Classical probability density

Last updated

The classical probability density is the probability density function that represents the likelihood of finding a particle in the vicinity of a certain location subject to a potential energy in a classical mechanical system. These probability densities are helpful in gaining insight into the correspondence principle and making connections between the quantum system under study and the classical limit. [1] [2] [3]

Contents

Mathematical background

Consider the example of a simple harmonic oscillator initially at rest with amplitude A. Suppose that this system was placed inside a light-tight container such that one could only view it using a camera which can only take a snapshot of what's happening inside. Each snapshot has some probability of seeing the oscillator at any possible position x along its trajectory. The classical probability density encapsulates which positions are more likely, which are less likely, the average position of the system, and so on. To derive this function, consider the fact that the positions where the oscillator is most likely to be found are those positions at which the oscillator spends most of its time. Indeed, the probability of being at a given x-value is proportional to the time spent in the vicinity of that x-value. If the oscillator spends an infinitesimal amount of time dt in the vicinity dx of a given x-value, then the probability P(x) dx of being in that vicinity will be

Since the force acting on the oscillator is conservative and the motion occurs over a finite domain, the motion will be cyclic with some period which will be denoted T. Since the probability of the oscillator being at any possible position between the minimum possible x-value and the maximum possible x-value must sum to 1, the normalization

is used, where N is the normalization constant. Since the oscillating mass covers this range of positions in half its period (a full period goes from A to +A then back to A) the integral over t is equal to T/2, which sets N to be 2/T.

Using the chain rule, dt can be put in terms of the height at which the mass is lingering by noting that dt = dx/(dx/dt), so our probability density becomes

where v(x) is the speed of the oscillator as a function of its position. (Note that because speed is a scalar, v(x) is the same for both half periods.) At this point, all that is needed is to provide a function v(x) to obtain P(x). For systems subject to conservative forces, this is done by relating speed to energy. Since kinetic energy K is 12mv2 and the total energy E = K + U, where U(x) is the potential energy of the system, the speed can be written as

Plugging this into our expression for P(x) yields

Though our starting example was the harmonic oscillator, all the math up to this point has been completely general for a particle subject to a conservative force. This formula can be generalized for any one-dimensional physical system by plugging in the corresponding potential energy function. Once this is done, P(x) is readily obtained for any allowed energy E.

Examples

Simple harmonic oscillator

The probability density function of the n = 30 state of the quantum harmonic oscillator. The solid plot represents the quantum mechanical probability density, while the dotted line represents the classical probability density. The dashed vertical lines indicate the classical turning points of the system. QHOn30pdf.svg
The probability density function of the n = 30 state of the quantum harmonic oscillator. The solid plot represents the quantum mechanical probability density, while the dotted line represents the classical probability density. The dashed vertical lines indicate the classical turning points of the system.

Starting with the example used in the derivation above, the simple harmonic oscillator has the potential energy function

where k is the spring constant of the oscillator and ω = 2π/T is the natural angular frequency of the oscillator. The total energy of the oscillator is given by evaluating U(x) at the turning points x = ±A. Plugging this into the expression for P(x) yields

This function has two vertical asymptotes at the turning points, which makes physical sense since the turning points are where the oscillator is at rest, and thus will be most likely found in the vicinity of those x values. Note that even though the probability density function tends toward infinity, the probability is still finite due to the area under the curve, and not the curve itself, representing probability.

Bouncing ball

Probability density functions of the quantum (red) and classical (black) quantum bouncing ball for n = 50. The turning point here is labeled zn (what this section refers to as h). Qmballprob050.svg
Probability density functions of the quantum (red) and classical (black) quantum bouncing ball for n = 50. The turning point here is labeled zn (what this section refers to as h).

For the lossless bouncing ball, the potential energy and total energy are

where h is the maximum height reached by the ball. Plugging these into P(z) yields

where the relation was used to simplify the factors out front. The domain of this function is (the ball does not fall through the floor at z = 0), so the distribution is not symmetric as in the case of the simple harmonic oscillator. Again, there is a vertical asymptote at the turning point z = h.

Momentum-space distribution

In addition to looking at probability distributions in position space, it is also helpful to characterize a system based on its momentum. Following a similar argument as above, the result is [2]

where F(x) = dU/dx is the force acting on the particle as a function of position. In practice, this function must be put in terms of the momentum p by change of variables.

Simple harmonic oscillator

Taking the example of the simple harmonic oscillator above, the potential energy and force can be written as

Identifying (2mE)1/2 = p0 as the maximum momentum of the system, this simplifies to

Note that this has the same functional form as the position-space probability distribution. This is specific to the problem of the simple harmonic oscillator and arises due to the symmetry between x and p in the equations of motion.

Bouncing ball

The example of the bouncing ball is more straightforward, since in this case the force is a constant,

resulting in the probability density function

where p0 = m(2gh)1/2 is the maximum momentum of the ball. In this system, all momenta are equally probable.

See also

Related Research Articles

In classical mechanics, a harmonic oscillator is a system that, when displaced from its equilibrium position, experiences a restoring force F proportional to the displacement x:

<span class="mw-page-title-main">Maxwell–Boltzmann distribution</span> Specific probability distribution function, important in physics

In physics, the Maxwell–Boltzmann distribution, or Maxwell(ian) distribution, is a particular probability distribution named after James Clerk Maxwell and Ludwig Boltzmann.

<span class="mw-page-title-main">Uncertainty principle</span> Foundational principle in quantum physics

In quantum mechanics, the uncertainty principle is any of a variety of mathematical inequalities asserting a fundamental limit to the accuracy with which the values for certain pairs of physical quantities of a particle, such as position, x, and momentum, p, can be predicted from initial conditions.

<span class="mw-page-title-main">Probability density function</span> Function whose integral over a region describes the probability of an event occurring in that region

In probability theory, a probability density function (PDF), or density of a continuous random variable, is a function whose value at any given sample in the sample space can be interpreted as providing a relative likelihood that the value of the random variable would be equal to that sample. Probability density is the probability per unit length, in other words, while the absolute likelihood for a continuous random variable to take on any particular value is 0, the value of the PDF at two different samples can be used to infer, in any particular draw of the random variable, how much more likely it is that the random variable would be close to one sample compared to the other sample.

<span class="mw-page-title-main">Quantum harmonic oscillator</span> Important, well-understood quantum mechanical model

The quantum harmonic oscillator is the quantum-mechanical analog of the classical harmonic oscillator. Because an arbitrary smooth potential can usually be approximated as a harmonic potential at the vicinity of a stable equilibrium point, it is one of the most important model systems in quantum mechanics. Furthermore, it is one of the few quantum-mechanical systems for which an exact, analytical solution is known.

<span class="mw-page-title-main">Schrödinger equation</span> Description of a quantum-mechanical system

The Schrödinger equation is a linear partial differential equation that governs the wave function of a quantum-mechanical system. It is a key result in quantum mechanics, and its discovery was a significant landmark in the development of the subject. The equation is named after Erwin Schrödinger, who postulated the equation in 1925, and published it in 1926, forming the basis for the work that resulted in his Nobel Prize in Physics in 1933.

In mathematics, the Hermite polynomials are a classical orthogonal polynomial sequence.

<span class="mw-page-title-main">Path integral formulation</span> Formulation of quantum mechanics

The path integral formulation is a description in quantum mechanics that generalizes the action principle of classical mechanics. It replaces the classical notion of a single, unique classical trajectory for a system with a sum, or functional integral, over an infinity of quantum-mechanically possible trajectories to compute a quantum amplitude.

In the quantum mechanics description of a particle in spherical coordinates, a spherically symmetric potential, is a potential that depends only on the distance between the particle and a defined center point. One example of a spherical potential is the electron within a hydrogen atom. The electron's potential only depends on its distance from the proton in the atom's nucleus. This spherical potential can be derived from Coulomb's law.

<span class="mw-page-title-main">Wave packet</span> Short "burst" or "envelope" of restricted wave action that travels as a unit

In physics, a wave packet is a short "burst" or "envelope" of localized wave action that travels as a unit. A wave packet can be analyzed into, or can be synthesized from, an infinite set of component sinusoidal waves of different wavenumbers, with phases and amplitudes such that they interfere constructively only over a small region of space, and destructively elsewhere. Each component wave function, and hence the wave packet, are solutions of a wave equation. Depending on the wave equation, the wave packet's profile may remain constant or it may change (dispersion) while propagating.

<span class="mw-page-title-main">Equipartition theorem</span> Theorem in classical statistical mechanics

In classical statistical mechanics, the equipartition theorem relates the temperature of a system to its average energies. The equipartition theorem is also known as the law of equipartition, equipartition of energy, or simply equipartition. The original idea of equipartition was that, in thermal equilibrium, energy is shared equally among all of its various forms; for example, the average kinetic energy per degree of freedom in translational motion of a molecule should equal that in rotational motion.

Creation operators and annihilation operators are mathematical operators that have widespread applications in quantum mechanics, notably in the study of quantum harmonic oscillators and many-particle systems. An annihilation operator lowers the number of particles in a given state by one. A creation operator increases the number of particles in a given state by one, and it is the adjoint of the annihilation operator. In many subfields of physics and chemistry, the use of these operators instead of wavefunctions is known as second quantization. They were introduced by Paul Dirac.

In theoretical physics, supersymmetric quantum mechanics is an area of research where supersymmetry are applied to the simpler setting of plain quantum mechanics, rather than quantum field theory. Supersymmetric quantum mechanics has found applications outside of high-energy physics, such as providing new methods to solve quantum mechanical problems, providing useful extensions to the WKB approximation, and statistical mechanics.

The old quantum theory is a collection of results from the years 1900–1925 which predate modern quantum mechanics. The theory was never complete or self-consistent, but was rather a set of heuristic corrections to classical mechanics. The theory is now understood as the semi-classical approximation to modern quantum mechanics. The main and final accomplishments of the old quantum theory were the determination of the modern form of the periodic table by Edmund Stoner and the Pauli Exclusion Principle which were both premised on the Arnold Sommerfeld enhancements to the Bohr model of the atom.

The relativistic Breit–Wigner distribution is a continuous probability distribution with the following probability density function,

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.

The Mehler kernel is a complex-valued function found to be the propagator of the quantum harmonic oscillator.

This is a glossary for the terminology often encountered in undergraduate quantum mechanics courses.

The phase-space formulation of quantum mechanics places the position and momentum variables on equal footing in phase space. In contrast, the Schrödinger picture uses the position or momentum representations. The two key features of the phase-space formulation are that the quantum state is described by a quasiprobability distribution and operator multiplication is replaced by a star product.

<span class="mw-page-title-main">Relativistic Lagrangian mechanics</span> Mathematical formulation of special and general relativity

In theoretical physics, relativistic Lagrangian mechanics is Lagrangian mechanics applied in the context of special relativity and general relativity.

References

  1. Griffiths, David J.; Schroeter, Darrel F. (2018). Introduction to Quantum Mechanics (3rd ed.). Cambridge University Press. pp. 12–13, 20, 53. ISBN   978-0-13-191175-8.
  2. 1 2 Robinett, R. W. (1995). "Quantum and classical probability distributions for position and momentum" . American Journal of Physics. 63 (9): 823–832. doi:10.1119/1.17807.
  3. Liboff, Richard L. (1980). Introductory Quantum Mechanics. Addison-Wesley Publishing Company, Inc. pp. 91, 194. ISBN   0-201-12221-9.