Fact-checked by Grok 2 weeks ago

Spherical pendulum

A spherical pendulum is a classical system consisting of a point attached to a fixed by a light, inextensible or of fixed length l, allowing the mass to move freely under on the surface of a of l centered at the pivot, without . This setup contrasts with the simple pendulum, which is constrained to oscillatory motion in a single vertical plane, by permitting three-dimensional trajectories that depend on two independent coordinates, typically the polar \theta (from the vertical) and the azimuthal angle \phi. The dynamics of the spherical pendulum are governed by Lagrangian mechanics, with the Lagrangian expressed in spherical coordinates as L = \frac{1}{2} m l^2 (\dot{\theta}^2 + \sin^2\theta \, \dot{\phi}^2) + m g l \cos\theta, where m is the mass and g is gravitational acceleration. The system conserves total energy and the z-component of angular momentum p_\phi = m l^2 \sin^2\theta \, \dot{\phi}, rendering the motion integrable and leading to a variety of periodic orbits. Notable trajectories include small-amplitude oscillations approximating independent simple harmonic motions in orthogonal directions with frequency \sqrt{g/l}, uniform circular motion at constant \theta (known as a conical pendulum), and more complex precessional paths. The spherical pendulum exemplifies key principles in , such as the use of and conserved quantities, and serves as a foundational model for studying nonlinear dynamics and stability. A prominent application is the , a long-period spherical pendulum designed to demonstrate through the apparent of its plane of oscillation at a rate \Omega \sin\lambda, where \Omega is Earth's and \lambda is the .

Introduction

Definition and Setup

A spherical pendulum consists of a point mass, referred to as the , attached to a fixed suspension point by an inextensible of fixed length l, which constrains the bob to move on the surface of a of l under the action of . This setup allows the bob to swing freely in three dimensions without restriction to a . In contrast to the simple pendulum, which possesses only one degree of freedom and oscillates in a fixed , the spherical pendulum has two , corresponding to independent angular displacements in polar and azimuthal directions. These enable complex trajectories, such as circular orbits or looping paths, depending on the input. The model assumes a massless and perfectly flexible , a bob treated as a with negligible size, a uniform downward , and the absence of dissipative effects like or air resistance. Typical initial conditions involve releasing the bob from rest at an from the vertical position, which determines the subsequent motion. A of the setup, illustrating the suspension point, , spherical path of the bob, and the angles defining its position relative to the vertical, aids in visualizing the .

Historical Context

The concept of the spherical pendulum originated in the late as an extension of studies on simple and s, building on ' foundational work in dynamics for timekeeping. In 1673, Huygens analyzed the —a special case of the spherical pendulum where the mass moves in a horizontal circle—in his treatise Horologium Oscillatorium, introducing the notion of to explain the under and . This analysis laid early groundwork for understanding multi-dimensional pendulum motions, with analogies drawn to for orbital stability. By the late 18th century, advanced the theoretical framework through , treating the spherical pendulum as a key example in his 1788 work Mécanique Analytique. Lagrange's variational approach generalized the system's dynamics, emphasizing coordinate transformations and constraints without relying on Newtonian forces directly. In the early 19th century, extended these ideas into formulations, incorporating Poisson brackets to describe conserved quantities in the spherical pendulum, as detailed in his multi-volume Traité de mécanique (1811–1833). These contributions solidified the spherical pendulum's role in formalizing . A significant milestone occurred in 1851 when employed a long spherical pendulum to demonstrate , influencing studies of rotational effects on pendular motion while focusing on the ideal unconstrained case. In the late , Poincaré's qualitative analysis around 1890, particularly in his appendix to Les méthodes nouvelles de la mécanique céleste (1899), introduced methods for studying nonlinear behaviors and stability in dynamical systems with multiple , foreshadowing applications in and influencing later analyses of pendular systems. The saw a revival of interest in the spherical pendulum within , driven by computational simulations from the 1960s onward that revealed transitions to chaotic motion for certain energy levels. Pioneering numerical studies, such as those distinguishing quasi-periodic from chaotic regimes in experimental setups, underscored its importance in nonlinear dynamics and served as a bridge to more complex systems like the .

Mathematical Description

Coordinate Systems

The position of the bob in a spherical pendulum is most naturally described using spherical coordinates, where the radial distance r is fixed at the length of the pendulum string l, the polar angle \theta measures the deviation from the vertical axis (with $0 \leq \theta \leq \pi), and the azimuthal angle \phi describes the rotation around the vertical axis (with $0 \leq \phi < 2\pi). This coordinate system aligns directly with the geometric constraint of motion on a sphere of radius l, simplifying the mathematical representation by eliminating the radial degree of freedom. The transformation from these to , assuming the suspension point at the origin and the positive pointing upward against , is given by: \begin{align} x &= l \sin \theta \cos \phi, \\ y &= l \sin \theta \sin \phi, \\ z &= -l \cos \theta. \end{align} Here, \theta = 0 corresponds to the equilibrium position directly below the suspension point, where z = -l, and increasing \theta reflects the bob's swing away from the vertical. The advantages of spherical coordinates for the spherical pendulum stem from their adaptation to the spherical constraint surface, where \theta quantifies the angular displacement from equilibrium and \phi captures the rotational freedom in the horizontal plane, thereby providing an intuitive parameterization of the two-dimensional motion. This choice reduces the system's description to two independent variables without loss of generality. In the Lagrangian formulation, \theta and \phi serve as the generalized coordinates q_1 = \theta and q_2 = \phi, with corresponding generalized velocities \dot{\theta} and \dot{\phi}, fully specifying the configuration and kinematics of the system. The fixed length constraint r = l is holonomic, expressible as an equality x^2 + y^2 + z^2 = l^2, which reduces the three-dimensional Cartesian space to two degrees of freedom on the sphere.

Energy Expressions

The kinetic energy T of a spherical pendulum, consisting of a point mass m attached to a massless rod of fixed length l, arises from its velocity components in spherical coordinates, where \theta is the polar angle from the vertical downward axis and \phi is the azimuthal angle. The position of the mass is given by Cartesian coordinates x = l \sin \theta \cos \phi, y = l \sin \theta \sin \phi, and z = -l \cos \theta (with the z-axis pointing upward). Differentiating these with respect to time yields the velocity squared as v^2 = l^2 \dot{\theta}^2 + l^2 \sin^2 \theta \, \dot{\phi}^2, so T = \frac{1}{2} m l^2 \left( \dot{\theta}^2 + \sin^2 \theta \, \dot{\phi}^2 \right). This expression separates the contributions from motion in the meridional plane (\dot{\theta}) and the azimuthal rotation (\sin \theta \, \dot{\phi}). The potential energy V is due to gravity and depends on the height of the mass above the lowest point. Taking V = 0 at the bottom (\theta = 0), the height is h = l (1 - \cos \theta), leading to V = m g l (1 - \cos \theta), where g is the acceleration due to gravity. Equivalently, setting V = 0 at the horizontal plane gives V = -m g l \cos \theta; both forms differ only by a constant and are used interchangeably in formulations. The total mechanical energy E = T + V is conserved in the absence of dissipative forces, as the system is conservative. The Lagrangian L = T - V provides the foundation for subsequent dynamical analyses using this energy framework, with parameters m, g, and l defining the system's scale.

Classical Mechanics Formulation

Lagrangian Approach

The Lagrangian formulation provides an analytical method to derive the equations of motion for the spherical pendulum by employing generalized coordinates, typically the polar angle \theta and azimuthal angle \phi. The Lagrangian L is constructed as the difference between the kinetic energy T and potential energy V, where T = \frac{1}{2} m l^2 (\dot{\theta}^2 + \sin^2 \theta \, \dot{\phi}^2) and V = -m g l \cos \theta, with m the mass, l the fixed length of the inextensible string or rod, g the gravitational acceleration, and the potential zero at the suspension point. The equations of motion follow from the Euler-Lagrange equations, \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}_i} \right) - \frac{\partial L}{\partial q_i} = 0, applied to the coordinates q_i = \theta, \phi. For \theta, this yields m l^2 \ddot{\theta} - m l^2 \sin \theta \cos \theta \, \dot{\phi}^2 + m g l \sin \theta = 0, which balances the inertial, centrifugal, and gravitational torques. For \phi, the equation simplifies to \frac{d}{dt} \left( m l^2 \sin^2 \theta \, \dot{\phi} \right) = 0, indicating that the angular momentum h = m l^2 \sin^2 \theta \, \dot{\phi} about the vertical axis is conserved due to the rotational symmetry of the system. Using the conserved angular momentum h, the dynamics can be reduced to an effective one-dimensional problem in \theta. The centrifugal term in the \theta equation is rewritten using \dot{\phi} = h / (m l^2 \sin^2 \theta), leading to an effective potential V_{\text{eff}}(\theta) = V(\theta) + \frac{h^2}{2 m l^2 \sin^2 \theta}, where the second term arises from the rotational kinetic energy. The total energy then takes the form E = \frac{1}{2} m l^2 \dot{\theta}^2 + V_{\text{eff}}(\theta), allowing the motion in \theta to be analyzed as that of a particle in this effective potential, solvable via separation of variables or quadrature.

Hamiltonian Approach

The Hamiltonian formulation of the spherical pendulum provides a phase space description of the dynamics, emphasizing the symplectic structure and conserved quantities inherent to the system. Starting from the Lagrangian L = T - V, where the kinetic energy is T = \frac{1}{2} m l^2 (\dot{\theta}^2 + \sin^2 \theta \, \dot{\phi}^2) and the potential energy is V = -m g l \cos \theta, the canonical momenta are obtained via the partial derivatives with respect to the velocities: p_\theta = \frac{\partial L}{\partial \dot{\theta}} = m l^2 \dot{\theta} and p_\phi = \frac{\partial L}{\partial \dot{\phi}} = m l^2 \sin^2 \theta \, \dot{\phi}. The momentum p_\phi represents the conserved z-component of angular momentum, often denoted as h. The Hamiltonian H, which equals the total energy in this time-independent case, is formed through the Legendre transform: H = p_\theta \dot{\theta} + p_\phi \dot{\phi} - L. Expressing the velocities in terms of the momenta yields H = \frac{p_\theta^2}{2 m l^2} + \frac{p_\phi^2}{2 m l^2 \sin^2 \theta} - m g l \cos \theta. This expression separates into kinetic and effective potential terms, with the second kinetic term incorporating a centrifugal contribution from the conserved angular momentum. Hamilton's equations govern the evolution in phase space: \dot{\theta} = \frac{\partial H}{\partial p_\theta}, \dot{p}_\theta = -\frac{\partial H}{\partial \theta}, \dot{\phi} = \frac{\partial H}{\partial p_\phi}, and \dot{p}_\phi = -\frac{\partial H}{\partial \phi}. Explicitly, these are \dot{\theta} = \frac{p_\theta}{m l^2}, \dot{\phi} = \frac{p_\phi}{m l^2 \sin^2 \theta}, \dot{p}_\phi = 0 (confirming conservation of h = p_\phi), and \dot{p}_\theta = -\frac{\partial H}{\partial \theta} = \frac{p_\phi^2 \cos \theta}{m l^2 \sin^3 \theta} - m g l \sin \theta. These first-order equations are equivalent to the second-order equations of motion from the Lagrangian formulation. The phase space is four-dimensional, parameterized by \theta, \phi, p_\theta, and p_\phi, with \phi being a cyclic (ignorable) coordinate that does not appear in H, thereby enforcing the conservation of p_\phi. Trajectories in this phase space lie on surfaces of constant H and constant p_\phi, reducing the effective dynamics to a two-dimensional problem in \theta and p_\theta. This approach offers advantages in contexts requiring preservation of the symplectic geometry, such as numerical integration schemes like symplectic integrators that maintain long-term stability, and in quantization procedures where the classical phase space directly informs the quantum Hilbert space structure.

Dynamics and Motion

Equations of Motion

The equations of motion for the spherical pendulum describe the dynamics of a point mass m attached to a massless rod of fixed length l, subject to gravity with acceleration g. These are derived from the Lagrangian formulation and take the form of two coupled second-order ordinary differential equations (ODEs) in the spherical coordinates \theta(t) (polar angle from the downward vertical) and \phi(t) (azimuthal angle). The equation for \ddot{\theta} incorporates both gravitational restoring torque and centrifugal effects from azimuthal motion: \ddot{\theta} = \sin\theta \cos\theta \, \dot{\phi}^2 - \frac{g}{l} \sin\theta The equation for \ddot{\phi} arises from the conservation of angular momentum about the vertical axis and is given by: \ddot{\phi} = -\frac{2 \cos\theta}{\sin\theta} \, \dot{\theta} \, \dot{\phi} These ODEs fully determine the system's evolution, with the \phi equation reflecting the rotational symmetry that conserves the z-component of angular momentum h = m l^2 \sin^2\theta \, \dot{\phi}. An alternative first-order form for the \theta dynamics can be obtained by substituting the conserved angular momentum h into the \theta equation, eliminating \dot{\phi}: \ddot{\theta} = -\frac{g}{l} \sin\theta + \frac{h^2 \cos\theta}{m^2 l^4 \sin^3\theta} Here, \dot{\phi} = h / (m l^2 \sin^2\theta), and the second term represents the centrifugal contribution scaled by the system's parameters. This form is particularly useful for analyzing effective potentials or reducing the problem to a single degree of freedom in \theta. The initial value problem requires specifying initial conditions \theta(0), \dot{\theta}(0), \phi(0), and \dot{\phi}(0), or equivalently \theta(0), \dot{\theta}(0), \phi(0), and h (computed from the initial \dot{\phi}(0)). These conditions determine the trajectory, with \phi(0) often set to zero without loss of generality due to rotational invariance. Numerical integration of these equations is challenging, particularly for motions where \theta approaches 0 or \pi, as the \sin^3\theta term in the denominator introduces stiffness from rapidly varying . Standard explicit integrators can exhibit instability or energy drift in such regimes, necessitating or energy-momentum-preserving methods to maintain long-term accuracy and conserve the system's invariants. The equations of motion can be derived using either the or . The system admits equilibrium solutions where \dot{\theta} = 0 and \ddot{\theta} = 0, \dot{\phi} = 0, \ddot{\phi} = 0. At \theta = 0 (downward position), the equilibrium is stable, corresponding to the minimum of the gravitational potential. At \theta = \pi (upward position), the equilibrium is unstable, as it lies at the potential maximum.

Conserved Quantities

The spherical pendulum possesses two independent conserved quantities arising from its symmetries and the nature of the conservative forces acting on it. The total mechanical energy E is conserved due to the time-independence of the Lagrangian, expressed after reduction as E = \frac{1}{2} m l^2 \dot{\theta}^2 + V_{\text{eff}}(\theta), where V_{\text{eff}}(\theta) = m g l (1 - \cos\theta) + \frac{h_z^2}{2 m l^2 \sin^2\theta} incorporates the gravitational potential and the centrifugal barrier from the conserved angular momentum. The second conserved quantity is the z-component of angular momentum h_z about the vertical axis, which remains constant owing to the rotational invariance of the system under rotations about this axis: h_z = m l^2 \sin^2\theta \, \dot{\phi}. This conservation stems from Noether's theorem applied to the ignorable coordinate \phi. These conserved quantities enable a significant simplification of the dynamics. Substituting h_z into the energy expression eliminates the azimuthal angle \phi, reducing the two-degree-of-freedom system to an effective one-dimensional motion in the polar angle \theta, akin to a particle moving in the effective potential V_{\text{eff}}(\theta). The value of \theta is then bounded between turning points \theta_{\min} and \theta_{\max}, determined by the intersections of the horizontal line at energy E with V_{\text{eff}}(\theta), ensuring periodic oscillation in \theta. With exactly two independent integrals of motion for a two-degree-of-freedom Hamiltonian system, and given that they Poisson commute due to the system's symmetries, the spherical pendulum is completely integrable. There are no additional exact integrals beyond energy and h_z. The integrability implies that the phase space is foliated into invariant tori, on which the motion is quasi-periodic, and the Poincaré invariants—specifically, the areas of these tori—are preserved by the Hamiltonian flow via Liouville's theorem.

Trajectories and Behavior

Types of Trajectories

The trajectories of a spherical pendulum are determined by the initial conditions, which set the values of the conserved total energy E and the z-component of angular momentum h, thereby bounding the possible motion in the configuration space. When the angular momentum h \approx 0, the motion confines to a vertical plane, resembling that of a simple pendulum, known as planar libration. In this case, the azimuthal angle \phi remains constant, and the polar angle \theta oscillates between turning points defined by the effective potential, without the bob completing full circles around the vertical axis. A special case occurs for constant \theta > 0 with uniform \dot{\phi}, resulting in a trajectory where the bob traces a circle at fixed height. The condition yields \cos \theta = g / (l \omega^2), with \omega = \dot{\phi}, balancing gravitational and centripetal forces. For general precession, \theta oscillates between minimum and maximum values while \phi advances at a nearly steady rate, producing quasiperiodic motion on invariant tori in . Projections of these paths onto the plane or the \theta-\phi plane often form patterns, resembling closed loops that fill densely over time due to incommensurate frequencies. The separatrix defines the boundary between librational and rotational regimes in \theta, occurring at total energy E = m g l (with potential V = -m g l \cos \theta), where the trajectory asymptotically approaches the unstable at \theta = \pi with zero . For total energies above the separatrix (E > m g l), the motion in θ becomes rotational, with θ advancing continuously through multiple full cycles, enabling the pendulum bob to pass over the unstable at θ = π with positive , while the azimuthal motion in φ continues to precess based on the h. These trajectories are also quasiperiodic and bounded by the constants of motion. In projection views, trajectories in the \theta-\phi plane appear as closed curves for commensurate frequencies (e.g., conical motion) or dense windings for incommensurate cases (e.g., precession), contrasting with the full three-dimensional paths that trace surfaces on the sphere bounded by the constants of motion.

Small-Angle Approximation

For small angles θ ≪ 1, the equations of motion for the spherical pendulum are linearized by approximating sin θ ≈ θ and cos θ ≈ 1 - θ²/2. This simplification is valid when the displacement from the vertical equilibrium is small, allowing the neglect of higher-order terms, including the centrifugal contribution if the conserved angular momentum about the vertical axis h is small. The resulting dynamics approximate simple harmonic motion in the horizontal plane. In Cartesian coordinates, with the suspension point at the origin and the vertical downward, the position of the bob is approximately x = l θ cos φ and y = l θ sin φ for small θ, where l is the of the . The under this approximation becomes that of two independent harmonic oscillators: \mathcal{L} \approx \frac{1}{2} m (\dot{x}^2 + \dot{y}^2) - \frac{1}{2} m \frac{g}{l} (x^2 + y^2), leading to the decoupled \ddot{x} + \frac{g}{l} x = 0, \quad \ddot{y} + \frac{g}{l} y = 0. The general solution is a linear combination of sinusoidal functions with angular frequency ω = √(g/l) for both x and y directions, corresponding to normal modes of oscillation along the principal axes. In the planar case (h = 0), the motion reduces to a single equation θ̈ + (g/l) θ = 0, with the same frequency. For h ≠ 0, the coordinates remain decoupled in this linear regime, but the overall trajectory is an ellipse in the xy-plane due to the phase difference between x and y oscillations; the frequency stays ω = √(g/l) as the centrifugal term contributes only at higher order. The conserved angular momentum h = m (x \dot{y} - y \dot{x}) determines the eccentricity of the ellipse but does not alter the oscillation frequency in the small-angle limit. The at θ = 0 is stable for all small motions, as the linearized equations describe bounded oscillations around this point, with the exhibiting a quadratic minimum. For moderate amplitudes where the small-angle assumption begins to fail but elliptic integrals are undesirable for the exact , an approximate expression avoids the full nonlinear solution while capturing the leading correction: T \approx 2\pi \sqrt{\frac{l}{g}} \left(1 + \frac{\theta_\mathrm{max}^2}{16}\right), where θ_max is the maximum angular displacement in radians. This series expansion, derived from the Taylor approximation of the elliptic integral of the first kind, provides accuracy within a few percent for θ_max up to approximately 0.3 radians (∼17°). For the spherical pendulum, this formula applies similarly to nearly planar elliptical paths with small eccentricity. A key application of the is the modeling of the , where the spherical geometry allows motion in any horizontal plane. In the non-inertial frame of the rotating , Coriolis terms couple the x and y equations, leading to of the oscillation plane at rate Ω sin λ (with Ω the and λ the ), while the oscillation frequency remains ≈ √(g/l). This demonstrates the without relying on the full nonlinear dynamics.

Chaotic Dynamics

The spherical pendulum, with two , possesses exactly two independent conserved quantities—total and the z-component of —sufficient for Liouville integrability in the unperturbed case, allowing solutions via and elliptic functions. However, under small perturbations such as harmonic forcing or damping, the system becomes non-integrable, as no additional global exists to ensure full solvability; instead, the Kolmogorov-Arnold-Moser (KAM) theorem dictates that most invariant tori persist as quasiperiodic motions, while others break down into regions. Chaos emerges in the perturbed spherical when the total energy exceeds the separatrix value of the , typically E > m g l (where m is the , g the , and l the pendulum length), or for sufficiently high h, leading to the formation of homoclinic tangles where stable and unstable manifolds intersect transversely. This transverse splitting, quantified by the Melnikov-Holmes-Marsden having real zeros, results in positive Lyapunov exponents, signifying divergence of infinitesimally close trajectories and the onset of sensitive dependence on initial conditions. Key indicators of chaotic behavior are evident in Poincaré sections, which slice the at fixed energy levels and angular positions, revealing isolated regular islands (surviving KAM tori) embedded in a broader sea of scattered points, contrasting with the closed curves of integrable motion. These sections highlight the system's sensitivity, where nearby initial conditions can lead to vastly different outcomes, such as versus near the separatrix. The exploration of such dynamics traces to Henri Poincaré's 1892 doctoral thesis on the , where he analyzed a periodically perturbed as an , uncovering homoclinic tangles and the impossibility of long-term predictions due to exponential divergence—foundational insights into Hamiltonian chaos overlooked until the mid-20th century. Modern studies rely on numerical simulations, such as Runge-Kutta integration, to map fractal basin boundaries separating coexisting attractors and confirm the absence of closed-form solutions for chaotic regimes, emphasizing the pendulum's role in validating KAM predictions and chaos diagnostics.

References

  1. [1]
    Spherical Pendulum - Richard Fitzpatrick
    This type of pendulum is usually called a conical pendulum, since the string attached to the pendulum bob sweeps out a cone as the bob rotates.Missing: physics | Show results with:physics
  2. [2]
  3. [3]
    [PDF] Physics 325 – Homework #12 due in 325 homework box by Fri, 1 pm
    The “spherical pendulum” is just a simple pendulum that is free to move in any sideways direction. (By contrast, the unqualified word “pendulum” or the explicit ...<|control11|><|separator|>
  4. [4]
    [PDF] Midterm
    The spherical pendulum is just a simple pendulum that is free to move in any sideways direction under the influence of gravity. The bob of a ...
  5. [5]
  6. [6]
    Turning points of the spherical pendulum and the golden ratio - ADS
    We study the turning point problem of a spherical pendulum. The special cases of the simple pendulum and the conical pendulum are noted.
  7. [7]
    MATHEMATICA TUTORIAL, Part 2.3: Spherical pendulum
    ### Summary of Coordinate Systems for the Spherical Pendulum
  8. [8]
    [PDF] Huygens and the Beginnings of Rational Mechanics
    As remarked before, Huygens constructed a conical pendulum clock in late 1659 and used it to measure g, obtaining a (rounded-off) value of 15.6 Rhenish ft ...
  9. [9]
    [PDF] 2. The Lagrangian Formalism - DAMTP
    The reason for this is the existence of two conservation laws for the spherical pendulum (energy and angular momentum) compared to just one (energy) for the ...
  10. [10]
    [PDF] 4. The Hamiltonian Formalism
    For small θ and small momentum, the pendulum oscillates back and forth, motion which appears as an ellipse in phase space. But for large momentum, the pendulum.
  11. [11]
    Foucault's pendulum - Panthéon
    The astronomer and physicist at the origin of this revolutionary experiment is French and is called Léon Foucault.
  12. [12]
    History of dynamical systems - Scholarpedia
    Oct 21, 2011 · The qualitative theory of dynamical systems originated in Poincaré's work on celestial mechanics (Poincaré 1899), and specifically in a 270-page ...
  13. [13]
    Distinguishing the transition to chaos in a spherical pendulum
    This discovery provides a convenient means for comparison of complex motions in the numerical and experimental air pendulum systems.Missing: theory | Show results with:theory
  14. [14]
    [PDF] Question 1: Spherical Pendulum
    The motion of the pendulum can therefore be described by the polar angle θ, the azimuthal angle φ, and their rates of change.
  15. [15]
    [PDF] Constraints. Noninertial coordinate systems
    8.1.1 Spherical Pendulum. Consider a spherical pendulum, a mass point suspending on a string of length l, so the particle is constrained to move on the ...
  16. [16]
    [PDF] EP 222: Classical Mechanics Tutorial Sheet 1
    Obtain the Lagrange equations of motion for a spherical pendulum, i.e., a point mass suspended by a rigid weightless rod. Soln: It is best to use spherical ...<|separator|>
  17. [17]
    [PDF] Intermediate Classical Mechanics Charles B. Thorn1 - UF Physics
    Now we would like to extend the concepts of kinetic energy and potential energy to more ... As an example consider the spherical pendulum. Introducing ...
  18. [18]
    [PDF] Phys 7221 Homework #2
    Sep 18, 2006 · its kinetic energy is. T = 1. 2 mv2 = 1. 2 ml2( ˙ψ2 + sin2 ψ. ˙ θ2), and the potential energy is. V = −mg · r = −mgl cosψ. The Lagrangian is. L ...
  19. [19]
    Example: A spherical pendulum
    This last equation, the equation of motion, shows a pseudopotential like that of angular momentum in the orbit problem. ... will be set by initial conditions.
  20. [20]
    None
    ### Summary of Spherical Pendulum in Hamiltonian Dynamics
  21. [21]
    [PDF] Chapter 11 Hamiltonian formulation - Physics
    PROBLEM 11-8: (a) Write the Hamiltonian for a spherical pendulum of length ` and mass m, using the polar angle θ and azimuthal angle ϕ as generalized ...
  22. [22]
    THE QUANTUM MECHANICAL SPHERICAL PENDULUM We begin ...
    As a Hamiltonian system the spherical pendulum has a configuration space. S2 = {q = (<?1,<Z2,<73) e R3 |l = q\ + q\ + q| = (q,q)} and a phase space. T*S2 ...
  23. [23]
    [PDF] Plane and Spherical Pendulums
    Note that the Lagrangian (25) can be also obtained from the Lagrangian of the spherical pendulum, Eq. (1), by setting ˙ϕ ≡ 0, meaning that the plane pendulum ...
  24. [24]
    [PDF] NONLINEAR DYNAMICS OF THE 3D PENDULUM - CCoM
    There are two conserved quantities, or integrals of motion, for the 3D pendulum. First, the total energy, which is the sum of the rotational kinetic energy and.
  25. [25]
    Introduction to Integrable Systems
    The differential equation describing the motion of the point mass is then ¨θ = − sin θ. ... Thus the spherical pendulum is a completely integrable system.
  26. [26]
    THE QUANTUM MECHANICAL SPHERICAL PENDULUM We begin ...
    Hence the spherical pendulum is completely integrable. Consider the energy momentum mapping. «#: T*S2 - R2 : (q,p) -. (E(q,p),L(q,p)). (see Figure 1). Suppose ...
  27. [27]
    [PDF] Spherical Pendulum, Actions, and Spin - Members
    Dec 16, 1996 · The classical and quantum mechanics of a spherical pendulum are worked out, including the dynamics of a suspending frame with moment of iner ...
  28. [28]
    [PDF] Nonlinear Mechanics - Oregon State University
    Jan 8, 2012 · 1.5.1 The Spherical Pendulum . . . . . . . . . . . . . . . . . 15. 2 ... the laws of physics are invariant under rotation. Despite its ...<|separator|>
  29. [29]
    [PDF] Challenge Problem Set 1, Math 292 Spring 2011 - Rutgers University
    Jan 28, 2011 · This challenge problem set concerns small amplitude oscillations of the spherical pendulum. The spherical pendulum consists of a mass moving ...
  30. [30]
    None
    ### Summary of Approximate Formula for Pendulum Period
  31. [31]
    On the isoenergetical non-degeneracy of the spherical pendulum
    For the system describing the motion of the spherical pendulum we find explicitly the set where the condition of isoenergetical non-degeneracy important for ...
  32. [32]
    On the Chaotic Dynamics of a Spherical Pendulum with a ...
    The existence of real zeros of the MHM integral implies the possible chaotic motion of the harmonically forced spherical pendulum as a result from the ...Missing: integrability | Show results with:integrability