Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Stochastic Methods in Quantum Mechanics
Stochastic Methods in Quantum Mechanics
Stochastic Methods in Quantum Mechanics
Ebook378 pages4 hours

Stochastic Methods in Quantum Mechanics

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Practical developments in such fields as optical coherence, communication engineering, and laser technology have developed from the applications of stochastic methods. This introductory survey offers a broad view of some of the most useful stochastic methods and techniques in quantum physics, functional analysis, probability theory, communications, and electrical engineering. Starting with a history of quantum mechanics, it examines both the quantum logic approach and the operational approach, with explorations of random fields and quantum field theory.
The text assumes a basic knowledge of functional analysis; although some experience with probability theory and quantum mechanics is helpful, necessary ideas and results from these two disciplines are developed as needed. A selection of exercises follows each chapter, and proofs to most of the theorems are included. A comprehensive bibliography allows researchers and students to continue in the direction of their individual interests.
LanguageEnglish
Release dateMay 5, 2014
ISBN9780486149189
Stochastic Methods in Quantum Mechanics

Related to Stochastic Methods in Quantum Mechanics

Related ebooks

Physics For You

View More

Related articles

Reviews for Stochastic Methods in Quantum Mechanics

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Stochastic Methods in Quantum Mechanics - Stanley P. Gudder

    Mechanics

    1

    The History of Quantum Mechanics

    This chapter has a twofold purpose. It gives the reader a historical perspective, and introduces some of the important concepts of quantum mechanics. As the reader will notice, probability theory did not enter into quantum mechanics at the outset. The pioneers of the field had no reason or intention to make quantum mechanics a probabilistic subject. The stochastic nature of quantum mechanics was reluctantly accepted later when it proved to be an intrinsic, inescapable part of the field.

    1.1. Early History

    The idea of a quantum of energy was first introduced by Max Planck on December 14, 1900 at a meeting of the German Physical Society. He invented this concept to explain an empirical formula for the energy per unit volume E(v) of radiation of frequency v emitted by a body heated to a temperature T.

    In the late 1800s, Rayleigh and Jeans derived a formula for E(v) using classical mechanics. Suppose heat radiation is emitted through a small door of a cubical furnace of length L which is heated to a temperature T. The energy of the radiation e(v) at frequency v is found by determining the number of waves n(v) with frequency v and multiplying n(v) by the energy associated with these waves. According to a time-honored principle of physics, the equipartition principle, each wave should carry KT units of energy on the average, where K is Boltzmann’s constant. Hence e(v) = n(v)KT. To find n(v), the waves are represented mathematically by plane waves exp[2πi(k·r vt)] where k is a vector of length k = v/c (c is the speed of light) in the direction of wave propagation with components (n1/L, n2/L, n3/L), nl, n2, n3 being integers. The integers appear since we must have a whole number of wavelengths in the furnace cavity. We now find the number of waves n(v) Δv with frequency between v and v + Δv. Draw a three-dimensional lattice with dots at each point with integer coordinates and draw a spherical shell whose inner and outer radii are k and k + Δk, respectively. Then the number of waves whose wavenumber is between k and k + Δk is the number of dots in this shell. This number is approximately the volume of the shell divided by the volume surrounding an individual dot and thus equals

    Hence

    and

    Formula (1.2), called the Rayleigh–Jeans law, agrees with experiment for low frequencies v. However, for high frequencies, (1.2) not only deviates substantially from experiment, it gives a contradiction called the ultraviolet catastrophy. According to Eq. (1.2), as the frequency increases so does the corresponding energy. The ultraviolet catastrophy concludes that the total energy per unit volume due to all frequencies is infinite.

    The ultraviolet catastrophy was one of the first instances in which classical mechanics was found to be inadequate. Planck’s energy quantum, invented to derive the correct formula for E(v) and thus avoid this catastrophy, ushered in a new era, the era of quantum mechanics. Planck’s idea was so unusual and seemed so grotesque at the time, that he himself could hardly believe it. Planck suggested that instead of allowing the energy of radiation waves to have arbitrary values, we make the following postulate.

    The energy of electromagnetic waves can exist only in the form of discrete packages, or quanta, the energy content of each package being directly proportional to its frequency.

    Using this postulate, Planck derived his radiation law. A wave with frequency v can have only a countable number of energy values e(n) = nhv, n = 0, 1, 2, …, where h is a constant. According to classical statistical mechanics, these energy values are distributed according to the Gibbs probability pn = c0 exp[– e(n)/KT], where c0 is a normalization constant. Since ∑ pn = 1 we obtain

    The average energy ev = ∑ e(n)pn of a wave with frequency v becomes

    Using our previous formula for n(v) the energy per unit volume E(v) = n(v)ev/L³ becomes

    Planck’s radiation law (1.5) agrees with the Rayleigh–Jeans law (1.2) at low frequencies. Moreover, (1.5) agrees remarkably with experiment for all frequencies if we let h = 6.62 × 10–27 erg sec.

    Planck’s basic postulate was also substantiated later by other phenomena such as Einstein’s photoelectric effect and the Compton effect. However, one of the most important applications appeared in Bohr’s quantum orbits. In 1911 Ernest Rutherford, using experiments involving the scattering of α particles through gold foil, discovered the model for the atom which is essentially accepted today. Rutherford concluded from his experiments that an atom must consist of a very small positively charged nucleus surrounded at a relatively large distance by enough negatively charged electrons to balance the nuclear charge. These findings again contradicted classical theory. The electrons could not be stationary since then they would be attracted to the nucleus, causing the atom to collapse, so the electrons must be orbiting the nucleus. But by classical electrodynamics, a charged particle which is accelerating must radiate energy, and one can calculate, using classical theory, that the electrons should radiate away all of their kinetic energy within 10–8 sec. Having lost all their kinetic energy, atomic electrons must fall into the nucleus and the atom ceases to exist!

    Another phenomenon concerning atoms which classical physics could not explain was the light spectrum emitted by heated elements. If an element is heated to a high temperature it will emit light consisting of a sequence of sharp spectral lines which show up on a photographic plate after the light is refracted by a prism. In the 1800s spectroscopists had amassed huge quantities of data concerning the frequencies of these spectral lines. They found, for example, that the spectral lines of light emitted by heated hydrogen had frequencies vm, n labeled by two positive integers n < m and given by

    where R = 3.289 × 10¹⁵ sec–1.

    In 1913, Niels Bohr, using Planck’s postulate, gave a theoretical explanation not only for why the atom does not collapse but for the formula (1.6) . Bohr reasoned that the light emitted by an excited hydrogen atom was due to a loss of energy by its single orbital electron. It follows that if no light is emitted then the electron’s energy remains constant. Moreover, since the energy carried by the light radiation must, according to Planck’s postulate, have only discrete values, the energy of the orbital electron must also have discrete energy values E1, E2, …. When the electron jumps from an energy level Em to a lower energy level En, the energy of the emitted light is Em En. Again, by Planck’s postulate, this energy must be proportional to the frequency of the emitted light. Hence, the frequency may be labeled by two positive integers n < m and is given by hvm, n = Em En or

    Guided by the above reasoning, Bohr announced the following postulates.

    1. The hydrogen electron moves in a circular orbit and the electron’s energy remains constant so long as it is in the same orbit.

    2. Only a discrete number of orbits are allowed. In the allowed orbits, the electron must have angular momentum equal to a multiple of h.

    3. When an electron shifts from one possible orbit to another, the energy difference Em En is transformed into radiation of frequency (Em En)/h.

    We now show that the spectroscopic data in Eq. (1.6) follows from Bohr’s three postulates. Suppose the electron is in the nth possible orbit. This orbit must be circular by Postulate 1. Let rn be the radius of the orbit and let μ, and e be the known mass and charge, respectively, of the electron. If νn /rn. Since the electron experiences a Coulomb attractive force of magnitude e. Hence vn = e(μrn)–l/2. Since the angular momentum has magnitude μνnrn, Postulate 2 implies that μνnrn = n . Solving these last two equations for rn we obtain rn = n²/μecm, which agrees with experiment.) Since the total energy En is the sum of the kinetic energy Kn and the potential energy Vn we have

    Hence, by Postulate 3

    From the known values of e, μ, and h one finds that 2π²eμh–3 = 3.289 × 10¹⁵ sec–1 and so Eq. (1.9) reduces to Eq. (1.6).

    Electromagnetic radiation is naturally associated with a wave-type motion. However, in certain phenomena such as the photoelectric effect and the Compton effect, radiation acts like a stream of particles or photons. In 1925, de Broglie postulated his wave–particle duality principle, which asserts that particles also play this dual role. In particular, electrons should have wavelike properties. He asserted that the hydrogen orbital electron in its nth quantum orbit should have associated with it a wave which fits in its orbit. Thus the wavelength of the wave must be λn = 2πrn/n. Substituting νn for rn according to our previous formula gives λn = h/μνn. De Broglie then concluded that an electron should have associated with it a wave of wavelength h/μν. This was substantiated experimentally in 1927 when Davisson and Germer showed that electron beams gave diffraction patterns similar to light waves and that these patterns correspond to waves of wavelength h/μν.

    In 1926, Schrodinger reasoned that since electromagnetic waves in a vacuum satisfy a wave equation c–2(∂² E/∂t²) = ∇²E [∇² is the Laplacian (∂/∂x)² + (∂/∂y)² + (∂/∂z)² and E is the electric field] then by wave–particle duality, electron waves (or de Broglie waves) should also satisfy a wave equation. He asserted that the equation which governs de Broglie waves in vacuum should be

    For the hydrogen atom an additional Coulomb potential term is added to give the equation

    where r = (x² + y² + z²)¹/².

    This equation was later generalized by Dirac to include relativistic effects. Moreover, Schrodinger postulated that the eigenvalues E of the eigenvalue equation

    with suitable boundary conditions give the allowed values for the electron’s energy. These eigenvalues turn out to be the values given in Eq. (1.8) and thus Schrödinger’s postulate leads to the same result as Bohr’s postulates.

    In 1927 Heisenberg announced his famous uncertainty principle. This principle states that the position and momentum of a physical object cannot be measured with arbitrary precision at the same time. More precisely, if Δp and Δq are the errors made in a momentum and coordinate measurement at the same time, then ΔpΔq h. This principle was substantiated by experiment. As a crude example, suppose an electron moves along the x axis and we want to measure its position and momentum. To find the electron’s position we must see it and hence bounce a light ray off of it. If the light ray has wavelength λ, then we cannot locate the electron more precisely than λ, so Δx ≥ λ. According to Planck’s postulate the light ray must carry an energy of at least hv, where v is the light ray’s frequency. Since the speed of the light ray is c, the light ray has an associated momentum of at least hv/c. Since λv = c, this becomes h/λ. Hence the light ray will impart a momentum h/λ to the electron, so Δp h/λ and Δx Δp h. This gives an example of interfering measurements. Since h is very small, this phenomenon is not noticeable for objects that are roughly human size or larger.

    1.2. Enter Probability

    The situation in 1926–1927 was confusing. Although quantum mechanics could explain phenomena that classical mechanics could not, the quantum mechanics of that time consisted of a collection of vaguely related assertions. The assertions were stated in terms of certain ad hoc postulates such as Planck’s energy quanta, Bohr’s quantization rules, de Broglie- Schrodinger waves, and Heisenberg’s uncertainty principle. The justification was that these postulates led to results which agreed with experiment. However, they applied only to specific phenomena, could not be easily generalized to include more complicated situations, and were not held together by a systematic unified theory. Moreover, there were some puzzling unanswered questions that needed resolving. What is the significance of the function ψ in Eqs. (1.10), (1.11), and (1.12)? Why do the eigenvalues of Eq. (1.12) give the allowed energy values of a hydrogen electron?

    It turned out that the unifying, underlying theory was a quantum probability theory. This was first realized by Born during the period 1926–1929. It is interesting that during this same period Kolmogorov developed the foundations of classical probability theory [160]. The relationships between these two types of probability theory were not systematically exploited until much later [172,173]. The Heisenberg uncertainty principle indicated to Born that quantum mechanics was a stochastic theory. In classical mechanics one can, in principle, simultaneously measure the position and momentum of a particle to arbitrary precision. The fact that one cannot do this for quantum mechanical particles indicates that these quantities have an intrinsic dispersion due to some kind of statistical fluctuations. When the position and momentum are measured one is really determining an average of these quantities and the uncertainty principle shows that there is a lower bound to the product of the dispersion of these quantities from their average values. Bom proposed that the de Broglie waves were not waves in a physical sense, but were probability waves in the sense that the de Broglie–Schrödinger wave function ψ(t, x, y, z) is a complex-valued function which determines the probability that the electron is at the point (x, y, z) at time tat time t is

    assuming that

    Born’s ideas were developed into a systematic theory by Dirac and von Neumann in the early 1930s. Although their theories were similar and Dirac’s approach was succinct and elegant, we shall follow von Neumann since his methods are more rigorous mathematically. Von Neumann recognized that the two important concepts in quantum mechanics are states and observables. The states correspond to a theoretically complete description of the system and the observables correspond to the measurable quantities such as position, momentum, and energy. He decided that the wave function ψ(t, x, y, z) describes the state of the system at time t. Equation (1.14) shows that for fixed t, ψ is a unit vector in the complex Hilbert space L³). Since the Schrödinger equation (1.12) is an operator eigenvalue equation = and the eigenvalues E are energy values, von Neumann and others reasoned that the operator H corresponds to the energy observable. On a suitable domain, H is self-adjoint, which is physically reasonable since the energy must have real values. Moreover, just as position in a certain direction is conjugate to momentum in that direction, energy is conjugate to time. Hence the energy operator H can be used to describe the time evolution of the system. This is further substantiated by the time-dependent Schrödinger equation (1.11) which can be written i (∂ψ/∂t) = . Putting these ideas together, von Neumann proposed the following axioms.

    .

    .

    A3.The probability that an observable A when the system is in the state ψ is 〈PA(Δ)ψ, ψ〉, where PA(·) is the resolution of the identity for A.

    A4.If the state at time t = 0 is ψ, then the state at time t is ψt = exp(– itH)ψ, where H is the energy observable.

    To be precise, von Neumann did not say that the states are unit vectors and the observables are self-adjoint operators, but that the states are described by unit vectors and the observables are described by self-adjoint operators. Also, it is only the pure states that are unit vectors; there are other types of states called mixed states which we shall discuss later.

    Axioms Al, A2, and A4 are descriptive. They tell how the physical concepts can be described by mathematical constructs and how the evolution of the system can be described mathematically. Axiom A3 is extremely important since it gives the contact between the mathematical structure and reality. It is this axiom which gives the distribution of values for an observable. This distribution is precisely what is measured in the laboratory, thus enabling one to test the theory and to make predictions about the behavior of the physical system. Since A3 is stochastic in nature we obtain a quantum probability theory. Not only are the above postulates elegant and general, but as we shall see, they contain the ideas of Planck, Bohr, de Broglie, Schrödinger, Heisenberg, Bom, and others as special cases.

    For generality, A2 does not prescribe which self-adjoint operators represent which observables. This must be decided upon according to the specific situation being considered. In the case of the hydrogen electron, Eq. (1.12) indicates that the energy observable is represented by the operator

    In classical mechanics the energy Hc has the form Hc = p²/2μ + V(x, y, zis the kinetic energy and V(x, y, z) is the potential energy. Using this analogy one is tempted to postulate that the x momentum observable is represented by the operator –i (∂/∂x) and the x coordinate observable is represented by the operator f(x, y, zxf(x, y, z). The y, z momenta and coordinate operators are defined analogously. This correspondence does indeed give the desired results.

    Differentiating A4 with respect to t gives

    Equation (1.16) is the general Schrödinger equation and it reduces to Eq. (1.11) for the H defined in Eq. (1.15). The resolution of the identity of the x , where χΔ is the indicator (or characteristic) function of Δ. Applying A3, the probability that the x coordinate of the electron is in Δ when the system is in state ψ becomes

    If we form similar expressions for the y, z coordinates, we obtain Bom’s equation (1.13). Now suppose we want to find the allowed energy values for the hydrogen electron. If H has the value E with certainty, then A3 and Schwarz’s inequality imply that

    It follows that PH({E})ψ = ψ and hence

    This is Schrödinger’s eigenvalue equation (1.12), which shows that the allowed energy values are the eigenvalues of H.

    Applying A3 and the spectral theorem, we can find the expectation E[A] of the observable A (if it exists) in the state ψ:

    The variance of A in the state ψ is defined by Var(A) = E[(A E[A])²]. The standard deviation ΔA =[Var(A)]¹/² gives a measure of the dispersion of the values of A about its expectation. Suppose A, B are self-adjoint operators and suppose ψ, , are in the domains of A and B.

    If 〈Aψ, ψ〉 = 〈, ψ〉 = 0 we have

    If 〈Aψ, ψ〉, 〈, ψ〉 ≠ 0 replace A by A – 〈, ψ〉 and B by B – 〈, ψ〉 to again obtain

    Now let A = px = – i (∂/∂x) and B = qx = multiplication by x be the x momentum and x coordinate observables, respectively. The operators px, qx satisfy the Heisenberg commutation relation

    (To make (1.19) rigorous one must specify the domains of the operators. We shall do this in a more satisfactory way later using the so-called Weyl form of the relation.) Substituting (1.19) into (1.18) gives

    This shows that (1.18) is a general form of the Heisenberg uncertainty relations.

    1.3. Some Comparisons

    Von Neumann’s axioms have stood the test of time and are essentially unchanged today. They form the basis of nonrelativistic quantum mechanics with a finite number of degrees of freedom used by most contemporary physicists. Even in relativistic quantum field theory with its infinite number of degrees of freedom much of von Neumann’s theory remains intact.

    Axiom A3 establishes quantum mechanics as a stochastic theory, but one which is quite different from the classical Kolmogorov theory. In order to compare the two, let us briefly review classical probability theory. In this theory, a basic role is played by a triple (Ω, ∑, μ), where (Ω, ∑) is a measurable space and μ is a nonnegative measure on the σ-algebra ∑ satisfying μ(Ω) = l. We call (Ω, ∑, μ) a probability space and μ a probability measure. The elements of Ω correspond to the possible outcomes of a random experiment, the sets in ∑ correspond to random events, and the measure μ(A) for A ∈ ∑ gives the probability that the event A occurs.

    A random variable is a measurable function f. For B and random variable f, f–1(B) is the event that f has a value in B and μ[f–1(B)] is the probability of that event. The probability measure μf defined by μf(B) = μ[f–1(B)] is called the distribution of f. The expectation of f (if it exists) is E[f] = ∫ f dμ and it is easily seen that

    Moreover, if uis a Borel function and f is a random variable, then u(f) = u f is a random variable and

    If f, g are random variables, it is important to know the probability of the simultaneous occurrence of events such as f–l(A) ∩ g–l(B), A, B . The joint distribution of f, g is defined as the probability measure μf, g satisfying

    for all A, B . It is easily shown that μf, g always exists and satisfies the consistency conditions

    . Thus μf, g determines the distributions μf and μg but one can give examples which show that μf, μg do not determine μf, g. Two random variables f and g are independent if

    for all A, B . Using (1.22) we see that f and g are independent if and only if μf, g = μf × μg. Intuitively, f and g are independent if and only if the probability that an event f–1(A) occurs is unchanged if it is known that an event g–l(B) occurs for all A, B . Notice that in the case of independent random variables μf and μg do determine μf, g.

    There are certain elementary comparisons that one can make between classical probability theory and quantum probability theory. Axiom A3 tells us that the probability that the observable A has a value in the Borel set B is 〈 PA(B)ψ, ψ 〉. Thus observables correspond to random variables and the projection operators PA(B) correspond to events. In fact, we interpret PA(B) as the event that the observable A has a value in the set Bas the set of quantum mechanical events. Since there is a natural one-to-one correspondence between orthogonal projections and

    Enjoying the preview?
    Page 1 of 1