Fact-checked by Grok 2 weeks ago

Hamilton–Jacobi equation

The Hamilton–Jacobi equation, named after William Rowan Hamilton and Carl Gustav Jacobi, is a first-order nonlinear partial differential equation that plays a central role in classical mechanics, providing a method to solve Hamilton's equations of motion through canonical transformations. It is formulated as \frac{\partial S}{\partial t} + H\left(q, \frac{\partial S}{\partial q}, t\right) = 0, where S(q, t) is Hamilton's principal function (or action), H is the Hamiltonian, q denotes the generalized coordinates, and t is time; solutions to this equation generate transformations that reduce the dynamics to separable or trivial forms. For time-independent Hamiltonians, the equation simplifies to H\left(q, \frac{\partial W}{\partial q}\right) = E, where W(q) is the characteristic function and E is the total energy. The equation originated in the work of , who introduced it in his 1834 essay "On a General Method in Dynamics" and expanded it in 1835 with the development of canonical formalism, drawing analogies from to mechanical systems. Jacobi extended 's theory in 1836–1837, generalizing it to time-dependent forces and non-conservative systems while deriving new integrals for problems like the , thus establishing the modern Hamilton–Jacobi framework. This historical development bridged variational principles from —such as of least time—with , laying groundwork for later advancements. In , the Hamilton–Jacobi equation enables the in integrable systems, allowing explicit solutions for trajectories via p = \frac{\partial S}{\partial q} and facilitating the identification of conserved quantities through action-angle variables. Beyond mechanics, it manifests as the in , describing wavefront propagation and ray paths, and serves as the semiclassical limit of the in when \hbar \to 0. Modern extensions include viscosity solutions for non-smooth cases in optimal control theory, such as the , with applications in path planning, , and differential games.

Introduction

Overview

The Hamilton–Jacobi equation is a in that governs the evolution of the principal function S, which depends on the q and time t. This function serves as a for canonical transformations in , enabling the reformulation of a dynamical system's coordinates and momenta. By solving the Hamilton–Jacobi equation, one obtains a that renders trivial, with the new coordinates constant and the new momenta linear in time. This approach is particularly powerful for integrable systems, as it converts the problem of finding trajectories into simpler integrations, known as quadratures. It applies effectively to both time-independent cases, where the is conserved, and time-dependent cases, where explicit time variation is present. Formulated within the framework, which encompasses positions and momenta, the equation facilitates the introduction of action-angle variables for periodic or quasi-periodic motions. These variables represent conserved action integrals and angles that advance uniformly, providing insight into system frequencies without requiring full trajectory computations.

Historical Development

The Hamilton–Jacobi equation originated in the work of William Rowan Hamilton during the 1830s, stemming from his investigations into optics and dynamics. In his 1834 paper "On a General Method in Dynamics," published in the Philosophical Transactions of the Royal Society, Hamilton introduced the characteristic function, a time-independent action integral that satisfies a partial differential equation derived from the principle of least action and analogies to Fermat's principle in optics, linking ray paths to mechanical trajectories. He extended this in his 1835 "Second Essay on a General Method in Dynamics," defining the principal function as a time-dependent action integral and deriving the core partial differential equation for conservative systems, emphasizing its role in simplifying perturbation problems in celestial mechanics. These contributions built on Hamilton's earlier studies of conical refraction, where optical characteristic functions inspired the dynamical formulation. Carl Gustav Jacob Jacobi advanced Hamilton's framework in 1837 through two papers in Crelle's Journal für die reine und angewandte Mathematik. In "Zur Theorie der Variations-Rechnung und der Differential-Gleichungen," Jacobi generalized the approach to time-dependent forces, deriving a single by integrating Hamilton's dual equations and introducing canonical transformations that preserve the form of Hamilton's equations. His second paper, "Über die Reduction der Integration der partiellen Differentialgleichungen erster Ordnung," showed how to reduce the integration of of the to equations. In his earlier 1836 work on the , Jacobi applied this framework, yielding Jacobi's integral as a analogous to , and extended the method to problems on surfaces by treating them as constrained mechanical systems. Jacobi's innovations emphasized the equation's utility in and variational principles, marking a shift toward a more algebraic and transformative structure in analytical dynamics. In the late 19th century, the equation gained prominence in celestial mechanics. Charles-Eugène Delaunay employed canonical transformations from the Hamilton–Jacobi framework in his 1860 Mémoire sur la théorie de la Lune, developing action-angle variables to analyze lunar perturbations and integrate the three-body problem perturbatively. Henri Poincaré further utilized the method in his 1892–1899 treatise Les Méthodes Nouvelles de la Mécanique Céleste, applying it to stability analysis in the three-body problem while highlighting its limitations for non-integrable systems, where small denominators arise in perturbation series. The 20th century saw the equation's adaptation to and . , in his 1926 papers in , drew on the optical-mechanical analogy—replacing the Hamilton–Jacobi equation's eikonal with a —to derive the time-independent , bridging and wave mechanics. Later, in the 1960s and 1970s, Peter Bergmann and Arthur Komar extended the formalism to , using the Hamilton–Jacobi equation to define covariant observables and super-Hamiltonians for constrained gravitational systems, facilitating quantization attempts.

Mathematical Formulation

Notation and Conventions

In the Hamilton–Jacobi framework, the configuration of a mechanical system is described using q = (q_1, \dots, q_n), where n denotes the number of , and their conjugate generalized momenta p = (p_1, \dots, p_n). These variables form the of the , a $2n-dimensional manifold denoted as (q, p) \in \mathbb{R}^{2n}. The Hamiltonian function H(q, p, t) represents the total energy of the system expressed in terms of the variables and possibly explicit time dependence. It is a smooth function H: \mathbb{R}^{2n} \times \mathbb{R} \to \mathbb{R}, typically C^2 in (q, p). The principal function, denoted S(q, t), is the time-dependent central to the Hamilton–Jacobi equation, obtained as the extremal value of integral along classical paths. The action integral is conventionally defined as S = \int_{t_0}^{t} L(q(\tau), \dot{q}(\tau), \tau) \, d\tau, where L is the of the system, and the integral is evaluated over paths from an initial at time t_0 to the q at time t. For systems with time-independent Hamiltonians, where is conserved, the principal function separates into a time-independent part W(q), known as Hamilton's , and a term linear in time: S(q, t) = W(q) - E t, with E the constant . This distinction simplifies the equation for conservative systems, reducing it to a time-independent in W. In multi-dimensional formulations, the Einstein summation convention is employed over repeated indices from 1 to n, such that expressions like p_i \dot{q}_i imply without explicit \sum. This convention streamlines the notation for systems with arbitrary while maintaining vectorial or tensorial interpretations of the variables.

Hamilton's Principal Function

Hamilton's principal function, denoted S(\mathbf{q}_f, \mathbf{q}_i, t), is defined as the value of functional S = \int_{t_i}^{t_f} L(\mathbf{q}, \dot{\mathbf{q}}, t) \, dt, where L is the of the system, evaluated along the classical connecting the initial configuration \mathbf{q}_i at time t_i to the final configuration \mathbf{q}_f at time t_f. This is the one that renders stationary in accordance with . The submanifold in associated with this is generated by S, representing the graph of the canonical relations between initial and final states. A key property of S arises from the stationarity condition \delta S = 0 for path variations with fixed endpoints, which directly implies the Euler-Lagrange equations governing the system's dynamics. In the context of canonical transformations, S serves as a type 2 F_2(\mathbf{q}, \mathbf{Q}, t) = S(\mathbf{q}, t) - \boldsymbol{\alpha} \cdot \mathbf{Q}, where \boldsymbol{\alpha} are constants related to the new momenta, facilitating a transformation to coordinates where the motion is trivialized. For given boundary conditions specifying the initial and final positions and times, S is unique up to an additive constant. In the time-independent case, where the Lagrangian does not explicitly depend on time (autonomous systems), the principal function takes the form S = -E t + W(\mathbf{q}), with W(\mathbf{q}) known as Hamilton's characteristic function and E the constant of the system. This separation allows W to capture the spatial dependence of the action, while the linear time term accounts for the conserved .

The Core Equation

The Hamilton–Jacobi equation is a first-order (PDE) that governs the evolution of Hamilton's principal function S, which serves as the for canonical transformations in . In its time-dependent form, for a system with Hamiltonian H(\mathbf{q}, \mathbf{p}, t), the equation is expressed as H\left(\mathbf{q}, \frac{\partial S}{\partial \mathbf{q}}, t\right) + \frac{\partial S}{\partial t} = 0, where \mathbf{q} denotes the , \mathbf{p} = \partial S / \partial \mathbf{q} represents the conjugate momenta, and S = S(\mathbf{q}, t) is the unknown function to be solved for. This form arises in systems where the Hamiltonian may explicitly depend on time, capturing the through the substitution of momenta in terms of spatial derivatives of S./15%3A_Advanced_Hamiltonian_Mechanics/15.04%3A_Hamilton-Jacobi_Theory) For time-independent Hamiltonians, where energy is conserved, the equation simplifies by assuming S(\mathbf{q}, t) = W(\mathbf{q}) - E t, with W being the time-independent and E the constant total energy. Substituting this yields the time-independent Hamilton–Jacobi equation: H\left(\mathbf{q}, \frac{\partial W}{\partial \mathbf{q}}\right) = E. This version is particularly useful for stationary problems, reducing the PDE to a form focused solely on spatial dependence. In multi-dimensional systems with n coordinates \mathbf{q} = (q_1, \dots, q_n), the equation generalizes straightforwardly to H\left(q_1, \dots, q_n, \frac{\partial S}{\partial q_1}, \dots, \frac{\partial S}{\partial q_n}, t\right) + \frac{\partial S}{\partial t} = 0, maintaining the first-order structure while accommodating the increased dimensionality of phase space. Although classified as a first-order PDE due to involving only first derivatives of S, the Hamilton–Jacobi equation is inherently nonlinear because the Hamiltonian H is typically a nonlinear function of the momenta \partial S / \partial \mathbf{q}. This nonlinearity distinguishes it from linear wave equations and poses challenges in solvability, often requiring specific techniques for exact solutions./15%3A_Advanced_Hamiltonian_Mechanics/15.04%3A_Hamilton-Jacobi_Theory) To ensure uniqueness of the solution, boundary conditions are imposed on S, typically specifying its value at initial or final configuration points in the trajectory, such as S(\mathbf{q}_0, t_0; \mathbf{q}_a, t_a) along extremal paths.

Momentum Relations

In the Hamilton–Jacobi formalism, the canonical momenta p_i are directly obtained from the principal function S as its partial derivatives with respect to the coordinates: p_i = \frac{\partial S}{\partial q_i}. This relation arises because S serves as the for a that simplifies the , mapping the original variables (q_i, p_i) to new constant values (Q_i, P_i). By solving the Hamilton–Jacobi equation for S, these expressions yield the momenta as functions of position and time, effectively parameterizing the flow. For time-independent Hamiltonians, the principal function separates as S(q, \alpha, t) = W(q, \alpha) - E t, where W is the time-independent and E is the total energy (a constant). In this case, the momenta simplify to p_i = \frac{\partial W}{\partial q_i}, and these p_i remain constant along the characteristic curves of the , reflecting the conservation of in the absence of explicit time dependence. This constancy facilitates the integration of trajectories, as the velocity \dot{q}_i = \partial H / \partial p_i can be directly computed from \nabla W. The practical implications of these momentum relations are profound: upon solving the Hamilton–Jacobi equation, the \mathbf{p} = \nabla S (or \nabla W) provides the tangent to the trajectories in configuration , allowing reconstruction of the full dynamical paths without integrating higher-order equations. Additionally, the new coordinates Q_i in the transformed system emerge as Q_i = \partial S / \partial P_i, where P_i are the constant new , completing the initiated by S. This structure underscores how S functions as a type-F₂ , ensuring the transformation preserves the structure of . A illustrative example is the in one dimension, where the is H = p^2 / 2m. The principal function takes the form S(q, p, t) = p q - \frac{p^2}{2m} t, yielding the constant p = \partial S / \partial q, which directly gives the linear trajectory q(t) = q_0 + (p/m) t. In higher dimensions, this generalizes to S = \sum_i p_i q_i - \left( \sum_i p_i^2 / 2m \right) t, with each p_i = \partial S / \partial q_i constant, demonstrating the method's efficiency for integrable systems.

Theoretical Derivations

From Lagrangian Mechanics

In Lagrangian mechanics, the dynamics of a system are governed by the principle of least action, where the action functional S is defined as the time integral of the Lagrangian L(q, \dot{q}, t), given by S[q(t)] = \int_{t_i}^{t_f} L(q, \dot{q}, t) \, dt. The stationarity condition \delta S = 0 for variations of the path q(t) that vanish at the endpoints yields the Euler-Lagrange equations, \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}} \right) = \frac{\partial L}{\partial q}, which describe the equations of motion. To connect this to the formulation, the Legendre transform is employed to introduce the canonical momenta p = \frac{\partial L}{\partial \dot{q}}, defining the as H(q, p, t) = p \dot{q} - L(q, \dot{q}, t), where \dot{q} is expressed as a function of q, p, t via the inverse of the momentum definition. This transformation shifts the description from velocity-dependent Lagrangians to phase-space coordinates. Consider Hamilton's principal function S(q_f, t_f; q_i, t_i), which is the value of integral along the true trajectory satisfying the Euler-Lagrange equations between fixed initial point (q_i, t_i) and variable final point (q_f, t_f). For a single-degree-of-freedom , perturb the final to q_f + \delta q_f while keeping t_f fixed; the first-order change in the action is \delta S = p_f \delta q_f, since the integral term vanishes by the Euler-Lagrange condition. Thus, \frac{\partial S}{\partial q_f} = p_f. Similarly, varying the final time to t_f + \delta t_f while adjusting q_f to q_f - \dot{q}_f \delta t_f to follow the trajectory yields \delta S = -H_f \delta t_f, implying \frac{\partial S}{\partial t_f} = -H_f. Substituting p = \frac{\partial S}{\partial q} into the gives the Hamilton–Jacobi equation, H\left(q, \frac{\partial S}{\partial q}, t\right) + \frac{\partial S}{\partial t} = 0. This derivation generalizes straightforwardly to a system with n , where the coordinates are \mathbf{q} = (q_1, \dots, q_n), momenta \mathbf{p} = (p_1, \dots, p_n), and the action S(\mathbf{q}_f, t_f; \mathbf{q}_i, t_i). The variation of the path leads to boundary terms \delta S = \sum_{j=1}^n p_{f,j} \delta q_{f,j} - H_f \delta t_f, yielding the relations \frac{\partial S}{\partial q_{f,j}} = p_{f,j} for each j and \frac{\partial S}{\partial t_f} = -H_f. The resulting Hamilton–Jacobi equation is the H\left(\mathbf{q}, \frac{\partial S}{\partial \mathbf{q}}, t\right) + \frac{\partial S}{\partial t} = 0, a PDE in S.

Via Canonical Transformations

The Hamilton–Jacobi equation arises in the context of , where the dynamics are governed by Hamilton's equations of motion: for a system with coordinates q_i and conjugate momenta p_i, \dot{q}_i = \frac{\partial H}{\partial p_i} and \dot{p}_i = -\frac{\partial H}{\partial q_i}, with H(q, p, t) the . A canonical transformation to new variables (Q_j, P_j) preserves the form of these equations, transforming the original H into a new one K(Q, P, t). To solve the system, one seeks a canonical transformation such that K = 0, rendering the new coordinates ignorable: the transformed Hamilton's equations then yield \dot{Q}_j = 0 and \dot{P}_j = 0, implying Q_j and P_j are constants of motion, denoted P_j = \alpha_j. Such transformations are generated by a type-2 F_2(q, Q, t), which relates the old and new variables through the conditions for canonicity: p_i = \frac{\partial F_2}{\partial q_i} and P_j = -\frac{\partial F_2}{\partial Q_j}, with the new coordinates given implicitly by the transformation relations. The transformation of the follows from the total time derivative, yielding K(Q, P, t) = H\left(q, \frac{\partial F_2}{\partial q}, t\right) + \frac{\partial F_2}{\partial t}. To achieve K = 0, assume the new momenta P_j = \alpha_j are constants, and express the as F_2(q, Q, t) = S(q, \alpha, t) - \sum_j \alpha_j Q_j, where S(q, \alpha, t) is Hamilton's principal function. Substituting this form gives the relation H\left(q, \frac{\partial S}{\partial q}, t\right) + \frac{\partial S}{\partial t} = K(\alpha); setting K(\alpha) = 0 for the trivial case produces the Hamilton–Jacobi equation: H\left(q_1, \dots, q_n; \frac{\partial S}{\partial q_1}, \dots, \frac{\partial S}{\partial q_n}; t\right) + \frac{\partial S}{\partial t} = 0. Solutions to this determine S, from which the original trajectories can be reconstructed via p_i = \frac{\partial S}{\partial q_i} and Q_j = \frac{\partial S}{\partial \alpha_j}. The canonicity of the transformation, ensuring preservation of the structure of , is verified using brackets. For the transformation to be , the fundamental brackets must hold: \{Q_i, Q_j\} = 0, \{P_i, P_j\} = 0, and \{Q_i, P_j\} = \delta_{ij}, where the is defined as \{f, g\} = \sum_k \left( \frac{\partial f}{\partial q_k} \frac{\partial g}{\partial p_k} - \frac{\partial f}{\partial p_k} \frac{\partial g}{\partial q_k} \right). These conditions are satisfied by the F_2, as the relations p_i = \frac{\partial F_2}{\partial q_i} and P_j = -\frac{\partial F_2}{\partial Q_j} guarantee the J is preserved under the transformation matrix M, via M^T J M = J. This symplectic invariance underpins the validity of the Hamilton–Jacobi approach for integrable systems.

Comparison to Hamiltonian and Lagrangian Approaches

The formulation of derives from the that extremizes \delta \int L(q, \dot{q}, t) \, dt = 0, where L is the function, yielding second-order ordinary differential equations (ODEs) known as the Euler-Lagrange equations for the q. This approach excels in incorporating constraints through Lagrange multipliers and is particularly suited for systems with , as it operates directly in configuration space without explicit momenta. However, it lacks a natural structure, making it less intuitive for analyzing conserved quantities like or for perturbations around equilibria. In contrast, the Hamiltonian formulation transforms the problem into first-order symplectic ODEs in phase space (q, p), where p are the conjugate momenta obtained via the Legendre transform p = \partial L / \partial \dot{q}, and the dynamics follow Hamilton's equations \dot{q} = \partial H / \partial p, \dot{p} = -\partial H / \partial q, with H the . This preserves the symplectic structure, facilitating the identification of integrals of motion and brackets, which is advantageous for and . Nonetheless, solving the resulting system of $2n coupled first-order ODEs analytically remains challenging for most non-integrable systems, often requiring numerical methods or approximations. The Hamilton–Jacobi equation (HJE) builds on the Hamiltonian framework by introducing Hamilton's principal function S(q, t), such that momenta are p = \partial S / \partial q and the equation reads \partial S / \partial t + H(q, \partial S / \partial q, t) = 0, effectively reducing the dynamics to a (PDE) in configuration space. A key advantage is that, for integrable systems with a complete integral of the HJE, the original equations decouple into n ODEs solvable by quadratures, yielding explicit solutions for trajectories without further . This is particularly effective for separable potentials, where additive or multiplicative separation allows global solutions via characteristics, providing a unified for all trajectories. In essence, the HJE transforms the problem into finding a to ignorable coordinates, simplifying integrable cases beyond the direct ODE approaches. Despite these benefits, the HJE's nonlinear PDE form poses significant challenges, as classical smooth solutions may fail to exist or be unique due to singularities and shock formation along characteristics, unlike the well-posedness of ODEs in or . In modern applications, such as , weak solutions via methods are employed to handle nonsmooth cases, though this adds conceptual complexity absent in the ODE formulations. All three approaches—Lagrangian, Hamiltonian, and Hamilton–Jacobi—are equivalent, stemming from the same least-action principle, but the HJE represents a "solved" variant where the action S serves as the , bridging variational origins to explicit integrability.

Solution Methods

Separation of Variables

The method provides a systematic approach to solving the Hamilton–Jacobi equation (HJE) for integrable systems by assuming the principal function decomposes into additive components along separable coordinates. This technique is particularly effective for natural Hamiltonians in orthogonal coordinate systems where the potential and metric allow such decomposition. By reducing the to a set of ordinary differential equations, the method yields constants of motion that facilitate integration of the . In the time-independent case, the HJE takes the form H(\mathbf{q}, \nabla W) = E, where W(\mathbf{q}) is the principal function and E is the total energy. The standard posits an additive separation: W(\mathbf{q}) = \sum_{i=1}^n W_i(q_i) , with possible additive constants absorbed into the functions W_i. Substituting this into the HJE requires the H to separate in the coordinates q_i, meaning each term involving \partial W / \partial q_i and the potential must isolate to individual equations. This leads to a of equations: \frac{1}{2m g_{ii}} \left( \frac{d W_i}{d q_i} \right)^2 + V_i(q_i) = \beta_i , where m is the particle , g_{ii} are the coefficients, V_i are separable potential components, and \beta_i are additive separation constants that serve as integrals of motion, with \sum \beta_i = E. The condition for such separation holds when the admits a Stäckel form, ensuring the off-diagonal terms vanish appropriately. For the time-dependent HJE, H(\mathbf{q}, \nabla S, t) + \frac{\partial S}{\partial t} = 0, where S(\mathbf{q}, t) is , the extends the time-independent form by separating time explicitly: S(\mathbf{q}, t) = -E t + W(\mathbf{q}). This reduces to the time-independent HJE for W, with the E acting as a separation constant between temporal and spatial parts. The spatial separation then proceeds analogously, yielding the same ordinary differential equations as in the time-independent case, provided the is time-independent. In certain systems, particularly those analogous to the , multiplicative separation may apply instead of additive. Here, the principal function satisfies a form like \nabla^2 \Psi + k^2 \Psi = 0, where separation assumes \Psi = \prod_i \Psi_i(q_i), leading to separated equations with multiplicative constants. This is linked to the additive case in the HJE through conformal transformations or Killing tensors, and it occurs in Stäckel systems where the metric supports both types of separability. The intrinsic connection arises because additive separation in the HJE corresponds to multiplicative separation in the associated Laplace-Beltrami equation on the configuration space. The precise criteria for separability in orthogonal coordinates are encapsulated in the Stäckel conditions, originally formulated for natural Hamiltonians. These require that the potential and metric components fit into a Stäckel matrix, a determinant form ensuring the HJE separates completely: for coordinates q_1, \dots, q_n, the condition is that there exist functions f_{ij}(q_k) such that the separated equations hold without cross-terms, specifically S_{ij}(g_{kk}) = 0 for i \neq j, k, where S_{ij} are components of the Stäckel operator. This guarantees n independent integrals of motion, confirming integrability in the Liouville-Arnold sense. These conditions apply to both (free particle) and general dynamical cases, generalizing to geodesics or charged particles under appropriate extensions.

Examples in Coordinate Systems

In Cartesian coordinates, the Hamilton–Jacobi equation for a single particle of m in a separable potential V(\mathbf{q}) = V_x(x) + V_y(y) + V_z(z) admits trivial , as the coordinates are orthogonal and the potential decomposes additively. The time-independent takes the form W(x, y, z; E, \boldsymbol{\beta}) = \sum_{i=x,y,z} \int^{q_i} \sqrt{2m (\beta_i - V_i(q_i'))} \, dq_i' , where E = \sum \beta_i is the total energy, the \beta_i are separation constants (partial energies) related to the momenta components, and each one-dimensional is \frac{1}{2m} \left( \frac{\partial W_i}{\partial q_i} \right)^2 + V_i(q_i) = \beta_i. For free motion with V = 0, the integrals simplify, yielding linear solutions W = \sum \sqrt{2m \beta_i} \, q_i and constant velocities \dot{q}_i = \sqrt{2 \beta_i / m}. This separation reduces the problem to independent quadratures for each coordinate, facilitating explicit integration of the . In spherical coordinates (r, \theta, \phi), the Hamilton–Jacobi equation separates for central potentials V(r) due to rotational invariance, yielding solutions parameterized by constants. The is W(r, \theta, \phi; E, L_z, L) = S_r(r) + S_\theta(\theta) + L_z \phi, where L_z is the azimuthal separation constant (z-component of ), and L^2 is the total constant from the \theta-equation: \frac{1}{2m r^2} \left[ \left( \frac{d S_\theta}{d \theta} \right)^2 + \frac{L_z^2}{\sin^2 \theta} \right] = \frac{L^2}{2m r^2}. The radial part follows as \frac{1}{2m} \left( \frac{d S_r}{d r} \right)^2 + V(r) + \frac{L^2}{2m r^2} = E, with solution S_r(r) = \int^r \sqrt{2m \left( E - V(r') - \frac{L^2}{2m {r'}^2} \right)} \, dr'. The motion reduces to quadratures, such as the radial turning points and r(t) = \int \frac{dr}{\sqrt{2m \left( E - V(r) - \frac{L^2}{2m r^2} \right)}}, which for V(r) = -k/r () yields conic sections. For cylindrical coordinates (\rho, \phi, z) with axial symmetry, where the potential V(\rho, z) is independent of the azimuthal angle \phi, the Hamilton–Jacobi equation separates in \phi but couples \rho and z unless V further decomposes. The characteristic function is W(\rho, \phi, z; E, m) = S_{\rho z}(\rho, z) + m \phi, with m the azimuthal separation constant (angular momentum about the z-axis). The separated equation becomes \frac{1}{2m} \left[ \left( \frac{\partial S_{\rho z}}{\partial \rho} \right)^2 + \left( \frac{\partial S_{\rho z}}{\partial z} \right)^2 + \frac{m^2}{\rho^2} \right] + V(\rho, z) = E. Full separation into S_\rho(\rho) + S_z(z) requires additive V(\rho, z) = V_\rho(\rho) + V_z(z), leading to independent quadratures analogous to the Cartesian case, with effective potentials incorporating the centrifugal term m^2 / (2m \rho^2). This form is useful for problems like charged particle motion in a uniform magnetic field along z. Elliptic coordinates, based on confocal quadrics, allow separation of the Hamilton–Jacobi equation for two-center potentials such as V = k_1 / r_1 + k_2 / r_2, where r_1, r_2 are distances to fixed centers; the separated equations involve elliptic integrals for the radial-like and angular-like functions. Similarly, separate the equation for potentials like the in , V = -k/r + F z, with coordinates (\xi, \eta, \phi) where \xi = r + z, \eta = r - z; the separation yields constants for azimuthal motion and independent quadratures for \xi and \eta, facilitating analysis of linear Stark splitting.

Physical and Geometric Interpretations

Optics and Wave Fronts

In the context of , the Hamilton–Jacobi equation finds its origins in William Rowan 's optico-mechanical analogy, developed in the early , where light rays are interpreted as geodesics that minimize the \int n \, ds, with n denoting the along the path ds. Here, the principal function S serves as the eikonal, representing the from a source to a point in space, analogous to the action in . This analogy posits that the propagation of light through inhomogeneous media follows paths of stationary optical length, bridging variational principles in and dynamics. The emerges as a specific form of the time-independent Hamilton–Jacobi equation in , given by |\nabla S|^2 = n^2(\mathbf{r}), where \nabla S = \mathbf{p} corresponds to the , and the is effectively H = \frac{1}{2} |\mathbf{p}|^2 scaled such that |\mathbf{p}| = n, reflecting the local c/n. This governs the eikonal S, whose level sets S(\mathbf{r}) = \text{constant} define the wave fronts—surfaces perpendicular to the rays at every point. The evolution of these wave fronts adheres to Huygens' principle, where each point on a wave front acts as a source of secondary wavelets, propagating the forward while the rays trace the orthogonal trajectories. The trajectories of light rays correspond to the characteristics of the , which in the full phase-space formulation are bicharacteristics of the Hamilton–Jacobi system, projecting onto the rays in configuration space. These characteristics satisfy Hamilton's equations \dot{\mathbf{r}} = \nabla_{\mathbf{p}} H = \mathbf{p}/n and \dot{\mathbf{p}} = -\nabla_{\mathbf{r}} H = \nabla n, ensuring rays bend according to the of the , as in . underpins this framework, stating that rays follow paths where the S = \int n \, ds is (\delta S = 0), minimizing or maximizing travel time in with varying n. In astigmatic optical systems, such as those with cylindrical lenses, the eikonal S becomes multi-valued near caustics, allowing multiple ray paths to converge and accounting for focal lines rather than points.

Connection to Quantum Mechanics

The connection between the Hamilton–Jacobi equation (HJE) and arises primarily through reformulations of the that reveal classical limits. In the Madelung transformation, the wave function is expressed in polar form as \psi = \sqrt{\rho} \exp(i S / \hbar), where \rho = |\psi|^2 is the and S is the phase function analogous to the classical . Substituting this into the time-dependent i \hbar \partial \psi / \partial t = \hat{H} \psi yields two coupled equations: a for \rho and a modified HJE for S, \partial S / \partial t + H(q, \nabla S / m) + Q = 0, where Q = -\hbar^2 / (2m) (\nabla^2 \sqrt{\rho}) / \sqrt{\rho} is the quantum potential accounting for non-classical effects. This Madelung formulation bridges and , with the quantum potential Q vanishing in the \hbar \to 0, recovering the standard HJE. In the WKB (Wentzel–Kramers–Brillouin) approximation, for slowly varying potentials, the wave function is approximated as \psi \approx A \exp(i S / \hbar), where S satisfies the HJE to leading order in \hbar. As \hbar \to 0, the semiclassical limit emerges, with quantum tunneling and bound states approximated via connection formulas linking oscillatory and evanescent regions. Bohmian mechanics interprets the deterministically, where particle trajectories are guided by the velocity field \mathbf{v} = \nabla S / m, satisfying the classical Euler-Lagrange equations modified only by the quantum potential. This provides a non-local, causal picture of quantum evolution, with the HJE for S dictating the guidance equation. In the stationary case, the time-independent \hat{H} \psi = E \psi leads to a time-independent HJE via the WKB \psi \approx \exp(i S / \hbar), where S incorporates the eigenvalue, enabling semiclassical quantization rules like the Bohr–Sommerfeld condition \oint p \, dq = (n + 1/2) h.

Applications

In Gravitational Fields

The Hamilton–Jacobi equation (HJE) provides an effective framework for analyzing particle motion in gravitational potentials, particularly for central forces where separability simplifies the solution. In the classical Kepler problem, which models the two-body interaction via an inverse-square law potential V(r) = -\frac{k}{r}, the time-independent HJE takes the form \frac{1}{2m} |\nabla W|^2 + V(r) = E, where W is the characteristic function, m is the particle mass, and E is the total energy. This equation is separable in spherical coordinates, yielding separation constants corresponding to the total energy E and the magnitude of angular momentum L. The resulting radial and angular equations integrate via quadratures, producing conic-section orbits—elliptical for bound states (E < 0)—with the eccentricity determined by the interplay of E and L. In general relativity, the HJE extends to geodesic motion in curved spacetime, describing the paths of test particles in gravitational fields. For a metric g_{\mu\nu}, the relativistic HJE for timelike geodesics is \frac{1}{2} g^{\mu\nu} \frac{\partial W}{\partial x^\mu} \frac{\partial W}{\partial x^\nu} = -\frac{m^2}{2}, where W is the principal function and the right-hand side is zero for null geodesics. Separation of variables in appropriate coordinates reveals conserved quantities analogous to energy and angular momentum, facilitating the integration of the geodesic equations. This approach highlights the geometric nature of gravity, reducing the problem to quadratures similar to the classical case. A prominent example is geodesic motion in the Schwarzschild metric, which describes the spacetime around a non-rotating, spherically symmetric mass M, given by ds^2 = -\left(1 - \frac{2GM}{c^2 r}\right) c^2 dt^2 + \left(1 - \frac{2GM}{c^2 r}\right)^{-1} dr^2 + r^2 d\Omega^2. In (which reduce to standard spherical for Schwarzschild), the HJE separates completely, introducing a separation constant K corresponding to the square of the total angular momentum, which governs the latitudinal motion. The separated equations yield effective potentials for radial and polar motions, enabling analytical solutions for orbits: stable circular orbits exist up to the innermost stable circular orbit at r = 6GM/c^2, beyond which plunging trajectories dominate near the event horizon. This separability underscores the integrability of Schwarzschild geodesics, contrasting with more complex spacetimes. For weak gravitational fields, post-Newtonian (PN) approximations adapt the HJE to incorporate relativistic corrections to Newtonian gravity, essential for systems like binary pulsars or inspiraling compact objects. The PN Hamiltonian includes terms up to order (v/c)^n, where v is the orbital velocity, and the corresponding HJE H\left( \mathbf{x}, \frac{\partial W}{\partial \mathbf{x}} \right) = E separates perturbatively, yielding orbit precession and energy loss via gravitational radiation at 2.5PN order. These solutions match observational data, such as the periastron advance in binary systems, and form the basis for effective one-body models in gravitational-wave physics. Seminal PN frameworks ensure coordinate-invariant results, bridging classical and fully relativistic regimes.

In Electromagnetic Fields

In the presence of electromagnetic fields, the Hamilton–Jacobi equation (HJE) describes the motion of a charged particle through the canonical momentum \vec{p} - e \vec{A}, where e is the particle charge and \vec{A} is the magnetic vector potential. In the relativistic formulation (with c = 1), the Hamiltonian takes the form H = \sqrt{m^2 + (\vec{p} - e \vec{A})^2} + e \phi = E, where m is the particle rest mass, \phi is the scalar electric potential, and E is the total energy, yielding the time-dependent HJE \frac{\partial S}{\partial t} + \sqrt{m^2 + (\nabla S - e \vec{A})^2} + e \phi = 0. For the non-relativistic case, the Hamiltonian simplifies to H = \frac{1}{2m} |\vec{p} - e \vec{A}|^2 + e \phi, leading to the HJE \frac{\partial S}{\partial t} + \frac{1}{2m} |\nabla S - e \vec{A}|^2 + e \phi = 0. Specific field configurations allow exact or separable solutions to the HJE. For a circularly polarized electromagnetic wave, where the vector potential \vec{A} oscillates with helical symmetry, the equation becomes separable in helical coordinates, enabling analytical treatment of the particle's resonant motion and energy gain. In the case of a monochromatic plane wave (including circular polarization as a special elliptical case), exact solutions are obtained via constants of motion such as transverse and longitudinal canonical momenta, with trajectories expressed parametrically using Jacobian elliptic functions that depend on wave intensity and initial conditions. For a solenoidal magnetic field, as in a magnetic bottle configuration, the HJE identifies adiabatic invariants—such as the magnetic moment \mu = \frac{m v_\perp^2}{2B}—that remain conserved under slow field variations, ensuring particle trapping and stability. The characteristics of the HJE in electromagnetic fields correspond to the classical particle trajectories, which integrate the Lorentz force \vec{F} = e (\vec{E} + \vec{v} \times \vec{B}) along the paths generated by . These characteristics provide a geometric framework for solving the equations of motion, reducing the problem to quadratures in integrable cases. An intriguing classical analogy to the emerges from the HJE, where the action S acquires a multi-valued component \frac{q}{\hbar} \oint \vec{A} \cdot d\vec{r} due to the vector potential enclosing a solenoid, manifesting as a phase shift interpretable via de Broglie wave-particle duality.

Modern Numerical and Control Theory Uses

In modern applications, the Hamilton–Jacobi equation (HJE) has been extended through the theory of to handle discontinuous weak solutions, particularly for non-convex Hamiltonians arising in complex dynamics. Introduced by in the 1980s, viscosity solutions provide a framework for proving existence, uniqueness, and stability of solutions to first-order nonlinear partial differential equations (PDEs) without requiring classical differentiability. This approach defines a solution u as a viscosity subsolution if, for any test function \phi touching u from above at a point, the HJE holds in the appropriate inequality sense, and similarly for supersolutions from below. Viscosity theory has enabled rigorous analysis of HJEs in scenarios with shocks or non-smooth data, such as front propagation in heterogeneous media. Recent advances in numerical methods leverage deep learning to address the curse of dimensionality in high-dimensional HJEs, where traditional grid-based solvers become computationally infeasible beyond a few dimensions. Post-2020 developments include neural network architectures inspired by representation formulas, such as those using the for convex Hamiltonians, achieving solutions in up to 20 dimensions with relative errors under 5%. Actor-critic methods combine policy iteration with deep operator networks () to approximate solutions for static HJEs, demonstrating scalability to 100 dimensions in mean-field games via generative adversarial network-like training. These techniques mitigate exponential growth in computational cost by parameterizing the solution directly with neural networks, as seen in () for diffusive Eikonal equations. A 2025 review highlights their efficacy in optimal control benchmarks, where neural solvers outperform classical finite difference methods by orders of magnitude in high dimensions. In optimal control theory, the Hamilton–Jacobi–Bellman (HJB) equation, a stochastic variant of the , defines the value function V(x,t) = \min \mathbb{E} \left[ \int_t^T L(x_s, u_s) ds + \phi(x_T) \right] as the infimum over admissible controls of the expected cost integral plus terminal payoff. This PDE arises in continuous-time stochastic optimization and yields the optimal control via u^* = \arg\min_u [L(x,u) + \nabla V \cdot f(x,u)]. Applications in robotics include trajectory optimization for underactuated systems, such as the , where the HJB reduces to a solved backward in time to compute feedback policies like u^*(x,t) = -R^{-1} B^T S(t) x. In finance, HJB equations model portfolio optimization under uncertainty, as in extended to robust control, where solutions guide dynamic asset allocation to minimize risk-adjusted costs. Differential games extend the HJE to adversarial settings via the Isaacs equation, a min-max formulation H(x, \nabla V, D^2 V) = \min_u \max_v [f(x,u,v) \cdot \nabla V + L(x,u,v)] = 0 for zero-sum two-player games. This PDE characterizes the value function as the equilibrium payoff, with lower and upper Isaacs equations ensuring consistency under incomplete information. Recent numerical solutions employ PINN-based policy iteration to handle nonconvex cases, approximating saddle-point policies in up to 10 dimensions for pursuit-evasion scenarios. Extensions to dissipative dynamics incorporate non-conservative forces into the HJE by modifying the with a damping term, such as \tilde{H} = e^{\lambda t} H + \frac{\partial S}{\partial t} \lambda, where \lambda > 0 accounts for energy loss. This approach derives directly from the principal function S for systems like damped oscillators, enabling integration of without auxiliary variables. Trends in special issues highlight applications to Hamiltonian systems, where the HJE identifies invariants in irreversible processes.

References

  1. [1]
    The Hamilton–Jacobi equation: An alternative approach
    May 1, 2020 · The Hamilton–Jacobi equation (HJE) is one of the most elegant approaches to Lagrangian systems such as geometrical optics and classical mechanics.Missing: applications | Show results with:applications
  2. [2]
    [PDF] An overview of the Hamilton-Jacobi equation
    Abstract. This paper is a survey of the Hamilton-Jacobi partial differential equation. We begin with its origins in Hamilton's formulation of classical ...
  3. [3]
    [PDF] The Early History of Hamilton-Jacobi Dynamics 1834–1837
    May 2, 2023 · As we observed above, Hamilton and Jacobi would have had to reproduce the whole derivation of the Hamilton-Jacobi equation using 7-derivatives ...
  4. [4]
    [PDF] Chapter 4 Canonical Transformations, Hamilton-Jacobi Equations ...
    We've made good use of the Lagrangian formalism. Here we'll study dynamics with the. Hamiltonian formalism. Problems can be greatly simplified by a good choice ...
  5. [5]
    XV. On a general method in dynamics - Journals
    Cite this article. Hamilton William Rowan. 1834XV. On a general method in dynamics; by which the study of the motions of all free systems of attracting or ...
  6. [6]
    Zur Theorie der Variations-Rechnung und der Differential ... - EuDML
    Jacobi, C.G.J.. "Zur Theorie der Variations-Rechnung und der Differential-Gleichungen.." Journal für die reine und angewandte Mathematik 17 (1837): 68-82.Missing: Crelle's | Show results with:Crelle's
  7. [7]
    Hamilton-Jacobi Version of General Relativity | Phys. Rev.
    The classical general relativistic Hamilton-Jacobi theory is then elaborated to the point where, via the correspondence principle, the preferred complementary ...Missing: 1960s | Show results with:1960s
  8. [8]
    Hamilton-Jacobi equation - Scholarpedia
    Oct 21, 2011 · The Hamilton-Jacobi equation is used to generate particular canonical transformations that simplify the equations of motion.History · Hamilton-Jacobi Equation and... · Action as a Solution of the...
  9. [9]
    4. Hamilton's Least Action Principle and Noether's Theorem
    Bottom Line: All of classical mechanics follows from the Principle of Least Action. But that Principle itself follows from quantum mechanics! Historical ...
  10. [10]
    [PDF] arXiv:1511.03302v1 [math-ph] 10 Nov 2015
    Nov 10, 2015 · The Lagrangian submanifold Π(EL) ⊂ T∗R × T∗R ∼= T∗(R × R) is ... submanifold with generating function given by Hamilton's principal function W.
  11. [11]
    [PDF] Chapter 6 Hamilton's Equations - Rutgers Physics
    The function S is known as Hamilton's principal function, while the function W is called Hamilton's characteristic function, and the equa- tions (6.17) and ( ...
  12. [12]
    [PDF] The Hamilton-Jacobi Equation: an intuitive approach. - HAL
    Oct 20, 2019 · The Hamilton-Jacobi equation (HJE) is one of the most elegant approach to Lagrangian systems such as geometrical optics and classical ...
  13. [13]
    [PDF] Notes on Classica Mechanics II 1 Hamilton–Jacobi Equations
    The use of action does not stop in obtaining Euler–Lagrange equation in classical mechanics. Instead of using the action to vary in order to obtain.
  14. [14]
    [PDF] Hamilton-Jacobi Equation [gam13]
    – Hamilton-Jacobi theory: Find a generating function of the F2-type which makes the transformed Hamiltonian vanish identically: E(q1,...,qn;p1,...,pn;t) + ∂ ∂t F2( ...Missing: F_2( | Show results with:F_2(
  15. [15]
    12. The Hamilton-Jacobi Equation - Galileo and Einstein
    The Hamilton-Jacobi equation is therefore a third complete description of the dynamics, equivalent to Lagrange's equations and to Hamilton's equations.
  16. [16]
    [PDF] 1 The Hamilton-Jacobi equation
    where A, φ are constants depending on initial conditions. What happens if we make a new coordi- nate as. Q = q − A0 sin(ωt + φ0).
  17. [17]
    [PDF] Hamiltonian Mechanics
    Hamiltonian mechanics uses generalized momenta (pi) and generalized coordinates (qi) as dynamical variables, resulting in symmetric equations of motion.
  18. [18]
    [PDF] 4. The Hamiltonian Formalism - DAMTP
    W0 is known as Hamilton's principal function. The spe- i r level surfaces of W. Figure 69: cial property of this solution to the Hamilton-Jacobi equation is ...
  19. [19]
    [PDF] An overview of the Hamilton–Jacobi theory: the classical and ... - arXiv
    Jan 11, 2021 · Assuming that the Lagrangian function is regular; that is, ωL is a sym- plectic form; the Lagrangian equation i(XL)ωL = dEL has a unique ...
  20. [20]
    [PDF] an introduction to viscosity solutions for fully nonlinear pde ... - arXiv
    Nov 11, 2014 · The name of this theory originates from the “vanishing viscosity method” developed first for 1st order fully nonlinear PDE (Hamilton-Jacobi PDE) ...
  21. [21]
    [PDF] Separation of Variables in the Null Hamilton–Jacobi Equation
    Our approach that leads to consideration of a new family of equations characterizing the separation of variables of a null Hamiltonian (1), is based upon three ...
  22. [22]
    [PDF] Separation of Variables in the Special Diagonal Hamilton-Jacobi ...
    For the Hamilton-Jacobi equation, there are infinitely many coordinate systems that have coeffi- cients a(*(q) that satisfy the conditions obtained by Stackel.
  23. [23]
    Killing Tensors and Variable Separation for Hamilton-Jacobi and ...
    The additive separation of variables in the Hamilton-Jacobi equation and the multiplicative separation of variables in the Laplace-Beltrami equation are ...
  24. [24]
    [PDF] Separation of variables in Hamilton-Jacobi equation for a ... - arXiv
    Dec 11, 2020 · In the case of the free Hamilton-Jacobi equation, such condition is met when there is a complete set of Killing fields in space.
  25. [25]
    When is a Hamiltonian system separable?
    It is known that if such a separation is possible, then it can take place only when the equation is expressed in terms of generalized elliptic coordinates or in ...
  26. [26]
    [PDF] Introducing quantum mechanics through its historical roots
    Sommerfeld used parabolic coordinates for his compari- sons to the Stark effect. He used the Hamilton–Jacobi method of separation of variables to develop three ...<|separator|>
  27. [27]
    Hamilton's Papers on Geometrical Optics
    (These partial differential equations are analogous to the `Hamilton-Jacobi equation' of dynamics.) Hamilton investigates the relations amongst the partial ...
  28. [28]
  29. [29]
  30. [30]
    Simple derivations of the Hamilton–Jacobi equation and the eikonal ...
    Jun 1, 2011 · We show that both the Hamilton–Jacobi equation and the eikonal equation can be derived by a common procedure using only elementary aspects of ...<|control11|><|separator|>
  31. [31]
    Quantentheorie in hydrodynamischer Form | Zeitschrift für Physik A ...
    Cite this article. Madelung, E. Quantentheorie in hydrodynamischer Form. Z. Physik 40, 322–326 (1927). https://doi.org/10.1007/BF01400372. Download citation.
  32. [32]
    On bi-Hamiltonian formulation of the perturbed Kepler problem
    Apr 8, 2015 · In the spherical variables, Hamiltonian H (1.3) takes the form. such that the Hamilton–Jacobi equation H = h has an additive separable solution.Abstract · Introduction · Bi-Hamiltonian formulation of... · Conclusion
  33. [33]
    [PDF] Integrating the geodesic equations in the Schwarzschild and Kerr ...
    Apr 3, 2008 · The application of this theorem to the Schwarzschild and Kerr metrics leads straightforwardly to the general solution of their geodesic.
  34. [34]
    Hamiltonian formulation of general relativity and post-Newtonian ...
    Hamiltonian formalisms provide powerful tools for the computation of approximate analytic solutions of the Einstein field equations. The post-Newtonian ...
  35. [35]
    72. Dynamics of a Relativistic Particle in an Electromagnetic Field
    *The Hamilton-Jacobi Equation. In classical dynamics, there are three standard approaches to the equations of motion: the Lagrangian, the Hamiltonian, and ...
  36. [36]
    Hamilton-Jacobi theory for Hamiltonian and non-Hamiltonian systems
    A1 Nonrelativistic particle in an electromagnetic field​​ The Lorentz force F = e ( E + 1 c [ v H ] ) acts on the particle, where e is the particle charge, E = − ...
  37. [37]
    Dynamics of a charged particle in a circularly polarized traveling ...
    Sep 1, 2000 · An equation is found for the particle energy when it is initially resonant. This equation is solved exactly, and the asymptotic solution is ...Missing: Jacobi | Show results with:Jacobi
  38. [38]
    [PDF] arXiv:0804.0731v1 [physics.class-ph] 4 Apr 2008
    Apr 4, 2008 · In particular, we show how a particle's trajectory in a monochromatic plane wave with a high intensity can be obtained from the “fundamental ...
  39. [39]
    [PDF] ACTION/ANGLE VARIABLES AND ADIABATIC INVARIANCE
    Just as the Hamilton-Jacobi equation is the short-wavelength limit of the Schrâodinger equation, the time-independent H-J equation is the same limit of the ...Missing: solenoidal | Show results with:solenoidal
  40. [40]
    THE HAMILTON-JACOBI EQUATIONS FOR A RELATIVISTIC ...
    It is the object of this paper to exhibit an H-J theory for relativistic particles which retains the symmetry of formulation and which yields certain known ...
  41. [41]
    [PDF] Aharonov-Bohm effect and Hamilton-Jacobi equation - arXiv
    Let us show that Aharonov-Bohm effect may be derived from the Hamilton-Jacobi equation using de Broglie idea (1924) on wave-particle duality. Schroedinger ...Missing: multi- valued
  42. [42]
    AMS :: Transactions of the American Mathematical Society
    Viscosity solutions of Hamilton-Jacobi equations. HTML articles powered by AMS MathViewer. by Michael G. Crandall and Pierre-Louis Lions: Trans. Amer. Math.
  43. [43]
    Existence of viscosity solutions of Hamilton-Jacobi equations
    Recently M. G. Crandall and P.-L. Lions (Trans. Amer. Math. Soc. 277 (1983), 1–42) introduced the class of “viscosity” solutions ...
  44. [44]
    [PDF] Recent Advances in Numerical Solutions for Hamilton-Jacobi PDEs
    Feb 28, 2025 · This review explores several recent approaches for numerically solving Hamilton-Jacobi partial differential equations, ranging from classical ...
  45. [45]
    [PDF] arXiv:2207.02947v2 [q-fin.MF] 22 Jun 2024
    Jun 22, 2024 · Subject to some regularity conditions on the solution of the HJB equation, we estab- lish the existence of the optimal investment strategy.
  46. [46]
    Hamilton Jacobi Isaacs equations for differential games with ...
    We investigate Hamilton Jacobi Isaacs equations associated to a two-players zero-sum differential game with incomplete information.
  47. [47]
    Jacobi--Isaacs equations with PINN-based policy iteration - arXiv
    Jul 21, 2025 · Abstract page for arXiv paper 2507.15455: Solving nonconvex Hamilton--Jacobi--Isaacs equations with PINN-based policy iteration.
  48. [48]
    [PDF] A HAMILTON-JACOBI TREATMENT OF DISSIPATIVE SYSTEMS
    The Hamilton-Jacobi treatment of dissipative systems uses the Hamilton-Jacobi equation, simplified with separation of variables, to find the principal function ...
  49. [49]
    Special Issue : New Trends in Hamilton-Jacobi Theory - MDPI
    The Hamilton–Jacobi equation is particularly useful in identifying conserved quantities for mechanical systems, which may be possible even when the ...