Hannay angle

Last updated

In classical mechanics, the Hannay angle is a mechanics analogue of the whirling geometric phase (or Berry phase). It was named after John Hannay of the University of Bristol, UK. Hannay first described the angle in 1985, extending the ideas of the recently formalized Berry phase to classical mechanics. [1]

Contents

Consider a one-dimensional system moving in a cycle, like a pendulum. Now slowly vary a slow parameter , like pulling and pushing on the string of a pendulum. We can picture the motion of the system as having a fast oscillation and a slow oscillation. The fast oscillation is the motion of the pendulum, and the slow oscillation is the motion of our pulling on its string. If we picture the system in phase space, its motion sweeps out a torus.

The adiabatic theorem in classical mechanics states that the action variable, which corresponds to the phase space area enclosed by the system's orbit, remains approximately constant. Thus, after one slow oscillation period, the fast oscillation is back to the same cycle, but its phase on the cycle has changed during the time. The phase change has two leading orders.

The first order is the "dynamical angle", which is simply . This angle depends on the precise details of the motion, and it is of order .

The second order is Hannay's angle, which surprisingly is independent of the precise details of . It depends on the trajectory of , but not how fast or slow it traverses the trajectory. It is of order . [2]

Hannay angle in classical mechanics

The Hannay angle is defined in the context of action-angle coordinates. In an initially time-invariant system, an action variable is a constant. After introducing a periodic perturbation , the action variable becomes an adiabatic invariant, and the Hannay angle for its corresponding angle variable can be calculated according to the path integral that represents an evolution in which the perturbation gets back to the original value [3]

where and are canonical variables of the Hamiltonian, and is the symplectic Hamiltonian 2-form.

Example

Foucault pendulum

The Foucault pendulum is an example from classical mechanics that is sometimes also used to illustrate the Berry phase. Below we study the Foucault pendulum using action-angle variables. For simplicity, we will avoid using the Hamilton–Jacobi equation, which is employed in the general protocol. [4]

We consider a plane pendulum with frequency under the effect of Earth's rotation whose angular velocity is with amplitude denoted as . Here, the direction points from the center of the Earth to the pendulum. The Lagrangian for the pendulum is

The corresponding motion equation is

We then introduce an auxiliary variable that is in fact an angle variable. We now have an equation for :

From its characteristic equation

we obtain its characteristic root (we note that )

The solution is then

After the Earth rotates one full rotation that is , we have the phase change for

The first term is due to dynamic effect of the pendulum and is termed as the dynamic phase, while the second term representing a geometric phase that is essentially the Hannay angle

Rotation of a rigid body

In the rigid body's frame, the direction of the angular momentum moves along one of the curves drawn here. It returns to its starting direction periodically. Contour plot of all solutions to Euler's equations.png
In the rigid body's frame, the direction of the angular momentum moves along one of the curves drawn here. It returns to its starting direction periodically.

A free rigid body tumbling in free space has two conserved quantities: energy and angular momentum vector . Viewed from within the rigid body's frame, the angular momentum direction is moving about, but its length is preserved. After a certain time , the angular momentum direction would return to its starting point.

Viewed in the inertial frame, the body has undergone a rotation (since all elements in SO(3) are rotations). A classical result states that during time , the body has rotated by angle

where is the solid angle swept by the angular momentum direction as viewed from within the rigid body's frame. [5]

Other examples

For the heavy top. [6] For the orbit of earth, periodically perturbed by the orbit of Jupiter. [7]

Related Research Articles

<span class="mw-page-title-main">Simple harmonic motion</span> To-and-fro periodic motion in science and engineering

In mechanics and physics, simple harmonic motion is a special type of periodic motion an object experiences due to a restoring force whose magnitude is directly proportional to the distance of the object from an equilibrium position and acts towards the equilibrium position. It results in an oscillation that is described by a sinusoid which continues indefinitely.

<span class="mw-page-title-main">Standing wave</span> Wave that remains in a constant position

In physics, a standing wave, also known as a stationary wave, is a wave that oscillates in time but whose peak amplitude profile does not move in space. The peak amplitude of the wave oscillations at any point in space is constant with respect to time, and the oscillations at different points throughout the wave are in phase. The locations at which the absolute value of the amplitude is minimum are called nodes, and the locations where the absolute value of the amplitude is maximum are called antinodes.

<span class="mw-page-title-main">Angular velocity</span> Direction and rate of rotation

In physics, angular velocity, also known as angular frequency vector, is a pseudovector representation of how the angular position or orientation of an object changes with time, i.e. how quickly an object rotates around an axis of rotation and how fast the axis itself changes direction.

<span class="mw-page-title-main">Transverse wave</span> Moving wave that has oscillations perpendicular to the direction of the wave

In physics, a transverse wave is a wave that oscillates perpendicularly to the direction of the wave's advance. In contrast, a longitudinal wave travels in the direction of its oscillations. All waves move energy from place to place without transporting the matter in the transmission medium if there is one. Electromagnetic waves are transverse without requiring a medium. The designation “transverse” indicates the direction of the wave is perpendicular to the displacement of the particles of the medium through which it passes, or in the case of EM waves, the oscillation is perpendicular to the direction of the wave.

<span class="mw-page-title-main">Moment of inertia</span> Scalar measure of the rotational inertia with respect to a fixed axis of rotation

The moment of inertia, otherwise known as the mass moment of inertia, angular/rotational mass, second moment of mass, or most accurately, rotational inertia, of a rigid body is a quantity that determines the torque needed for a desired angular acceleration about a rotational axis, akin to how mass determines the force needed for a desired acceleration. It depends on the body's mass distribution and the axis chosen, with larger moments requiring more torque to change the body's rate of rotation by a given amount.

In physics, the CHSH inequality can be used in the proof of Bell's theorem, which states that certain consequences of entanglement in quantum mechanics cannot be reproduced by local hidden-variable theories. Experimental verification of the inequality being violated is seen as confirmation that nature cannot be described by such theories. CHSH stands for John Clauser, Michael Horne, Abner Shimony, and Richard Holt, who described it in a much-cited paper published in 1969. They derived the CHSH inequality, which, as with John Stewart Bell's original inequality, is a constraint—on the statistical occurrence of "coincidences" in a Bell test—which is necessarily true if an underlying local hidden-variable theory exists. In practice, the inequality is routinely violated by modern experiments in quantum mechanics.

In classical and quantum mechanics, geometric phase is a phase difference acquired over the course of a cycle, when a system is subjected to cyclic adiabatic processes, which results from the geometrical properties of the parameter space of the Hamiltonian. The phenomenon was independently discovered by S. Pancharatnam (1956), in classical optics and by H. C. Longuet-Higgins (1958) in molecular physics; it was generalized by Michael Berry in (1984). It is also known as the Pancharatnam–Berry phase, Pancharatnam phase, or Berry phase. It can be seen in the conical intersection of potential energy surfaces and in the Aharonov–Bohm effect. Geometric phase around the conical intersection involving the ground electronic state of the C6H3F3+ molecular ion is discussed on pages 385–386 of the textbook by Bunker and Jensen. In the case of the Aharonov–Bohm effect, the adiabatic parameter is the magnetic field enclosed by two interference paths, and it is cyclic in the sense that these two paths form a loop. In the case of the conical intersection, the adiabatic parameters are the molecular coordinates. Apart from quantum mechanics, it arises in a variety of other wave systems, such as classical optics. As a rule of thumb, it can occur whenever there are at least two parameters characterizing a wave in the vicinity of some sort of singularity or hole in the topology; two parameters are required because either the set of nonsingular states will not be simply connected, or there will be nonzero holonomy.

<span class="mw-page-title-main">Sine wave</span> Wave shaped like the sine function

A sine wave, sinusoidal wave, or sinusoid is a periodic wave whose waveform (shape) is the trigonometric sine function. In mechanics, as a linear motion over time, this is simple harmonic motion; as rotation, it corresponds to uniform circular motion. Sine waves occur often in physics, including wind waves, sound waves, and light waves, such as monochromatic radiation. In engineering, signal processing, and mathematics, Fourier analysis decomposes general functions into a sum of sine waves of various frequencies, relative phases, and magnitudes.

In physics, a wave vector is a vector used in describing a wave, with a typical unit being cycle per metre. It has a magnitude and direction. Its magnitude is the wavenumber of the wave, and its direction is perpendicular to the wavefront. In isotropic media, this is also the direction of wave propagation.

In differential geometry, the four-gradient is the four-vector analogue of the gradient from vector calculus.

In physical systems, damping is the loss of energy of an oscillating system by dissipation. Damping is an influence within or upon an oscillatory system that has the effect of reducing or preventing its oscillation. Examples of damping include viscous damping in a fluid, surface friction, radiation, resistance in electronic oscillators, and absorption and scattering of light in optical oscillators. Damping not based on energy loss can be important in other oscillating systems such as those that occur in biological systems and bikes. Damping is not to be confused with friction, which is a type of dissipative force acting on a system. Friction can cause or be a factor of damping.

<span class="mw-page-title-main">Pendulum (mechanics)</span> Free swinging suspended body

A pendulum is a body suspended from a fixed support so that it swings freely back and forth under the influence of gravity. When a pendulum is displaced sideways from its resting, equilibrium position, it is subject to a restoring force due to gravity that will accelerate it back towards the equilibrium position. When released, the restoring force acting on the pendulum's mass causes it to oscillate about the equilibrium position, swinging it back and forth. The mathematics of pendulums are in general quite complicated. Simplifying assumptions can be made, which in the case of a simple pendulum allow the equations of motion to be solved analytically for small-angle oscillations.

The theoretical and experimental justification for the Schrödinger equation motivates the discovery of the Schrödinger equation, the equation that describes the dynamics of nonrelativistic particles. The motivation uses photons, which are relativistic particles with dynamics described by Maxwell's equations, as an analogue for all types of particles.

<span class="mw-page-title-main">Stokes drift</span> Average velocity of a fluid parcel in a gravity wave

For a pure wave motion in fluid dynamics, the Stokes drift velocity is the average velocity when following a specific fluid parcel as it travels with the fluid flow. For instance, a particle floating at the free surface of water waves, experiences a net Stokes drift velocity in the direction of wave propagation.

In mathematics, the method of steepest descent or saddle-point method is an extension of Laplace's method for approximating an integral, where one deforms a contour integral in the complex plane to pass near a stationary point, in roughly the direction of steepest descent or stationary phase. The saddle-point approximation is used with integrals in the complex plane, whereas Laplace’s method is used with real integrals.

<span class="mw-page-title-main">Envelope (waves)</span> Smooth curve outlining the extremes of an oscillating signal

In physics and engineering, the envelope of an oscillating signal is a smooth curve outlining its extremes. The envelope thus generalizes the concept of a constant amplitude into an instantaneous amplitude. The figure illustrates a modulated sine wave varying between an upper envelope and a lower envelope. The envelope function may be a function of time, space, angle, or indeed of any variable.

In optics, the Fraunhofer diffraction equation is used to model the diffraction of waves when the diffraction pattern is viewed at a long distance from the diffracting object, and also when it is viewed at the focal plane of an imaging lens.

<span class="mw-page-title-main">Trochoidal wave</span> Exact solution of the Euler equations for periodic surface gravity waves

In fluid dynamics, a trochoidal wave or Gerstner wave is an exact solution of the Euler equations for periodic surface gravity waves. It describes a progressive wave of permanent form on the surface of an incompressible fluid of infinite depth. The free surface of this wave solution is an inverted (upside-down) trochoid – with sharper crests and flat troughs. This wave solution was discovered by Gerstner in 1802, and rediscovered independently by Rankine in 1863.

In fluid dynamics, Beltrami flows are flows in which the vorticity vector and the velocity vector are parallel to each other. In other words, Beltrami flow is a flow in which the Lamb vector is zero. It is named after the Italian mathematician Eugenio Beltrami due to his derivation of the Beltrami vector field, while initial developments in fluid dynamics were done by the Russian scientist Ippolit S. Gromeka in 1881.

Tau functions are an important ingredient in the modern mathematical theory of integrable systems, and have numerous applications in a variety of other domains. They were originally introduced by Ryogo Hirota in his direct method approach to soliton equations, based on expressing them in an equivalent bilinear form.

References

  1. Hannay, J H (1985-02-01). "Angle variable holonomy in adiabatic excursion of an integrable Hamiltonian". Journal of Physics A: Mathematical and General. 18 (2): 221–230. doi:10.1088/0305-4470/18/2/011. ISSN   0305-4470.
  2. Robbins, J M (2016-10-28). "The Hannay angle, thirty years on". Journal of Physics A: Mathematical and Theoretical. 49 (43): 431002. doi:10.1088/1751-8113/49/43/431002. hdl: 1983/2992186e-5dde-4a3f-a2a9-67377afcadf9 . ISSN   1751-8113.
  3. Toshikaze Kariyado; Yasuhiro Hatsugai (2016). "Hannay Angle: Yet Another Symmetry-Protected Topological Order Parameter in Classical Mechanics". J. Phys. Soc. Jpn. 85 (4): 043001. arXiv: 1508.06946 . Bibcode:2016JPSJ...85d3001K. doi:10.7566/JPSJ.85.043001. S2CID   119297582.
  4. Khein, Alexander; Nelson, D. F. (1993-02-01). "Hannay angle study of the Foucault pendulum in action-angle variables". American Journal of Physics. 61 (2): 170–174. doi:10.1119/1.17332. ISSN   0002-9505.
  5. Montgomery, Richard (1991-05-01). "How much does the rigid body rotate? A Berry's phase from the 18th century". American Journal of Physics. 59 (5): 394–398. doi:10.1119/1.16514. ISSN   0002-9505.
  6. Park, Changsoo (2023-05-01). "Heavy symmetric tops and the Hannay angle". American Journal of Physics. 91 (5): 357. doi:10.1119/5.0101149. ISSN   0002-9505.
  7. Berry, M V; Morgan, M A (1996-05-01). "Geometric angle for rotated rotators, and the Hannay angle of the world". Nonlinearity. 9 (3): 787–799. doi:10.1088/0951-7715/9/3/009. ISSN   0951-7715.