Fact-checked by Grok 2 weeks ago

Hamilton's principle

Hamilton's principle is a foundational in , stating that the actual path of a between two fixed points in configuration space over a specified time interval renders the action functional , where the action is defined as the time of the L = T - V, with T denoting and V . This stationarity condition, derived from the , yields the Euler-Lagrange equations, which are mathematically equivalent to Newton's second law for conservative systems. Formulated by Irish mathematician and astronomer in his 1834 paper "On a General Method in Dynamics," the principle builds on earlier ideas such as of least action (1744) and Joseph-Louis Lagrange's (1788), but Hamilton extended it to a more general framework using a to solve for system trajectories. Mathematically, for a system with q_i, the principle asserts \delta \int_{t_1}^{t_2} L(q_i, \dot{q}_i, t) \, dt = 0, where the variation \delta is taken over admissible paths with fixed endpoints. This approach unifies diverse mechanical phenomena, from particle motion in gravitational fields to , by reducing the problem to extremizing a single scalar functional rather than solving vectorial differential equations directly. Beyond , Hamilton's principle underpins , where the is transformed via a Legendre transform into the , facilitating phase-space analysis and quantization in . It also extends to field theories, where the action principle governs stationary paths for fields, and to non-conservative systems through modifications like Rayleigh's dissipation function. The principle's elegance lies in its predictive power: for instance, it naturally derives conservation laws via when the exhibits symmetries.

Historical background

Origins in variational calculus

The calculus of variations originated in the early as mathematicians sought to determine curves or paths that extremize certain quantities, such as length or time. A pivotal early problem was the brachistochrone, posed by in 1696, which asked for the curve of fastest descent between two points under ; this challenge stimulated variational techniques, with Leonhard Euler providing a systematic solution in his 1744 treatise Methodus inveniendi lineas curvas maximi minimive proprietate gaudentes. Euler's work established methods for optimizing integrals depending on unknown functions, marking the field's formal inception. In 1766, Euler advanced these ideas by coining the term "" and applying it to geodesics on surfaces, where the shortest path between points satisfies a variational condition. Central to this development was the concept of functionals—mappings from functions to real numbers, typically expressed as integrals—and paths, which are functions that render the functional invariant under small perturbations. During the late 18th and early 19th centuries, mathematicians like Euler and Lagrange refined these notions, emphasizing paths where variations yield zero change in the functional value, thus achieving maxima, minima, or saddle points. The basic form of such a functional is J = \int_{a}^{b} F(x, y, y') \, dx, with the stationarity condition \delta J = 0, indicating an extremal path. Joseph-Louis Lagrange further generalized variational principles to mechanics in his 1788 work Mécanique Analytique, where he introduced the principle of least action for constrained systems, positing that natural motions minimize an action integral derived from kinetic and potential energies. This extension transformed variational calculus from geometric optimization to a foundational tool for dynamical laws, accommodating holonomic constraints through multipliers and enabling derivations of equations of motion without reference to forces. Lagrange's formulation unified earlier scattered results, providing a rigorous analytical framework that influenced subsequent developments in classical mechanics.

Hamilton's contributions and formulation

William Rowan , an Irish mathematician and astronomer, developed his foundational work on dynamics during the early 1830s, building on his prior research in from the 1820s. As Andrews Professor of Astronomy at , Hamilton introduced the in his 1827–1828 paper "Theory of Systems of Rays," published in the Transactions of the Royal Irish Academy, where it represented the in ray trajectories governed by of least time. This optical framework, involving partial differential equations for ray propagation, inspired Hamilton to draw an explicit analogy to mechanical systems, viewing particle motions as analogous to light rays in a higher-dimensional configuration space, thereby unifying principles of optics and mechanics through variational methods. In his seminal 1834 paper, "On a General Method in Dynamics," published in the Philosophical Transactions of the Royal Society, Hamilton extended this optical analogy to general dynamics by introducing the principal function, initially denoted as the characteristic function V, which encapsulates the action for systems of attracting or repelling points. Hamilton formulated the core principle as the stationarity of the action integral, stating that for the true path of motion between fixed endpoints in time and configuration space, the variation of the definite integral V = \int_{t_1}^{t_2} (T - U) \, dt vanishes, where T is the kinetic energy and U the potential energy—a condition he termed the "law of stationary action." This reformulation reduced the study of complex dynamical systems to the search and differentiation of a single central relation, V, satisfying two first-order partial differential equations, and predated the full integration of Lagrangian mechanics by emphasizing a unified variational approach over coordinate-specific equations. Hamilton further refined his theory in the 1835 "Second Essay on a General Method in ," also in the Philosophical Transactions, where he introduced the time-explicit principal function S, defined as S = \int_{t_1}^{t_2} L(q, \dot{q}, t) \, dt with L = T - U, asserting that this is stationary for the actual path among virtual paths connecting specified initial and final states. This extension applied the method to perturbed motions in , such as binary systems, and proposed a " of principal functions" as a new tool for physics. 's innovations, which built briefly on the variational of Euler and Lagrange without deriving new equations, profoundly influenced contemporaries, notably , who in 1837 generalized the theory to time-dependent forces and developed the Hamilton-Jacobi equation for in his papers "Zur Theorie der Variationsrechnung" and "Über die Reduction."

Core concepts in classical mechanics

The Lagrangian and action integral

In classical mechanics, the framework of generalized coordinates, introduced by Joseph-Louis Lagrange, allows the description of a system's configuration using a set of independent parameters q_i (where i = 1, 2, \dots, n, with n being the number of degrees of freedom), rather than Cartesian coordinates. These coordinates q_i(t) and their time derivatives \dot{q}_i(t) fully specify the positions and velocities of the system's components, enabling a compact formulation of the dynamics. The Lagrangian L, a central quantity in this formulation, is defined as the difference between the T and the V of the system: L(q_i, \dot{q}_i, t) = T(q_i, \dot{q}_i) - V(q_i, t). Here, T typically depends quadratically on the velocities \dot{q}_i, reflecting the system's , while V captures the conservative forces and may explicitly depend on time t for non-autonomous systems. This form of the ensures that the derived from it reproduce Newton's laws for standard mechanical systems. The action integral S, also known as the action functional, is constructed by integrating the over a time from fixed initial time t_1 to final time t_2: S[q(t)] = \int_{t_1}^{t_2} L(q_i(t), \dot{q}_i(t), t) \, dt, with the endpoints of the path constrained such that q_i(t_1) and q_i(t_2) are specified. This integral quantifies the "total action" along a particular q(t) in the space of . According to Hamilton's principle, the true physical of the is the one among all possible paths connecting the fixed endpoints that renders stationary, meaning its first variation vanishes. This stationarity condition implies that the actual path is a or extremum of S with respect to small perturbations in the , distinguishing it from nearby virtual paths while preserving the endpoints. Such a variational unifies diverse phenomena under a single principle, extending beyond simple minimization to include cases where may be maximized or neither strictly.

Stationarity condition and virtual paths

In Hamilton's principle, the stationarity condition asserts that the true of a mechanical system, connecting fixed initial and final over a specified time , renders stationary with respect to small perturbations. This means that among all possible paths in the configuration space that satisfy the boundary conditions, the physical path is one for which the first-order variation in vanishes. Central to this variational approach are admissible virtual displacements, denoted as \delta q(t), which represent infinitesimal deviations from the true path q(t) that must satisfy kinematic constraints and vanish at the endpoints of the time interval, i.e., \delta q(t_1) = \delta q(t_2) = 0. These virtual displacements are "admissible" in the sense that they preserve the fixed boundary conditions while allowing arbitrary smooth variations in between, ensuring that the perturbed paths still connect the same initial and final states. The stationarity condition is then expressed as \delta S = 0, where S is the action integral S = \int_{t_1}^{t_2} L(q, \dot{q}, t) \, dt and L is the , indicating that the first-order change in the action due to any such virtual displacement is zero. Mathematically, this is formalized by considering a family of perturbed paths q_\epsilon(t) = q(t) + \epsilon \eta(t), where \epsilon is a small and \eta(t) is an arbitrary smooth function satisfying \eta(t_1) = \eta(t_2) = 0. The action along this perturbed path is S(\epsilon) = \int_{t_1}^{t_2} L(q + \epsilon \eta, \dot{q} + \epsilon \dot{\eta}, t) \, dt, and the stationarity condition requires that the derivative \frac{dS}{d\epsilon} \bigg|_{\epsilon=0} = 0 for all admissible \eta(t). This condition implies that the true path corresponds to an extremum—or more generally, a —in the infinite-dimensional of all admissible paths, where the action functional achieves a value locally. For short trajectories, this extremum is typically a minimum, while for longer paths it may be a , reflecting the geometric and dynamical properties of the system without implying a global minimum across all possible paths.

Derivations and key results

Euler-Lagrange equations from variation

To obtain the from Hamilton's principle, the variation of the action functional S = \int_{t_1}^{t_2} L(q, \dot{q}, t) \, dt is computed, where L is the and the variation satisfies \delta q(t_1) = \delta q(t_2) = 0 to ensure stationarity along the true path. The first variation yields \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 = 0. Integrating the second term by parts gives \int_{t_1}^{t_2} \frac{\partial L}{\partial \dot{q}} \delta \dot{q} \, dt = \left[ \frac{\partial L}{\partial \dot{q}} \delta q \right]_{t_1}^{t_2} - \int_{t_1}^{t_2} \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}} \right) \delta q \, dt. The boundary term vanishes due to the fixed endpoints, resulting in \int_{t_1}^{t_2} \left( \frac{\partial L}{\partial q} - \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}} \right) \right) \delta q \, dt = 0. Since \delta q is arbitrary, the integrand must be zero, yielding the Euler-Lagrange equation \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}} \right) - \frac{\partial L}{\partial q} = 0 $$ for a single [generalized coordinate](/page/Generalized_coordinates) $ q $.[](https://www.maths.tcd.ie/pub/HistMath/People/Hamilton/Dynamics/GenMeth.pdf) For a system with $ n $ [degrees of freedom](/page/Degrees_of_freedom) and [generalized coordinates](/page/Generalized_coordinates) $ q_i $, the [Lagrangian](/page/Lagrangian) $ L(q_i, \dot{q}_i, t) $ leads to the set of equations \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}_i} \right) - \frac{\partial L}{\partial q_i} = 0, \quad i = 1, \dots, n. [](https://archive.org/details/mcaniqueanalyt01lagr) In the presence of $ m $ [holonomic constraints](/page/Holonomic_constraints) $ f_j(q_i, t) = 0 $, $ j = 1, \dots, m $, the equations are modified using Lagrange multipliers $ \lambda_j $, incorporating constraint forces as \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}i} \right) - \frac{\partial L}{\partial q_i} = \sum{j=1}^m \lambda_j \frac{\partial f_j}{\partial q_i}, \quad i = 1, \dots, n, ### Canonical momenta and [Noether's theorem](/page/Noether's_theorem) In [Lagrangian mechanics](/page/Lagrangian_mechanics), the canonical [momentum](/page/Momentum) conjugate to a generalized coordinate $ q_i $ is defined as $ p_i = \frac{\partial L}{\partial \dot{q}_i} $, where $ L $ is the [Lagrangian](/page/Lagrangian) and $ \dot{q}_i $ is the corresponding generalized velocity./07%3A_Symmetries_Invariance_and_the_Hamiltonian/7.02%3A_Generalized_Momentum) This quantity generalizes the classical [momentum](/page/Momentum) and arises naturally from the structure of the [Lagrangian](/page/Lagrangian), facilitating the [transition](/page/Transition) to [Hamiltonian mechanics](/page/Hamiltonian_mechanics). The [Hamiltonian](/page/Hamiltonian) $ H $ is obtained via the Legendre transform of the [Lagrangian](/page/Lagrangian) as $ H = \sum_i p_i \dot{q}_i - L $, and for time-independent Lagrangians, it represents the total energy of the system.[](https://www.damtp.cam.ac.uk/user/tong/dynamics/four.pdf) This formulation shifts the description from velocities to [momenta](/page/Momentum), preserving the [dynamics](/page/Dynamics) derived from Hamilton's principle. A profound connection between symmetries and conservation laws emerges through [Noether's theorem](/page/Noether's_theorem), established in 1918, which states that every [continuous symmetry](/page/Continuous_symmetry) of the action integral implies a corresponding [conserved quantity](/page/Conserved_quantity) for the system's motion.[](https://arxiv.org/abs/physics/0503066) For instance, invariance under time translations yields [conservation of energy](/page/Conservation_of_energy), expressed as $ \frac{dH}{dt} = 0 $ when the [Lagrangian](/page/Lagrangian) lacks explicit time dependence. Other symmetries yield analogous results: spatial translation invariance conserves linear [momentum](/page/Momentum), while rotational invariance conserves angular [momentum](/page/Momentum).[](http://eduardo.physics.illinois.edu/phys582/582-chapter3-edited.pdf) In cases with time-dependent [Lagrangians](/page/Lagrangian), quasi-symmetries can lead to modified conservation laws, such as time-varying conserved quantities. A specific case arises for ignorable coordinates $ q_j $, where the [Lagrangian](/page/Lagrangian) has no explicit dependence on $ q_j $; here, the conjugate [momentum](/page/Momentum) $ p_j $ is constant along the trajectory./07%3A_Symmetries_Invariance_and_the_Hamiltonian/7.02%3A_Generalized_Momentum) ## Illustrative examples ### Free particle in Cartesian coordinates A fundamental illustrative example of Hamilton's principle is the motion of a [free particle](/page/Free_particle) in three-dimensional Cartesian coordinates, where no external forces act, so the [potential energy](/page/Potential_energy) is zero. The [Lagrangian](/page/Lagrangian) for such a particle of [mass](/page/Mass) $ m $ is purely kinetic: L = \frac{1}{2} m (\dot{x}^2 + \dot{y}^2 + \dot{z}^2), where $ x(t) $, $ y(t) $, and $ z(t) $ are the position coordinates as functions of time $ t $, and the dots denote time derivatives.[](http://ppc.inr.ac.ru/uploads/476_Hamill.pdf) This form arises because the [kinetic energy](/page/Kinetic_energy) $ T = \frac{1}{2} m v^2 $ with [velocity](/page/Velocity) $ \mathbf{v} = (\dot{x}, \dot{y}, \dot{z}) $ fully describes the system's dynamics in the absence of a potential $ V = 0 $.[](https://galileoandeinstein.phys.virginia.edu/7010/CM_04_HamiltonsPrinciple.pdf) The [action](/page/Action) functional is then the time [integral](/page/Integral) of this [Lagrangian](/page/Lagrangian) over a fixed [interval](/page/Interval) from $ t_1 $ to $ t_2 $: S = \int_{t_1}^{t_2} L , dt = \int_{t_1}^{t_2} \frac{1}{2} m (\dot{x}^2 + \dot{y}^2 + \dot{z}^2) , dt = \int_{t_1}^{t_2} \frac{1}{2} m v^2 , dt. According to Hamilton's principle, the physical path $ \mathbf{r}(t) = (x(t), y(t), z(t)) $ between fixed endpoints $ \mathbf{r}(t_1) $ and $ \mathbf{r}(t_2) $ renders this [action](/page/Action) [stationary](/page/Stationary), i.e., $ \delta S = 0 $ for [infinitesimal](/page/Infinitesimal) variations $ \delta \mathbf{r}(t) $ that vanish at the endpoints. The [stationary](/page/Stationary) path corresponds to a straight line in space traversed at constant speed, as this configuration minimizes the [integral](/page/Integral) of the squared speed over the fixed time [interval](/page/Interval).[](https://kestrel.nmt.edu/~raymond/classes/ph321/notes/hamilton/hamilton.pdf)[](https://www.physics.rutgers.edu/~shapiro/507/book3.pdf) To derive the equations of motion, consider the first variation of the action. For a variation $ \delta \mathbf{r}(t) $, integration by parts yields \delta S = \int_{t_1}^{t_2} m \ddot{\mathbf{r}} \cdot \delta \mathbf{r} , dt, after boundary terms vanish. Setting $ \delta S = 0 $ for arbitrary $ \delta \mathbf{r} $ implies $ m \ddot{\mathbf{r}} = 0 $, or $ \ddot{\mathbf{r}} = 0 $, meaning the acceleration is zero and the velocity $ \dot{\mathbf{r}} $ is constant—recovering Newton's first law for an unforced particle.[](https://galileoandeinstein.phys.virginia.edu/7010/CM_04_HamiltonsPrinciple.pdf)[](http://ppc.inr.ac.ru/uploads/476_Hamill.pdf) Geometrically, this stationary path represents the shortest "time path" in the configuration space (spanned by $ x, y, z $), where the trajectory minimizes the accumulated kinetic energy over the specified duration.[](https://www.physics.rutgers.edu/~shapiro/507/book3.pdf) ### Harmonic oscillator The harmonic oscillator serves as a fundamental example of applying Hamilton's principle to a system featuring a quadratic potential, illustrating how the principle derives periodic motion from energy considerations. For a particle of mass $ m $ constrained to one-dimensional motion along the $ x $-axis, attached to a spring with constant $ k $, the kinetic energy is $ \frac{1}{2} m \dot{x}^2 $ and the potential energy is $ \frac{1}{2} k x^2 $, so the Lagrangian takes the form L = \frac{1}{2} m \dot{x}^2 - \frac{1}{2} k x^2. [](https://www.math.unl.edu/~scohn1/8423/cvar7.pdf) The action is the time integral of this Lagrangian over an interval from $ t_1 $ to $ t_2 $. Hamilton's principle requires that the true path extremizes the action, meaning its first variation vanishes: $ \delta S = 0 $. Performing the variation yields the Euler-Lagrange equation \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{x}} \right) = \frac{\partial L}{\partial x}, which reduces to the second-order differential equation m \ddot{x} + k x = 0. The general solution is x(t) = A \cos(\omega t + \phi), with amplitude $ A $, phase $ \phi $, and angular frequency $ \omega = \sqrt{k/m} $, capturing the system's sinusoidal oscillation.[](https://www.math.unl.edu/~scohn1/8423/cvar7.pdf) In phase space, the canonical momentum $ p = m \dot{x} $ parametrizes the trajectory alongside $ x $, revealing a closed elliptical orbit due to constant total energy. The true path that renders the action stationary exhibits symmetric excursions in the position and momentum directions of this phase space ellipse, embodying the symmetry inherent to the quadratic potential. For time intervals less than half the oscillation period, this stationary point corresponds to a minimum of the action.[](https://pubs.aip.org/aapt/ajp/article/85/8/633/1057963/Condition-for-minimal-harmonic-oscillator-action) This stationarity aligns with energy-based analyses, as the action principle produces [equations of motion](/page/Equations_of_motion) satisfying the [virial theorem](/page/Virial_theorem) for the [harmonic oscillator](/page/Harmonic_oscillator): the time average of the [kinetic energy](/page/Kinetic_energy) equals that of the [potential energy](/page/Potential_energy), each comprising half the total [energy](/page/Energy). Thus, Hamilton's principle not only derives the [dynamics](/page/Dynamics) but also connects directly to this theorem's prediction of equipartition in the harmonic case.[](https://courses.physics.ucsd.edu/2014/Fall/physics200a/LECTURES/MECHANICS.pdf) ## Applications to mechanical systems ### Rigid body dynamics Hamilton's principle provides a variational framework for deriving the [equations of motion](/page/Equations_of_motion) for [rigid bodies](/page/Rigid_body_dynamics), which maintain fixed shape and size but possess rotational [degrees of freedom](/page/Degrees_of_freedom). These systems are typically parameterized using [Euler angles](/page/Euler_angles)—θ ([nutation](/page/Nutation)), φ ([precession](/page/Precession)), and ψ (spin)—to describe the orientation relative to a fixed frame. The principle is applied by formulating the action integral over virtual paths in configuration space, leading to stationarity conditions that yield the dynamics.[](https://itp.uni-frankfurt.de/~hees/gen-phys/spinning-top.pdf) The Lagrangian for a rigid body is constructed from its kinetic and potential energies, expressed in the body-fixed frame. The kinetic energy arises from the rotational motion, given by $ T = \frac{1}{2} \boldsymbol{\omega} \cdot \mathbf{I} \boldsymbol{\omega} $, where $ \boldsymbol{\omega} $ is the angular velocity vector and $ \mathbf{I} $ is the inertia tensor, assumed diagonal for principal axes with components $ (I_1, I_2, I_3) $. The potential energy $ V $ depends on the orientation angles, such as gravitational effects for a heavy top: $ V = V(\theta, \phi, \psi) $. Thus, the Lagrangian takes the form $ L = \frac{1}{2} \boldsymbol{\omega} \cdot \mathbf{I} \boldsymbol{\omega} - V(\theta, \phi, \psi) $.[](https://mathweb.ucsd.edu/~mleok/pdf/samplechap.pdf)[](https://itp.uni-frankfurt.de/~hees/gen-phys/spinning-top.pdf) To apply Hamilton's principle, the angular velocity must be expressed in terms of the Euler angles and their time derivatives, imposing kinematic relations that connect the velocities. In the body-fixed frame, the components are $ \boldsymbol{\omega} = (\dot{\theta} \cos \phi + \dot{\psi} \sin \theta \sin \phi, \, -\dot{\theta} \sin \phi + \dot{\psi} \sin \theta \cos \phi, \, \dot{\phi} + \dot{\psi} \cos \theta) $. These expressions reflect the non-holonomic nature of the velocity constraints, as the angular velocities are linear combinations of the generalized velocities $ \dot{\theta}, \dot{\phi}, \dot{\psi} $, but the orientation evolves holonomically on the rotation group SO(3). Substituting into the [Lagrangian](/page/Lagrangian) allows variation with respect to the angles.[](https://itp.uni-frankfurt.de/~hees/gen-phys/spinning-top.pdf)[](https://www.mdpi.com/2227-7390/11/12/2727) The Euler-Lagrange equations, derived from the stationarity of [the action](/page/The_Action) $ S = \int_{t_1}^{t_2} L \, dt $, are applied to the [generalized coordinates](/page/Generalized_coordinates) θ, φ, ψ: $ \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}_i} \right) - \frac{\partial L}{\partial q_i} = 0 $ for $ q_i = \theta, \phi, \psi $. This yields the Euler equations of [rigid body](/page/Rigid_body) motion in the body frame: \mathbf{I} \dot{\boldsymbol{\omega}} + \boldsymbol{\omega} \times \mathbf{I} \boldsymbol{\omega} = -\nabla V, where the [torque](/page/Torque) term $ -\nabla V $ accounts for external potentials, and the cross-product term captures the rotational coupling. For a general [inertia](/page/Inertia) tensor, the equations become component-wise, such as $ I_1 \dot{\omega}_1 + (I_3 - I_2) \omega_2 \omega_3 = -\frac{\partial V}{\partial \alpha_1} $, with similar forms for the other axes, using body-frame angles. These equations govern the evolution of [angular velocity](/page/Angular_velocity) under the [variational principle](/page/Variational_principle).[](https://www.mdpi.com/2227-7390/11/12/2727)[](https://mathweb.ucsd.edu/~mleok/pdf/samplechap.pdf)[](https://itp.uni-frankfurt.de/~hees/gen-phys/spinning-top.pdf) A key example is torque-free motion, where $ V = 0 $ or is independent of orientation, simplifying to $ \mathbf{I} \dot{\boldsymbol{\omega}} + \boldsymbol{\omega} \times \mathbf{I} \boldsymbol{\omega} = 0 $. In this case, the angular momentum $ \mathbf{h} = \mathbf{I} \boldsymbol{\omega} $ is conserved due to the rotational invariance of the [Lagrangian](/page/Lagrangian), as per [Noether's theorem](/page/Noether's_theorem) applied to the [symmetry group](/page/Symmetry_group) SO(3). For a symmetric top with $ I_1 = I_2 \neq I_3 $, the motion features steady [precession](/page/Precession), with the angular velocity tracing a polhode on the constant-energy [ellipsoid](/page/Ellipsoid), while the angular momentum vector remains fixed in space. This conservation establishes both the magnitude $ |\mathbf{h}|^2 $ and kinetic energy $ T $ as invariants, constraining the dynamics to intersections of ellipsoids in angular momentum space.[](https://mathweb.ucsd.edu/~mleok/pdf/samplechap.pdf)[](https://itp.uni-frankfurt.de/~hees/gen-phys/spinning-top.pdf) ### Deformable solids and continua Hamilton's principle is extended to deformable solids and continua by formulating the dynamics in terms of a [Lagrangian](/page/Lagrangian) [density](/page/Density) that accounts for the distributed nature of the material. The [displacement](/page/Displacement) field $\mathbf{u}(\mathbf{x}, t)$ describes the deformation of the body from a reference configuration, with $\mathbf{x}$ denoting the position in the reference volume $V$. The kinetic energy [density](/page/Density) is $\frac{1}{2} \rho \dot{\mathbf{u}}^2$, where $\rho$ is the mass [density](/page/Density) and $\dot{\mathbf{u}}$ is the time derivative of the [displacement](/page/Displacement). The potential energy is stored as [strain energy](/page/Strain_energy), with [density](/page/Density) $W(\nabla \mathbf{u})$, where $\nabla \mathbf{u}$ represents the [displacement](/page/Displacement) [gradient](/page/Gradient), often through the [infinitesimal](/page/Infinitesimal) [strain](/page/Strain) tensor $\mathbf{e} = \frac{1}{2} (\nabla \mathbf{u} + (\nabla \mathbf{u})^T)$ in [linear elasticity](/page/Linear_elasticity). The [Lagrangian](/page/Lagrangian) [density](/page/Density) is thus given by \mathcal{L} = \frac{1}{2} \rho \dot{\mathbf{u}}^2 - W(\nabla \mathbf{u}). The action integral over the time interval $[t_1, t_2]$ and volume $V$ is S = \int_{t_1}^{t_2} \int_V \mathcal{L} , dV , dt. Requiring the stationarity of the action, $\delta S = 0$, for admissible virtual displacements $\delta \mathbf{u}$ vanishing at the endpoints $t_1$ and $t_2$, yields the Euler-Lagrange equations for the continuum. After integration by parts and applying the divergence theorem, assuming suitable boundary conditions, this results in the equations of linear momentum balance for the deformable body: \rho \ddot{\mathbf{u}} = \nabla \cdot \boldsymbol{\sigma}, where $\boldsymbol{\sigma}$ is the Cauchy stress tensor derived from the strain energy as $\boldsymbol{\sigma} = \frac{\partial W}{\partial \mathbf{e}}$ (symmetrized appropriately for linear theory).[](https://link.springer.com/book/10.1007/978-3-030-90306-3) Applications of Hamilton's principle to continuum mechanics gained prominence in the 19th century, following Hamilton's original work. Lord Kelvin (William Thomson) employed variational methods in his treatments of elastic deformations and stability, while Lord Rayleigh extensively used the principle in deriving dynamic equations for elastic media in his seminal work *The Theory of Sound* (1877), bridging acoustics and solid mechanics. A special case arises in the study of [waves](/page/Waves) propagating through solids, where the principle directly derives the elastic [wave equation](/page/Wave_equation). For isotropic [linear elasticity](/page/Linear_elasticity), substituting the quadratic [strain energy](/page/Strain_energy) $W = \frac{1}{2} \lambda (\nabla \cdot \mathbf{u})^2 + \frac{\mu}{4} |\nabla \mathbf{u} + (\nabla \mathbf{u})^T|^2 $ (with Lamé constants $\lambda$ and $\mu$) into the varied [action](/page/Action) yields the vector [wave equation](/page/Wave_equation) \rho \ddot{\mathbf{u}} = (\lambda + 2\mu) \nabla (\nabla \cdot \mathbf{u}) - \mu \nabla \times (\nabla \times \mathbf{u}), describing the coupled longitudinal (P-wave) and [shear](/page/Shear) (S-wave) motions in the [solid](/page/Solid). This [formulation](/page/Formulation) highlights the principle's utility in capturing wave [dispersion](/page/Dispersion) and [attenuation](/page/Attenuation) in deformable continua without relying on Newton's [laws](/page/Law) directly.[](https://link.springer.com/book/10.1007/978-3-030-90306-3) ## Comparisons with related principles ### Maupertuis' principle of least [action](/page/Action) Pierre-Louis Maupertuis introduced the principle of least [action](/page/Action) in 1744, presenting it as a unifying [law](/page/Law) for [optics](/page/Optics) and [mechanics](/page/Mechanics) that reflected [nature](/page/Nature)'s tendency to act with maximal simplicity and economy.[](http://www.scholarpedia.org/article/Principle_of_least_action) In his memoir "Accord de différentes loix de la nature qui avoient jusqu'à présent paru incompatibles," read to the Académie des Sciences, Maupertuis proposed a teleological [interpretation](/page/Interpretation), suggesting that [nature](/page/Nature) chooses paths minimizing a [quantity](/page/Quantity) he termed "[action](/page/Action)," thereby implying a purposeful design in physical [laws](/page/Law). This view sparked significant controversy in the [1740s](/page/1740s) and 1750s, particularly with Samuel König, who accused Maupertuis of plagiarizing the idea from Leibniz and questioned its metaphysical implications; Maupertuis responded by charging König with forgery, leading to a heated debate involving Euler and [Voltaire](/page/Voltaire) that highlighted tensions between mathematical rigor and philosophical [interpretation](/page/Interpretation).[](https://mathshistory.st-andrews.ac.uk/HistTopics/Forgery_2/) For conservative systems, Maupertuis' principle states that the actual path taken by a particle minimizes the abbreviated action $\int p \, ds$, where $p = mv$ is the [momentum](/page/Momentum) and $ds$ is the path element, among all paths connecting fixed endpoints with a specified fixed [energy](/page/Energy) $E$.[](http://www.scholarpedia.org/article/Principle_of_least_action) Mathematically, for a one-dimensional [conservative system](/page/Conservative_system) with potential $V(q)$, this abbreviated action takes the form W = \int_{q_1}^{q_2} \sqrt{2m(E - V(q))} , dq, where the [integral](/page/Integral) is [stationary](/page/Stationary) for the true [trajectory](/page/Trajectory), yielding geodesics in a [metric](/page/Metric) determined by the energy surface.[](http://www.scholarpedia.org/article/Principle_of_least_action) This formulation leads to the same extremal paths as Hamilton's principle but constrains the variation to fixed energy, focusing on spatial geometry rather than [time evolution](/page/Time_evolution).[](http://www.scholarpedia.org/article/Principle_of_least_action) The principle assumes [energy conservation](/page/Energy_conservation), making it applicable only to conservative systems without time-dependent potentials or dissipative forces.[](http://www.scholarpedia.org/article/Principle_of_least_action) It thus lacks the generality of Hamilton's full [action](/page/Action) integral, which incorporates time and applies more broadly to non-conservative cases.[](http://www.scholarpedia.org/article/Principle_of_least_action) ### Differences from Lagrange's [equations of motion](/page/Equations_of_motion) Lagrange introduced his [equations of motion](/page/Equations_of_motion) in the 1788 treatise *Mécanique Analytique*, deriving them directly from [d'Alembert's principle](/page/D'Alembert's_principle) of [virtual](/page/Virtual) work, which extends the statics of virtual displacements to [dynamics](/page/Dynamics) by incorporating inertial forces. This approach yields the second-order differential equations \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}_i} \right) - \frac{\partial L}{\partial q_i} = Q_i without invoking the calculus of variations, where $L = T - V$ is the Lagrangian, $q_i$ are generalized coordinates, and $Q_i$ are generalized forces accounting for non-conservative effects and constraints. The method relies on instantaneous virtual displacements $\delta \mathbf{r}$ consistent with constraints, ensuring $\sum (\mathbf{F} - m \mathbf{a}) \cdot \delta \mathbf{r} = 0$ at each time, transforming Newton's laws into a coordinate-independent form suitable for complex systems. In contrast, Hamilton's principle, formulated in [1834](/page/1834), posits that the physical path minimizes or makes stationary [the action](/page/The_Action) [integral](/page/Integral) $\int_{t_1}^{t_2} L \, dt$, leading to the same Euler-Lagrange equations through explicit variation over paths in configuration space. This variational framework offers advantages in handling [holonomic constraints](/page/Holonomic_constraints) systematically via Lagrange multipliers and in revealing symmetries through the invariance of [the action](/page/The_Action), facilitating derivations like [Noether's theorem](/page/Noether's_theorem). Additionally, Hamilton's approach unifies [classical mechanics](/page/Classical_mechanics) with [geometrical optics](/page/Geometrical_optics) by analogy, employing a [characteristic function](/page/Characteristic_function) to describe both ray paths and mechanical trajectories as solutions to similar variational problems. The core methodological difference lies in their foundations: Lagrange's direct use of [virtual work](/page/Virtual_work) provides an instantaneous, algebraic condition without path integration, making it more straightforward for deriving equations in [generalized coordinates](/page/Generalized_coordinates), whereas Hamilton's indirect variational [calculus](/page/Calculus) integrates over time, emphasizing global path optimization. For [holonomic](/page/Holonomic) scleronomic systems, both methods produce identical [equations of motion](/page/Equations_of_motion). However, they diverge in treating non-[holonomic](/page/Holonomic) constraints, where velocity-dependent restrictions (e.g., no slipping in rolling) are naturally incorporated in Lagrange's framework through [d'Alembert's principle](/page/D'Alembert's_principle) via modified virtual displacements, while Hamilton's principle requires extensions, such as the non-variational Lagrange-d'Alembert equations, to avoid inconsistencies in the action variation. ## Generalizations to fields and modern physics ### Classical field theories Hamilton's principle generalizes to classical field theories by formulating the dynamics of continuous fields distributed over [spacetime](/page/Spacetime), rather than discrete particles or rigid bodies. In this framework, the action $ S $ is defined as the [integral](/page/Integral) over a four-dimensional spacetime volume of a Lagrangian density $ \mathcal{L} $, which depends on the field $ \phi(x) $ and its first derivatives $ \partial_\mu \phi(x) $: S[\phi] = \int d^4 x , \mathcal{L}(\phi, \partial_\mu \phi). The principle states that the physical configuration of the field extremizes this action, so that the first variation vanishes: $ \delta S = 0 $.[](https://www.sas.rochester.edu/pas/assets/pdf/undergraduate/Introduction_to_Lagrangian_Field_Theory.pdf)[](https://web2.ph.utexas.edu/~vadim/Classes/2022f/classical.pdf) Performing the variation with respect to the field $ \phi $ yields the Euler-Lagrange equations for fields: \frac{\partial \mathcal{L}}{\partial \phi} - \partial_\mu \left( \frac{\partial \mathcal{L}}{\partial (\partial_\mu \phi)} \right) = 0. These partial differential equations govern the evolution of [the field](/page/The_Field), analogous to the ordinary differential equations obtained for [mechanical](/page/Mechanical) systems. For fields with multiple components or more complex dependencies, the equations extend accordingly, providing a unified variational [derivation](/page/Derivation) of field equations across diverse physical systems.[](https://www.sas.rochester.edu/pas/assets/pdf/undergraduate/Introduction_to_Lagrangian_Field_Theory.pdf) A canonical example is the [electromagnetic field](/page/Electromagnetic_field) in [vacuum](/page/Vacuum), described by the antisymmetric [field strength](/page/Field_strength) tensor $ F_{\mu\nu} = \partial_\mu A_\nu - \partial_\nu A_\mu $, where $ A_\mu $ is the four-potential. The [Lagrangian](/page/Lagrangian) density is \mathcal{L} = -\frac{1}{4} F_{\mu\nu} F^{\mu\nu}, in units where the speed of light and permittivity are set to 1. Varying the action $ S = \int d^4 x \, \mathcal{L} $ with respect to $ A_\nu $ produces the source-free Maxwell equations: \partial_\mu F^{\mu\nu} = 0, along with the Bianchi identity $ \partial_\lambda F_{\mu\nu} + \partial_\mu F_{\nu\lambda} + \partial_\nu F_{\lambda\mu} = 0 $, which follows from the antisymmetry of $ F_{\mu\nu} $. When coupled to charged matter with current $ j^\mu $, the equations become $ \partial_\mu F^{\mu\nu} = j^\nu $, maintaining the variational structure.[](https://www.phys.ufl.edu/~thorn/homepage/emlectures1.pdf)[](https://bohr.physics.berkeley.edu/classes/209/f02/classemf.pdf)[](https://math.uchicago.edu/~may/REU2012/REUPapers/Yu.pdf) Another fundamental example is general relativity, where gravity is described as a curvature of spacetime. The Einstein-Hilbert action for the gravitational field is S = \frac{c^4}{16\pi G} \int d^4 x , \sqrt{-g} , R, with $ g $ the determinant of the metric tensor $ g_{\mu\nu} $ and $ R $ the Ricci scalar. Varying this action with respect to the metric yields the Einstein field equations R_{\mu\nu} - \frac{1}{2} R g_{\mu\nu} = \frac{8\pi G}{c^4} T_{\mu\nu}, relating spacetime curvature to the stress-energy tensor $ T_{\mu\nu} $, thus deriving the dynamics of gravity variationally from Hamilton's principle. The electromagnetic Lagrangian exhibits gauge symmetry under local transformations $ A_\mu \to A_\mu + \partial_\mu \Lambda $, where $ \Lambda $ is an arbitrary scalar function. Noether's theorem associates continuous symmetries of the action with conserved currents; for the global (constant $ \Lambda $) subgroup of this gauge symmetry, it yields the conservation of the electric charge current $ \partial_\mu j^\mu = 0 $ when matter is included. For the full local gauge invariance in the pure field sector, the theorem implies the on-shell vanishing of the Noether current, enforcing the homogeneous Maxwell equations as identities. This connection highlights how symmetries underpin conservation laws in field theories derived from Hamilton's principle.[](https://web2.ph.utexas.edu/~vadim/Classes/2022f/Noether.pdf) ### Role in quantum mechanics and path integrals In the semiclassical limit of [quantum mechanics](/page/Quantum_mechanics), the classical trajectory determined by Hamilton's principle—characterized by stationary [action](/page/Action)—becomes the leading contribution to the quantum [propagator](/page/Propagator), as rapid oscillations in the [phase](/page/Phase) factors for non-classical paths lead to destructive [interference](/page/Interference). This transition bridges classical and quantum descriptions, where the [action](/page/Action) functional from Hamilton's principle plays a central role in weighting quantum amplitudes.[](http://www-f1.ijs.si/~ramsak/km1/feynman.pdf) The historical foundation for this connection traces back to Paul Dirac's 1933 insight, which linked the [Lagrangian](/page/Lagrangian) formulation of [classical mechanics](/page/Classical_mechanics) to [quantum mechanics](/page/Quantum_mechanics) by proposing that the quantum transition amplitude could be expressed in terms of an exponential involving the classical [action](/page/Action), drawing on the Hamilton-Jacobi equation's principal function as the [phase](/page/Phase) of the wave function.[](https://www.informationphilosopher.com/solutions/scientists/dirac/Lagrangian_1933.pdf) Dirac's suggestion emphasized the analogy between the stationary [action](/page/Action) in [classical mechanics](/page/Classical_mechanics) and a sum over [phase](/page/Phase) factors in [quantum theory](/page/Quantum_theory), setting the stage for a path-based reformulation.[](https://www.informationphilosopher.com/solutions/scientists/dirac/Lagrangian_1933.pdf) Richard Feynman fully developed this idea in his 1948 formulation of the path integral approach to non-relativistic quantum mechanics. The probability amplitude for a particle to evolve from position $q_1$ at time $t_1$ to $q_2$ at $t_2 = t_1 + \tau$ is given by \langle q_2 | e^{-i H \tau / \hbar} | q_1 \rangle = \int \mathcal{D}q(t) , e^{i S/\hbar}, where the integral is a sum over all possible paths $q(t)$ connecting the endpoints, weighted by the phase factor $e^{i S/\hbar}$ with $S$ being the classical action functional along each path, directly extending Hamilton's principle to quantum amplitudes.[](http://www-f1.ijs.si/~ramsak/km1/feynman.pdf) This formulation generalizes the stationary action condition: instead of a single classical path, quantum mechanics incorporates contributions from all paths, but the classical path extremizing $S$ provides the constructive interference backbone.[](http://www-f1.ijs.si/~ramsak/km1/feynman.pdf) Applying the [stationary phase approximation](/page/Stationary_phase_approximation) to the [path integral](/page/Path_integral) in the limit $\hbar \to 0$ (or large [action](/page/Action) scales) recovers the classical result, as the [integral](/page/Integral) is dominated by paths near the one where the phase $S/\hbar$ is stationary, yielding the Hamilton-Jacobi principal function and thus Hamilton's principle itself.[](http://www-f1.ijs.si/~ramsak/km1/feynman.pdf) This semiclassical recovery underscores how Hamilton's principle underlies both classical [determinism](/page/Determinism) and the probabilistic nature of [quantum mechanics](/page/Quantum_mechanics) through path [interference](/page/Interference).[](http://www-f1.ijs.si/~ramsak/km1/feynman.pdf) ### Extensions in quantum field theory In quantum field theory (QFT), Hamilton's principle is extended to the action functional $ S = \int d^4 x \, \mathcal{L}(\phi, \partial_\mu \phi, \text{interactions}) $, where $\mathcal{L}$ is the [Lagrangian](/page/Lagrangian) density for fields $\phi$ (such as scalars, fermions, or [gauge](/page/Gauge) fields) and their derivatives, integrated over four-dimensional [spacetime](/page/Spacetime). The principle of stationary action, $\delta S = 0$, yields the Euler-Lagrange field equations that govern the [dynamics](/page/Dynamics), analogous to [classical mechanics](/page/Classical_mechanics) but for relativistic fields with interactions. This formulation underpins the derivation of equations like the [Dirac equation](/page/Dirac_equation) for fermions or the Yang-Mills equations for [gauge](/page/Gauge) fields.[](https://web.physics.ucsb.edu/~mark/ms-qft-DRAFT.pdf) Furthermore, symmetries of the [action](/page/Action) lead to Ward identities, which encode conservation laws and constrain scattering amplitudes in perturbative QFT.[](https://web.physics.ucsb.edu/~mark/ms-qft-DRAFT.pdf) The quantum extension incorporates Feynman's [path integral formulation](/page/Path_integral_formulation), where the partition function is $ Z = \int \mathcal{D}\phi \, e^{i S / \hbar} $, summing over all field configurations weighted by the phase factor from [the action](/page/The_Action); this generates correlation functions and enables diagrammatic [perturbation theory](/page/Perturbation_theory). [Renormalization](/page/Renormalization) procedures, essential for handling infinities in loop diagrams, are systematically applied to [the action](/page/The_Action) to yield finite predictions, as in [quantum electrodynamics](/page/Quantum_electrodynamics) or non-Abelian gauge theories.[](https://web.physics.ucsb.edu/~mark/ms-qft-DRAFT.pdf) Modern applications include effective field theories (EFTs), where the [Standard Model](/page/Standard_Model) emerges as a low-energy EFT defined by its gauge-invariant action principle, incorporating electroweak and strong interactions via the SU(3) × SU(2) × U(1) gauge group, with the [Higgs mechanism](/page/Higgs_mechanism) breaking symmetry spontaneously. In holographic duality, such as the [AdS](/page/Ads)/CFT [correspondence](/page/Correspondence), the stationary condition on the bulk gravitational action in [anti-de Sitter space](/page/Anti-de_Sitter_space) implies specific correlator structures in the dual [conformal field theory](/page/Conformal_field_theory) on the boundary, facilitating computations of strongly coupled phenomena like quark-gluon [plasma](/page/Plasma) properties. Post-2000 developments feature applications to condensed matter systems, where the action principle describes topological insulators through an axion-like term $\theta \int d^4 x \, \frac{e^2}{32\pi^2} F_{\mu\nu} \tilde{F}^{\mu\nu}$ in the effective QFT, capturing magnetoelectric effects and protected surface states. No fundamental paradigm shifts have occurred by 2025, but refinements in lattice QFT simulations—discretizing the action on a grid for non-perturbative computations—have advanced, enabling higher-precision determinations of hadron masses and decay constants using improved algorithms on supercomputers.[](https://pdg.lbl.gov/2025/reviews/rpp2024-rev-lattice-qcd.pdf)

References

  1. [1]
    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.<|control11|><|separator|>
  2. [2]
    [PDF] ON A GENERAL METHOD IN DYNAMICS By William Rowan Hamilton
    This edition is based on the original publication in the Philosophical Transactions of the. Royal Society, part II for 1834. The following errors in the ...
  3. [3]
    Methodus inveniendi lineas curvas maximi minimive proprietate ...
    May 25, 2018 · Methodus inveniendi lineas curvas maximi minimive proprietate gaudentes, sive, Solutio problematis isoperimetrici latissimo sensu accepti.
  4. [4]
    [PDF] A Brief Survey of the Calculus of Variations - arXiv
    The geometrical technique used by Newton was later adopted by Jacob Bernoulli in his solution of the Brachistochrone and was also later systematized by Euler.Missing: 1766 | Show results with:1766
  5. [5]
    Mécanique analytique : Lagrange, J. L. (Joseph Louis), 1736-1813
    Jan 18, 2010 · Mécanique analytique. by: Lagrange, J. L. (Joseph Louis), 1736 ... PDF download · download 1 file · SINGLE PAGE ORIGINAL JP2 TAR download.Missing: online | Show results with:online
  6. [6]
    [PDF] J. L. Lagrange's early contributions to the principles and methods of ...
    The problem consisted of deducing the motion of a particle from a variational law, a law which by the 1750's was known as the principle of least action. We ...
  7. [7]
    The history of the Méchanique analitique | Lettera Matematica
    Jun 6, 2014 · We note that Lagrange does not speak of the “principle of least action”—a term associated with the name of Maupertuis and to troublesome ...
  8. [8]
    [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.
  9. [9]
  10. [10]
    [PDF] SECOND ESSAY ON A GENERAL METHOD IN DYNAMICS By ...
    Mathematical Papers of Sir William Rowan Hamilton, Volume II: Dynamics, edited for the. Royal Irish Academy by A. W. Conway and A. J. McConnell, and ...
  11. [11]
    VII. Second essay on a general method in dynamics - Journals
    Second essay on a general method in dynamics. William Rowan Hamilton. Google Scholar · Find this author on PubMed · Search for more papers by this author.Missing: RIA | Show results with:RIA
  12. [12]
    9.1: Introduction to Hamilton's Action Principle - Physics LibreTexts
    Mar 14, 2021 · Hamilton's Action Principle provides the foundation for building Lagrangian mechanics that had been pioneered years earlier.
  13. [13]
    13.4: The Lagrangian Equations of Motion - Physics LibreTexts
    Aug 7, 2022 · The quantity L = T − V is known as the lagrangian for the system, and Lagrange's equation can then be written. (13.4.16) d d ⁢ t ⁢ ∂ L ...
  14. [14]
    [PDF] CHM 532 Notes on Classical Mechanics Lagrange's and Hamilton's ...
    (13) We now give a defining relation for the classical Lagrangian: Definition: The classical Lagrangian L is given by. L = T − V. (14)
  15. [15]
    [PDF] Chapter 2 Lagrange's and Hamilton's Equations - Rutgers Physics
    Lagrange's and Hamilton's equations are reformulations of Newtonian mechanics. Lagrange's uses configuration space, while Hamilton's uses phase space. Lagrange ...
  16. [16]
    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 ...
  17. [17]
    Principle of least action - Scholarpedia
    Jun 5, 2015 · In Hamilton's principle the conceivable or trial trajectories are not constrained to satisfy energy conservation, unlike the case for Maupertuis ...
  18. [18]
  19. [19]
    [PDF] 4. The Hamiltonian Formalism - DAMTP
    We define the Hamiltonian to be the Legendre transform of the. Lagrangian with respect to the ˙qi variables,. H(qi,pi,t) = n. X i=1 pi ˙qi. L(qi, ˙qi,t). (4.11).
  20. [20]
    [physics/0503066] Invariant Variation Problems - arXiv
    Mar 8, 2005 · Authors:Emmy Noether, M. A. Tavel. View a PDF of the paper titled Invariant Variation Problems, by Emmy Noether and M. A. Tavel. View PDF.
  21. [21]
    [PDF] 3 Classical Symmetries and Conservation Laws
    We have used the existence of symmetries in a physical system as a guiding principle for the construction of their Lagrangians and energy functionals.
  22. [22]
    [PDF] A Student's Guide to Lagrangians and Hamiltonians
    This book covers Lagrangian and Hamiltonian systems, Lagrange's equations, generalized coordinates, Euler-Lagrange equations, Hamilton's principle, and  ...
  23. [23]
    [PDF] 4. Laws of Dynamics: Hamilton's Principle and Noether's Theorem
    Derivation of Hamilton's Principle from Newton's Laws in Cartesian Co-ordinates: Calculus of Variations Done Backwards! We've shown how, given an integrand, we ...
  24. [24]
    [PDF] Chapter 1 Hamilton's mechanics
    We now use the calculus of variations to see if Hamilton's principle implies the Lagrange equations. If it does, then the circle is closed and we know that the ...
  25. [25]
    [PDF] Hamilton's Equations 1. Let T be the kinetic energy and U the ...
    You can get the equations of motion in Lagrangian form from (11): Lq − d dt L˙q = −(kq − mq)=0, or q+ k m q = 0, which is the familiar simple harmonic ...
  26. [26]
    Condition for minimal harmonic oscillator action - AIP Publishing
    Aug 1, 2017 · The path that extremizes the action is often assumed to be the one that minimizes it, but this is not guaranteed by the first-order variation.
  27. [27]
    [PDF] Lecture Notes on Classical Mechanics (A Work in Progress)
    Jan 13, 2022 · ... Hamilton's Principle ... Virial Theorem. The virial theorem is a statement about the time-averaged motion of a mechanical system. Define the.
  28. [28]
    [PDF] Theory of rigid-body motion
    Dec 17, 2021 · 3 The dynamics. To describe the dynamics of the rigid body we use Hamilton's principle of least action, according to which the equations of ...
  29. [29]
    [PDF] Lagrangian and Hamiltonian Dynamics on SO(3) - UCSD Math
    This chapter treats the Lagrangian dynamics and Hamiltonian dynamics of a rotating rigid body. A rigid body is a collection of mass particles whose.Missing: Euler's | Show results with:Euler's
  30. [30]
    Deriving Euler's Equation for Rigid-Body Rotation via Lagrangian ...
    This paper uses Lagrangian dynamics to derive Euler's equation in terms of generalized coordinates. This is done by parameterizing the angular velocity vector.
  31. [31]
    Page Not Found
    - **Status**: Insufficient relevant content.
  32. [32]
    Forgery 2 - MacTutor History of Mathematics - University of St Andrews
    The paper which König gave to Maupertuis both attacked the validity of the principle of least action and also claimed that the principle was due to Leibniz and ...
  33. [33]
    [PDF] Lagrangian Field Theory
    Apr 26, 2017 · For an arbitrary region Ω in spacetime, the action integral is defined by. S(Ω) := ∫. Ω d4x多 (φr,φr,α). (2.2). By imposing the usual ...
  34. [34]
    [PDF] CLASSICAL FIELD THEORY - UT Physics
    The 4D volume element d4x is invariant under Lorentz transformations, so a Lorentz- invariant action calls for a Lorentz-invariant (or rather Lorentz scalar) ...
  35. [35]
    [PDF] Classical Electrodynamics Charles B. Thorn1 - UF Physics
    7.4 Action Principle for Maxwell's Equations. To formulate an action principle, it is best to introduce potentials so that two of the equations are ...
  36. [36]
    [PDF] Physics 221B Spring 1997 Notes 32 Lagrangian and Hamiltonian ...
    These notes cover material on the classical electromagnetic field which is preliminary to the quantization discussed in the next set of notes.
  37. [37]
    [PDF] lagrangian formulation of the electromagnetic field - UChicago Math
    Jul 16, 2012 · This paper will, given some physical assumptions and experimen- tally verified facts, derive the equations of motion of a charged particle in an ...
  38. [38]
    [PDF] Noether Theorem - UT Physics
    Back in 1915, Emmy Noether proved the theorem: For every generator of a continuous symmetry of a mechanical system there is a conserved quantity.
  39. [39]
    [PDF] Space-Time Approach to Non-Relativistic Quantum Mechanics
    APRIL, 1948. Space-Time Approach to Non-Relativistic. Quantum Mechanics. R. P. FEYNMAN. Cornell University, Ithaca, New York. Non-relativistic quantum mechanics ...
  40. [40]
    [PDF] THE LAGRANGIAN IN QUANTUM MECHANICS. By PAM Dirac.
    Quantum mer.hanics was built up on a foundation of ana- logy with the Hamiltonian theory of classical mechanics. This is because the classical notion of ...
  41. [41]
    [PDF] Quantum Field Theory - UCSB Physics
    Quantum field theory is the basic mathematical language that is used to ... action principle. δS. δϕa(x). = 0 ,. (22.2) where S = R d4y L(y) is the action ...Missing: seminal | Show results with:seminal
  42. [42]
    [PDF] 17. Lattice Quantum Chromodynamics - Particle Data Group
    May 31, 2024 · For high precision, however, dynamical charm quarks are necessary, and some of the most recent simulations now include them. The b quark ...