Lagrange equation. If one or both endpoints are free (i.e., y(a) or y(b) not fixed), the corresponding boundary term in the integration by parts does not vanish. For a free endpoint at x = b, the condition \frac{\partial L}{\partial y'}(b) \eta(b) = 0 holds for arbitrary \eta(b), implying the natural boundary condition \frac{\partial L}{\partial y'}(b) = 0. A similar condition applies at x = a. These ensure the variation is stationary without prescribed boundary values. The Euler–Lagrange equation is a second-order ordinary differential equation in y(x), as the derivative term generally involves second derivatives of y. For well-posed problems with suitable boundary conditions (fixed or natural), existence and uniqueness of solutions follow from standard ODE theory, assuming L satisfies regularity conditions like continuity and differentiability. A simple example is the problem of finding the shortest path between two points (a, y_a) and (b, y_b) in the plane, where the functional is the arc length J = \int_a^b \sqrt{1 + (y')^2} \, dx, so L = \sqrt{1 + (y')^2}. Here, \frac{\partial L}{\partial y} = 0 and \frac{\partial L}{\partial y'} = \frac{y'}{\sqrt{1 + (y')^2}}, leading to \frac{d}{dx} \left( \frac{y'}{\sqrt{1 + (y')^2}} \right) = 0. Integrating gives \frac{y'}{\sqrt{1 + (y')^2}} = c (constant), which simplifies to y' = k (constant slope), or y'' = 0. Thus, the extremal is a straight line, the geodesic in Euclidean space.

Classical Results and Identities

Beltrami's Identity

Beltrami's identity provides a first integral for variational problems where the Lagrangian L depends only on the dependent variable y and its derivative y', but not explicitly on the independent variable x, satisfying \frac{\partial L}{\partial x} = 0. This condition implies an autonomous system, analogous to conservation laws in mechanics, and reduces the second-order Euler–Lagrange equation to a first-order differential equation. The identity simplifies solving for extremals by eliminating the need to integrate a higher-order ODE directly. To derive the identity, start with the Euler–Lagrange equation for such Lagrangians: \frac{d}{dx}\left(\frac{\partial L}{\partial y'}\right) - \frac{\partial L}{\partial y} = 0. The total derivative of L is \frac{dL}{dx} = \frac{\partial L}{\partial y} y' + \frac{\partial L}{\partial y'} y''. Substituting \frac{\partial L}{\partial y} = \frac{d}{dx}\left(\frac{\partial L}{\partial y'}\right) yields \frac{dL}{dx} = y' \frac{d}{dx}\left(\frac{\partial L}{\partial y'}\right) + \frac{\partial L}{\partial y'} y'' = \frac{d}{dx}\left(y' \frac{\partial L}{\partial y'}\right). Rearranging gives \frac{d}{dx}\left(L - y' \frac{\partial L}{\partial y'}\right) = 0, so integrating produces Beltrami's identity: L - y' \frac{\partial L}{\partial y'} = C, where C is a constant determined by boundary conditions. This identity often allows explicit solution for y' as a function of y, transforming the problem into a separable first-order ODE for x(y). A classic application is the brachistochrone problem, minimizing travel time for a particle sliding under gravity from (0,0) to (a,b). The Lagrangian is L = \frac{\sqrt{1 + (y')^2}}{\sqrt{y}} (up to constants). Applying Beltrami's identity: \frac{\partial L}{\partial y'} = \frac{y'}{\sqrt{y(1 + (y')^2)}}, \quad y' \frac{\partial L}{\partial y'} = \frac{(y')^2}{\sqrt{y(1 + (y')^2)}}, L - y' \frac{\partial L}{\partial y'} = \frac{1}{\sqrt{y(1 + (y')^2)}} = C. Thus, \sqrt{y(1 + (y')^2)} = 1/C = k, so y' = \sqrt{\frac{k^2}{y} - 1}. Integrating dx = \frac{dy}{\sqrt{k^2/y - 1}} yields the cycloid parametric equations x = \frac{k^2}{2} (\theta - \sin \theta), y = \frac{k^2}{2} (1 - \cos \theta), with k fitted to boundary points.

Euler–Poisson Equation

The Euler–Poisson equation arises in the calculus of variations when the functional depends on derivatives of the unknown function up to order n, generalizing the standard Euler–Lagrange equation (which corresponds to n=1). Consider a functional I = \int_a^b L\left(x, y, y', \dots, y^{(n)}\right) \, dx, where y^{(k)} = \frac{d^k y}{dx^k} for k = 0, 1, \dots, n, with y^{(0)} = y. For y to extremize I, it must satisfy the Euler–Poisson equation \sum_{k=0}^n (-1)^k \frac{d^k}{dx^k} \left( \frac{\partial L}{\partial y^{(k)}} \right) = 0. \tag{1}\label{eq:ep} This is a higher-order differential equation of order $2n. To derive \eqref{eq:ep}, compute the first variation of I under an admissible variation \eta(x) with \eta(a) = \eta(b) = 0 (and higher derivatives vanishing at endpoints if specified). The first-order change is \delta I = \int_a^b \sum_{k=0}^n \frac{\partial L}{\partial y^{(k)}} \eta^{(k)} \, dx = 0. Integrate the k-th term by parts k times: \int_a^b \frac{\partial L}{\partial y^{(k)}} \eta^{(k)} \, dx = (-1)^k \int_a^b \frac{d^k}{dx^k} \left( \frac{\partial L}{\partial y^{(k)}} \right) \eta \, dx + \left[ \text{boundary terms involving } \eta, \eta', \dots, \eta^{(k-1)} \right]_a^b. Summing over k and requiring \delta I = 0 for all admissible \eta yields \eqref{eq:ep} by the fundamental lemma of calculus of variations (after the interior integral vanishes), provided the boundary terms are handled appropriately. For the second-order case (n=2), where L = L(x, y, y', y''), equation \eqref{eq:ep} specializes to \frac{\partial L}{\partial y} - \frac{d}{dx} \left( \frac{\partial L}{\partial y'} \right) + \frac{d^2}{dx^2} \left( \frac{\partial L}{\partial y''} \right) = 0. \tag{2}\label{eq:ep2} This is a fourth-order ordinary differential equation. In problems with unspecified boundary values, natural boundary conditions ensure the boundary terms from integration by parts vanish. For general n, if y^{(m)} is free at an endpoint for m = 0, 1, \dots, n-1, then \sum_{k=m+1}^n (-1)^{k-m-1} \frac{d^{k-m-1}}{dx^{k-m-1}} \left( \frac{\partial L}{\partial y^{(k)}} \right) \bigg|_{endpoint} = 0. For n=2, if y(b) is free, the condition is \frac{\partial L}{\partial y'} - \frac{d}{dx} \frac{\partial L}{\partial y''} \big|_{x=b} = 0; if y'(b) is also free, an additional condition \frac{\partial L}{\partial y''} \big|_{x=b} = 0 applies. These supplement \eqref{eq:ep2} to form a complete boundary value problem. A classic application appears in the bending of a thin elastic beam under no load, minimizing the strain energy functional I = \frac{1}{2} \int_a^b (y'')^2 \, dx (with unit flexural rigidity). Here L = \frac{1}{2} (y'')^2, so \frac{\partial L}{\partial y} = \frac{\partial L}{\partial y'} = 0 and \frac{\partial L}{\partial y''} = y'', yielding \frac{d^2}{dx^2} (y'') = 0 or y^{(4)} = 0 from \eqref{eq:ep2}. Solutions are cubic polynomials, describing the beam's deflection shape.

Existence and Regularity

Du Bois-Reymond's Theorem

Du Bois-Reymond's theorem, also known as the fundamental lemma of the calculus of variations, asserts that if f is a continuous function on the closed interval [a, b] and \int_a^b f(x) \eta(x) \, dx = 0 for every smooth test function \eta \in C_c^\infty(a, b) with compact support in the open interval (a, b), then f(x) = 0 for all x \in (a, b). This result establishes a necessary condition for weak extrema of functionals, implying that the Euler–Lagrange equation holds almost everywhere when the Lagrangian is continuous. The proof proceeds by contradiction. Suppose there exists x_0 \in (a, b) such that f(x_0) > 0. By continuity of f, there is a subinterval (c, d) \subset (a, b) where f(x) > f(x_0)/2 > 0 for all x \in (c, d). Construct a nonnegative test function \eta that is identically 1 on (c, d) and zero outside a slightly larger compact interval within (a, b), smoothed appropriately to be in C_c^\infty(a, b). Then the integral becomes strictly positive, contradicting the assumption. A similar argument applies if f(x_0) < 0. For the more general case where f \in L^1(a, b), the result follows by approximating f with continuous functions via density arguments in L^1 spaces and applying the continuous case. This theorem has significant implications for the regularity of solutions in variational problems. For a weak extremum of a functional J = \int_a^b L(x, y, y') \, dx with continuous Lagrangian L, the first variation vanishing leads to the integrated form of the Euler–Lagrange equation, and by the theorem, the pointwise equation holds almost everywhere. If L is sufficiently smooth (e.g., C^2), classical bootstrap arguments show that extremals are C^1, though the theorem itself accommodates weak solutions satisfying the equation only almost everywhere. Historically, Paul du Bois-Reymond developed this result in the late 1870s to rigorously justify the derivation of the Euler–Lagrange equation for minimizers potentially exhibiting discontinuities, building on earlier attempts like Stegmann's 1854 effort and resolving foundational issues in the field. His proof appeared in the 1879 paper "Erläuterungen zu den Anfangsgründen der Variationsrechnung" published in Mathematische Annalen. The theorem's limitations lie in its focus on necessary conditions alone; it does not address the existence of extrema nor distinguish between weak and strong minima, leaving room for pathologies in non-regular settings.

Lavrentiev Phenomenon

The Lavrentiev phenomenon arises in the calculus of variations when there is a strict discrepancy between the infimum of a functional J = \int_a^b L(t, y, y') \, dt over an admissible set A (typically smooth functions like C^1[a,b] with fixed boundary conditions) and the infimum over the closure of A in a weaker norm (such as the Sobolev space W^{1,p}(a,b)). This gap occurs because minimizers in the weaker topology may not belong to the original set A, leading to non-attainment of the infimum in stronger topologies. A classic construction demonstrating this was provided by Mikhail Lavrentiev in 1926, involving a non-convex Lagrangian where the infimum is not attained in the class of continuously differentiable functions but is lower in a Sobolev space. A simplified version of this example, due to Mania in 1934, considers the functional J = \int_0^1 (t - y^3)^2 (y')^6 \, dt with boundary conditions y(0) = 0 and y(1) = 1. Here, the infimum over C^1[0,1] functions is positive (specifically 1), yet over W^{1,6}(0,1), it drops to 0, achieved by a limiting sequence of functions that oscillate wildly near the boundary, preventing smooth approximation. This example highlights how non-convexity allows for "pathological" minimizers that escape smooth classes. The phenomenon has significant implications for approximation in numerical schemes, where discretizations based on smooth or piecewise linear functions fail to converge to the true infimum, often overestimating the minimum energy. To address this, relaxation methods are employed, replacing the original functional with a lower semicontinuous convex envelope that attains the infimum in the weaker space. Resolution approaches include the use of Young measures, which capture fine-scale oscillations in limiting sequences of minimizers, and gamma-convergence, a variational convergence notion that ensures the relaxed functional approximates the original infimum correctly. A key result is that the Lavrentiev phenomenon manifests precisely when the Lagrangian L lacks quasi-convexity, a condition weaker than convexity that ensures lower semicontinuity under weak convergence in Sobolev spaces. This underscores the necessity of smoothness assumptions in theorems like Du Bois-Reymond's to guarantee attainment without such gaps.

Multivariable and Boundary Value Problems

Functions of Several Variables

In the calculus of variations, functionals depending on functions of several variables take the form J = \int_{\Omega} L(x, y(x), \nabla y(x)) \, dx, where \Omega \subset \mathbb{R}^n is an open domain, y: \Omega \to \mathbb{R}^m is a vector-valued function with m \geq 1, and L is a smooth scalar function depending on the position x, the function value y, and its gradient \nabla y (the Jacobian matrix). This formulation generalizes the classical single-variable case (where n=1, m=1) to higher dimensions, allowing for problems in fields like elasticity and field theories. To derive the conditions for critical points of J, consider a smooth variation y_\epsilon = y + \epsilon \eta, where \eta: \Omega \to \mathbb{R}^m is a test function vanishing near the boundary (for interior extrema), and compute the first variation \frac{d}{d\epsilon} J[y_\epsilon] \big|_{\epsilon=0} = 0. This yields \int_{\Omega} \left( \frac{\partial L}{\partial y} \cdot \eta + \sum_{i=1}^m \sum_{j=1}^n \frac{\partial L}{\partial (\partial_j y_i)} \partial_j \eta_i \right) dx = 0 for all admissible \eta. Applying multivariable integration by parts via the divergence theorem (Green's theorem in higher dimensions) to the second term gives \int_{\Omega} \eta \cdot \left( \frac{\partial L}{\partial y} - \div \left( \frac{\partial L}{\partial \nabla y} \right) \right) dx + \text{boundary integral} = 0, where \div \left( \frac{\partial L}{\partial \nabla y} \right) denotes the vector whose i-th component is \sum_{j=1}^n \partial_j \left( \frac{\partial L}{\partial (\partial_j y_i)} \right). Since \eta is arbitrary, the Euler–Lagrange system follows: \div \left( \frac{\partial L}{\partial \nabla y} \right) = \frac{\partial L}{\partial y}, a vector equation holding componentwise. Boundary conditions specify the behavior on \partial \Omega. For Dirichlet conditions, y is prescribed on the boundary, so \eta = 0 there, and the boundary integral vanishes without additional constraints. For natural (Neumann) conditions, the variation \eta is free on \partial \Omega, requiring the boundary term to vanish independently, which imposes \frac{\partial L}{\partial \nabla y} \cdot \nu = 0 on \partial \Omega, where \nu is the outward unit normal and \cdot denotes the appropriate contraction (e.g., the flux of the rows of \frac{\partial L}{\partial \nabla y}). A representative example is Plateau's problem, which seeks a surface of minimal area spanning a given boundary curve. For a parametrized surface y: \Omega \subset \mathbb{R}^2 \to \mathbb{R}^3, the area functional is J = \int_{\Omega} \sqrt{\det(g)} \, dx, where g_{kl} = \nabla y_k \cdot \nabla y_l is the induced metric tensor. The corresponding Euler–Lagrange system yields the minimal surface equation, expressing that the mean curvature vanishes, \Delta y \perp T_y \mathbb{R}^3 (where \Delta is the Laplace–Beltrami operator on the surface). A simpler graph case (m=1, n=2) reduces to J = \int_{\Omega} \sqrt{1 + |\nabla y|^2} \, dx, with Euler–Lagrange equation \div \left( \frac{\nabla y}{\sqrt{1 + |\nabla y|^2}} \right) = 0.

Dirichlet's Principle and Generalizations

Dirichlet's principle, introduced in the 1850s by Peter Gustav Lejeune Dirichlet, posits that the solution to the Dirichlet problem for Laplace's equation—finding a function u satisfying \Delta u = 0 in a bounded domain \Omega \subset \mathbb{R}^n with prescribed boundary values u = g on \partial \Omega—is obtained by minimizing the Dirichlet energy functional J(u) = \int_\Omega |\nabla u|^2 \, dx over all functions in a suitable space that agree with g on the boundary. This variational approach reformulates the partial differential equation (PDE) as an optimization problem, leveraging the framework of multivariable calculus of variations where the integrand depends on the gradient but not explicitly on u itself. The equivalence between minimization and the PDE follows directly from the Euler–Lagrange equation applied to the quadratic Lagrangian L(\nabla u) = \frac{1}{2} |\nabla u|^2. For such a functional with fixed boundary values, the stationarity condition yields \operatorname{div}(\nabla L / \partial \nabla u) = 0, which simplifies to \Delta u = 0 in \Omega. Although intuitively appealing and rooted in physical analogies, the principle initially lacked a fully rigorous existence proof due to potential issues with the minimizing sequence converging to a non-smooth function; David Hilbert provided such a proof in 1900 using the direct method in the calculus of variations, incorporating compactness arguments to ensure a minimizer exists in appropriate Sobolev spaces under suitable domain regularity assumptions. A prominent physical interpretation arises in electrostatics, where the electric potential u in a region \Omega with given boundary potentials g minimizes the electrostatic energy functional \frac{1}{8\pi} \int_\Omega |\nabla u|^2 \, dx = \frac{1}{8\pi} \int_\Omega | \mathbf{E} |^2 \, dV, with \mathbf{E} = -\nabla u denoting the electric field; this minimizer satisfies Laplace's equation, representing equilibrium in the absence of charges. Generalizations extend the principle to other boundary conditions and higher-order PDEs. For Robin boundary conditions, which combine Dirichlet and Neumann aspects as \frac{\partial u}{\partial \nu} + \alpha u = \beta on \partial \Omega (where \nu is the outward normal), the variational formulation incorporates a boundary integral term, such as J(u) = \int_\Omega \frac{1}{2} |\nabla u|^2 \, dx + \int_{\partial \Omega} \frac{\alpha}{2} u^2 \, dS, yielding Laplace's equation interiorly and the Robin condition as the natural boundary condition upon variation. Similarly, for the biharmonic equation \Delta^2 u = 0 modeling thin plate deflection under clamped boundaries (u = \frac{\partial u}{\partial \nu} = 0 on \partial \Omega), the principle applies by minimizing the bending energy functional \int_\Omega (\Delta u)^2 \, dx over functions vanishing with their normal derivative on the boundary, with the Euler–Lagrange equation producing the biharmonic PDE. These extensions preserve the energy-minimization paradigm while adapting to more complex boundary value problems in PDE theory.

Eigenvalue Problems

Sturm–Liouville Problems

Sturm–Liouville problems arise in the calculus of variations as eigenvalue problems for self-adjoint ordinary differential operators in one dimension, where the eigenvalues and eigenfunctions are characterized variationally through minimization principles. These problems typically take the form of finding eigenvalues \lambda and non-trivial solutions y to the boundary value problem -(p y')' + q y = \lambda r y on an interval [a, b], with suitable boundary conditions ensuring self-adjointness, such as separated or periodic conditions. The variational formulation links this directly to functionals in the calculus of variations, where the eigenfunctions extremize certain quadratic forms subject to constraints. Central to this approach is the Rayleigh quotient, defined for admissible functions y (twice differentiable and satisfying the boundary conditions) as R = \frac{\int_a^b \left[ p (y')^2 + q y^2 \right] \, dx}{\int_a^b r y^2 \, dx}, where p > 0, r > 0, and q are given coefficient functions, assumed sufficiently smooth. This quotient originates from Lord Rayleigh's work on the theory of sound, where it was used to approximate fundamental frequencies of vibrating systems. The critical points of R correspond to the eigenvalues, obtained by applying the Euler–Lagrange equation to the associated variational problem of minimizing the numerator subject to the denominator being fixed (or vice versa). Specifically, varying the functional \int_a^b \left[ p (y')^2 + q y^2 - \lambda r y^2 \right] \, dx = 0 yields the Sturm–Liouville equation -(p y')' + q y = \lambda r y. The variational characterization provides a complete ordering of the eigenvalues: the smallest eigenvalue \lambda_1 is the minimum of R over all non-zero admissible y, with the minimizer being the corresponding eigenfunction. Higher eigenvalues \lambda_k are characterized by the Courant–Fischer min-max theorem, which states that \lambda_k = \min_{\dim V = k} \max_{y \in V, \, y \neq 0} R = \max_{\dim W = k-1} \min_{\substack{y \perp W \\ y \neq 0}} R, where the minima and maxima are taken over subspaces V and W of admissible functions. This principle, developed by Courant and Hilbert, ensures that the eigenvalues form an increasing sequence \lambda_1 < \lambda_2 < \cdots tending to infinity, and it facilitates numerical approximations via the Rayleigh–Ritz method. The self-adjoint nature of the Sturm–Liouville operator guarantees that eigenfunctions corresponding to distinct eigenvalues are orthogonal with respect to the weight r, i.e., \int_a^b y_m y_n r \, dx = 0 for m \neq n. This orthogonality follows from integrating the difference of the equations for two eigenfunctions against each other and applying integration by parts, leveraging the boundary conditions to eliminate boundary terms. A classic example is the vibrating string problem, modeling transverse vibrations of a uniform string of length L under tension T with fixed ends, where p(x) = T, q(x) = 0, and r(x) = \rho (linear density). The equation becomes -T y'' = \lambda \rho y, or y'' + \frac{\lambda \rho}{T} y = 0 with y(0) = y(L) = 0. The eigenvalues are \lambda_k = \left( \frac{k \pi}{L} \right)^2 \frac{T}{\rho} for k = 1, 2, \dots, with eigenfunctions y_k(x) = \sin \left( \frac{k \pi x}{L} \right). The Rayleigh quotient here becomes R = \frac{T \int_0^L (y')^2 \, dx}{\rho \int_0^L y^2 \, dx}, whose minimum \lambda_1 = \left( \frac{\pi}{L} \right)^2 \frac{T}{\rho} gives the square of the fundamental angular frequency.

Eigenvalue Problems in Several Dimensions

Eigenvalue problems in several dimensions arise in the calculus of variations when seeking critical points of Rayleigh quotients associated with elliptic partial differential equations, such as the Dirichlet Laplacian on a bounded domain \Omega \subset \mathbb{R}^n. The prototypical problem is to find \lambda > 0 and u \not\equiv 0 in H_0^1(\Omega) satisfying -\Delta u = \lambda u in \Omega, with u = 0 on \partial \Omega. This is variationally characterized by extremizing the functional J = \frac{\int_\Omega |\nabla u|^2 \, dx}{\int_\Omega u^2 \, dx}, where the eigenvalues \{\lambda_k\}_{k=1}^\infty correspond to the critical values of J, ordered increasingly, and the eigenfunctions \{e_k\} form an orthonormal basis of L^2(\Omega). The min-max theorem provides a variational principle for these eigenvalues, extending the one-dimensional Sturm–Liouville theory to higher dimensions. Specifically, the k-th eigenvalue satisfies \lambda_k = \min \max J, where the minimum is taken over all k-dimensional subspaces of H_0^1(\Omega) orthogonal to the span of the first k-1 eigenfunctions \{e_1, \dots, e_{k-1}\}, and the maximum is over u in that subspace with \int_\Omega u^2 \, dx = 1. This characterization, known as the Courant–Fischer min-max principle, ensures that \lambda_k is the best possible upper bound obtainable from Rayleigh quotients over suitable finite-dimensional trial spaces. A key application of these variational principles is the Faber–Krahn inequality, which asserts that among all domains \Omega \subset \mathbb{R}^n of fixed volume |\Omega| = V, the first Dirichlet eigenvalue \lambda_1(\Omega) is minimized uniquely by the ball B of the same volume, with \lambda_1(\Omega) \geq \lambda_1(B) = \left( j_{(n/2)-1,1} / r \right)^2, where r = (V / \omega_n)^{1/n} is the radius of B and j_{\nu,1} is the first positive zero of the Bessel function J_\nu. This isoperimetric result follows from symmetrization techniques applied to the minimizers of J and has implications for shape optimization in spectral geometry. For numerical approximation, the Rayleigh–Ritz method projects the infinite-dimensional problem onto a finite-dimensional subspace spanned by trial functions \{\phi_1, \dots, \phi_m\} \subset H_0^1(\Omega), leading to a matrix eigenvalue problem whose eigenvalues provide upper bounds to the first m eigenvalues of the Laplacian. The method, originally developed by Rayleigh for approximate solutions to vibration problems and extended by Ritz to more general boundary value problems, converges to the true spectrum as the basis dimension increases, with error estimates depending on the choice of basis functions. A concrete example occurs for the unit disk \Omega = D = \{ (x,y) \in \mathbb{R}^2 : x^2 + y^2 < 1 \}, where separation of variables in polar coordinates yields eigenfunctions involving Bessel functions J_m(\sqrt{\lambda} \rho) \cos(m\theta) or \sin(m\theta), with eigenvalues \lambda_{m,k} = j_{m,k}^2 for m = 0,1,2,\dots and k = 1,2,\dots, where j_{m,k} is the k-th positive zero of J_m. The first eigenvalue is thus \lambda_1 = j_{0,1}^2 \approx 5.783, corresponding to the radial mode J_0(j_{0,1} \rho). For a disk of radius r, the eigenvalues scale as \lambda_{m,k} = j_{m,k}^2 / r^2.

Advanced Conditions and Variations

Sufficient Conditions for Extrema

In the calculus of variations, sufficient conditions for a local extremum of a functional J = \int_a^b L(x, y, y') \, dx at an extremal y rely on analyzing the second-order behavior through the second variation and related criteria. The first variation \delta J being zero is necessary for stationarity, but to confirm a minimum or maximum, higher-order terms provide the requisite guarantees. These conditions distinguish between weak extrema (for variations tangent to the extremal) and strong extrema (for arbitrary nearby curves). The second variation \delta^2 J[y; \eta] for a variation \tilde{y} = y + \epsilon \eta is obtained by expanding J[\tilde{y}] in a Taylor series around \epsilon = 0, yielding the quadratic form \delta^2 J[y; \eta] = \int_a^b \left( \frac{\partial^2 L}{\partial y^2} \eta^2 + 2 \frac{\partial^2 L}{\partial y \partial y'} \eta \eta' + \frac{\partial^2 L}{\partial (y')^2} (\eta')^2 \right) dx, where the partial derivatives are evaluated along the extremal y. For a weak local minimum, a sufficient condition is that \delta^2 J[y; \eta] > 0 for all admissible variations \eta \not\equiv 0 with \eta(a) = \eta(b) = 0. To ensure this positivity, the Legendre condition requires that \frac{\partial^2 L}{\partial (y')^2} > 0 along the entire extremal, guaranteeing the quadratic form is positive definite in the derivative term. Complementing this, the Jacobi condition states that the associated Jacobi equation—a linear second-order ODE derived from the Euler-Lagrange equation for the second variation—has no nontrivial solutions vanishing at two distinct points in [a, b], known as the absence of conjugate points. These conditions together imply that the second variation is positive, confirming a weak local minimum. For strong local minima, where variations may not be tangent, the Weierstrass condition applies: the Weierstrass excess function E(x, y, p; \bar{y}, \bar{p}) = L(x, y, \bar{p}) - L(x, y, p) - (\bar{p} - p) \frac{\partial L}{\partial y'}(x, y, p) must satisfy E \geq 0 for all x \in [a, b], y near the extremal, and admissible slopes p, \bar{p}, with equality only when \bar{p} = p. This non-negativity ensures no nearby curve yields a smaller functional value. A representative example is the quadratic functional J = \frac{1}{2} \int_a^b \left[ (y')^2 + q(x) y^2 \right] dx with q(x) > 0, subject to fixed endpoints. The Euler-Lagrange equation yields a unique extremal, and the second variation coincides with J[\eta], which is positive definite by the properties of the quadratic form, confirming a global minimum.

General Variations

In the calculus of variations, general variations extend the analysis of functionals beyond fixed endpoints and smooth paths, incorporating perturbations that allow for variable boundaries and constraints. These techniques are essential for handling problems where the extremal may intersect prescribed curves or satisfy integral constraints, ensuring the first variation vanishes appropriately. Transversality conditions arise when the endpoint of the extremal lies on a movable curve, such as \phi(x, y) = 0, requiring the variation to be tangent to this curve at the endpoint. For a functional J = \int_a^b L(x, y, y') \, dx, the condition for a stationary point is \delta J + \left( \frac{\partial L}{\partial y'} \delta y \right) \big|_{b} = 0, where \delta y satisfies the tangency \frac{\partial \phi}{\partial x} + \frac{\partial \phi}{\partial y} \delta y = 0. This ensures the boundary term in the integration by parts of the first variation aligns with the constraint, as derived in standard treatments of variable endpoint problems. A classic application of transversality conditions is the derivation of the law of reflection in optics via Fermat's principle, where the time of light travel is minimized with a free reflection point on a mirror surface. For light passing from point A to point B via reflection on the line y=0, the extremal path satisfies \sin i = \sin r, with the transversality condition at the reflection point enforcing equal angles of incidence and reflection. For isoperimetric problems, where the extremal must satisfy an additional integral constraint \int_a^b M(x, y, y') \, dx = C alongside minimizing J, variations incorporate Lagrange multipliers to enforce the constraint. The augmented functional becomes J + \lambda \left( \int_a^b M \, dx - C \right), leading to modified Euler-Lagrange equations \frac{\partial (L + \lambda M)}{\partial y} - \frac{d}{dx} \frac{\partial (L + \lambda M)}{\partial y'} = 0, with \lambda determined by the constraint. This method, introduced by Lagrange, unifies constrained optimization in the calculus of variations. Corner conditions apply to piecewise smooth extremals, where the path has discontinuities in the derivative at junction points x = c. For the extremal to be optimal, the Weierstrass-Erdmann conditions require continuity of the partial derivative \frac{\partial L}{\partial y'}(c^-, y(c), y'(c^-)) = \frac{\partial L}{\partial y'}(c^+, y(c), y'(c^+)) and of the expression L - y' \frac{\partial L}{\partial y'}, ensuring the first variation vanishes across the corner without introducing boundary terms. These conditions, formulated by Weierstrass and Erdmann, allow analysis of broken extremals in problems like minimal surfaces or geodesics on polyhedra. Higher-order variations provide tools for global analysis of functionals by expanding J[y + \epsilon h] in a Taylor series around the extremal y: J[y + \epsilon h] = J + \epsilon \delta J[y; h] + \frac{\epsilon^2}{2} \delta^2 J[y; h] + \cdots, where higher terms beyond the second variation assess convergence or stability. These expansions, analogous to finite-dimensional Taylor series, are used to study the behavior near local extrema and inform sufficient conditions for minima.

Applications

Optics and Geometrical Problems

In optics, the calculus of variations provides a mathematical framework for deriving the paths of light rays based on Fermat's principle, which posits that light travels between two points along the path that minimizes the time of travel. This principle can be expressed variationally by minimizing the optical path length functional \int n \, ds, where n is the refractive index along the ray and ds is the differential arc length, equivalent to (1/c) \int n \, ds for the time integral with c the speed of light in vacuum. For a ray in two dimensions parametrized as y = y(x) in an inhomogeneous medium where n = n(y), the functional becomes \int_{x_1}^{x_2} n(y) \sqrt{1 + (y')^2} \, dx. Applying the Euler-Lagrange equation \frac{d}{dx} \left( \frac{\partial L}{\partial y'} \right) = \frac{\partial L}{\partial y} with L = n(y) \sqrt{1 + (y')^2} yields the ray equation \frac{d}{dx} \left( n(y) \frac{y'}{\sqrt{1 + (y')^2}} \right) = n'(y) \sqrt{1 + (y')^2}, describing curved paths in graded-index media. A key derivation using this approach is Snell's law of refraction at an interface between two media. Consider light passing from a medium with refractive index n_1 (for y > 0) to n_2 (for y < 0), crossing the interface at some point (x, 0) along a path from (0, a) to (d, -b). The total optical path length is \int_0^x n_1 \sqrt{1 + (y_1')^2} \, dx' + \int_x^d n_2 \sqrt{1 + (y_2')^2} \, dx', and minimizing with respect to the crossing point x (using the condition for stationary paths at discontinuities, akin to the Weierstrass-Erdmann corner condition) gives n_1 \sin \theta_1 = n_2 \sin \theta_2, where \theta_1 and \theta_2 are the angles of incidence and refraction with the normal to the interface. This result follows directly from the variational principle without assuming wave optics, highlighting the boundary conditions in variational problems for refraction. For constant n in each layer, Beltrami's identity simplifies the paths to straight lines on either side. In three dimensions, the variational formulation generalizes to rays in inhomogeneous media through the eikonal equation. The minimal time surface T(\mathbf{r}) from a source satisfies the functional \int n \, ds = \min, leading to the first-order partial differential equation |\nabla T| = n (up to the constant $1/c), where T is the eikonal or travel time function. Solutions to this Hamilton-Jacobi equation describe ray trajectories as orthogonal curves to the level sets of T, with the refractive index n(\mathbf{r}) governing bending in stratified or varying media. This equation underpins geometrical optics and can be solved numerically or asymptotically for complex index profiles. Geometrical optics problems, such as refraction at a planar interface, further illustrate variational boundary conditions. For a ray refracting from medium 1 to 2 at y=0, the fixed endpoints impose natural boundary conditions on the variations, ensuring continuity of the ray but a jump in the derivative governed by Snell's law, as derived variationally. This setup extends to curved interfaces by incorporating appropriate transversality conditions at the boundary. Beyond ray optics, the calculus of variations addresses geometrical problems like Plateau's soap film problem, which seeks surfaces of minimal area spanning a given boundary curve. The area functional for a parametric surface \mathbf{r}(u,v) is \iint_D \sqrt{EG - F^2} \, du \, dv, where E = \mathbf{r}_u \cdot \mathbf{r}_u, G = \mathbf{r}_v \cdot \mathbf{r}_v, and F = \mathbf{r}_u \cdot \mathbf{r}_v are the coefficients of the first fundamental form. Stationary surfaces satisfy the Euler-Lagrange equations for this functional, yielding minimal surfaces such as catenoids or helicoids for specific boundaries, mimicking the equilibrium shapes of soap films under surface tension. Solving this nonlinear problem requires advanced techniques, but it exemplifies area-minimizing geometries in variational calculus.

Mechanics and Physics

The calculus of variations plays a foundational role in classical mechanics through the principle of least action, which posits that the path taken by a physical system between two points in configuration space extremizes the action functional S = \int_{t_1}^{t_2} L(q, \dot{q}, t) \, dt, where L is the Lagrangian, typically given by the difference between kinetic energy T and potential energy V, so L = T - V. Applying the Euler-Lagrange equation to this functional yields the equations of motion; for a particle in a potential, it produces Newton's second law in the form m \ddot{q} = -\partial V / \partial q. This variational approach culminates in Hamilton's principle, which states that the actual trajectory of the system renders the variation of the action zero, \delta \int_{t_1}^{t_2} L \, dt = 0, for fixed endpoints in time./06:_Lagrangian_Dynamics/6.03:_Lagrange_Equations_from_dAlemberts_Principle) The principle extends naturally to systems with constraints, such as holonomic constraints expressible as f(q, t) = 0. Here, D'Alembert's principle of virtual work is incorporated via Lagrange multipliers, modifying the Euler-Lagrange equations to \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}_i} \right) - \frac{\partial L}{\partial q_i} = \sum_k \lambda_k \frac{\partial f_k}{\partial q_i}, where the multipliers \lambda_k account for constraint forces. A classic example is the simple pendulum, where the Lagrangian is L = \frac{1}{2} m l^2 \dot{\theta}^2 - m g l (1 - \cos \theta), with \theta as the angle from the vertical, m the mass, l the length, and g gravity. The Euler-Lagrange equation then derives the nonlinear equation of motion \ddot{\theta} + \frac{g}{l} \sin \theta = 0. In relativistic mechanics, the calculus of variations extends to spacetime paths via the proper time functional, which for timelike curves is \tau = \int \sqrt{-ds^2}, and geodesics are the curves that maximize this proper time between events, analogous to the action principle but yielding the geodesic equation through the Euler-Lagrange formalism. This framework also underpins field theories in multivariable settings, where action integrals over fields lead to equations like those in electromagnetism.

Other Scientific Fields

The calculus of variations finds significant application in economics, particularly in optimal control problems for economic growth models. A seminal example is the Ramsey problem, formulated in 1928, which seeks to maximize intertemporal utility by minimizing a functional of the form \int_0^\infty u(c(t)) e^{-\rho t} \, dt subject to the capital accumulation constraint \dot{k}(t) = f(k(t)) - c(t), where c(t) is consumption, k(t) is capital, f is the production function, and \rho is the discount rate. This infinite-horizon variational problem, solved using Euler-Lagrange equations or Pontryagin's maximum principle, determines optimal saving and consumption paths to achieve Pareto-efficient allocations in neoclassical growth theory. In biology, the calculus of variations is employed to model shape optimization in evolutionary processes, such as minimizing bending energy for cell membranes. The Helfrich model, introduced in 1973, describes the elastic energy of lipid bilayers through a functional E = \int \left( \frac{\kappa}{2} (2H - C_0)^2 + \bar{\kappa} K \right) dA + \sigma \int dA, where H and K are mean and Gaussian curvatures, \kappa and \bar{\kappa} are bending rigidities, C_0 is spontaneous curvature, and \sigma is surface tension; minimizing this under volume or area constraints yields equilibrium shapes like spheres or bilayers observed in vesicles and red blood cells. This variational approach provides insights into membrane stability and morphological transitions driven by biophysical forces. Engineering applications leverage the calculus of variations for structural optimization, notably in minimizing compliance for truss designs to enhance load-bearing efficiency. In minimum compliance problems, the objective is to minimize the strain energy functional \int_V \sigma : \epsilon \, dV (or equivalently, compliance c = \mathbf{f}^T \mathbf{u}) subject to a volume constraint on the material distribution, often formulated within topology optimization frameworks using density-based methods or ground structures. Variational principles, combined with evolutionary algorithms, enable the derivation of optimal topologies that reduce weight while maintaining stiffness, as demonstrated in form-finding for elastic structures under static loads. In image processing, the total variation (TV) model addresses denoising by minimizing a functional that balances fidelity to the observed image with smoothness. The Rudin-Osher-Fatemi (ROF) formulation, proposed in 1992, solves \min_u J = \int_\Omega |\nabla u| \, dx + \frac{\lambda}{2} \int_\Omega (u - f)^2 \, dx, where u is the denoised image, f is the noisy input, \Omega is the domain, and \lambda > 0 is a regularization parameter; the Euler-Lagrange equation for this variational problem preserves edges while removing noise through total variation regularization. This approach has become foundational for inverse problems in computer vision, influencing extensions like anisotropic TV and higher-order models. Finally, in machine learning, variational methods inspired by the calculus of variations appear in probabilistic models such as variational autoencoders (VAEs), which minimize a functional involving KL divergence to approximate posterior distributions. The seminal 2013 framework encodes data into latent spaces by optimizing the evidence lower bound (ELBO), effectively minimizing \mathbb{E}_{q(z|x)}[\log p(x|z)] - D_{KL}(q(z|x) || p(z)), enabling generative modeling with structured latent representations.

References

  1. [1]
    [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 ...
  2. [2]
    [PDF] A Brief Survey of the Calculus of Variations - arXiv
    Abstract. In this paper, we trace the development of the theory of the calculus of variations. From its roots in the work of Greek thinkers and continuing ...
  3. [3]
    A History of the Calculus of Variations from the 17th through the 19th ...
    The calculus of variations is a subject whose beginning can be precisely dated. It might be said to begin at the moment that Euler coined the name calculus of ...
  4. [4]
    Calculus of Variations - Richard Fitzpatrick
    It is a well-known fact, first enunciated by Archimedes, that the shortest distance between two points in a plane is a straight-line.
  5. [5]
    2. The Calculus of Variations - Galileo and Einstein
    This method of solving the problem is called the calculus of variations: in ordinary calculus, we make an infinitesimal change in a variable, and compute the ...
  6. [6]
    [PDF] APPLICATIONS OF THE CALCULUS OF VARIATIONS
    General equations which describe optimum flight trajectories in a plane for an assumed model aircraft are obtained in this paper by the use of the methods of ...
  7. [7]
    [PDF] The Calculus of Variations - College of Science and Engineering
    Jan 7, 2022 · The history of the calculus of variations is tightly interwoven with the history of math- ematics, [27]. The field has drawn the attention ...
  8. [8]
    [PDF] Calculus of Variation
    The calculus of variations is a modern treatment of its elements, with physical applications, and is easily understandable.
  9. [9]
    [PDF] Calculus of Variations - Mathematics
    These lecture notes are written for the Mastermath course “Calculus of Variations” at Radboud. University in Spring 2023. Calculus of variations is an active ...
  10. [10]
    [PDF] Lecture notes for Calculus of Variations
    Aug 8, 2023 · Calculus of variations studies energy minimisation problems and properties of the minimizers, linking it to the Euler-Lagrange equation.
  11. [11]
    [PDF] Calculus of Variations - Physics Courses
    Oct 1, 2009 · So a curve is just a function defined on some interval, and a functional is a “function of a function.” For example, let y(x) be a real valued ...
  12. [12]
    [PDF] Calculus of variations - Aerostudents
    The term εη(x) is known as the variation. It is often also written as δy∗(x), with δ being the variational operator. We then also write I[εη] = I[δy∗] = δI.
  13. [13]
    [PDF] Calculus of Variations - IIST
    ǫ η(x) in y(x) as a function is called the variation of y and is denoted by δy. That is δy = ǫ η(x). Similarly we have δy/ = ǫ η/(x). In F = F(x, y, y/) for ...
  14. [14]
    [PDF] Calculus of variations - Assets - Cambridge University Press
    By taking δy(x) = εη(x) to be zero outside this interval, and of one sign within it, we obtain a non-zero δJ – in contradiction to stationarity. In making ...
  15. [15]
    [PDF] The Gâteaux Variation of a Functional 1. Let X be a ... - UNL math
    In the calculus of variations, one seeks the local minimizers of a functional J among its extremals. You find those extremals by solving (6) for x. 11. Here is ...
  16. [16]
    [PDF] II. Variation of functionals
    Thus, the Gateaux variation extends the concept of the directional derivative concept of finite multi-variable calculus to infinite dimensional vector spaces, ...
  17. [17]
    [PDF] CALCULUS OF VARIATIONS - EPFL
    As is to be expected from its definition, the. Gâteaux derivative is a linear operation on the functional J (by the linearity of the ordinary derivative): δ(J1 ...
  18. [18]
    [PDF] Waves and Imaging, Calculus of Variations, Functional Derivatives
    Functional calculus is used to formulate linearized forward models for imaging, as well as higher-order terms in Born series. It is also useful for finding ...<|control11|><|separator|>
  19. [19]
    2.5 Variational problems with constraints - Daniel Liberzon
    In this section we generalize that result to situations where equality constraints are imposed on the admissible curves.
  20. [20]
    [PDF] Constrained VARIATIONAL PROBLEMS - UCI Mathematics
    Jun 3, 2024 · Constrained variational problems include constrained optimization, integral constraints, and pointwise constraints.
  21. [21]
    [PDF] Constrainted variational problems
    Feb 21, 2019 · Constrained variational problems, like the isoperimetric problem, minimize one integral functional while another integral functional is fixed.
  22. [22]
    Why You Don't Believe Fermat's Principle - 21st Century
    In 1657, in a letter to Cureau de la Chambre, Fermat stated for the first time his idea that the law of refraction might be deduced from a minimum principle, ...
  23. [23]
    Brachistochrone problem - MacTutor History of Mathematics
    Five solutions were obtained, Newton, Jacob Bernoulli, Leibniz and de L'Hôpital solving the problem in addition to Johann Bernoulli. Now Johann Bernoulli and ...
  24. [24]
    "Methodus inveniendi lineas curvas maximi minimive proprietate ...
    Sep 25, 2018 · This work is concerned with the calculus of variations. Euler's main contribution to this subject was to change it from a discussion of ...
  25. [25]
    [PDF] Introduction to the Modern Calculus of Variations
    can be seen as the time of birth of the Calculus of Variations (the name, however, is from. Leonhard Euler's 1766 treatise Elementa calculi variationum).
  26. [26]
    Intended and Unintended Mathematics: The Case of the Lagrange ...
    Oct 24, 2019 · ... changes his notation by using lower case delta 'ó' to denote ... Lagrange, J.L. (1788). Mécanique Analytique. (1sted.). Paris:La Veuve ...
  27. [27]
    Beltrami Identity -- from Wolfram MathWorld
    An identity in calculus of variations discovered in 1868 by Beltrami. The Euler-Lagrange differential equation is ...Missing: 19th | Show results with:19th
  28. [28]
    [PDF] Calculus of Variations: The Direct Approach
    Jun 7, 2010 · Dirichlet was mistaken in assuming a minimizer must exist. Weierstrass found the flaw and Hilbert finally proved the principle rigorously (but ...
  29. [29]
    [PDF] The Lavrentiev phenomenon in the Calculus of Variations - cvgmt
    A measure of the importance of the choice of X is given by the Lavrentiev phenomenon (1926, [34]): a Lagrangian L exhibits the Lavrentiev phenomenon if the ...
  30. [30]
    Calculus Of Variations And Partial Differential Equations Of The First ...
    He came from arespected Greek family which had lived in Constantinople since the beginning oi the nineteenth century.Missing: holonomic | Show results with:holonomic
  31. [31]
    Optimal Control Theory - jstor
    Introduction. The development of the mathematical theory of optimal control began in the early to mid 1950's, partially in response to problems in various ...
  32. [32]
    A Brief History of Finite Element Method and Its Applications to ...
    Nov 14, 2022 · The development of the finite element method is traced, from its deepest roots, reaching back to the birth of calculus of variations in the 17th century.
  33. [33]
    [PDF] CALCULUS OF VARIATIONS
    Until now, in studying extrema of functionals, we have only considered a ... first variation), <p2[/i] is a quad ratic functional, and e->0 as ||/;|| 0 ...<|separator|>
  34. [34]
    [PDF] cvoc.pdf - Daniel Liberzon
    Aug 9, 2011 · 1.3.3 Second variation and second-order conditions . ... Lectures on the Calculus of Variations and Optimal Control Theory.
  35. [35]
    The Calculus of Variations and Optimal Control
    The second method involves the employment of sufficient conditions. ... HESTENES, M. R., Calculus of Variations and Optimal Control Theory, John Wiley and.
  36. [36]
    Direct Methods in the Calculus of Variations - SpringerLink
    Direct Methods in the Calculus of Variations ; Authors: Bernard Dacorogna ; Softcover Book USD 199.99. Price excludes VAT (USA) ; Hardcover Book USD 199.99.
  37. [37]
    [PDF] The Calculus of Variations - UC Davis Math
    The calculus of variations studies the extreme and critical points of functions. It has its roots in many areas, from geometry to optimization to mechanics, ...
  38. [38]
    Euler-Lagrange Equation - Richard Fitzpatrick
    This condition is known as the Euler-Lagrange equation. is the equation of a straight-line. Thus, the shortest distance between two fixed points in a plane is ...
  39. [39]
    [PDF] 7.2 Calculus of Variations - MIT Mathematics
    This is the Euler-Lagrange equation ATCA = f, or −∇ · c∇u = f. For constant c it is Poisson. If the y variable is removed, we are back to a one-dimensional rod.
  40. [40]
    Calculus of variations. Functionals. Euler s equation. Beltrami ...
    For this situation it is the equivalent of Euler's equation and greatly reduces the work involved. Derivation of the Beltrami Identity.<|control11|><|separator|>
  41. [41]
    [PDF] CDS 112: Winter 2014/2015
    First, let's derive the cost function which will govern the Calculus of Variations problem ... The Beltrami Identity is derived as follows. d dx. L(y, y. 0. ) ...
  42. [42]
    Calculus Of Variations : I.M. Gelfand; S.V. Fomin - Internet Archive
    Feb 8, 2022 · This book is a modern introduction to the calculus of variations and certain of its ramifications, and I trust that its fresh and lively point of view will ...Missing: seminal | Show results with:seminal
  43. [43]
    Erläuterungen zu den Anfangsgründen der Variationsrechnung
    ... Anfangsgründen der Variationsrechnung ... du Bois-Reymond, P. Erläuterungen zu den Anfangsgründen der Variationsrechnung. Math. Ann. 15, 283–314 (1879).
  44. [44]
    [PDF] Calculus of variations
    The reason we had to use the du Bois-. Reymond Lemma in the proof of Theorem 3.1 is that we wanted to derive the Euler–Lagrange equation under minimal ...
  45. [45]
    [PDF] A Bird's Eye View on Calculus of Variations - UTK Math
    At face value, Calculus of Variations (CV) is about finding minima or maxima of certain expressions; in contradistinction to single- and multi-variable ...
  46. [46]
    [PDF] 1.3 The Lemma of DuBois-Reymond
    Figure 1.5: For the lemma of DuBois-Reymond, we mollify a piecewise smooth function. Another proof of the lemma of DuBois-Raymond. Again the authors of BGH ...Missing: variations statement
  47. [47]
    [PDF] Fundamental Lemmas of Calculus of Variations
    For the sake of completeness, let us look at a proof of this lemma. It will be proved by contradiction—a legitimate method of proving things. Incidentally, ...Missing: Reymond | Show results with:Reymond
  48. [48]
    [2404.02901] The Lavrentiev phenomenon - arXiv
    Apr 3, 2024 · Our aim here is to introduce the reader to the calculus of variations, to illustrate the Lavrentiev phenomenon with the simplest known example.
  49. [49]
    [PDF] Nonoccurrence of the Lavrentiev phenomenon for nonconvex ...
    The Lavrentiev phenomenon in the calculus of variations was discovered in 1926 by M. ... Mania [11] simplified the original example of Lavrentiev. Ball and Mizel.
  50. [50]
    The Lavrentiev Phenomenon: The American Mathematical Monthly
    The Lavrentiev phenomenon is when numerical methods cannot approximate the infimum of energy in calculus of variations, despite the existence of a solution.
  51. [51]
    A New Example of the Lavrentiev Phenomenon - SIAM.org
    The Lavrentiev phenomenon in the calculus of variations is viewed and handled as a value Hadamard illposedness problem. Regularization is obtained by a ...
  52. [52]
    [PDF] Interpretation of the Lavrentiev Phenomenon by Relaxation - CORE
    multidimensional problems permits a clear discussion of the Lavrentiev phenomenon for general integral functionals of the calculus of variations. Page 4 ...<|control11|><|separator|>
  53. [53]
    The Lavrentiev phenomenon in calculus of variations with ...
    Feb 22, 2024 · In this article we study convex non-autonomous variational problems with differential forms and corresponding function spaces.
  54. [54]
    [PDF] Euler equations for multiple integrals - Math
    Jan 22, 2013 · To derive the Euler equation, we consider the variation δu of the minimizer u and the difference δI = I(u + δu) - I(u).Missing: theorem | Show results with:theorem
  55. [55]
    [PDF] From Euler, Ritz, and Galerkin to Modern Computing
    Revival of Dirichlet's Principle by Hilbert. Precisely half a century after Rie- mann's thesis, D.Hilbert undertook the task of putting Dirichlet's principle ...<|separator|>
  56. [56]
    The Dirichlet problem for the biharmonic equation in a Lipschitz ...
    In this paper we study and give optimal estimates for the Dirichlet problem for the biharmonic operator Δ 2 , on an arbitrary bounded Lipschitz domain D in R n ...
  57. [57]
    [PDF] 6 Eigenvalues of the Laplacian
    This theorem is known as the minimax principle. It will allow us to prove a relationship between eigenvalues of sets contained within larger sets. In particular ...
  58. [58]
    [1306.0392] Faber-Krahn inequalities in sharp quantitative form - arXiv
    Jun 3, 2013 · The classical Faber-Krahn inequality asserts that balls (uniquely) minimize the first eigenvalue of the Dirichlet-Laplacian among sets with given volume.
  59. [59]
    Rayleigh–Ritz–Galerkin Methods for Multidimensional Problems
    This paper is concerned with the application of the Ritz–Galerkin method to the numerical solution of singular boundary value problems of the type arising when ...
  60. [60]
  61. [61]
    [PDF] classic theory of calculus of variation
    Jun 4, 2025 · Strong minimum: perturbation lies in a larger -ball of a weaker norm. • Weak minimum: perturbation lies in a smaller -ball of a stronger norm.Missing: Karl 1870s
  62. [62]
    on the sufficient conditions in the calculus of vaeiations.
    We shall then show that Jacobi's condition (together with Lagrange's and Legendre's) is a suf- ficient condition for weak minima, the Weierstrass condition.
  63. [63]
    [PDF] CALCULUS OF VARIATIONS MA 4311 LECTURE NOTES - Faculty
    Jul 9, 2002 · Bliss - Calculus of Variations, Carus monograph ... Since the fundamental lemma allowed for this, then the proof of the integrated form is.
  64. [64]
    [PDF] The Problem of Lagrange in the Calculus of Variations
    Feb 23, 2005 · It is in fact a problem of Lagrange with second end-point variable. The so-called isoperimetric problems form a very large class, and all of.
  65. [65]
    [PDF] The Calculus of Variations - College of Science and Engineering
    Mar 21, 2021 · The history of the calculus of variations is tightly interwoven with the history of math- ematics, [12]. The field has drawn the attention of a ...
  66. [66]
    [PDF] Calculus of Variations - Physics Courses
    EXAMPLES FROM THE CALCULUS OF VARIATIONS. 7. Figure 5.4: Minimal surface ... and extract the term of order ǫ1 ǫ2 in the double Taylor expansion. One ...
  67. [67]
    [PDF] Chapter 3 The Variational Principle - Physics
    In this chapter we explore Fermat's ap- proach, and then go on to introduce techniques in variational calculus used to implement this approach, and to solve a ...Missing: 1657 | Show results with:1657
  68. [68]
    [PDF] A variational formulation of the fast marching eikonal solver
    Nov 9, 2000 · Based originally on the eikonal equation, it can be derived equally well from Fermat's principle. The new variational formulation has two ...
  69. [69]
    [PDF] What is Plateau's problem? - OSU Math
    Jun 26, 2012 · Plateau's problem is the problem of finding a minimal surface with a given boundary contour. It was first formulated by Lagrange (1760) in a ...
  70. [70]
    Principle of least action - Scholarpedia
    Jun 5, 2015 · The principle of least action finds a system's true trajectory by selecting the one with the least action, a functional of the trajectory.Statements of Hamilton and... · Euler-Lagrange Equations · Relation to Quantum...
  71. [71]
    19: The Principle of Least Action - Feynman Lectures
    The principle of least action states that the average kinetic energy minus the average potential energy is as little as possible for an object's path.
  72. [72]
    d'Alembert–Lagrange analytical dynamics for nonholonomic systems
    Mar 18, 2011 · The d'Alembert–Lagrange principle (DLP) is designed primarily for dynamical systems under ideal geometric constraints.
  73. [73]
    [PDF] a classical problem in the calculus of variations leading to an implicit ...
    Mar 23, 2010 · Abstract. In 1928, motivated by conversations with Keynes, Ramsey formulated an infinite-horizon problem in the calculus of variations.
  74. [74]
    [PDF] Calculus of Variations - Mathematics and Statistics
    Biological membranes and their shapes have attracted attention from researchers across many areas of mathematics. For example, membranes connect classical ...Missing: optimization | Show results with:optimization
  75. [75]
    [PDF] On the Structural Shape Optimization through Variational Methods ...
    We employ the variational theory of optimal control problems and evolutionary algorithms to investigate the form finding of min- imum compliance elastic ...
  76. [76]
    Nonlinear total variation based noise removal algorithms
    The algorithm minimizes image total variation with noise constraints, using a gradient-projection method, solving a partial differential equation to denoise ...Missing: paper | Show results with:paper
  77. [77]
    [1312.6114] Auto-Encoding Variational Bayes - arXiv
    Dec 20, 2013 · Authors:Diederik P Kingma, Max Welling. View a PDF of the paper titled Auto-Encoding Variational Bayes, by Diederik P Kingma and 1 other authors.