Fact-checked by Grok 2 weeks ago

Variational principle

The variational principle is a core methodology in and that determines the , motion, or of a by identifying the path or state that extremizes—typically minimizes or renders stationary—a specific functional, such as or energy functional. This approach leverages the to derive governing equations, providing an elegant framework for understanding natural phenomena by positing that systems follow "optimal" trajectories among possible alternatives. The historical development of variational principles began in the 17th century with Pierre de Fermat's principle of least time, which posits that light travels along the path requiring the minimal time between two points, laying early groundwork for variational thinking in optics. Systematic formalization emerged in the 18th century through the contributions of Leonhard Euler and Joseph-Louis Lagrange, who applied variational methods to mechanics, deriving equations of motion from extremizing integrals related to energy or work. In the 19th century, William Rowan Hamilton advanced the field with Hamilton's principle, stating that the actual path of a system makes the action integral—defined as the time integral of the Lagrangian—stationary, thus unifying and generalizing prior formulations. This progression, as chronicled in foundational texts, transformed mechanics from Newtonian vectorial methods to analytical frameworks emphasizing symmetry and conservation laws. Variational principles find broad applications across physics, underpinning derivations in (yielding Lagrange's equations for constrained systems), (via the Rayleigh-Ritz method, which provides upper bounds on ground-state energies for approximate wavefunctions), and (through formulations like the principle of least action for field equations). In modern contexts, they extend to (via the Einstein-Hilbert action), , and even nonequilibrium thermodynamics, where they facilitate modeling dissipative processes and multiscale phenomena. These principles not only simplify complex derivations but also reveal deep symmetries, influencing numerical methods like variational integrators for simulations.

Fundamentals

Definition

The variational principle, a of the , asserts that the solution to certain optimization problems in and physics is obtained by finding functions that extremize—either minimize or maximize—a functional, which is a from a space of functions to the real numbers. Functionals typically take the form of integrals over paths or fields, such as J = \int_a^b F(x, y(x), y'(x)) \, dx, where y(x) is the function being varied, y'(x) = \frac{dy}{dx}, and F is a given integrand depending on the independent variable x, the function y, and its derivative. Central to the principle are variations, defined as small perturbations \delta y around a candidate function y, which allow assessment of how the functional changes under infinitesimal deformations while respecting boundary conditions, such as fixed endpoints y(a) = A and y(b) = B. The first variation \delta J[y; \delta y] quantifies this change to first order, analogous to the differential in ordinary calculus. At an extremum, the functional is , meaning the first variation vanishes: \delta J = 0 for all admissible variations \delta y. This condition ensures that the function y is a critical point, neither increasing nor decreasing the functional value locally. Historically, such principles arose from physics problems seeking paths of least , providing an optimization framework for natural laws. The variational principle encompasses two distinct problem types: the direct problem, which involves computing the extremizing for a given functional, and the , which seeks a functional whose extremal functions satisfy a prescribed set of equations. In the direct case, one solves for the path or field that achieves the extremum; in the case, one determines whether the equations admit a variational , often via conditions like those of Helmholtz.

Motivation

The variational principle draws its intuition from familiar optimization problems in , such as finding the shortest path between two points, which in flat space is a straight line, or determining the shape that encloses the maximum area for a given perimeter, like a . These examples illustrate how often selects paths or configurations that minimize or maximize certain quantities, such as distance or area, leading to fundamental laws that govern physical systems. A key advantage of variational principles lies in their ability to provide a unified framework for deriving the across diverse physical domains, from to field theories, without directly solving complex differential equations. This approach simplifies the formulation of physical laws by extremizing functionals—mappings from functions to real numbers—that encode the system's behavior. Moreover, symmetries in these functionals naturally give rise to conservation laws, as encapsulated in , which connects continuous symmetries to conserved quantities like energy or momentum. Historically, variational methods were driven by the need to address boundary value problems that resisted direct solution via differential equations, particularly isoperimetric problems requiring the optimization of one quantity subject to a constraint on another, such as maximizing enclosed area under fixed . These challenges highlighted the limitations of traditional approaches and spurred the of techniques to handle constrained extrema. Philosophically, the variational principle appeals to the idea that operates along "efficient" paths, minimizing or effort, which echoes teleological views in early positing that physical processes follow optimal routes as if guided by purpose. This perspective, often termed the principle of natural , suggests an inherent in natural laws, where systems evolve toward states of least deviation from .

History

Origins in Physics

The variational principle traces its origins to early , where intuitive notions of nature's efficiency began to formalize paths of least or stationary quantities in natural phenomena. In 1657, proposed the principle of least time for propagation, asserting that a ray travels from one point to another along the path that minimizes the travel time, which laid the groundwork for understanding and in . This idea, inspired by earlier geometric and Hero of Alexandria's work on , marked an initial shift toward extremal principles in physical laws, though Fermat's formulation was more than rigorously derived. Building on Fermat's insight, the principle of least action emerged in the 1740s through the efforts of Pierre-Louis Moreau de Maupertuis and Leonhard Euler, extending extremal ideas from optics to mechanics. In his 1744 paper "Accord des différentes lois de la nature qui avoient jusqu'à présent paru incompatibles," presented to the Académie des sciences de Paris, Maupertuis introduced the action as the integral \int p \, dq, where p is momentum and dq is an infinitesimal displacement, positing that nature acts along paths minimizing this quantity to embody an "economy" in physical processes. Euler collaborated closely with Maupertuis, refining the principle in works like his 1744 "Methodus inveniendi lineas curvas maximi minimive proprietate gaudentes," applying it to derive laws of motion and optics while emphasizing its mechanistic universality over teleological interpretations. Their formulation sparked debates on whether the principle implied divine purpose—Maupertuis viewing it as evidence of God's frugality—or purely mechanical efficiency, contrasting with Leibnizian teleology and Newtonian mechanism. By the 1760s, advanced these ideas into a more systematic framework for mechanics, reformulating dynamics through the stationary principle of action to unify and motion. In his 1760 presentation to the , Lagrange demonstrated that mechanical systems evolve such that the action integral remains stationary, bridging variational methods with and eliminating reliance on forces in favor of energy-based functionals. This work, later expanded in his Mécanique analytique (1788), solidified the variational approach as a cornerstone of physics. In the , further advanced the field with (1834), which states that the motion of a system is such that the action—defined as the integral of the over time—is stationary, providing a unified framework that emphasized symmetries and conservation laws.

Development in Mathematics

Leonhard Euler laid the foundational work for the with his 1744 book Methodus inveniendi lineas curvas maximi minimive proprietate gaudentes, building on earlier problems like the brachistochrone posed by in 1696, which Euler solved using variational methods, and further developed the field in his 1766 work Elementa calculi variationum. Euler's approach treated these as optimization problems over functions, exemplified by his solutions to isoperimetric problems, where he sought curves enclosing maximum area for a given perimeter, introducing systematic techniques for deriving extremal paths. Joseph-Louis Lagrange advanced this framework in the late 1700s through his "delta method," introduced in letters to Euler around 1755 and formalized in his 1788 Mécanique Analytique, where he employed infinitesimal variations δy to derive necessary conditions for extrema by setting the first variation of a functional to zero. This method shifted the focus from geometric intuition to algebraic manipulation of variations, providing a more rigorous derivation of equations for extremals and influencing the field's transition to a purely mathematical discipline, albeit still inspired by physical action principles. In the , elevated the theory's rigor in his lectures of the 1870s, developing sufficient conditions for weak and strong extrema through the introduction of the excess function and the concept of fields of extremals, which addressed limitations in earlier necessary conditions by ensuring local minimality via second-order tests. further propelled the field in 1904 by tackling the Dirichlet principle in his foundational paper, restoring its validity through variational methods and posing problems on the existence and regularity of solutions to variational integrals, which highlighted the need for global theorems beyond local extremals. The 20th century brought increased mathematical rigor with theorems, as Leonida Tonelli's direct method in the 1920s relaxed smoothness assumptions on admissible functions to prove the of minimizers for certain functionals using arguments, detailed in his 1921 book Fondamenti di Calcolo delle Variazioni (Vol. 1). Charles B. Morrey extended this in the 1950s to multiple integrals, establishing for problems in Sobolev spaces via his 1952 on quasi-convexity, which ensured minimizers for non-convex integrands under growth conditions; however, these efforts also revealed gaps, such as non- of minimizers for certain functionals like the Plateau problem without additional constraints.

Mathematical Framework

Functionals

In the calculus of variations, a functional is a mapping that assigns a real number to each function in a suitable space of admissible functions, often arising as an integral that encodes an optimization objective. The most common form is the integral functional, defined for a function y: [a, b] \to \mathbb{R} as J = \int_a^b L(x, y(x), y'(x)) \, dx, where L(x, y, y') is the Lagrangian density, a given smooth function, and y' = dy/dx. This structure generalizes to higher dimensions, where the functional operates on a vector-valued function \mathbf{y}: \Omega \to \mathbb{R}^m over a domain \Omega \subset \mathbb{R}^n, taking the form J[\mathbf{y}] = \int_\Omega L(\mathbf{x}, \mathbf{y}(\mathbf{x}), \nabla \mathbf{y}(\mathbf{x})) \, d\mathbf{x}, with \nabla \mathbf{y} denoting the gradient. Integral functionals predominate in variational problems, as they aggregate contributions over a , but pointwise functionals also exist, evaluating the objective directly at specific points without , such as J = y(c) for some fixed c in the . A representative example of an integral functional is the functional, which measures the length of a y(x) from x = a to x = b: J = \int_a^b \sqrt{1 + (y'(x))^2} \, dx. This functional attains its minimum for the straight line connecting the endpoints, illustrating how variational principles seek extremal paths. For well-posedness, functionals are typically defined on appropriate function spaces that ensure the integral converges and variations are meaningful. Classical spaces include C^1[a, b], the set of continuously differentiable functions on [a, b], equipped with the \|y\|_1 = \max |y(x)| + \max |y'(x)|. In more advanced settings, Sobolev spaces W^{1,p}(\Omega) (for $1 \leq p \leq \infty) are used, comprising functions whose weak derivatives in L^p(\Omega); these spaces provide the reflexivity and compactness needed for existence theorems via the . Variations of functionals are analyzed using generalized derivatives in infinite-dimensional spaces. The Gateaux derivative at y in direction h is the linear part of the increment: \delta J(y; h) = \lim_{\epsilon \to 0} \frac{J[y + \epsilon h] - J}{\epsilon} = \int_a^b \left( L_y h + L_{y'} h' \right) dx, capturing directional sensitivity. The stronger requires uniformity over bounded directions, defined as a bounded linear DJ(y): h \mapsto \delta J(y; h) such that \lim_{\|h\| \to 0} \frac{|J[y + h] - J - DJ(y)h|}{\|h\|} = 0, ensuring the functional behaves like a differentiable map in Banach spaces. Key properties of functionals influence the existence and uniqueness of minimizers. Convexity holds if J[\lambda y + (1-\lambda) z] \leq \lambda J + (1-\lambda) J for \lambda \in [0,1], often imposed on the L with respect to y and y', guaranteeing that local minima are global. Coercivity requires J \to \infty as \|y\| \to \infty, preventing minimizing sequences from escaping to infinity and enabling compactness arguments in reflexive spaces like Sobolev spaces. Together, these ensure a minimizer exists under lower . Additionally, the Legendre transform connects variational functionals to formulations, converting the L(x, y, p) (with p = y') to the H(x, y, p) = p \cdot y' - L(x, y, y'), facilitating phase-space analysis.

Euler–Lagrange Equation

The arises as the necessary condition for a to extremize a functional defined on a of admissible curves, building on the concept of functionals as integrals over paths in the variational framework.

Derivation for the One-Variable Case

Consider a functional of the form J = \int_a^b L(x, y(x), y'(x)) \, dx, where L is the Lagrangian density, assumed twice continuously differentiable, and y'(x) = dy/dx. To identify points, the first variation must vanish: \delta J = 0. This is obtained by considering a perturbed y(x) + \epsilon \eta(x), where \epsilon is and \eta(x) is an arbitrary smooth variation vanishing at the endpoints, \eta(a) = \eta(b) = 0. The first-order change in the functional is \delta J = \left. \frac{d}{d\epsilon} J[y + \epsilon \eta] \right|_{\epsilon=0} = \int_a^b \left( \frac{\partial L}{\partial y} \eta + \frac{\partial L}{\partial y'} \eta' \right) dx = 0. 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 = - \int_a^b \frac{d}{dx} \left( \frac{\partial L}{\partial y'} \right) \eta \, dx, since the term vanishes due to \eta(a) = \eta(b) = 0. Substituting back gives \int_a^b \left[ \frac{\partial L}{\partial y} - \frac{d}{dx} \left( \frac{\partial L}{\partial y'} \right) \right] \eta(x) \, dx = 0. By the fundamental lemma of the , which states that if the of a times an arbitrary \eta vanishes, 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 for a single dependent variable, first derived by Euler in his foundational work on the .

General Forms

For functionals depending on multiple dependent variables y_i(x), i=1,\dots,n, the Lagrangian takes the form L(x, y_1, \dots, y_n, y_1', \dots, y_n'), and the stationary condition yields a system of coupled Euler–Lagrange equations: \frac{\partial L}{\partial y_i} - \frac{d}{dx} \left( \frac{\partial L}{\partial y_i'} \right) = 0, \quad i=1,\dots,n. Each equation governs one variable independently in form, though coupling arises through L. This multivariable extension was systematized by Lagrange in his , where it applies to systems with multiple . For higher-order derivatives, if L depends on y^{(k)}(x) up to the k-th derivative, the generalized becomes \frac{\partial L}{\partial y} - \sum_{m=1}^k \frac{d^m}{dx^m} \left( \frac{\partial L}{\partial y^{(m)}} \right) = 0, balancing the direct partial with higher-order total derivatives. In the presence of constraints, such as isoperimetric conditions \int_a^b G(x, y, y') \, dx = c or g(x, y) = 0, Lagrange multipliers \lambda are introduced to form an augmented \tilde{L} = L + \lambda (G - c) or \tilde{L} = L + \lambda g. The resulting s incorporate the multiplier: \frac{\partial \tilde{L}}{\partial y} - \frac{d}{dx} \left( \frac{\partial \tilde{L}}{\partial y'} \right) = 0, along with the constraint equation, yielding a system solvable for both y and \lambda. This method, originating with Lagrange, handles restricted variational problems without parameterizing the constraints explicitly.

Boundary Conditions

Boundary conditions classify variational problems based on whether endpoints are fixed or free. For fixed (essential or Dirichlet) boundaries, y(a) and y(b) are prescribed, requiring \eta(a) = \eta(b) = 0 in the variation, so the Euler–Lagrange equation alone determines the extremal. For free (natural or Neumann) boundaries, where y(b) (say) is unspecified, the variation \eta(b) need not vanish. The integration by parts then leaves a boundary term \left. \frac{\partial L}{\partial y'} \eta \right|_a^b, and for \delta J = 0 to hold for arbitrary \eta(b), the transversality () condition must apply: \frac{\partial L}{\partial y'}(b) = 0. This specifies the derivative at the free endpoint. Broken extremals occur when the optimal discontinuously joins segments satisfying different Lagrangians or constraints, with corner conditions ensuring of \frac{\partial L}{\partial y'} across the junction to minimize the total functional. These conditions were developed in the classical to handle variable endpoint problems.

Sufficient Conditions

The Euler–Lagrange equation provides a necessary condition for stationarity but not sufficiency for a minimum or maximum. To confirm a weak local minimum, the second variation \delta^2 J must be positive definite. For the one-variable case, expanding to second order in \epsilon gives \delta^2 J = \int_a^b \left( P \eta^2 + 2 Q \eta \eta' + R (\eta')^2 \right) dx, where P = \frac{\partial^2 L}{\partial y^2}, Q = \frac{\partial^2 L}{\partial y \partial y'}, and R = \frac{\partial^2 L}{\partial (y')^2}, evaluated along the extremal. For \delta^2 J > 0 for all admissible \eta \neq 0, the quadratic form must be positive definite, often checked via the Legendre condition R > 0 and the Jacobi accessory equation. The Jacobi equation, \frac{d}{dx} \left( R \frac{d \zeta}{dx} \right) - Q \frac{d \zeta}{dx} + P \zeta = 0, arises from varying the extremal itself and governs Jacobi fields \zeta(x), which are displacements to nearby extremals. A sufficient for a minimum is the absence of conjugate points—points where a nontrivial vanishes—ensuring no smaller nearby functional value. This stability analysis, involving the second variation and Jacobi fields, distinguishes true minima from inflection points in the variational landscape.

Applications in Physics

Classical Mechanics

In , the variational principle finds its primary application through , which asserts that the actual trajectory of a between two fixed times t_1 and t_2 renders the action functional stationary. The action S is given by S = \int_{t_1}^{t_2} L \, dt, where L = T - V is the , with T denoting the and V the of the . This principle, formulated by in 1834, reformulates Newtonian mechanics in terms of a variational extremum rather than force balances. Applying the to S yields the Euler-Lagrange equations, specialized here to q_i as \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}_i} \right) - \frac{\partial L}{\partial q_i} = 0, known as Lagrange's equations, which provide the for the . These equations, derived by in 1788, eliminate the need for explicit constraint forces in many cases by incorporating constraints directly into the choice of coordinates. A classic illustration of the variational approach is the brachistochrone problem, posed by in 1696, which seeks the curve connecting two points that minimizes the descent time of a particle under gravity, assuming no friction. The time functional to minimize is t = \int_{A}^{B} \frac{ds}{\sqrt{2gy}}, where ds is the arc length element and y the vertical drop, leading via the Euler-Lagrange equation to a as the solution curve—surprisingly faster than the straight line despite being longer. This problem highlighted the power of variational methods in identifying non-intuitive optima, influencing the development of the . For the simple pendulum, consisting of a mass m attached to a massless rod of length l pivoting from a fixed point, the Lagrangian in terms of the angle \theta from the vertical is L = \frac{1}{2} m l^2 \dot{\theta}^2 - m g l (1 - \cos \theta). Substituting into Lagrange's equation produces the nonlinear equation of motion \ddot{\theta} + \frac{g}{l} \sin \theta = 0, which for small angles \theta approximates with frequency \sqrt{g/l}. This derivation demonstrates how the variational principle systematically yields dynamics from energy expressions, applicable to both conservative and more complex systems. Systems often involve constraints that restrict possible motions, classified as holonomic or non-holonomic. expressible as functions of coordinates and time, such as f(q_i, t) = 0, reduce the and can be eliminated by selecting appropriate , preserving the standard form of Lagrange's equations. Non-holonomic constraints, typically involving velocities like \sum a_i \dot{q}_i + a_t = 0, cannot be integrated to position constraints and require modifications, such as Lagrange multipliers. For both types, provides a variational foundation by extending the principle of to : for a system of particles, \sum_i (\mathbf{F}_i - m_i \ddot{\mathbf{r}}_i) \cdot \delta \mathbf{r}_i = 0, where \mathbf{F}_i are applied forces and \delta \mathbf{r}_i virtual displacements consistent with constraints, effectively treating inertial forces as equilibrating elements. This principle, introduced by in 1743, bridges and in constrained systems. The variational formulation also reveals deep connections between symmetries and conservation laws through , established in 1918. If the is invariant under a continuous transformation of the coordinates q_i \to q_i + \epsilon \xi_i (with \epsilon ), then the quantity Q = \sum_i \frac{\partial L}{\partial \dot{q}_i} \xi_i - \left( \sum_i \dot{q}_i \frac{\partial L}{\partial \dot{q}_i} - L \right) \eta (for time-dependent transformations with t \to t + \epsilon \eta) is conserved along the trajectory. For instance, time-translation invariance (\eta = 1, \xi_i = 0) conserves , spatial translation invariance conserves linear momentum, and rotational invariance conserves , providing a systematic derivation of these laws from the structure of the . This theorem underscores the foundational role of symmetries in .

Field Theories

In field theories, the variational principle provides a foundational framework for deriving the for continuous distributions of , such as scalar, vector, or tensor fields defined over . The action functional for a field theory is expressed as S = \int \mathcal{L}(\phi, \partial_\mu \phi) \, d^4 x, where \phi represents the field (or fields), \partial_\mu denotes derivatives, and \mathcal{L} is the depending on the field and its first derivatives. The principle of stationary action, requiring \delta S = 0 under admissible variations \delta \phi that vanish at boundaries, yields the Euler–Lagrange equations for fields: \partial_\mu \left( \frac{\partial \mathcal{L}}{\partial (\partial_\mu \phi)} \right) - \frac{\partial \mathcal{L}}{\partial \phi} = 0. This generalization of the particle Euler–Lagrange equation accommodates both non-relativistic and relativistic contexts, enabling the unification of dynamics across extended systems. A canonical example is the real scalar field theory, where the Lagrangian density takes the form \mathcal{L} = \frac{1}{2} \partial_\mu \phi \partial^\mu \phi - V(\phi), with V(\phi) = \frac{1}{2} m^2 \phi^2 for a free massive field. Applying the Euler–Lagrange equation produces the Klein–Gordon equation, (\square + m^2) \phi = 0, which describes the propagation of a relativistic scalar field with mass m. Similarly, for classical electrodynamics, the Lagrangian density is \mathcal{L} = -\frac{1}{4} F_{\mu\nu} F^{\mu\nu} - A_\mu J^\mu, where F_{\mu\nu} = \partial_\mu A_\nu - \partial_\nu A_\mu is the electromagnetic field strength tensor and J^\mu the current. Variation with respect to the vector potential A_\mu yields Maxwell's equations in covariant form: \partial_\mu F^{\mu\nu} = J^\nu (inhomogeneous) and \partial_\lambda F_{\mu\nu} + \partial_\mu F_{\nu\lambda} + \partial_\nu F_{\lambda\mu} = 0 (homogeneous). These derivations highlight how the variational approach systematically recovers the fundamental field equations from a single action principle. Gauge invariance plays a central role in modern field theories, ensuring the physical content remains unchanged under local transformations of the fields. In non-Abelian gauge theories, such as Yang–Mills theory, the Lagrangian \mathcal{L} = -\frac{1}{4} F^a_{\mu\nu} F^{a\mu\nu} (with a labeling the gauge group generators) is constructed to be invariant under local gauge transformations \phi \to U(x) \phi, where U(x) is a spacetime-dependent group element. The field strength generalizes to F^a_{\mu\nu} = \partial_\mu A^a_\nu - \partial_\nu A^a_\mu + g f^{abc} A^b_\mu A^c_\nu, incorporating non-linear self-interactions. Matter fields couple minimally via the covariant derivative D_\mu = \partial_\mu - i g A^a_\mu T^a, preserving gauge invariance in the full action. The Euler–Lagrange equations then produce the Yang–Mills equations \partial_\mu F^{a\mu\nu} + g f^{abc} A^b_\mu F^{c\mu\nu} = J^{a\nu}, describing the dynamics of gluons in quantum chromodynamics or W/Z bosons in electroweak theory. This structure, introduced in the seminal work on isotopic gauge invariance, underpins the Standard Model of particle physics. In general relativity, the variational principle achieves a geometric unification of gravity with spacetime structure. David Hilbert proposed in 1915 an action S = \frac{1}{16\pi G} \int R \sqrt{-g} \, d^4 x + S_m, where R is the Ricci scalar, g the metric determinant, and S_m the matter action. Varying with respect to the metric g_{\mu\nu} (while treating it as the dynamical field) leads to the Einstein field equations G_{\mu\nu} = 8\pi G T_{\mu\nu}, with G_{\mu\nu} the Einstein tensor and T_{\mu\nu} the stress–energy tensor derived from S_m. This formulation reveals gravity as the curvature response to energy-momentum, derived variationally without presupposing the field equations. Hilbert's approach, presented alongside electromagnetic terms in his unified theory, established the modern basis for gravitational field dynamics.

Applications in Mathematics

Geometry

In differential geometry, variational principles characterize geodesics as the shortest paths on a , obtained by minimizing the length functional or, equivalently, its square, the functional E(\gamma) = \frac{1}{2} \int_a^b g_{ij} \dot{x}^i \dot{x}^j \, dt, where g_{ij} denotes the components and the curve \gamma: [a,b] \to M is piecewise smooth. The critical points of this functional satisfy the geodesic equation, derived from the condition that the first variation of the vanishes for all admissible variations, leading to the covariant acceleration \nabla_{\dot{\gamma}} \dot{\gamma} = 0. Minimal surfaces, which locally minimize area among nearby surfaces, arise as critical points of the area functional A(\Sigma) = \iint_D \sqrt{EG - F^2} \, du \, dv, where E, F, G are the coefficients of the induced on parameters (u,v) for the surface parametrization over domain D. seeks a spanning a given closed \Gamma in \mathbb{R}^3, formalized as minimizing the area subject to the boundary condition. This problem was solved independently in 1930 by Jesse Douglas, who established existence for arbitrary Jordan curves using a direct method in the , and by Tibor Radó, who employed conformal mappings and approximation by polyhedral surfaces to prove the result for smooth boundaries. Soap films provide physical realizations of minimal surfaces, where surface tension minimizes the film's area for a fixed enclosed volume, aligning with the variational principle of zero mean curvature. In capillarity models, soap films are approximated as thin regions of small volume minimizing a perimeter-type functional, converging in the limit to classical minimal surfaces as the thickness approaches zero. Bernstein's theorem states that any complete minimal graph over \mathbb{R}^2 in \mathbb{R}^3 must be a plane, implying that non-flat entire solutions do not exist in this setting; this result, proved using elliptic regularity and growth estimates, highlights the rigidity of minimal hypersurfaces in low dimensions. In , curvature admits a variational characterization through the second variation of the energy functional along , where the index form measures stability and involves the via Jacobi fields—solutions to the J'' + R(J, \dot{\gamma})\dot{\gamma} = 0 along a \gamma—with conjugate points (where non-trivial Jacobi fields vanish) signaling positive curvature and limiting completeness.

Optimization

In optimization theory, variational principles provide a framework for finding extrema of functionals, which are mappings from infinite-dimensional spaces of functions to the real numbers. These principles extend the classical calculus to address problems where the objective is to minimize or maximize integrals representing costs, energies, or other quantities subject to constraints. A key application arises in optimal control, where the goal is to determine control functions that steer a dynamical system from an initial state to a desired terminal state while minimizing a performance criterion. In optimal control problems, the cost functional is typically formulated as J[x, u] = \int_{t_0}^{t_f} L(x(t), u(t), t) \, dt + \Phi(x(t_f)), where x(t) is the state trajectory satisfying the \dot{x}(t) = f(x(t), u(t), t), u(t) is the control input, L is the running cost (often quadratic in state and control to penalize deviations and efforts), and \Phi is the terminal cost. serves as a necessary condition for optimality, stating that an optimal control maximizes the H(x, u, \lambda, t) = L(x, u, t) + \lambda^T f(x, u, t) pointwise, where \lambda(t) is the costate satisfying \dot{\lambda} = -\frac{\partial H}{\partial x}. This principle can be viewed as the dual of direct variational methods, transforming the infinite-dimensional optimization into a two-point solvable via indirect methods like . The original formulation addressed fixed-time problems with bounded controls, enabling solutions to "bang-bang" controls in many engineering applications, such as rocket . When inequality constraints are present, such as state or control bounds g(x(t), u(t), t) \leq 0, the incorporates with non-negativity requirements. Complementary slackness conditions ensure that the multiplier \mu(t) \geq 0 is zero when the constraint is inactive (g < 0), and \mu g = 0 holds, analogous to finite-dimensional Karush-Kuhn-Tucker conditions. These extend the Euler-Lagrange equations to include jumps or arcs where constraints bind, as seen in problems like obstacle avoidance in path planning. This framework maintains the variational structure by augmenting the with inequality terms, ensuring stationarity only along feasible paths. A foundational example in variational optimization is the Dirichlet principle, which posits that solutions to \Delta u = 0 in a \Omega with prescribed values u|_{\partial \Omega} = \phi are minimizers of the functional E = \int_{\Omega} |\nabla u|^2 \, dV over admissible functions. Harmonic functions thus achieve the minimal energy configuration, interpreting physical equilibria like electrostatic potentials or steady-state heat distributions as global minimizers. Initially proposed by Dirichlet in the context of and later justified by Riemann for value problems, the principle relies on the quadratic nature of the functional, ensuring convexity and existence via the direct method in the . For practical approximations, the discretizes variational problems by expanding the trial solution as u_n(x) = \sum_{k=1}^n c_k \psi_k(x), where \{\psi_k\} are basis functions satisfying conditions, and minimizing the functional over the coefficients c_k. This reduces the infinite-dimensional problem to a finite system of algebraic equations from setting partial derivatives to zero, converging to the exact solution as n \to \infty under suitable completeness assumptions. Developed by Walther for value problems in , the approach underpins spectral methods and finite element precursors, offering efficient numerical solutions for elliptic variational inequalities.

Extensions

Quantum Variational Methods

In , the variational principle provides an upper bound on the ground state of a system through the Rayleigh-Ritz method. For a operator \hat{H}, the states that for any normalized wave function \psi, the satisfies E_0 \leq \langle \psi | \hat{H} | \psi \rangle, where E_0 is the ground state . This arises from the positive semi-definiteness of \hat{H} - E_0 and allows approximate solutions by minimizing the functional over a parameterized family of functions. An early application to the used a function incorporating electron correlation, yielding an accurate to within 1.7% of the . Feynman's extends the variational principle to a sum over all possible paths in configuration space, weighted by the e^{i S / \hbar}, where S is the classical . This approach, introduced in , reformulates the as the continuum limit of a , providing a probabilistic of quantum and enabling variational approximations by restricting to dominant path ensembles. In the classical limit \hbar \to 0, it reduces to the of . In quantum field theory, the generating functional Z[J] = \int \mathcal{D}\phi \, e^{i (S[\phi] + \int J \phi)} encodes all correlation functions via functional derivatives with respect to the source J. Introduced in Schwinger's functional formalism in 1951, it generalizes the path integral to fields and allows variational minimization of an effective action \Gamma[\phi], obtained as the Legendre transform of W[J] = -i \ln Z[J], to approximate the ground state or vacuum energy. This framework facilitates perturbative expansions and non-perturbative approximations in interacting theories. Key applications include the Hartree-Fock approximation, which variationally optimizes a trial for many-electron systems, leading to self-consistent single-particle orbitals. Developed by Fock in 1930 as an antisymmetrized extension of Hartree's mean-field method, it captures exchange effects and provides a foundation for post-Hartree-Fock corrections, with typical errors of a few percent in binding energies for light atoms. Another is variational , which stochastically evaluates the energy expectation value for complex trial functions, such as those with Jastrow correlations for quantum liquids. Pioneered by Kalos in 1964 for the , it achieves near-exact energies for few-body systems by sampling the probability density |\psi|^2. A modern extension is the (VQE), introduced in 2014, which applies variational principles to . VQE uses a hybrid quantum-classical algorithm to approximate the of a by optimizing parameterized quantum circuits (ansatze) to minimize the energy expectation value measured on near-term quantum hardware. This approach is particularly suited for simulating and materials problems beyond classical capabilities.

Computational Approaches

Computational approaches to variational principles leverage numerical techniques to approximate solutions to complex optimization problems arising from functionals, particularly in solving partial differential equations (PDEs) and probabilistic modeling. One foundational method is the (FEM), which discretizes continuous domains into finite elements and employs variational formulations to approximate solutions. In the Galerkin projection approach, the weak form of a PDE is projected onto a finite-dimensional subspace of trial functions, minimizing the residual in a variational sense. This method, rooted in the works of Galerkin and further developed for FEM, enables efficient numerical solutions for boundary value problems by transforming the variational principle into a system of algebraic equations solvable via matrix assembly. In , variational principles underpin variational inference (), a scalable technique for approximating posterior distributions in Bayesian models by maximizing the (ELBO). The ELBO, defined as the expected log-likelihood minus the KL between the approximate and true posterior, provides a tractable lower bound on the model , allowing optimization via gradient-based methods. Post-2010 developments, particularly in , have popularized VI for high-dimensional data; for instance, stochastic VI uses mini-batches to scale inference for large datasets. This approach addresses the intractability of exact while maintaining probabilistic rigor. Adjoint methods extend variational principles to in optimization, particularly for PDE-constrained problems, by solving a variational problem to compute gradients efficiently. In this framework, the , derived from the of the , propagates sensitivities backward, enabling the evaluation of how parameters affect objective functions with computational cost independent of the number of parameters. This technique is widely used in aerodynamic design and , where it facilitates gradient-based optimization without repeated forward simulations. Modern software bridges these computational methods with practical implementation, exemplified by FEniCS, an open-source platform for automated FEM solution of variational problems. FEniCS allows users to specify PDEs in variational form using a high-level syntax close to , automating , assembly, and solving via libraries like PETSc. This democratizes access to advanced numerical variational methods for engineering and scientific simulations. As a neural extension, variational autoencoders (VAEs) integrate VI into deep generative models, optimizing an ELBO to learn latent representations that balance reconstruction fidelity and regularization, enabling applications in data generation and since their introduction in 2013.

References

  1. [1]
    Variational principles and nonequilibrium thermodynamics - PMC
    Variational principles are the ultimate tools for deriving the evolution equations of physics, in particular the equations of motion in mechanics. They work ...
  2. [2]
    [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 ...
  3. [3]
    [PDF] Variational Principles in Classical Mechanics - Digital Showcase
    Aug 9, 2017 · Powerful variational techniques in mathematics, that underlie much of modern physics, are introduced and problem solving skills are ...
  4. [4]
    [PDF] THIRD EDITION - Variational Principles in Classical Mechanics
    Aug 19, 2021 · ... history of classical ... principles that underlie Hamilton's Principle, and the related Lagrangian and Hamiltonian analytical formulations.
  5. [5]
    [PDF] cornelius lanczos - Variational Principles of Mechanics
    There is a tremendous treasure of philosophical meaning behind the great theories of Euler and Lagrange, and of Hamilton and. Jacobi, which is completely ...
  6. [6]
    [PDF] Chapter 13 The Variational Method
    The variational method is an approximation procedure that we will use to establish an upper bound for a ground state energy.
  7. [7]
    [PDF] Variational Principles in Physics (Particularly in Quantum Mechanics)
    This approach generalizes in a straightforward way to functions of several variables (several particles, moving in two or three dimensions), and in a less- ...
  8. [8]
    [PDF] AN OVERVIEW OF VARIATIONAL INTEGRATORS - Matthew West
    The idea of variational integrators is very simple: one obtains algorithms by forming a discrete version of Hamilton's variational principle. For conservative ...<|control11|><|separator|>
  9. [9]
    [PDF] Calculus of Variation
    Our aim is to give a treatment of the ele ments of the calculus of variations in a form which is both easily understandable and sufficiently modem. Considerable ...
  10. [10]
    [PDF] Inverse Variational Problem and Symmetry in Action - arXiv
    INVERSE VARIATIONAL PROBLEM AND SYMMETRY IN. ACTION: THE RELATIVISTIC ... calculus of variations. The main goal is to introduce a combined approach ...
  11. [11]
    Calculus of Variations - Richard Fitzpatrick
    Calculus of Variations. It is a well-known fact, first enunciated by Archimedes, that the shortest distance between two points in a plane is a straight-line.Missing: intuition | Show results with:intuition
  12. [12]
    [PDF] Variational methods - Galileo
    The calculus of variations originated with attempts to solve Queen Dido's problem1, known to mathematicians as the isoperimetric problem: to deter- mine the ...<|separator|>
  13. [13]
    [PDF] applications of variational principles to dynamics and conservation ...
    Aug 12, 2012 · Abstract. Much of physics can be condensed and simplified using the principle of least action from the calculus of variations.
  14. [14]
    [PDF] 6 Variational Principles
    In the standard definition used for rigid body motion, there is a rotation of ϕ around the third axis, a rotation of θ around the new position of the first axis ...
  15. [15]
    Variational Principles in Physics
    Fermat called it the "principle of natural economy," and he added the remark that "Nature always acts by the shortest paths." As we said, this principle had ...
  16. [16]
  17. [17]
    3.Fermat's Principle of Least Time - Galileo and Einstein
    Fermat famously stated in the 1630's that a ray of light going from point A to point B always takes the route of least time.Missing: origins | Show results with:origins
  18. [18]
    Basic Seismology 12—Heron of Alexandria and Fermat's principle of ...
    May 1, 2006 · Fermat had studied the works of the ancient mathematicians in the original Greek and, consequently, was inspired to derive the law of refraction ...Missing: origins primary<|control11|><|separator|>
  19. [19]
    History of Two Fundamental Principles of Physics: Least Action and ...
    The LAP was discussed by Maupertuis in his paper entitled Accord des Différentes Lois de la Nature qui Avaient Jusqu'ici Paru Incompatibles presented to the ...Missing: sources | Show results with:sources
  20. [20]
    Principle of least action - Scholarpedia
    Jun 5, 2015 · Maupertuis' principle is older than Hamilton's principle by about a century (1744 vs 1834). The original formulation of Maupertuis was vague ...Statements of Hamilton and... · History · Practical Use of Action Principles
  21. [21]
  22. [22]
    [PDF] J. L. Lagrange's early contributions to the principles and methods of ...
    In his treatise of 1760 on least action LAGRANGE refers only to the earlier appendix. There is, however, considerable evidence to suggest he was also ...Missing: reformulation stationary
  23. [23]
    The history of the Méchanique analitique | Lettera Matematica
    Jun 6, 2014 · In effect, we get the impression that, from the end of 1755 until about 1760, Lagrange devoted himself above all to mechanics. His memoirs of ...Missing: stationary sources
  24. [24]
    Brachistochrone problem - MacTutor History of Mathematics
    The idea is to find a function which maximises or minimises a certain quantity where the function is constrained to satisfy certain constraints. For example ...
  25. [25]
    About the history of Calculus of variations
    [19] "The Lagrange equations arise immediately from the variational problem delta int L dt = 0 (the limits of the integral kept fixed). Remarkably, this ...Missing: 1700s | Show results with:1700s
  26. [26]
    [PDF] Sufficient conditions, fields and the calculus of variations - Craig Fraser
    A new approach to the subject was initiated by Karl Weierstrass (1815–1897) in his Berlin lec- tures of the late 1870s and early 1880s. Although Weierstrass's ...
  27. [27]
    [PDF] A Brief Survey of the Calculus of Variations - arXiv
    Beginning in the 1870s, Weierstrass gave the theory of the calculus of variations a complete overhaul. It took quite some time for these results to become ...
  28. [28]
    [PDF] History of Riemann Mapping Theorem - Stony Brook University
    In his [1904] Hilbert first restored Dirichlet's principle, at least. ON THE ... in the calculus of variations seemed to have solutions which were much ...<|separator|>
  29. [29]
    [PDF] SCUOLA NORMALE SUPERIORE DI PISA Classe di Scienze
    OF THE CALCULUS OF VARIATIONS. (PART I) by EDWARD JAMES MCSHANE (Göttingen). The search for existence theorems for the parametric problem of the calculus of ...Missing: Morrey 1920s 1930s
  30. [30]
    Multiple Integral Peoblems in the Calculus of Variations and Related ...
    The first part of this research reported in these lectures is an extension of Tonelli's work on single and double integral problems in which he employed the so ...
  31. [31]
    [PDF] The Calculus of Variations - College of Science and Engineering
    Jun 6, 2024 · 4.2 Brief review of Sobolev spaces. Sobolev spaces are the natural function spaces in which to look for minimizers of many functionals in the ...
  32. [32]
    [PDF] Functional Analysis, Sobolev Spaces and Partial Differential Equations
    The Sobolev spaces occur in a wide range of questions, in both pure ... calculus of variations and in optimization. Theorem 3.24. Let E and F be two ...
  33. [33]
    [PDF] 1 1D Calculus of Variations.
    Definitions of Frechet and Gateaux derivatives can be easily transferred to the case of functions of two variables f : R × R → R, in which case Frechet ...
  34. [34]
    [PDF] 7 Calculus of Variations
    That is, we will assume L is convex in the variable p. 7.4.2 Coercivity. In order to prove that the functional I has a minimum, we would need to know that I is.
  35. [35]
    [PDF] Existence of global minimum for calculus of variation
    Jun 6, 2025 · To ensure the existence of a minimizer, we need conditions on L that guarantee: • coercivity to ensure the boundedness of a minimizing sequence ...
  36. [36]
  37. [37]
    [PDF] Methodus inveniendi lineas curvas maximi minimive proprietate ...
    Google is proud to partner with libraries to digitize public domain materials and make them widely accessible. Public domain books belong to the.
  38. [38]
    [PDF] The original Euler's calculus-of-variations method - Edwin F. Taylor
    Leonhard Euler's original version of the calculus of variations (1744) used elementary mathematics and was intuitive, geometric, and easily visualized. In.Missing: delta 1700s
  39. [39]
    [PDF] Hamilton's Principle and Lagrange's Equation - Duke Physics
    The contributions of Galileo to the development of classical mechanics are underplayed in many textbooks. In fact they were crucial.
  40. [40]
    Mécanique analytique : Lagrange, J. L. (Joseph Louis), 1736-1813
    Jan 18, 2010 · Mécanique analytique. by: Lagrange, J. L. (Joseph Louis), 1736 ... download 10 Original · SHOW ALL. IN COLLECTIONS. Thomas Fisher Rare Book ...
  41. [41]
    [PDF] The Calculus of Variations - College of Science and Engineering
    Mar 21, 2021 · The Brachistochrone Problem. The most famous classical variational principle is the so-called brachistochrone problem. The compound Greek ...
  42. [42]
    Ch. 2 - The Simple Pendulum - Underactuated Robotics
    Feb 13, 2024 · The Lagrangian derivation of the equations of motion (as described in the appendix) of the simple pendulum yields: m l 2 θ ¨ ( t ) + m g l sin ⁡
  43. [43]
    [PDF] PHY411 Lecture notes on Constraints
    Jan 4, 2019 · Constraints in the form of equation 20 that can be written like equation 23 are holonomic, otherwise they are non-holonomic. Not all constraints ...
  44. [44]
    [PDF] Physics 5153 Classical Mechanics D'Alembert's Principle and The ...
    D'Alembert's principle introduces the force of inertia, converting dynamics to statics, and is an extension of the principle of virtual work to dynamical ...
  45. [45]
    [PDF] D'Alembert's Principle - Craig Fraser - University of Toronto
    Introduction. In 1743 Jean d'Alembert published his work “Treatise on Dynamics, in which the Laws of Equilibrium and Motion of Bodies are reduced.
  46. [46]
    English trans. of E. Noether Paper - UCLA
    Wiss. zu Göttingen 1918, pp235-257. English translation: M.A. Tavel, Reprinted from "Transport Theory and Statistical Mechanics" 1(3), 183-207 (1971).
  47. [47]
    [PDF] Emmy Noether and Symmetry
    Noether's (First) Theorem states that to each continuous symmetry group of the action functional there is a corresponding conservation law of the physical ...
  48. [48]
    [PDF] 13.1 Field theories from Lagrangians - MIT
    This approach is based on developing a Lagrangian density which describes gravitation. Such an approach is now standard in much of field theory. The basic idea ...
  49. [49]
    [PDF] basic differential geometry: variational theory of geodesics
    May 25, 2003 · We assume throughout that M is a Riemannian manifold. For points p ... Variation Formula determines this second derivative in terms of the ...
  50. [50]
    [PDF] Lecture Notes on Minimal Surfaces and Plateau's Problem
    Motivated by the corollary we now define minimal surfaces as surfaces with vanishing mean curvature, that is, essentially, as the critical points of the Plateau ...
  51. [51]
    Solution of the Problem of Plateau - jstor
    We conclude this paper with the following brief proof, based on the formulas of Part III, that the minimal surface whose existence is proved in Part I has the ...
  52. [52]
    On the Problem of Plateau - PNAS
    The solution of Plateau's problem has recently been the subject of a series of important papers by Gamier, Douglas, Rad6, McShane and others. ... Generalizations ...
  53. [53]
    [PDF] entire solutions to equations of minimal surface type in six dimensions
    Introduction. A well-known theorem of Bernstein says that entire minimal graphs in R3 are planes. Building on work of Fleming [6], De Giorgi [5], ...
  54. [54]
    [PDF] Comparison Theorems in Riemannian Geometry
    Definition 2.1 A vector field J along a geodesic γ is called a Jacobi field if it satisfies the Jacobi equation. Remark 2.2 J is a Jacobi field along γ if and ...<|separator|>
  55. [55]
    [PDF] Optimization-Based Control
    The term L(x, u) is referred to as the integral cost and V (x(T)) is the final (or terminal) cost. There are many variations and special cases of the optimal ...
  56. [56]
    [PDF] Lagrange multipliers and optimality - UW Math Department
    Others in the calculus of variations had earlier considered inequality constraints ... Requirements of the form (4.7) or (2.9) are known as complementary ...
  57. [57]
    Hilbert on Dirichlet's Principle - Original Sources
    Every problem of the calculus of variations has a solution, provided restrictive assumptions appropriate to the nature of the given boundary conditions are ...Missing: 1904 | Show results with:1904
  58. [58]
    The historical bases of the Rayleigh and Ritz methods - ScienceDirect
    The method of Ritz. The first paper that Ritz wrote to describe his method for solving boundary value and eigenvalue problems was published in 1908 [3] (“On ...
  59. [59]
    Space-Time Approach to Non-Relativistic Quantum Mechanics
    The probability that a particle will be found to have a path x ⁡ ( t ) lying somewhere within a region of space time is the square of a sum of contributions, ...
  60. [60]
    The Theory of Quantized Fields. I | Phys. Rev.
    The theory is developed in terms of the model supplied by localizable fields. A short review is first presented of the general quantum-mechanical scheme of ...
  61. [61]
    [PDF] Chapter 3: Classical Variational Methods and the Finite Element ...
    Unlike the Rayleigh-Ritz and. Galerkin methods that required approximation functions with sufficient smoothness and satiable boundary conditions over the entire ...
  62. [62]
    [PDF] Eighty Years of the Finite Element Method - arXiv
    This began the use of mixed variational principles to formulate Galerkin finite element methods, which was followed by the assumed strain finite element ...
  63. [63]
    [PDF] Variational Inference: A Review for Statisticians - arXiv
    May 9, 2018 · In this paper, we review variational inference (VI), a method from machine learning that approximates probability densities through optimization ...
  64. [64]
    Variational Methods in Design Optimization and Sensitivity Analysis ...
    Apr 1, 1997 · Variational methods (VM) sensitivity analysis employed to derive the costate (adjoint) equations, the transversality conditions, ...
  65. [65]
    Variational approach by means of adjoint systems to structural ...
    Variational approach by means of adjoint systems to structural optimization and sensitivity analysis—I: Variation of material parameters within fixed domain.Missing: methods | Show results with:methods
  66. [66]
    [PDF] Automated Solution of Differential Equations by the Finite ... - FEniCS
    1. Identify the PDE and its boundary conditions. 2. Reformulate the PDE problem as a variational problem. 3. Make a Python program where ...<|separator|>
  67. [67]
    FEniCS | FEniCS Project
    FEniCS is a popular open-source computing platform for solving partial differential equations (PDEs) with the finite element method (FEM).Download · Documentation · FEniCS 2025 · FEniCSx DocumentationMissing: variational | Show results with:variational
  68. [68]
    [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.