Fact-checked by Grok 2 weeks ago

Analytical mechanics

Analytical mechanics is a reformulation of that employs variational principles and energy-based methods to derive the for mechanical systems, primarily through the and formalisms, enabling the treatment of constraints and without direct reference to forces. It provides a coordinate-independent framework for analyzing systems with multiple , from particles to rigid bodies and continuous media, and serves as the foundation for extensions into , , and field theories. The development of analytical mechanics traces its origins to the 17th century with Isaac Newton's Philosophiæ Naturalis Principia Mathematica (1687), which established the foundational laws of motion, but it truly emerged in the through efforts to mathematize and generalize these principles. Key milestones include Leonhard Euler's introduction of variational methods in the 1740s, Jean le Rond d'Alembert's in 1743, and Joseph-Louis Lagrange's comprehensive synthesis in Mécanique Analytique (1788), which formalized the approach using the principle that the variation of the action integral is zero. In the 19th century, advanced the field with his 1834–1835 papers on the principal function and canonical equations, introducing and momentum variables, while refined these ideas in the 1840s. Central to analytical mechanics is the Lagrangian formulation, where the Lagrangian L = T - V (with T as and V as ) leads to the Euler–Lagrange equations: \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}_j} \right) - \frac{\partial L}{\partial q_j} = 0, for q_j. This method excels in handling (expressible as functions of coordinates) by reducing , as in the case of a where the constraint simplifies the system to one coordinate. The Hamiltonian formulation complements it by defining the Hamiltonian H = \sum p_j \dot{q}_j - L (where p_j = \frac{\partial L}{\partial \dot{q}_j} are conjugate momenta), yielding Hamilton's equations: \dot{q}_j = \frac{\partial H}{\partial p_j} and \dot{p}_j = -\frac{\partial H}{\partial q_j}, which are particularly useful for conservative systems where H represents total energy and is conserved. Analytical mechanics offers significant advantages over Newtonian vectorial methods, including invariance under coordinate transformations, natural incorporation of constraints via (\sum (m_k \dot{\mathbf{v}}_k - \mathbf{F}_k) \cdot \delta \mathbf{r}_k = 0), and facilitation of symmetry-based conservation laws through . It applies to diverse fields such as , , and , where it simplifies multi-body dynamics, and underpins by providing the phase-space structure essential for quantization. Later extensions by figures like Paul Appell and Georg Hamel in the early addressed nonholonomic constraints, while connections to have further enriched its scope.

Introduction

Motivation

Analytical mechanics emerged as a response to the limitations of Newtonian mechanics in handling complex systems, particularly those involving and non-inertial or variable-mass scenarios. Newtonian formulations, reliant on vector-based force balances and geometric constructions, often become cumbersome for systems where explicit computation of constraint forces is required or where coordinates are not naturally Cartesian. , in his seminal work Mécanique Analytique (1788), sought to reformulate analytically, eliminating the need for geometric diagrams and providing a unified framework applicable to any . This shift addressed the challenges of applying Newton's laws to systems like rockets with variable , where the standard form \mathbf{F} = m \mathbf{a} assumes and requires modifications for mass variation. A primary advantage of analytical mechanics lies in its reformulation of using concepts rather than forces, expressed through the L = T - V, where T is and V is . This scalar approach simplifies the treatment of constraints by incorporating them directly into the choice of , obviating the need to solve for constraint forces explicitly, which can complicate Newtonian equations. For non-Cartesian coordinates, such as spherical or cylindrical systems, analytical methods allow seamless adaptation, exploiting symmetries to reduce the number of equations without altering the fundamental form. This energy-based description not only enhances computational elegance but also reveals conserved quantities more readily, providing a more general and insightful perspective on mechanical systems. Illustrative examples highlight these benefits. In a simple pendulum, Newtonian mechanics requires resolving tension forces perpendicular to the motion, leading to coupled differential equations; analytical mechanics, using the angle as a generalized coordinate, yields a single equation focused on energy differences, bypassing tension entirely. For multi-body systems like a double pendulum, where interconnections impose holonomic constraints, Newtonian analysis demands explicit inclusion of joint tensions and results in a proliferation of vector components, whereas the Lagrangian approach condenses the dynamics into a manageable set of scalar equations in generalized coordinates. Such simplifications make analytical mechanics indispensable for intricate configurations, such as beads constrained to helical wires or masses sliding on conical surfaces, where force computations are prohibitive.

Historical overview

The development of analytical mechanics began in the mid-18th century with foundational principles that shifted from geometric to analytical methods. In 1743, introduced his principle in the Traité de dynamique, which reformulated Newton's laws using virtual displacements to handle constraints and equilibrium, laying groundwork for later formulations by treating dynamic problems as static ones with inertial forces. Shortly after, in 1744, Pierre-Louis Maupertuis proposed the principle of least , positing that nature minimizes a quantity proportional to action along the path of light or particles, influencing variational approaches. That same year, Leonhard Euler advanced variational principles in his Methodus inveniendi lineas curvas maximi minimive proprietate gaudentes, developing methods to solve isoperimetric problems and deriving Euler-Lagrange equations for optimizing functionals, marking a key step in the applied to mechanics. Joseph-Louis Lagrange synthesized these ideas in his seminal 1788 work Mécanique Analytique, where he presented a fully analytical framework using and the function, eliminating the need for geometric figures and enabling treatment of complex systems like rigid bodies and fluids. This text formalized into Lagrange's , emphasizing variational principles without reference to forces, and became the cornerstone of analytical mechanics. In the early , William Rowan Hamilton extended this in his 1834 paper "On a General Method in Dynamics" and the 1835 "Second Essay on a General Method in Dynamics," published in the Philosophical Transactions of the Royal Society, introducing the formulation with and momenta, facilitating analysis and optical-mechanical analogies. The 19th century saw further evolution through canonical transformations, introduced by in 1837 during his work on the Hamilton-Jacobi equation, which allowed simplification of systems via coordinate changes preserving the form of Hamilton's equations, enhancing solvability for periodic motions and perturbations. In the , analytical mechanics expanded to handle constrained systems, notably through Paul Dirac's contributions in papers from 1950 to 1964, where he developed the and constraint classification for singular Lagrangians, bridging with applications. These extensions solidified analytical mechanics as a versatile tool for .

Foundations

Generalized coordinates and constraints

In analytical mechanics, are a set of independent parameters q_i (where i = 1, 2, \dots, n) that completely specify the configuration of a , without being restricted to Cartesian or other orthogonal systems. These coordinates can represent any suitable variables, such as lengths, angles, or other measurable quantities, allowing for a flexible description of the 's position in its configuration space. Unlike , which are spatial transformations of Cartesian coordinates (e.g., polar or spherical systems that still describe points in ), generalized coordinates are broader and may not correspond to spatial positions at all; for instance, they can include non-spatial parameters like rotation angles for rigid bodies. This generality enables the reduction of the system's description to the minimal number of variables needed, accounting for inherent symmetries or constraints. The configuration space of a is the manifold formed by all possible values of the , with its dimension equal to the number of . Constraints on the reduce this dimension: for a with $3N Cartesian coordinates (where N is the number of particles), the n are given by n = 3N - m, where m is the number of independent constraints. Constraints in analytical mechanics are classified into holonomic and non-holonomic types. are integrable relations of the form f(q_i, t) = 0, which can be used to eliminate variables and reduce the number of directly. Non-holonomic constraints, in contrast, involve velocities and cannot be integrated to position constraints alone, for example, the for a rolling on a plane, which constrains the velocities but allows various paths on the plane while coupling position to in a path-dependent way. Holonomic constraints are further divided into scleronomic (time-independent, e.g., fixed lengths of rigid rods) and rheonomic (time-dependent, e.g., a bob constrained to move on a surface whose shape varies with time). Scleronomic constraints preserve the structure of the configuration space over time, while rheonomic ones introduce explicit time dependence, complicating the dynamics. A classic example is the , consisting of two point masses connected by massless rods of fixed lengths, subject to and scleronomic constraints from the rod lengths and . Here, the system has two , described by \theta_1 and \theta_2, the angles each pendulum makes with the vertical, reducing the six Cartesian coordinates (three per mass) by four constraints (two fixed lengths and two for planarity). This choice simplifies the formulation while capturing the full configuration space as a two-dimensional .

D'Alembert's principle and virtual work

In analytical mechanics, the concept of provides a foundational tool for analyzing systems with constraints. A is an change in the position of a particle, denoted as \delta \mathbf{r}_i, that is consistent with the system's constraints at a given instant but does not involve any actual motion or passage of time. These displacements are orthogonal to the constraint forces, ensuring that the constraint forces perform no , and they satisfy the variations \delta q_j associated with . The principle of states that for a system in under applied forces, the total virtual work done by all forces through any is zero: \delta W = \sum_i \mathbf{F}_i \cdot \delta \mathbf{r}_i = 0, where \mathbf{F}_i are the applied forces. For ideal constraints, the virtual work performed by constraint forces is zero, allowing the principle to focus solely on applied forces while automatically accounting for constraints. D'Alembert's principle extends the principle of virtual work from statics to dynamics by incorporating inertial effects, generalizing Newton's second law for constrained systems. It asserts that \sum_i (\mathbf{F}_i - m_i \mathbf{a}_i) \cdot \delta \mathbf{r}_i = 0 for all virtual displacements \delta \mathbf{r}_i, where m_i \mathbf{a}_i is the inertial force; this treats the dynamics as an equivalent static problem where inertial forces balance the applied forces. This formulation, originally proposed by in 1743, eliminates the need to explicitly solve for constraint forces, simplifying the analysis of complex mechanical systems. The derivation of follows directly from Newton's laws. Starting with \mathbf{F}_i = m_i \mathbf{a}_i for each particle, introduce a fictitious inertial -\ m_i \mathbf{a}_i to rewrite the equation as \mathbf{F}_i - m_i \mathbf{a}_i = 0. Taking the with an arbitrary \delta \mathbf{r}_i and summing over all particles yields the principle, as the virtual work of the zero vanishes. One key advantage of is its automatic handling of , as the virtual displacements are chosen to satisfy them, rendering constraint forces irrelevant in the equation. For example, consider Atwood's machine, consisting of two masses m and M (M > m) connected by an inextensible string over a frictionless . The enforces equal-magnitude displacements \delta s for both masses but in opposite directions. The virtual work due to is -mg \delta s + Mg \delta s = (M - m)g \delta s, while the inertial virtual work is -(m + M) \ddot{s} \delta s. Setting the total to zero gives \ddot{s} = \frac{(M - m)g}{M + m}, without needing to compute the in the string. In terms of generalized coordinates q_j, which describe the system's configuration while incorporating constraints, D'Alembert's principle takes the form \sum_j \left( Q_j - \frac{d}{dt} \left( \frac{\partial T}{\partial \dot{q}_j} \right) + \frac{\partial T}{\partial q_j} \right) \delta q_j = 0, where T is the , Q_j are the generalized forces, and the sum holds for all virtual variations \delta q_j. Since the \delta q_j are arbitrary, the coefficient of each must vanish individually.

Lagrangian mechanics

Lagrange's equations

Lagrangian mechanics is formulated using the Lagrangian function L, defined as the difference between the T and the V of the system: L = T - V. This function depends on the q_i, their time derivatives \dot{q}_i (generalized velocities), and possibly time t. For systems with n , the are given by the Euler-Lagrange equations. For conservative systems where all forces derive from a potential, these take the form \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}_i} \right) - \frac{\partial L}{\partial q_i} = 0, \quad i = 1, \dots, n. These second-order differential equations determine the system's dynamics once the Lagrangian is specified. For systems subject to non-conservative forces, the Euler-Lagrange equations are generalized by including generalized forces Q_i on the right-hand side: \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}_i} \right) - \frac{\partial L}{\partial q_i} = Q_i, where Q_i represent the work done by non-potential forces through virtual displacements in the q_i direction. The kinetic energy T in generalized coordinates is typically a quadratic form T = \frac{1}{2} \sum_{i,j=1}^n m_{ij}(q, t) \dot{q}_i \dot{q}_j, with m_{ij} as the mass matrix elements depending on coordinates and time. The derivation of these equations stems from D'Alembert's principle, which equates the virtual work of applied and inertial forces to zero, expressed in generalized coordinates to yield the Euler-Lagrange form without explicit force calculations. Holonomic constraints, expressible as f_k(q_1, \dots, q_n, t) = 0 for k = 1, \dots, m, reduce the to n - m. One approach is : solve the constraints for m coordinates and express L in terms of the independent n - m coordinates. Alternatively, retain all n coordinates and incorporate constraints using Lagrange multipliers \lambda_k, leading to the modified equations \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}_i} \right) - \frac{\partial L}{\partial q_i} = \sum_{k=1}^m \lambda_k \frac{\partial f_k}{\partial q_i}, \quad i = 1, \dots, n, along with the m constraint equations f_k = 0. The multipliers \lambda_k account for constraint forces. This systematically handles constraints without eliminating variables. A example is the one-dimensional harmonic oscillator, with coordinate q = x, kinetic energy T = \frac{1}{2} m \dot{x}^2, and potential V = \frac{1}{2} k x^2, so L = \frac{1}{2} m \dot{x}^2 - \frac{1}{2} k x^2. The Euler-Lagrange equation yields \frac{d}{dt} (m \dot{x}) - (-k x) = 0, or m \ddot{x} + k x = 0, recovering the familiar equation of motion. For central force problems, such as a particle in a spherically symmetric potential, the in spherical coordinates includes an angular coordinate \phi that is cyclic—meaning \partial L / \partial \phi = 0—implying of the conjugate momentum p_\phi = \partial L / \partial \dot{\phi}, which is the about the force center. In general, if \partial L / \partial q_i = 0 for a coordinate q_i, then the generalized momentum p_i = \partial L / \partial \dot{q}_i is conserved.

Properties of the Lagrangian

The Lagrangian L(q, \dot{q}, t) in analytical mechanics exhibits homogeneity of degree two in the generalized velocities \dot{q}_i for standard systems where the T is a in \dot{q}. By for homogeneous functions applied to the , this property implies \sum_i \dot{q}_i \frac{\partial L}{\partial \dot{q}_i} = 2T. For scleronomic constraints and potentials where L = T - V with T , the relation simplifies to \sum_i \dot{q}_i \frac{\partial L}{\partial \dot{q}_i} = 2T. This homogeneity underpins the conservation of total E = T + V when L is independent of time, as the time derivative of E vanishes along trajectories satisfying Lagrange's equations. The remains invariant under , such as boosts to a frame moving with constant velocity \mathbf{v} relative to an inertial frame. For a transformation q_i' = q_i + v_i t and \dot{q}_i' = \dot{q}_i + v_i, the form of L is preserved up to a total time , ensuring the are unchanged and linear is conserved for isolated systems. To incorporate dissipative forces linear in velocities, such as , the R(q, \dot{q}, t) is defined, typically quadratic in \dot{q} for viscous (e.g., R = \frac{1}{2} \sum_i b_i \dot{q}_i^2 where b_i > 0). The associated generalized forces are then Q_i = -\frac{\partial R}{\partial \dot{q}_i}, which modify Lagrange's equations without altering the conservative structure of L itself. A key gauge invariance holds: if L' = L + \frac{dF}{dt} for an arbitrary differentiable function F(q, t), the Euler-Lagrange equations derived from L' are identical to those from L, as the additional term integrates to boundary contributions in the . For standard mechanical systems, the Lagrangian satisfies the Legendre condition of strict convexity in the velocities, meaning the \frac{\partial^2 L}{\partial \dot{q}_i \partial \dot{q}_j} is positive definite. This ensures the Legendre transform to the is well-defined and invertible.

Hamiltonian mechanics

Hamilton's equations

Hamiltonian mechanics reformulates the dynamics of a system in terms of generalized coordinates q_i and their conjugate momenta p_i, providing a canonical structure that emphasizes the phase space evolution. This formulation arises from the Legendre transformation of the Lagrangian L(q_i, \dot{q}_i, t), where the momenta are defined as p_i = \frac{\partial L}{\partial \dot{q}_i}. The Hamiltonian function H(q_i, p_i, t) is then given by H = \sum_i p_i \dot{q}_i - L, with the velocities \dot{q}_i expressed as functions of q_i and p_i to yield H solely in terms of these canonical variables. For standard mechanical systems where the Lagrangian separates into kinetic energy T (quadratic in velocities) and potential energy V (velocity-independent), the Hamiltonian simplifies to H = T + V, representing the total energy in phase space coordinates./08:_Hamiltonian_Mechanics/8.02:_Legendre_Transformation_between_Lagrangian_and_Hamiltonian_mechanics) The equations of motion in this framework, known as Hamilton's equations, take the symmetric first-order form: \dot{q}_i = \frac{\partial H}{\partial p_i}, \quad \dot{p}_i = -\frac{\partial H}{\partial q_i}. These $2n equations replace the n second-order Lagrange equations, describing the of the system as a in the $2n-dimensional , a manifold coordinatized by \{q_i, p_i\}. Trajectories in phase space are curves parameterized by time, with the generating the through its partial derivatives, ensuring conservation of H for time-independent systems. This structure was originally developed by in his 1834 paper, where he introduced the characteristic function leading to these canonical equations. Phase space provides a geometric interpretation of dynamics, where the non-crossing nature of trajectories highlights deterministic evolution and stability. The Hamiltonian formulation offers key advantages, including its suitability for perturbative analyses due to the linear structure in velocities and its direct pathway to quantization in quantum mechanics, where q_i and p_i become operators satisfying commutation relations. It also facilitates the study of symmetries and conserved quantities through Noether's theorem in canonical variables./15:_Advanced_Hamiltonian_Mechanics/15.08:_Comparison_of_the_Lagrangian_and_Hamiltonian_Formulations) A simple example is the one-dimensional , with L = \frac{1}{2} m \dot{q}^2 - \frac{1}{2} k q^2, yielding p = m \dot{q} and H = \frac{p^2}{2m} + \frac{1}{2} k q^2. Hamilton's equations then become \dot{q} = \frac{p}{m} and \dot{p} = -k q, reproducing the familiar . For the describing planetary motion under inverse-square gravity, using polar coordinates r, \theta with \mu, the effective is H = \frac{p_r^2}{2\mu} + \frac{l^2}{2\mu r^2} - \frac{G M \mu}{r}, where l is the conserved ; the equations \dot{r} = \frac{p_r}{\mu} and \dot{p}_r = \frac{l^2}{\mu r^3} - \frac{G M \mu}{r^2} (with \dot{\theta} = \frac{l}{\mu r^2}) yield elliptical orbits. In systems with constraints, the Hamiltonian formulation requires careful handling to maintain consistency. Classically, are incorporated by choosing appropriate that eliminate the constraints from the outset, reducing the dimensionality. For more general cases, including non-holonomic or singular Lagrangians, Dirac's method classifies constraints into primary ones (arising directly from the , such as when \frac{\partial^2 L}{\partial \dot{q}_i^2} = 0) and secondary ones (generated by requiring time persistence of primary constraints via the total time derivative). These are further distinguished as first-class (commuting with all constraints, generating symmetries) or second-class (not commuting, fixing variables), with Dirac brackets modifying the Poisson structure to enforce them. This approach ensures the equations remain well-defined on the constrained .

Poisson brackets and properties of the Hamiltonian

In , the provides an for describing the dynamics of functions on . For two smooth functions f and g depending on q_i and momenta p_i, the is defined as \{f, g\} = \sum_i \left( \frac{\partial f}{\partial q_i} \frac{\partial g}{\partial p_i} - \frac{\partial f}{\partial p_i} \frac{\partial g}{\partial q_i} \right). This definition, introduced by in his 1809 work on perturbations, encodes the of . The Poisson bracket exhibits several key properties that make it a Lie bracket on the algebra of functions. It is bilinear in its arguments, meaning \{af + bg, h\} = a\{f, h\} + b\{g, h\} and similarly for the second argument, where a and b are constants. It is antisymmetric, satisfying \{f, g\} = -\{g, f\}, and obeys the Leibniz rule for products: \{fg, h\} = f\{g, h\} + g\{f, h\}. Additionally, it satisfies the Jacobi identity: \{f, \{g, h\}\} + \{g, \{h, f\}\} + \{h, \{f, g\}\} = 0, ensuring consistency with the Lie algebra structure. The fundamental Poisson brackets among the canonical variables are \{q_i, q_j\} = 0, \{p_i, p_j\} = 0, and \{q_i, p_j\} = \delta_{ij}, where \delta_{ij} is the Kronecker delta. A central role of the Poisson bracket is in governing the time evolution of observables. For any function f(q, p, t), the total time derivative is given by \frac{df}{dt} = \frac{\partial f}{\partial t} + \{f, H\}, where H is the Hamiltonian function. This formulation unifies Hamilton's equations as \dot{q}_i = \{q_i, H\} and \dot{p}_i = \{p_i, H\}. If the Hamiltonian has no explicit time dependence, \partial H / \partial t = 0, then \{H, H\} = 0 implies dH/dt = 0, so the Hamiltonian is conserved and represents the total energy. Poisson brackets play a crucial role in canonical transformations, which are changes of variables (q_i, p_i) \to (Q_i, P_i) that preserve the form of Hamilton's equations. Such transformations maintain the canonical structure if the Poisson brackets satisfy \{Q_i, Q_j\} = 0, \{P_i, P_j\} = 0, and \{Q_i, P_j\} = \delta_{ij}. The invariance of the under these transformations ensures that the symplectic form of is preserved, implying that canonical maps are area-preserving in two-dimensional phase space projections. transformations are conveniently generated by scalar functions F, known as generating functions, of which there are four standard types depending on the mixed variables: F_1(q, Q), F_2(q, P), F_3(p, Q), and F_4(p, P). For instance, for F_1(q, Q, t), the relations are p_i = \partial F_1 / \partial q_i and P_i = -\partial F_1 / \partial Q_i, with the new Hamiltonian K = H + \partial F_1 / \partial t. In standard non-relativistic systems, the Hamiltonian exhibits homogeneity in the momenta due to the form of the . Since the is quadratic in velocities and momenta are linear in velocities, H = T(p) + V(q) is homogeneous of degree two in the p_i, satisfying p_i \partial H / \partial p_i = 2H - 2V(q). This property facilitates scaling arguments in analysis and underscores the quadratic nature of free-particle motion encoded in the brackets.

Variational principles

Principle of least action

The principle of least action, also known as Hamilton's principle, posits that the path taken by a physical system between two points in configuration space over a specified time interval is the one that makes the action functional stationary. The action S is defined as the time integral of the Lagrangian L, which typically represents the difference between kinetic and potential energy: S[q(t)] = \int_{t_1}^{t_2} L(q, \dot{q}, t) \, dt, where q(t) denotes the generalized coordinates and \dot{q} = dq/dt. According to the principle, the actual trajectory q(t) satisfies \delta S = 0 for variations \delta q(t) that vanish at the endpoints t_1 and t_2. This variational condition ensures that nearby paths yield second-order changes in S, establishing the true path as an extremum (minimum, maximum, or saddle point) among admissible paths. The derivation of the from this principle relies on the . Consider a small variation \delta q(t) in the path, leading to the first-order change in action: \delta S = \int_{t_1}^{t_2} \left( \frac{\partial L}{\partial q} \delta q + \frac{\partial L}{\partial \dot{q}} \delta \dot{q} \right) dt. Integrating the second term by parts gives \delta S = \int_{t_1}^{t_2} \left( \frac{\partial L}{\partial q} - \frac{d}{dt} \frac{\partial L}{\partial \dot{q}} \right) \delta q \, dt + \left[ \frac{\partial L}{\partial \dot{q}} \delta q \right]_{t_1}^{t_2}. Since the boundary term vanishes due to fixed endpoints, setting \delta S = 0 for arbitrary \delta q yields the Euler-Lagrange equation: \frac{d}{dt} \frac{\partial L}{\partial \dot{q}} - \frac{\partial L}{\partial q} = 0. This equation governs the system's and reproduces Newton's laws for appropriate choices of L. Historically, precursors to emerged in the 18th century, with Pierre-Louis Maupertuis proposing an abbreviated for systems of fixed total energy E, defined as S' = \int \sqrt{2m(E - V)} \, ds, where ds is the element, m is the mass, and V is the potential. This form, minimizing the integral along the path rather than over time, aligns with in optics and applies to conservative systems where time is not fixed. later refined this into a formulation, interpreting the as the length in a conformally transformed space with ds^2 = 2m(E - V) dq^2, which extremizes geodesics and aids in solving separable Hamilton-Jacobi equations. These extensions highlight the principle's adaptability to constrained or energy-fixed scenarios. Illustrative examples demonstrate the principle's power. For a free particle, where L = T = \frac{1}{2} m \dot{q}^2, the stationary action implies straight-line geodesics in configuration space at constant velocity, matching inertial motion. The brachistochrone problem, seeking the curve of fastest descent under gravity from point A to B, is solved by minimizing the time functional \int \sqrt{\frac{1 + (dy/dx)^2}{y}} \, dx (up to constants), yielding a cycloid path that outperforms straight lines due to velocity buildup. This variational approach is analogous to the action principle under reparameterization. In the quantum domain, Richard Feynman's generalizes the principle, summing contributions from all paths weighted by e^{iS/\hbar}; in the classical limit as \hbar \to 0, interference effects concentrate around the stationary-action path, recovering deterministic mechanics.

Connection to Lagrangian and Hamiltonian formulations

The principle of stationary action serves as the variational foundation connecting the abstract principle to the specific and formulations in analytical mechanics. In the approach, the action is expressed as the integral of the over time, S = \int_{t_1}^{t_2} L(q, \dot{q}, t) \, dt, where L = T - V for conservative systems, with T the and V the . Requiring the first variation of the action to vanish, \delta S = 0, for admissible paths with fixed endpoints directly produces the Euler-Lagrange equations, \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}_i} \right) - \frac{\partial L}{\partial q_i} = 0, which encapsulate the . This derivation, originating from Lagrange's reformulation of Newtonian mechanics, emphasizes and constraints without reference to forces, unifying diverse mechanical systems under a single variational law. The connection to the Hamiltonian formulation extends this variational structure to phase space through a Legendre transformation. Defining the canonical momentum as p_i = \frac{\partial L}{\partial \dot{q}_i} and the as H(q, p, t) = \sum_i p_i \dot{q}_i - L(q, \dot{q}, t), the action becomes S = \int_{t_1}^{t_2} \left( \sum_i p_i \dot{q}_i - H(q, p, t) \right) dt, often rewritten in as S = \int \left( \sum_i p_i dq_i - H dt \right). Stationarity \delta S = 0 with respect to independent variations in both coordinates q_i and momenta p_i, subject to fixed conditions in the extended , yields Hamilton's canonical equations: \dot{q}_i = \frac{\partial H}{\partial p_i} and \dot{p}_i = -\frac{\partial H}{\partial q_i}. This phase-space , while evocative of quantum path integrals, provides a classical derivation that highlights the of , where paths connect specified initial and final states in (q, p). Hamilton pioneered this transformation in his general method for , bridging characteristic functions and variational principles. Beyond deriving the equations, the action principle reveals deeper structural ties, particularly through its role in generating canonical transformations and preserving the symplectic form. The time-evolution map from t_1 to t_2, obtained by extremizing , constitutes a canonical transformation that maintains the structure \{q_i, p_j\} = \delta_{ij} and the symplectic two-form \omega = \sum_i dq_i \wedge dp_i./15%3A_Advanced_Hamiltonian_Mechanics/15.03%3A_Canonical_Transformations_in_Hamiltonian_Mechanics) In this framework, the action functional acts as a for such transformations, ensuring the invariance of the form under coordinate changes. This unification underscores how the provides a common variational origin for both formulations, enabling their equivalence while facilitating extensions to more complex systems like fields and .

Advanced formulations

Hamilton-Jacobi theory

The Hamilton–Jacobi theory provides a powerful method for solving through a canonical transformation generated by a principal function S(q, t; \alpha), where q are the original coordinates and \alpha are constants labeling the solutions. This approach transforms the original H(q, p, t) into a new Hamiltonian K = 0, rendering the transformed equations trivial and solvable directly, with both the new coordinates \beta and momenta \alpha constant. The theory originates from the independent contributions of , who introduced the in his 1834 paper on dynamics, and , who extended it in the 1840s to multi-dimensional systems. The core of the theory is the Hamilton–Jacobi equation, \frac{\partial S}{\partial t} + H\left(q, \frac{\partial S}{\partial q}, t\right) = 0, where S is the principal function, serving as a type 1 generating function for the canonical transformation p_i = \frac{\partial S}{\partial q_i} and \beta_j = -\frac{\partial S}{\partial \alpha_j}. A complete integral of this equation, S(q, t; \alpha), must depend on the n coordinates q and n arbitrary constants \alpha, ensuring the transformation covers the full phase space for an n-degree-of-freedom system. For time-independent Hamiltonians, S separates as S = W(q; \alpha) - \alpha_0 t, reducing the equation to H(q, \frac{\partial W}{\partial q}; \alpha_0) = \alpha_0, where \alpha_0 is the energy constant. The resulting motion follows geodesics in the characteristic manifold defined by the HJ equation, confirming the preservation of canonical structure under Poisson brackets. Separation of variables is a key technique for solving the HJ equation, applicable when the Hamiltonian allows additive or multiplicative separability in orthogonal . For additive separation, assume S = \sum_i S_i(q_i, \alpha), leading to independent ordinary differential equations for each S_i if the Hamiltonian decomposes accordingly, such as H = \sum_i T_i(p_i) + V_i(q_i). Multiplicative separation, common in central force problems, assumes S = \sum_i \sqrt{f_i(q_i)} S_i(\alpha), yielding separated constants of motion. This method succeeds for systems like the or , where the coordinate system aligns with the symmetries. In integrable systems with n independent constants of motion, the theory yields action-angle variables, introduced by Charles-Eugène Delaunay in 1860 for applications. The action variables J_i are defined as J_i = \frac{1}{2\pi} \oint p_i \, dq_i, the areas enclosed by the periodic orbits in the i-th invariant torus divided by $2\pi, while the angle variables \theta_i = \frac{\partial S}{\partial J_i} evolve linearly as \dot{\theta_i} = \omega_i(J), with frequencies \omega_i = \frac{\partial H}{\partial J_i}. The actions J_i are conserved, reflecting the system's integrability, and the transformation simplifies the to H(J), independent of angles. This coordinate system highlights the quasi-periodic motion on tori and underpins adiabatic invariants. A representative example is the one-dimensional with H = \frac{p^2}{2m} + \frac{1}{2} m \omega^2 q^2. The time-independent HJ equation becomes \frac{1}{2m} \left( \frac{\partial W}{\partial q} \right)^2 + \frac{1}{2} m \omega^2 q^2 = E, yielding W(q; E) = \int \sqrt{2m \left( E - \frac{1}{2} m \omega^2 q^2 \right)} \, dq, which evaluates to a involving inverse sine functions. The complete solution S = W - E t generates elliptical orbits in , with J = \frac{E}{\omega} and \theta = \omega t + \phi. For the Kepler problem, describing motion in a $1/r potential, the HJ equation separates in spherical coordinates, with the radial part \frac{1}{2m} \left( \frac{\partial S_r}{\partial r} \right)^2 + \frac{l^2}{2m r^2} - \frac{k}{r} = E (where l is the constant) and angular parts yielding conserved quantities. The complete integral leads to conic section trajectories, with action variables related to and , demonstrating bounded elliptical orbits for negative energies. The Hamilton–Jacobi theory forms the foundation for canonical perturbation theory in nearly integrable systems, where small perturbations to a solvable are handled by expanding the in powers of the parameter, enabling averaging over angles to isolate secular effects. This approach, rooted in Delaunay's , has been pivotal in for long-term stability analysis.

Routhian and Appellian mechanics

Routhian mechanics provides a hybrid formulation between the and approaches, particularly suited for systems possessing cyclic coordinates, where certain do not appear explicitly in the . Developed by Edward John Routh in the late , the Routhian function R is defined as R = \sum_i p_i \dot{q}_i - L, where the sum is over the cyclic coordinates q_i, p_i = \partial L / \partial \dot{q}_i are their conjugate momenta (which are conserved), and L = T - V is the with T and potential V. This construction allows partial , treating cyclic variables in a Hamiltonian-like manner while retaining Lagrangian equations for the remaining non-cyclic coordinates, thereby reducing the system's from n to n - s, with s being the number of cyclic coordinates. The in consist of two sets: for non-cyclic coordinates q_k, \frac{d}{dt} \left( \frac{\partial R}{\partial \dot{q}_k} \right) - \frac{\partial R}{\partial q_k} = Q_k, resembling Lagrange's equations with generalized forces Q_k, and for cyclic coordinates, \dot{p}_i = -\frac{\partial R}{\partial q_i}, which enforce since \partial R / \partial q_i = 0 for true cyclic variables. Often, R simplifies to R = T_2 + V - \sum p_i \dot{q}_i, where T_2 is the quadratic part of the in velocities; this form bridges to when all coordinates are cyclic. The method is advantageous for exploiting symmetries, such as ignorable rotations, without full transformation to . A representative example is the central force problem, where a particle moves in a potential V(r) depending only on radial distance r, with spherical coordinates (r, \theta, \phi); the azimuthal angle \phi is cyclic, yielding conserved angular momentum p_\phi = \partial L / \partial \dot{\phi} = m r^2 \sin^2 \theta \, \dot{\phi}. The Routhian is R = p_\phi \dot{\phi} - L. Substituting \dot{\phi} = p_\phi / (m r^2 \sin^2 \theta) yields R = -\frac{1}{2} m (\dot{r}^2 + r^2 \dot{\theta}^2) + V_{\rm eff}(r, \theta), where the effective potential is V_{\rm eff}(r, \theta) = V(r) + \frac{p_\phi^2}{2 m r^2 \sin^2 \theta}. The reduced equations then govern radial and polar motion, simplifying to a one-dimensional effective radial problem if equatorial motion (\theta = \pi/2) is assumed, with equation m \ddot{r} = -\partial V_{\text{eff}} / \partial r. This reduction highlights angular momentum conservation without solving the full three-dimensional system. Appellian mechanics, also known as the Gibbs-Appell , extends to systems with non- constraints by introducing quasi-coordinates, which are functions of the generalized velocities rather than integrable position variables. Independently developed by J. Willard Gibbs and Paul Appell around 1879–1899, it addresses limitations of standard methods for constraints like velocity-dependent relations (e.g., no-slipping conditions) that cannot be integrated to form. The is described by q ( $3N - M, with N particles and M ), while quasi-coordinates \rho ( $3N - M - L, with L non-holonomic constraints) represent velocity components, related via \dot{q}_i = \sum_j a_{ij} \dot{\rho}_j + b_i. The Appellian function A is the expressed in these terms: A = T(\mathbf{q}, \dot{\rho}, t) - V(\mathbf{q}, t), where T is the quadratic in \dot{\rho}. The Gibbs-Appell equations of motion for the quasi-coordinates are \sum_k \frac{d}{dt} \left( \frac{\partial A}{\partial \dot{\rho}_k} \right) \dot{\rho}_k - \frac{\partial A}{\partial \rho_j} + \sum_{k,s} \left[ \frac{d}{dt} \left( \frac{\partial A}{\partial \dot{\rho}_s} \right) - \frac{\partial A}{\partial \rho_s} \right] a_{jk} \dot{\rho}_s = Q_j, where Q_j are generalized forces; a simplified form for certain systems is \partial A / \partial \dot{\rho}_j = 0 after substitution. This formulation incorporates constraint forces naturally through the quasi-velocity structure, avoiding explicit multipliers, and is particularly effective for or systems with constraints. It reduces the problem to independent velocity equations while preserving the . An illustrative example is the rolling disk without slipping on a horizontal plane, a classic non-holonomic system. For a homogeneous disk of m and a, the generalized coordinates are the center position (X, Y), lean angle \theta, and rotations \phi, \psi; non-holonomic constraints enforce \dot{X} = a (\dot{\psi} \sin \phi \sin \theta - \dot{\theta} \cos \phi), \dot{Y} = a (\dot{\psi} \cos \phi \sin \theta + \dot{\theta} \sin \phi), and no-slip relating linear and angular velocities. Quasi-coordinates are chosen as angular velocity components \omega_1, \omega_2, \omega_3 in body frame. The Appellian function A = \frac{1}{2} m a^2 (\omega_1^2 + \omega_2^2 + \omega_3^2 \sin^2 \theta) - m g a (1 - \cos \theta) (adjusted for inertia), and the equations yield relations like \partial A / \partial \omega_1 = I \omega_1 + cross terms = 0, solving for \dot{\psi}, \dot{\theta}, \omega_3 in terms of initial conditions, enabling prediction of wobbling or steady rolling without integrating non-integrable constraints directly. This demonstrates Appellian mechanics' utility for contact problems in rigid body dynamics.

Field theory extensions

Lagrangian field theory

Lagrangian field theory extends the principles of Lagrangian mechanics from discrete particles to continuous systems, such as scalar, vector, or tensor fields distributed over spacetime. In this formalism, the dynamics of a field \phi(x) are described by a Lagrangian density \mathcal{L}(\phi, \partial_\mu \phi, x), which depends on the field, its first spacetime derivatives, and possibly the position x. The action functional is then given by the spacetime integral S = \int \mathcal{L} \, d^4x, where the integral is typically over a four-dimensional Minkowski spacetime volume. The arise from the , requiring the to be stationary with respect to arbitrary variations of the \delta \phi, to appropriate conditions. This leads to the field-theoretic Euler-Lagrange equations: \partial_\mu \left( \frac{\partial \mathcal{L}}{\partial (\partial_\mu \phi)} \right) - \frac{\partial \mathcal{L}}{\partial \phi} = 0, which generalize the particle case to partial derivatives over indices \mu = 0,1,2,3. These equations determine the configurations that extremize the , analogous to the principle of least for particles but applied to histories. A canonical example is the real obeying the Klein-Gordon , with Lagrangian density \mathcal{L} = \frac{1}{2} \partial_\mu \phi \partial^\mu \phi - \frac{1}{2} m^2 \phi^2, where m is the mass parameter and the is (+,-,-,-). Substituting this into the Euler-Lagrange yields (\square + m^2) \phi = 0, where \square = \partial_\mu \partial^\mu is the d'Alembertian operator, describing a free relativistic . This form captures the relativistic invariance essential for fields in . For vector fields, such as the , the Lagrangian density is \mathcal{L} = -\frac{1}{4} F_{\mu\nu} F^{\mu\nu}, where F_{\mu\nu} = \partial_\mu A_\nu - \partial_\nu A_\mu is the field strength tensor derived from the four-potential A_\mu. The Euler-Lagrange equations from this reproduce in vacuum, \partial_\mu F^{\mu\nu} = 0, highlighting the gauge invariance of the theory under A_\mu \to A_\mu + \partial_\mu \Lambda. In theories with gauge symmetries, such as , the is invariant under local transformations, but the four-potential A_\mu has redundant . Constraints are imposed either by , such as the Lorenz gauge \partial_\mu A^\mu = 0, which simplifies the equations while preserving physical content, or by introducing Lagrange multipliers to enforce the constraints directly in the action. This approach ensures the theory remains consistent with the underlying symmetries.

Hamiltonian field theory

Hamiltonian field theory extends the canonical formalism of analytical mechanics to continuous systems, treating fields as infinite collections of degrees of freedom in an infinite-dimensional phase space. In this formulation, a field \phi(\mathbf{x}, t) and its canonically conjugate momentum density \pi(\mathbf{x}, t) serve as the fundamental variables, where \pi = \frac{\partial \mathcal{L}}{\partial (\partial_0 \phi)} and \mathcal{L} is the Lagrangian density. The Hamiltonian density \mathcal{H} is obtained via the Legendre transform \mathcal{H} = \pi \partial_0 \phi - \mathcal{L}, and the total Hamiltonian H = \int \mathcal{H} \, d^3x generates time evolution through functional derivatives. The dynamics follow Hamilton's equations adapted to fields: \partial_0 \phi = \frac{\delta H}{\delta \pi}, \quad \partial_0 \pi = -\frac{\delta H}{\delta \phi}. These equations describe the evolution of the field configuration \phi(\mathbf{x}) and momentum \pi(\mathbf{x}) on equal-time hypersurfaces, with the comprising all possible functional pairs (\phi(\mathbf{x}), \pi(\mathbf{x})). The classical structure is extended to this infinite-dimensional setting using functional derivatives, \{A, B\} = \int \left( \frac{\delta A}{\delta \phi} \frac{\delta B}{\delta \pi} - \frac{\delta B}{\delta \phi} \frac{\delta A}{\delta \pi} \right) d^3x. For a real scalar field with Lagrangian density \mathcal{L} = \frac{1}{2} \partial^\mu \phi \partial_\mu \phi - \frac{1}{2} m^2 \phi^2 (in Minkowski metric with signature (+,-,-,-)), the momentum is \pi = \partial_0 \phi, and the Hamiltonian density is \mathcal{H} = \frac{1}{2} \pi^2 + \frac{1}{2} (\nabla \phi)^2 + \frac{1}{2} m^2 \phi^2. This yields the Klein-Gordon equation upon substitution into Hamilton's equations. For the Dirac field, described by a spinor \psi, the Lagrangian density \mathcal{L} = \bar{\psi} (i \gamma^\mu \partial_\mu - m) \psi (with \bar{\psi} = \psi^\dagger \gamma^0) leads to the conjugate momentum density \pi = i \psi^\dagger. The Hamiltonian density is \mathcal{H} = \bar{\psi} (-i \boldsymbol{\alpha} \cdot \boldsymbol{\nabla} + \beta m) \psi, where \boldsymbol{\alpha} = \gamma^0 \boldsymbol{\gamma} and \beta = \gamma^0. The resulting equations reproduce the Dirac equation (i \gamma^\mu \partial_\mu - m) \psi = 0. In gauge theories, such as electromagnetism, the Hamiltonian formulation often introduces constraints; for instance, the primary constraint from the temporal gauge component yields Gauss's law \boldsymbol{\nabla} \cdot \mathbf{E} = 0 (in the absence of sources), which is preserved under time evolution. Such second-class constraints are handled by introducing Dirac brackets, which modify the Poisson structure to enforce the constraints consistently: \{A, B\}_D = \{A, B\} - \{A, \phi_i\} (C^{-1})_{ij} \{\phi_j, B\}, where \phi_i are the constraints and C_{ij} = \{\phi_i, \phi_j\}. This ensures the theory's gauge invariance and physical consistency in the constrained phase space.

Symmetries and conservation

Noether's theorem

Noether's theorem, first articulated by mathematician in her 1918 paper "Invariant Variation Problems," provides a general framework linking continuous of the action principle in analytical mechanics to corresponding conservation laws. In the context of , the theorem asserts that if the action S = \int_{t_1}^{t_2} L(\mathbf{q}, \dot{\mathbf{q}}, t) \, dt, where L is the , remains invariant under an infinitesimal transformation of the coordinates and time, then there exists a associated with that symmetry. This invariance is understood in the sense that the variation of the Lagrangian satisfies \delta L = \frac{d}{dt} F(\mathbf{q}, t) for some function F, ensuring \delta S = 0 up to boundary terms. To formalize this in analytical mechanics, consider an transformation \delta q^i = \epsilon [K](/page/K)^i(\mathbf{q}, t), where \epsilon is an parameter and K^i defines the of the , alongside a possible transformation of time \delta t = \epsilon \Xi(t, \mathbf{q}). The condition for the transformation to be a symmetry of is that the change in the is a total time : \delta L = \epsilon \left( \frac{\partial L}{\partial q^i} K^i + \frac{\partial L}{\partial \dot{q}^i} \dot{K}^i + \frac{\partial L}{\partial t} \Xi \right) + \epsilon L \frac{d \Xi}{dt} = \frac{d}{dt} \left( \epsilon G(\mathbf{q}, t) \right), for some G. This ensures the varied action \delta S = \left[ \frac{\partial L}{\partial \dot{q}^i} \delta q^i \right]_{t_1}^{t_2}, which vanishes for fixed endpoints. The proof proceeds by considering the variation of under this while incorporating the . Substituting the transformation into \delta S = 0 yields \delta S = \int_{t_1}^{t_2} \left( \frac{\partial L}{\partial q^i} \delta q^i + \frac{\partial L}{\partial \dot{q}^i} \delta \dot{q}^i \right) dt = 0. Using the Euler-Lagrange equations \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}^i} \right) - \frac{\partial L}{\partial q^i} = 0 on solutions (on-shell), transforms the variation to a : \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}^i} \delta q^i - G \right) = 0. Thus, the quantity Q = p_i \delta q^i - G, where p_i = \frac{\partial L}{\partial \dot{q}^i} is the generalized , is conserved: \frac{dQ}{dt} = 0. This holds for any group parameterized by \epsilon, establishing the direct between symmetries and integrals of motion. Illustrative examples in classical mechanics highlight the theorem's power. For time-translation invariance, where \delta t = \epsilon and \delta q^i = 0, the is independent of explicit time dependence (\Xi = 1, G = L), yielding conservation of the H = \sum_i p_i \dot{q}^i - L, interpreted as . Spatial translation symmetry, \delta q^i = \epsilon^a (constant \epsilon^a in direction a), with \Xi = 0 and G = 0, conserves linear momentum p_a = \sum_i p_i \frac{\partial q^i}{\partial x^a}. Rotational invariance, \delta q^i = \epsilon^{ab} (x_a \frac{\partial q^i}{\partial x_b} - x_b \frac{\partial q^i}{\partial x_a}), leads to conservation of angular momentum J^{ab} = \sum_i p_i (q^i_a \frac{\partial q^i}{\partial x_b} - q^i_b \frac{\partial q^i}{\partial x_a}) - \text{corresponding } G. These cases demonstrate how spacetime symmetries underpin fundamental conservations in isolated systems. The theorem extends naturally to field theories, where the action is S = \int \mathcal{L}(\phi, \partial_\mu \phi, x) \, d^4 x over . For an infinitesimal \delta \phi = \epsilon K(\phi, x), symmetry requires \delta \mathcal{L} = \partial_\mu (\epsilon \Lambda^\mu). The proof mirrors the mechanical case: varying the action and using the Euler-Lagrange equations \partial_\mu \left( \frac{\partial \mathcal{L}}{\partial (\partial_\mu \phi)} \right) - \frac{\partial \mathcal{L}}{\partial \phi} = 0 yields a j^\mu = \frac{\partial \mathcal{L}}{\partial (\partial_\mu \phi)} \delta \phi - \Lambda^\mu, \quad \partial_\mu j^\mu = 0 on-shell, with the conserved charge Q = \int j^0 \, d^3 x. Internal symmetries, such as the U(1) \delta \phi = i \epsilon \phi for a complex scalar , where \mathcal{L} is invariant and \Lambda^\mu = 0, produce the j^\mu = i \left( \phi^* \partial^\mu \phi - \phi \partial^\mu \phi^* \right), corresponding to electric charge conservation in gauge theories.

Conserved quantities and symmetry groups

In analytical mechanics, conserved quantities emerge as direct consequences of symmetries in the system's Lagrangian or Hamiltonian, providing invariants that simplify the integration of equations of motion. According to Noether's theorem, continuous symmetries of the action lead to corresponding conserved quantities, such as energy from time-translation invariance and momentum from spatial translations. These quantities remain constant along trajectories, enabling reductions in the degrees of freedom and revealing underlying structures in dynamical systems. Energy conservation arises from the invariance of the under time translations, implying that the H (or the L for time-independent cases) is constant. For a with L(q, \dot{q}, t), if \frac{\partial L}{\partial t} = 0, the total energy E = \sum_i \dot{q}_i \frac{\partial L}{\partial \dot{q}_i} - L is conserved. This principle underpins the stability of mechanical s, as seen in isolated conservative forces where converts to kinetic without dissipation. Linear momentum conservation follows from translational invariance of the Lagrangian under spatial shifts, yielding the total momentum \mathbf{p} = \sum_i m_i \mathbf{v}_i as a constant vector. In a system free from external forces, this implies the center of mass moves with uniform velocity, a key feature in multi-particle dynamics. Angular momentum conservation stems from rotational invariance, conserving the total angular momentum \mathbf{L} = \sum_i \mathbf{r}_i \times \mathbf{p}_i. For central force problems, such as gravitational or electrostatic interactions, this symmetry confines motion to a plane, with \mathbf{L} perpendicular to the orbital plane and constant in magnitude and direction. Broader symmetry groups classify these conservations systematically. The Galilean group, encompassing translations, rotations, and boosts in non-relativistic , conserves the center-of-mass motion under boosts, linking to the uniformity of inertial frames. In relativistic contexts, the extends this to spacetime symmetries, conserving the energy-momentum and tensor in field theories, essential for consistent relativistic formulations. A notable example of hidden symmetry is the Runge-Lenz vector in the Kepler problem, conserved for inverse-square potentials like the hydrogen atom or planetary orbits. Defined as \mathbf{A} = \mathbf{p} \times \mathbf{L} - m k \hat{\mathbf{r}}, where k is the force constant, it points toward the periapsis and arises from an underlying SO(4) group, closing the algebra with to yield closed elliptical orbits. This non-obvious conservation integrates the system fully, beyond the basic from rotational invariance. Discrete symmetries, such as (spatial reflection), do not yield continuous conserved currents in but influence the form of potentials and trajectories; for instance, parity-invariant systems exhibit mirror-symmetric solutions, though violations can occur in certain non-central forces. In the framework of analytical mechanics, conserved quantities from actions are captured by the momentum map J: (M, \omega) \to \mathfrak{g}^*, where M is the , \omega the form, and \mathfrak{g} the . This map associates infinitesimal symmetries to vector fields, generating conserved functions that Poisson-commute with the dynamics, providing a geometric unification for reductions like those in central force problems.

References

  1. [1]
    None
    Below is a merged summary of Analytical Mechanics based on all provided segments. To retain all information in a dense and organized manner, I will use a combination of narrative text and a table in CSV format for key details (definitions, principles, formulations, equations, and URLs). This ensures comprehensive coverage while maintaining clarity and avoiding redundancy.
  2. [2]
    [PDF] ANALYTICAL MECHANICS
    This is a corrected reprint of a work first published in early 2002, by Oxford University. Press, and which went out of print shortly thereafter.
  3. [3]
    None
    ### Summary of Analytical Mechanics from MEEN 617 Handout 4
  4. [4]
    [PDF] The Origins of Analytic Mechanics in the 18th century - HAL-SHS
    This is the history of the transformation of mechanics from a geometric science into an analytical one, or even, according to the Lagrange's dictum, into “a ...
  5. [5]
    Mécanique analytique | work by LaGrange - Britannica
    In them Lagrange tried to substitute an algebraic foundation for the existing and problematic analytic foundation of calculus—although ultimately unsuccessful, ...
  6. [6]
    Variable Mass in Newtonian Mechanics - Physics Stack Exchange
    Oct 27, 2020 · In Newtonian mechanics mass cannot disappear, it is assumed to be constant. To imagine why it is not a good idea to have variable mass, ...Missing: holonomic constraints<|separator|>
  7. [7]
    [PDF] Chapter 1 A Review of Analytical Mechanics - MIT OpenCourseWare
    A few examples related to the concepts of inertial frames are: 1. In a rotating frame, the transformation is given by x/ y/. = cos(θ) sin(θ) x. −sin(θ) cos(θ).
  8. [8]
    Lagrangian Mechanics – Engineering Systems Dynamics, Modelling ...
    ... mechanics (1788). Lagrange's approach has advantages over that of Newton's, specifically for analyzing complex multi-domain, multi-component systems.2 Lagrangian Mechanics · 2.3 Lagrange's Equations For... · 2.20 Example: An Extensible...
  9. [9]
    EOM for Double Pendulum using Newton's Laws of Motion
    Feb 1, 2024 · Using Newton's laws of motion, for a double pendulum, you would have to explicitly include the tension in the equations of motion.
  10. [10]
    D'Alembert's Principle: The Original Formulation and Application in ...
    D'Alembert's Principle: The Original Formulation and Application in Jean d'Alembert's Tradé de Dynamique (1743) ... First published: July 1985. https://doi ...
  11. [11]
    Principle of least action - Scholarpedia
    Jun 5, 2015 · The principle of least action is the basic variational principle of particle and continuum systems.
  12. [12]
    [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.
  13. [13]
    The history of the Méchanique analitique | Lettera Matematica
    Jun 6, 2014 · La Méchanique analitique was published in 1788, when its author was established in Paris. It was immediately clear to all experts that it was a ...<|control11|><|separator|>
  14. [14]
    XV. On a general method in dynamics - Journals
    On a general method in dynamics; by which the study of the motions of all free systems of attracting or repelling points is reduced to the search and ...
  15. [15]
    [PDF] Canonical transformations from Jacobi to Whittaker - Craig Fraser
    Jun 21, 2022 · The idea of a canonical transformation emerged in 1837 in the course of Carl Jacobi's researches in analytical dynamics.
  16. [16]
    [PDF] Lectures on Constrained Systems - arXiv
    Oct 11, 2010 · These lecture notes have been prepared as a basic introduction to the theory of constrained systems which is how basic forces of nature ...
  17. [17]
    [PDF] Generalized Coordinates, Lagrange's Equations, and Constraints
    1 Cartesian Coordinates and Generalized Coordinates. The set of coordinates used to describe the motion of a dynamic system is not unique.
  18. [18]
    [PDF] Lagrangian Dynamics: Generalized Coordinates and Forces Lecture ...
    The generalized coordinates of a system (of particles or rigid body or rigid bodies) is the natural, minimal, complete set of parameters by which you can.<|control11|><|separator|>
  19. [19]
  20. [20]
  21. [21]
    [PDF] Single and Double plane pendulum
    In the double pendulum We know there should be only two generalized coordinates, since there are 3N=6 coordinates, and m=4 constraints, so n=3N-m=6-4=2. We ...
  22. [22]
    [PDF] Lecture #9 Virtual Work And the Derivation of Lagrange's Equations
    Principle of virtual work: The necessary and sufficient conditions for the static equilibrium of an initially motionless scleronomic system.
  23. [23]
    [PDF] CHAPTER 6 LAGRANGE'S EQUATIONS (Analytical Mechanics)
    3: Double pendulum: it consists of two particles and two massless rigid ... • generalized coordinates - any number of variables needed to completely ...
  24. [24]
    [PDF] Physics 5153 Classical Mechanics D'Alembert's Principle and The ...
    In this lecture, we will discuss the extension of the principle of virtual work to dynamical system. This extension is based on the work of D'Alembert, and is ...Missing: analytical | Show results with:analytical
  25. [25]
    [PDF] Chapter 1 D'Alembert's principle and applications
    Feb 1, 2014 · 1.1 D'Alembert's principle. The principle of virtual work states that the sum of the incremental virtual works done by all external forces Fi ...Missing: analytical | Show results with:analytical
  26. [26]
    [PDF] Energy Methods: Lagrange's Equations - MIT OpenCourseWare
    They are the beginning of a complex, more mathematical approach to mechanics called analytical dynamics. In this course we will only deal with this method at an ...
  27. [27]
    [PDF] Chapter 2. Lagrangian Analytical Mechanics
    The goal of this chapter is to describe the Lagrangian formalism of analytical mechanics, which is extremely useful for obtaining the differential equations of ...
  28. [28]
    [PDF] Analytical Dynamics: Lagrange's Equation and its Application
    Equation (28) is the Lagrange equation for systems where the virtual work may be expressed as a variation of a potential function, V . In the frequent cases ...
  29. [29]
    [PDF] The Lagrangian formalism for classical mechanics: 8/31/17
    Dec 16, 2017 · These notes were taken in UT Austin's M393C (Topics in Mathematical Physics) class in Fall 2017, taught by. Thomas Chen.
  30. [30]
    [PDF] Classical Mechanics - UC Homepages
    Nov 30, 2023 · where the two terms in the parentheses cancel due to Euler's equation (9.62). ... This is a special case of Euler's theorem: For a homogeneous ...
  31. [31]
    [PDF] Lecture Notes on Classical Mechanics (A Work in Progress)
    Jan 13, 2022 · These lecture notes are based on material presented in both graduate and undergraduate mechanics classes which I have taught on several ...
  32. [32]
    Galilean invariance in Lagrangian mechanics - AIP Publishing
    Oct 1, 2015 · The importance of requiring such invariance comes from its connection with the conservation of linear momentum of an isolated system. Within ...
  33. [33]
    [PDF] Lagrange's Equations
    Friction in Lagrange's Formulation. Massachusetts Institute of Technology ... Rayleigh's Dissipation Function. • For systems with conservative and non ...
  34. [34]
    [PDF] Rayleigh's dissipation function at work - arXiv
    Feb 28, 2015 · It is shown that the Rayleigh's dissipation function can be successfully ap- plied in the solution of mechanical problems involving friction ...
  35. [35]
    [PDF] 2. The Lagrangian Formalism - DAMTP
    ... time derivatives and (possibly) time t is called a constant of motion (or a conserved quantity) if the total time derivative vanishes. dF dt. = n. X j=1 ✓@F.
  36. [36]
    [PDF] Chapter 2 Lagrange's and Hamilton's Equations - Rutgers Physics
    which is known as D'Alembert's Principle. This gives an equation which ... Atwood's machine consists of two blocks of mass m1 and m2 attached by an.
  37. [37]
    [PDF] Part IB - Variational Principles - Dexter Chua
    The Legendre transform is an important tool in classical dynamics and thermo- dynamics. In classical dynamics, it is used to transform between the Lagrangian.
  38. [38]
    [PDF] Classical Mechanics Review - Math (Princeton)
    Feb 24, 2022 · the Legendre transform applied to a convex function yields a convex function and applying it twice results in the original function. We also ...
  39. [39]
    [PDF] 4. The Hamiltonian Formalism - DAMTP
    This is currently an active area of research. 4.4 Canonical Transformations. There is a way to write Hamilton's equations so that they look even more symmetric.Missing: seminal sources
  40. [40]
    [PDF] ON A GENERAL METHOD IN DYNAMICS By William Rowan Hamilton
    The paper On a General Method in Dynamics has also been republished in The Mathe- matical Papers of Sir William Rowan Hamilton, Volume II: Dynamics, edited for ...
  41. [41]
    [PDF] L03: Kepler problem & Hamiltonian dynamics - MIT Mathematics
    4 Kepler's problem and Hamiltonian dynamics. Why do we study applied ... In plane polar coordinates dr dt. = ˙rr + r. ˙ θˆθ,. (57a) d. 2 r dt2. = (r- r.
  42. [42]
    [PDF] Lecture I: Constrained Hamiltonian systems - Cosmo-ufes
    Dec 15, 2014 · constitutes primary and secondary constraints, first-class and second-class constraints, and Dirac brackets. In quantum theory, the operator ...
  43. [43]
    [PDF] NL2735 Poisson brackets
    The Poisson bracket defined this way satisfies the properties (i), (ii), and (iv). The operator B is chosen so that the Jacobi identity (iii) is also satisfied.
  44. [44]
    [PDF] Hamiltonian Mechanics
    The resulting 2N Hamiltonian equations of motion for qi and pi have an elegant symmetric form that is the reason for calling them canonical equations.Missing: seminal sources
  45. [45]
    19: The Principle of Least Action - Feynman Lectures
    The rule says that in going from one point to another in a given amount of time, the kinetic energy integral is least, so it must go at a uniform speed.
  46. [46]
    [PDF] The Principle of Least Action in Dynamics - DAMTP
    Apr 25, 2013 · The first law states that motion of a body at constant velocity is self-sustaining, and no force is needed. Velocity is the time derivative of ...
  47. [47]
    [PDF] Calculus of Variations 6: Hamilton's Principle - UNL Math
    By Hamilton's principle, the equation of motion is. ∂L. ∂θ. − d dt. ∂L. ∂. ˙ θ. = 0, which reduces to. ¨ θ + g. Λ sinθ = 0. Page 3. In this example the ...
  48. [48]
    [PDF] History of Two Fundamental Principles of Physics: Least Action and ...
    Mar 9, 2014 · Later, the LAP would be developed by Pierre-Louis Moreau de Maupertuis (1698-1759), Euler and. Lagrange2. The LAP is indissolubly linked to the ...<|control11|><|separator|>
  49. [49]
    [PDF] Jacobi's Action and the Density of States J. David Brown - arXiv
    2 INTRODUCTION. Jacobi's form of the action principle involves variations at fixed energy, rather than the variations at fixed time used in Hamilton's principle ...
  50. [50]
    Jacobi's principle and the disappearance of time | Phys. Rev. D
    Feb 23, 2010 · 1. Action and equations of motion. In general, Jacobi's principle is the vanishing of the variation of an action of the form.
  51. [51]
    From conservation of energy to the principle of least action: A story line
    An eight sentence history of Newtonian mechanics shows how much the subject has developed since Newton introduced F = dp/dt in the second half of the 1600s.
  52. [52]
    The Principle of Least Action for Reversible Thermodynamic ...
    Therefore, the brachistochrone problem is, in fact, the application of the principle of least action with the harmonic average velocity, S A B / ( ∫ A B d s / ...<|separator|>
  53. [53]
    [PDF] Path Integrals and the Classical Limit
    Feynman: Time evolution as sum over paths. Phases evolve by. Why does this give the classical limit? Big, heavy objects: S >> hbar. S varies rapidly as x(t) ...
  54. [54]
    Structure and Interpretation of Classical Mechanics: Canonical ...
    We can show directly from the action principle that time evolution generates a symplectic transformation. Recall that the Lagrangian action S is. S[q](t1,t2)=∫ ...
  55. [55]
    Hamilton-Jacobi equation - Scholarpedia
    Oct 21, 2011 · The Hamilton-Jacobi equation is used to generate particular canonical transformations that simplify the equations of motion.
  56. [56]
    [PDF] The Early History of Hamilton-Jacobi Dynamics 1834–1837
    May 2, 2023 · Hamilton's method as set forth in the 1834 paper was superseded in later me- chanics by the theory elaborated in his 1835 paper. In the next ...
  57. [57]
    [2101.03830] An overview of the Hamilton--Jacobi theory - arXiv
    Jan 11, 2021 · This work is devoted to review the modern geometric description of the Lagrangian and Hamiltonian formalisms of the Hamilton--Jacobi theory.
  58. [58]
    [PDF] geometric derivation of the delaunay variables and ... - Caltech
    The Delaunay variables were first introduced in Delaunay (1860) and have been frequently used as canonical variables in celestial mechanics. In particular, they.
  59. [59]
    [PDF] Chapter 5 Perturbation Theory - MIT OpenCourseWare
    In this chapter we will discuss time dependent perturbation theory in classical mechanics. Many problems we have encountered yield equations of motion that ...
  60. [60]
    [PDF] solutions - CLASSE (Cornell)
    Prove that the problem is reduced to N - 1 degrees of freedom by using the Routhian as a new ... CENTRAL FORCE problem,. We know how to reduce this to a 1D system ...
  61. [61]
    [PDF] The Gibbs–Appell equations of motion - HAL
    Nov 20, 2016 · The relationship between the Gibbs-Appell equations of motion and Lagrange's equations of motion is discussed. Auxiliary results thal facilitate ...
  62. [62]
    The Gibbs–Appell equations of motion | American Journal of Physics
    Sep 1, 1988 · The relationship between the Gibbs–Appell equations of motion and Lagrange's equations of motion is discussed. Auxiliary results that facilitate ...
  63. [63]
    Lagrangian formalism for fields - Scholarpedia
    Aug 30, 2010 · The Lagrangian formalism is one of the main tools of the description of the dynamics of a vast variety of physical systems.Scalar field · Vector field · Spinor fields · Supersymmetric field theories
  64. [64]
    [PDF] The Classical Theory of Fields
    Page 1. Landau. Lifshitz. The Classical. Theory of Fields. Third Revised English Edition. Course of Theoretical Physics. Volume 2. CD CD o CO. J.. fl*. CD —. E.
  65. [65]
    [PDF] Classical Hamiltonian Field Theory
    Aug 26, 2009 · This is a set of notes introducing Hamiltonian field theory, with focus on the scalar field. Contents. 1 Lagrangian Field Theory. 1. 2 ...
  66. [66]
  67. [67]
    [PDF] Lagrangian constraint analysis of first-order classical field theories ...
    Sep 21, 2020 · This is the familiar Gauss law, telling us that, in the absence of sources, the electric field is divergenceless. Note that this is an on ...<|control11|><|separator|>
  68. [68]
    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 ...
  69. [69]
    [PDF] Noether's theorem in classical mechanics revisited - arXiv
    Aug 26, 2006 · Noether's[1] theorem, presented in 1918, is one of the most beautiful theorems in physics. It relates symmetries of a theory with its laws ...
  70. [70]
    [PDF] What is Noether's Theorem? | OSU Math
    Abstract. Noether's theorem states that given a physical system, for every infinitesimal symmetry, there is a corresponding law of symme-.
  71. [71]
    [PDF] Noether's Two Theorems
    Proof : Noether's Identity = Integration by Parts pr v(L) + LDiv ξ = QE(L) ... Both have their origins in the classical mechanics of the nine- teenth ...
  72. [72]
    [PDF] A short review on Noether's theorems, gauge symmetries and ... - arXiv
    Aug 30, 2017 · Noether's theorem is often associated to field theory, but it is a property of any system that can be derived from an action and possesses some ...
  73. [73]
    [PDF] HAMILTONIAN MECHANICS, NOETHER'S THEOREM
    Classical Mechanics. LECTURE 28: HAMILTONIAN. MECHANICS,. NOETHER'S THEOREM ... conserved quantities. ▻ From before, conjugate momentum : pk = ∂L. ∂ ˙qk.
  74. [74]
    17 Symmetry and Conservation Laws - Feynman Lectures - Caltech
    In classical physics there are a number of quantities which are conserved—such as momentum, energy, and angular momentum. Conservation theorems about ...
  75. [75]
    [PDF] Classical Dynamics - DAMTP
    The fundamental principles of classical mechanics were laid down by Galileo and New- ton in the 16th and 17th centuries. In 1686, Newton wrote the Principia ...
  76. [76]
    [2212.04931] Runge-Lenz Vector as a 3d Projection of SO(4 ... - arXiv
    Dec 9, 2022 · We show, using the methods of geometric algebra, that Runge-Lenz vector in the Kepler problem is a 3-dimensional projection of SO(4) moment map ...
  77. [77]
    On the global symmetry of the classical Kepler problem
    These actions are generated by the Hamiltonian function, the angular momentum and the Runge-Lenz vector. The symmetry group is SO(4) for negative and SO(1,3)0 ...
  78. [78]
    [PDF] The law of parity conservation and other symmetry laws of physics
    Whereas the continuous symmetries always lead to conservation laws in classical mechanics, a discrete symmetry does not.Missing: quantities | Show results with:quantities
  79. [79]
    [PDF] On Symplectic Reduction in Classical Mechanics - PhilSci-Archive
    Jul 21, 2005 · that a momentum map representing a conserved quantity has components. In the symplectic case, differentiating Φ. ∗ gω = ω implies that the ...