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

Only $11.99/month after trial. Cancel anytime.

Quantum Theory
Quantum Theory
Quantum Theory
Ebook1,097 pages21 hours

Quantum Theory

Rating: 4 out of 5 stars

4/5

()

Read preview

About this ebook

This superb text by David Bohm, formerly Princeton University and Emeritus Professor of Theoretical Physics at Birkbeck College, University of London, provides a formulation of the quantum theory in terms of qualitative and imaginative concepts that have evolved outside and beyond classical theory. Although it presents the main ideas of quantum theory essentially in nonmathematical terms, it follows these with a broad range of specific applications that are worked out in considerable mathematical detail.
Addressed primarily to advanced undergraduate students, the text begins with a study of the physical formulation of the quantum theory, from its origin and early development through an analysis of wave vs. particle properties of matter. In Part II, Professor Bohm addresses the mathematical formulation of the quantum theory, examining wave functions, operators, Schrödinger's equation, fluctuations, correlations, and eigenfunctions.
Part III takes up applications to simple systems and further extensions of quantum theory formulation, including matrix formulation and spin and angular momentum. Parts IV and V explore the methods of approximate solution of Schrödinger's equation and the theory of scattering. In Part VI, the process of measurement is examined along with the relationship between quantum and classical concepts.
Throughout the text, Professor Bohm places strong emphasis on showing how the quantum theory can be developed in a natural way, starting from the previously existing classical theory and going step by step through the experimental facts and theoretical lines of reasoning which led to replacement of the classical theory by the quantum theory.

LanguageEnglish
Release dateApr 25, 2012
ISBN9780486134888
Quantum Theory

Related to Quantum Theory

Titles in the series (100)

View More

Related ebooks

Physics For You

View More

Related articles

Reviews for Quantum Theory

Rating: 3.9666699999999997 out of 5 stars
4/5

15 ratings1 review

What did you think?

Tap to rate

Review must be at least 10 words

  • Rating: 2 out of 5 stars
    2/5
    I didn't find this book very clear, and it meanders around somewhat without seeming to have a point. I suppose it was meant as a survey of the subject, but it doesn't present its logic well and some of the conclusions get muddled as a result.

Book preview

Quantum Theory - David Bohm

THEORY

PART I

PHYSICAL FORMULATION OF THE QUANTUM THEORY

MODERN QUANTUM THEORY is unusual in two respects. First, it embodies a set of physical ideas that differ completely with much of our everyday experience, and also with most experiments in physics on a macroscopic scale. Second, the mathematical apparatus needed to apply this theory to even the simplest examples is much less familiar than that required in corresponding problems of classical physics. As a result, there has been a tendency to present the quantum theory as being inseparable from the mathematical problems that arise in its applications. This approach might be likened to introducing Newton’s laws of motion to a student of elementary physics, as problems in the theory of differential equations. In this book, special emphasis is placed on developing the guiding physical principles that are useful not only when it is necessary to apply our ideas to a new problem, but also when we wish to forsee the general properties of the mathematical solutions without carrying out extensive calculations. The development of the special mathematical techniques that are necessary for obtaining quantitative results in complex problems should take place, for the most part, either in a mathematics course or in a special course concerned with the mathematics of quantum theory. It seems impossible, however, to develop quantum concepts extensively without Fourier analysis. It is, therefore, presupposed that the reader is moderately familiar with Fourier analysis.

In the first part of this book, an unusual amount of attention is given to the steps by which the quantum theory may be developed, starting with classical theory and with specific experiments that led to the replacement of classical theory by the quantum theory. The experiments are presented not in historical order, but rather in what may be called a logical order. An historical order would contain many confusing elements that would hide the inherent unity that the quantum theory possesses. In this book, the experimental and theoretical developments are presented in such a way as to emphasize this unity and to show that each new step is either based directly on experiment or else follows logically from the previous steps. In this manner, the quantum theory can be made to seem less like a strange and somewhat arbitrary prescription, justified only by the fact that the results of its abstruse mathematical calculations happen to agree with experiment.

As an integral part of our plan for developing the theory on a basis that is not too abstract for a beginner, a complete account of the relation between quantum theory and the previously existing classical theory is given. Wherever possible the meaning of the quantum theory is illustrated in simple physical terms. Moreover, the final chapter of Part I points out broad regions of everyday experience in which we continually use ways of thinking that are closer to quantum-theoretical than to classical concepts. In this chapter, we also discuss in detail some of the philosophical implications of the quantum theory, and show that these lead to a striking modification in our general view of the world, as compared with that suggested by classical theory.

The reader will notice that most of the problems are interspersed throughout the text. These problems should be read as part of the text, because the results obtained from them are often used directly in the development of ideas. It is usually possible to understand the significance of the results without solving the problems, but the reader is strongly urged to try to solve them. The main advantage of the interspersed problems is that they make the reader think more specifically about the subject previously discussed, thus facilitating his understanding of the subject.

Supplementary References

The following list of supplementary texts will prove very helpful to the reader and will be referred to throughout various parts of this book:

Bohr, N., Atomic Theory and the Description of Nature. London: Cambridge University Press, 1934.

Born, M., Atomic Physics. Glasgow: Blackie & Son, Ltd., 1945.

Born, M., Mechanics of the Atom. London: George Bell & Sons, Ltd., 1927.

Dirac, P. A. M., The Principles of Quantum Mechanics. Oxford : Clarendon Press, 1947.

Heisenberg, W., The Physical Principles of the Quantum Theory. Chicago: University of Chicago Press, 1930.

Kramers, H. A., Die Grundlagen der Quantentheorie. Leipzig: Akademische Verlagsgesellschaft, 1938.

Mott, N. F., An Outline of Wave Mechanics. London: Cambridge University Press, 1934.

Mott, N. F., and I. N. Sneddon, Wave Mechanics and Its Applications. Oxford: Clarendon Press, 1948.

Pauli, W., Die Allgemeinen Prinzipen der Wellenmechanik. Ann Arbor, Mich.: Edwards Bros., Inc., 1946. Reprinted from Handbuch der Physik, 2. Aufl., Band 24. 1. Teil.

Pauling, L., and E. Wilson, Introduction to Quantum Mechanics. New York: McGraw-Hill Book Company, Inc., 1935.

Richtmeyer, F. K., and E. H. Kennard, Introduction to Modern Physics. New York: McGraw-Hill Book Company, Inc., 1933.

Rojansky, V., Introductory Quantum Mechanics. New York: Prentice-Hall, Inc., 1938.

Ruark, A. E., and H. C. Urey, Atoms, Molecules, and Quanta. New York: McGraw-Hill Book Company, Inc., 1930.

Schiff, L., Quantum Mechanics. New York: McGraw-Hill Book Company, Inc., 1949.

CHAPTER 1

The Origin of the Quantum Theory

The Rayleigh-Jeans Law

1. Blackbody Radiation in Equilibrium. Historically, the quantum theory began with the attempt to account for the equilibrium distribution of electromagnetic radiation in a hollow cavity. We shall, therefore, begin with a brief description of the characteristics of this distribution of radiation. The radiant energy originates in the walls of the cavity, which continually emit waves of every possible frequency and direction, at a rate which increases very rapidly with the temperature. The amount of radiant energy in the cavity does not, however, continue to increase indefinitely with time, because the process of emission is opposed by the process of absorption that takes place at a rate proportional to the intensity of radiation already present in the cavity. In the state of thermodynamic equilibrium, the amount of energy U(v)dv, in the frequency range between v and v + dv, will be determined by the condition that the rate at which the walls emit this frequency shall be balanced by the rate at which they absorb this frequency. It has been demonstrated both experimentally and theoretically,* that after equilibrium has been reached, U(v) depends only on the temperature of the walls, and not on the material of which the walls are made nor on their structure.

, where c is the velocity of light.

Measurements disclose that, at a particular temperature, the function U(v) follows a curve resembling the solid curve of Fig. 1. At low frequencies the energy is proportional to v², while at high frequencies it drops off exponentially. As the temperature is raised, the maximum is shifted in the direction of higher frequencies; this accounts for the change in the color of the radiation emitted by a body as it gets hotter.

By thermodynamic arguments† Wien showed that the distribution must be of the form U(v) = v³f(v/T). The function f, however, cannot be determined from thermodynamics alone. Wien obtained a fairly good, but not perfect, fit to the empirical curve with the formula

Here κ is Boltzmann’s constant, and h is an experimentally determined constant (which later turned out to be the famous quantum of action).*

FIG. 1

Classical electrodynamics, on the other hand, leads to a perfectly definite and quite incorrect form for U(v). This theoretical distribution, which will be derived in subsequent sections, is given by

Reference to Fig. 1 shows that the Rayleigh-Jeans law is in agreement with experiment at low frequencies, but gives too much radiation for high frequencies. In fact, if we attempt to integrate over all frequencies to find the total energy, the result diverges, and we are led to the absurd conclusion that the cavity contains an infinite amount of energy. Experimentally, the correct curve begins to deviate appreciably from the Rayleigh-Jeans law where hv becomes of the order of κT. Hence, we must try to develop a theory that leads to the classical results for hv < κT, but which deviates from classical theory at higher frequencies.

Before we proceed to discuss the way in which the classical theory must be modified, however, we shall find it instructive to examine in some detail the derivation of the Rayleigh-Jeans law. In the course of this deviation we shall not only gain insight into the ways in which classical physics fails, but we shall also be led to introduce certain classical physical concepts that are very helpful in the understanding of the quantum theory. In addition, the introduction of Fourier analysis to deal with this classical problem will also constitute some preparation for its later use in the problems of quantum theory.

2. Electromagnetic Energy. According to classical electrodynamics, empty space containing electromagnetic radiation possesses energy. In fact, this radiant energy is responsible for the ability of a hollow cavity to absorb heat. In terms of the electric field, ε(x, y, z, t, the energy can be shown to be*

where signifies integration over all the space available to the fields.

Our problem, then, is to determine the way in which this energy is distributed among the various frequencies present in the cavity when the walls are at a given temperature. The first step will be to use Fourier analysis for the fields and to express the energy as a sum of contributions from each frequency. In so doing, we shall see that the radiation field behaves, in every respect, like a collection of simple harmonic oscillators, the so-called radiation oscillators. We shall then apply statistical mechanics to these oscillators and determine the mean energy of each oscillator when it is in equilibrium with the walls at the temperature T. Finally, we shall determine the number of oscillators in a given frequency range and, by multiplying this number by the mean energy of an oscillator, we shall obtain the equilibrium energy corresponding to this frequency, i.e., the Rayleigh-Jeans law.

3. Electromagnetic Potentials. We begin with a brief review of electrodynamics. The partial differential equations of the electromagnetic field, according to Maxwell, are given by

where j is the current density and ρ is the charge density. We can show from (4) and (5) that the most general electric and magnetic field can be expressed in terms of the vector and scalar potentials, a and ϕ, in the following way:

and

When ε are expressed in this form, (4) and (5) are satisfied identically, and the equations for a and ϕ are then obtained by the substitution of relations (8) and (9) into (6) and (7).

Now, eqs. (8) and (9) do not define the potentials uniquely in terms of the fields. If, for example, we add an arbitrary vector, –∇ψ, to the vector potential, the magnetic field is not changed because ∇ × ∇ψ to the scalar potential, the electric field is also unchanged. Thus, we find that the electric and magnetic fields remain invariant under the following transformation of the potentials:*

The above is called a gauge transformation.

We can utilize the invariance of the fields to a gauge transformation for the purpose of simplifying the expressions for ε . A common choice is to make div a = 0. To show that this is always possible, suppose that we start with an arbitrary set of potentials, a(x, y, z, t) and ψ(x, y, z, t). We then make the gauge transformation of eq. (10) to a new set of potentials, A′ and ψ′. In order to obtain div a′ = 0, we must choose ψ such that

But the above is just Poisson’s equation defining ψ in terms of the specified function, div a. Its solution can always be obtained and is, in fact, equal to

Thus, we prove that a gauge transformation that yields div a′ = 0 can always be carried out.

We now show that in empty space the choice div a = 0 also leads to ϕ = 0 and, therefore, to a considerable simplification in the representation of the electric field. To do this, we substitute eq. (9) into (7), setting ρ = 0 since, by hypothesis, there are no charges in empty space. The result is

But since div a = 0, we obtain

This is, however, simply Laplace’s equation. It is well known that the only solution of this equation that is regular over all of space is ϕ = 0. (All other solutions imply the existence of charge at some points in space and, therefore, a failure of Laplace’s equation at these points.) We should note, however, that the condition ψ = 0 follows only in empty space because, in the presence of charge, eq. (7) leads to ∇²ψ = –4πρ, which is Poisson’s equation. This equation has nonzero regular solutions, provided that ρ is not everywhere zero.

We conclude, then, that in empty space we obtain the following expressions for the fields:

subject to the condition that

Finally, we obtain the partial differential equation defining a in empty space by inserting (11), (12), and (13) into (6), provided that we also assume that j = 0, as is necessary in the absence of matter. We obtain

Equations (11), (12), (13), and (14), together with the boundary conditions, completely determine the electromagnetic fields in a cavity that contains no charges or currents.

4. Boundary Conditions. As pointed out in Sec. 1, it has been demonstrated both experimentally and theoretically* that the equilibrium distribution of energy density in a hollow cavity does not depend on the shape of the container or on the material in the walls. Hence, we are at liberty to choose the simplest possible boundary conditions consistent with equilibrium. We shall choose a set of boundary conditions that are somewhat artificial from an experimental point of view, but that greatly simplify the mathematical treatment.

Let us imagine a cube of side L with very thin walls of some material that is not an electrical conductor. We then imagine that this structure is repeated periodically through space in all directions, so that space is filled up with cubes of side L. Let us suppose, further, that the fields are the same at corresponding points of every cube.

We now assert that these boundary conditions will yield the same equilibrium radiation density as will any other boundary conditions at the walls.† To prove this, we need only ask why the equilibrium conditions are independent of the type of boundary. The answer is that, from the thermodynamic viewpoint, the wall merely serves to prevent the system from gaining or losing energy. Making the fields periodic must have, the same effect because each cube can neither gain energy from the other cubes nor lose it to them; if this were not so, the system would cease to be periodic. Thus, we have a boundary condition that serves the essential function of keeping the energy in any individual cube constant. Although artificial, it must give the right answer, and it will make the calculations easier by simplifying the Fourier analysis of the fields.

5. Fourier Analysis. Now, a(x, y, z, t) may be any conceivable solution of Maxwell’s equations, with the sole restriction, imposed by our boundary conditions, that it must be periodic in space with period L/n, where n is an integer.* It is a well-known mathematical theorem that an arbitrary periodic function,† f(x, y, z, t), can be represented by means of a Fourier series in the following manner:

where l, m, n are integers running from –∞ to ∞, including zero. Any choice of a’s and b’s leading to a convergent series defines a function, f(x, y, z, t), which is periodic in the sense that it takes on the same value each time x, y, or z changes by L. For a given function, f(x, y, z, t) it can be shown that the al,m,n(t) and the bl,m,n(t) are given by the following formulas:

These formulas illustrate the fact that only the sum of the a’s and the difference of the b’s are determined by the function f.

From the above, we conclude that f may be specified completely in terms of the quantities al,m,n + al,−m,−n and bl,m,n + bl,−m,−n, but we prefer to retain the specification in terms of the al,m,n and bl,m,n because of the simpler mathematical expressions to which they lead.

Equations (16) are obtained with the aid of the following orthogonality relations:*

unless

in which case it is L³/2, except when l = m = n = 0, in which case it is L³.

unless

in which case it is L³/2. [It is suggested that the reader prove (17a and b) as an exercise, and use the results to obtain (16).]

Fourier analysis, in the preceding form, enables us to represent an arbitrary function as a sum of standing plane waves of all possible wavelengths and amplitudes. The entire treatment is essentially the same as that used with waves in strings and organ pipes, except that it is three-dimensional.

Let us now expand the vector potential in a Fourier series. Because a is a vector, involving three components, each al,m,n and bl,m,n also has three components and, hence, must be represented as a vector:

We assume that a0,0,0 is zero in the above series.

We now introduce the propagation vector k, defined as follows:

By orienting our co-ordinate axes in such a way that the z axis is directed along the k vector, we obtain l = m = 0, and k = 2π/L. From the definition of k, it follows that k/2π is the number of waves in the distance L; hence the wavelength is λ = 2π/k, or

In this co-ordinate system a typical wave takes the form cos 2πnz/L. Thus, the vector k is in the direction in which the phase of the wave changes. Going back to arbitrary co-ordinate axes, we conclude that k is a vector in the direction of propagation of the wave. Its magnitude is 2π/λ, and it is allowed to take on only the values permitted by integral l, m, and n in eq. (18).

With this simplification of notation, we obtain

where the summation extends over all permissible values of k.

6. Polarization of Waves. Let us now apply the condition div a = 0 to (20). We have

It is a well-known theorem that if a Fourier series is identically zero, then all of the coefficients, ak and bk, must vanish.

Problem 1: Prove the above theorem, using the orthogonality relations (17).

From the above it follows that k · ak(t) = k · bk(t) = 0. Thus, ak(t) and bk(t) are perpendicular to k, as are also the electric and magnetic fields belonging to the kth wave. Since the vibrations are normal to the direction of propagation, the waves are transverse. The direction of the electric field is also called the direction of polarization.

To describe the orientation of ak let us return to the set of co-ordinate axes in which the z axis is in the direction of k. The vector ak can have only x and y components, and if we specify the values of these, we shall have specified both the magnitude and the direction of ak.

We designate the direction of the vector ak by the subscript μ, writing ak,μ, where μ is allowed to take on the values 1 and 2. For μ = 1, ak,μ is in the x direction; but for μ = 2, it is in the y direction. All possible ak vectors can then be represented as a sum of some ak,1 vector, and some other ak,2 vector. Hence, the most general vector potential, subject to the condition that div a = 0, is given by

Here the summation extends over all permissible k vectors and over the two possible values of μ.

It can be verified from (14) and (21) that the satisfies the following differential equation:

which shows that the ak,μ terms oscillate with simple harmonic motion and with angular frequency, ω = kc.

7. Evaluation of the Electromagnetic Energy. The first step in evaluating the electromagnetic energy is to express ε in terms of the Fourier series for a. These expressions are:

Problem 2: Derive the above expressions for ε .

Let us now evaluate the following over the cube of side L:

With the aid of eqs. (17) we see that all integrals vanish except when k = kunless μ = μ′. When μ μ′, the two vectors are, by definition, perpendicular to each other. Thus, the above expression reduces to

With a similar method, which involves somewhat more algebra, we obtain

.

Thus, the electromagnetic energy in the cavity is (with L³ = V)

8. Meaning of Preceding Result for Electromagnetic Energy. The following are the most important properties of eq. (23) :

(1) The energy is a sum of separate terms, one for each ak,μ, and one for each bk,μ. This means that different wavelengths and polarizations do not interact with each other, because the interaction of any two systems always requires that the energy of one should depend on the state of the other. Here we see that the energy in each wave of propagation vector k and polarization direction μ and ak,μ, and not to any of the other a’s or b’s. A similar result holds for each of the b’s.

(2) The energy associated with each ak,μ (or bk,μ) has the same mathematical form as that of a material harmonic oscillator. A harmonic oscillator of mass m, angular frequency ω, has energy

By analogy, we can write

The frequency is then f = ω/2ω = kc/2π = c/λ. We know, of course, that an electromagnetic wave of wavelength λ has just the above frequency.* This shows that our harmonic oscillator analogy gives the right description of the way in which the a’s oscillate.

. Here the momentum is

We can then introduce a Hamiltonian function

For the ak,μ we get

Similar terms may be introduced for the bk,μ.

The correct equations of motion are obtained from the Hamiltonian equations

which yield eq. (22), obtained originally by direct substitution into Maxwell’s equations.

and similarly for the b’s.

The ak,μ and bk,μ are, as we have seen, analogous to the co-ordinates of separate noninteracting harmonic oscillators. In a sense, the ak,μ and bk,μ may also be regarded as the co-ordinates of the radiation field. This is because, once they are given, the field is specified everywhere through eq. (20). There are an infinite number of these co-ordinates, because there are an infinite number of possible values of k. But the infinity is discrete, or countable, as distinguished from the continuous infinity of points on a line. The main advantage of the Fourier series is that it enables us to describe the fields over a continuous region of space by means of a discrete infinity of co-ordinates.

How many independent co-ordinates are there for each permissible value of k? First, there are two polarization directions; then we have, for each k and , an ak,μ and a bk,μ. Thus, it would seem, at first sight, that we need four independent co-ordinates for each value of k. But, from eq. (16), we see that it is necessary to specify only the combinations ak,μ + a−k,μ and bk,μ b−k,μ, so that the number of variables necessary is reduced by a factor of two. This means that for each k there are two independent co-ordinates.

9. Number of Oscillators. We must now find the number of oscillators with frequencies between v and (v + dv). Since v = kc/2π, the problem is equivalent to that of finding the number between k and k + dk.

Now, for any reasonable value of k, the number of waves fitting into a box is usually very large. For example, at moderate temperatures, most of the radiation is in the infrared, with wavelengths ∼10−4 cm. Hence, when k changes in such a way that one more wavelength fits into the box, only a very small fractional shift of k results. It is possible, therefore, to choose the interval dk so small that no important physical quantity (such as the mean energy) changes appreciably within it, yet so large that very many radiation oscillators are included. This means that the number of oscillators can be treated as virtually continuous, so that we can represent it in terms of a density function.

We must now find the number of oscillators in the volume dkx dky dkz. If we imagine a space in which the co-ordinates are l, m, and n, there will be one oscillator every time l, m, and n take on separate integral values. Hence, there is one oscillator per unit cube of l, m, n space, so that the density in this space is unity. To go to k space, we use eq. (18), obtaining

It is now convenient to adopt polar co-ordinates in k . Then the element of volume becomes k²dkdΩ; where dΩ is the element of solid angle. Since we are not interested in the direction of k, we integrate over dΩ, obtaining 4πk² dk for the element of volume, and

Writing v = kc/2π, we obtain

This gives the number of permissible values of k in the range between v and v + dv. As shown in the section discussing the significance of the a’s and b’s, there are two independent coordinates for each k, corresponding to the two directions of polarization. Thus, for the total number of oscillators between v and v + dv, we find

10. Equipartition of Energy. To calculate the mean energy possessed by each oscillator when it is in thermodynamic equilibrium with the walls, we shall apply classical statistical mechanics to these oscillators. Although this theory was derived for material oscillators alone, the derivation involved only the formal properties of the equations of motion. Any other systems acting formally like material oscillators must, therefore, have the same equilibrium distribution of energy. It is shown in classical statistical mechanics* that in any assembly of independent, noninteracting systems (such as our assembly of radiation oscillators), the probability that a co-ordinate lies between q and q + dq, and that the corresponding momentum lies between p and p + dp, is equal to

E denotes total energy, kinetic and potential ; and A denotes a normalizing factor, defined by the requirement that the total probability integrate out to unity or

For a perfect gas, E = p²/2m, and we obtain the familiar Maxwell-Boltzmann distribution of velocities

For the harmonic oscillator, we have

It is convenient to transform these equations to new variables defined by

This yields E = P² + Q².

The probability that the system lies between P and P + dP, and Q and Q + dQ, is

Let us transform to polar co-ordinates R, ϕ, in phase space, where

The element of area is now

Since we are not interested in the angle ϕ, we may integrate over it. We then obtain, for the normalized probability that the energy lies between E and E + dE :

The mean value of the energy Ē is obtained by integration of EW(E) over all energies. This means that we weight each energy according to its probability. We get

. Thus, we prove that the average energy of each oscillator is κT. This is an example of the theorem of equipartition of energy.*

Collecting the information obtained from (29) and (30), we get the Rayleigh-Jeans law:

Because this law disagrees with experiments, we conclude that the concepts of classical physics are in some way inadequate to describe the interaction of matter and radiation.

Planck’s Hypothesis

11. The Quantization of the Radiation Oscillators. In searching for a modification of the above treatment that would reduce the contribution of the high frequencies to the energy, Planck was led to make an assumption equivalent to the following: The energy of an oscillator of natural frequency v is restricted to integral multiples of a basic unit hv. This basic unit is not the same for all oscillators, since it is proportional to the frequency. The energy of an oscillator is, then, E = nhv, where n is any integer from 0 to ∞. With this assumption, Planck obtained an exact fit, within experimental error, to the observed distribution of radiation.

According to classical mechanics, there should be no restrictions whatever on the energy an oscillator may possess. Our experience with oscillators, such as radio waves, clock springs, and pendulums, seem to verify this prediction. How, then, can Planck’s hypothesis be consistent with all these well-known results? The answer is that h is a very small quantity, equal to about 6.6 × 10−27 erg-sec. Hence, even for microwaves having a frequency as high as 10¹⁰ cps, the basic unit of energy is only 6.6 × 10−17 erg which is not detectable except by use of the most sensitive apparatus now available. With clock springs and pendulums of period of the order of 1 sec, the basic unit of energy is obviously so small that in such relatively gross observations as can now be made, the allowed values of energy seem to be continuous. With light waves, however, v erg, a value that can be detected with sensitive instruments. Hence, as we go to higher frequencies, where the basic unit becomes larger, quantization of the energy levels is easier to observe.

To obtain Planck’s distribution of energy, we need to know what the probability is that the oscillator has an energy corresponding to its nth allowed value. Now, when n is very large, so that the discrete character of the energy becomes unimportant (as with radio waves, for example) we must obtain a result that is consistent with classical mechanics, which is known to be correct in this region. The simplest way of obtaining agreement is to choose a probability that is the same function of the energy as in the classical theory,* namely, eE/κT. For a given energy, En = nhv, the probability is then

To normalize this, we write*

The mean energy is

To evaluate this sum, we can write

We can apply this result, obtaining

Multiplying by δN, we find the Planck distribution

12. Discussion of Results. For small hv/κT the exponentials can be expanded, and retention of only the first terms yields Ē = κT, the classical result, in agreement with the fact that the Rayleigh-Jeans law is correct for small hv/κT. As hv/κT becomes large, then Ē hvehv/κT. This leads to the Wien law. In between, there is excellent agreement with experiment at all temperatures. Hence, despite the strangeness of Planck’s hypothesis, there is evidently something to it.

The decrease in mean energy of the high-frequency oscillators arises because of the great amount of energy required to bring them to the first excited state, which is a state of rare occurrence. As v is lowered or T raised, it becomes more likely that the oscillator will gain a quantum of energy. After the oscillator is excited to a high quantum number n, its behavior will be essentially classical, because the basic unit of energy is then much less than the mean available energy κT.

13. Material vs. Radiation Oscillators. Planck’s original idea was not to quantize the radiation oscillators as we have done previously. Instead, he assumed that the radiation was in equilibrium with material oscillators in the walls of the container, and that these material oscillators could give up or absorb radiant energy only in quanta with E = nhv. With this assumption, he obtained exactly the same distribution of radiant energy. The quantization of the radiation oscillators was a later idea that has, as we shall see, many far-reaching consequences. Moreover, the step of quantizing the radiation oscillators is almost imperative to explain the fact that the blackbody spectrum is independent of the materials of which the walls are composed.

The only alternative possibility is that all matter, and not only harmonic oscillators, can accept or emit radiation only in quanta of size, E = hv. But this means that all radiation ever emitted has energy restricted to E = hv; even if some were present with other energies it could not, by hypothesis, interact with matter and, hence, would be undetectable. This hypothesis is, then, equivalent to the statement that all radiation oscillators have their energies restricted to E = nhv.

14. Quantization of Material Oscillators. We can consider the specific heats of solids, to determine whether Planck’s quantum hypothesis applies to material oscillators. In a crystal, for example, each atom is in equilibrium when it lies in its proper lattice position and, if disturbed, it can oscillate about the equilibrium position with a motion that is approximately simple harmonic for small oscillations.

FIG. 2

To a first approximation, the oscillators can be regarded as independent. The frequency of the oscillation can be computed in terms of the mass of the atom and the elastic constants of the crystal (see Richtmeyer and Kennard). According to the classical equipartition theorem, each oscillator possesses energy κT and, therefore, makes a contribution to the specific heat of κ per atom. Experimentally, it is found that the specific heat approaches zero at absolute zero, and rises asymptotically to κ per atom at high temperatures, as shown in Fig. 2. Thus, the classical theory is certainly wrong at low temperatures.

Einstein proposed that this curve could be explained by assuming that the molecular oscillators are quantized with E = nhv. In contrast to the radiation oscillators, which can have all possible frequencies, the material oscillators have only one frequency, which is the characteristic frequency of the substance. Applying Planck’s result for a given frequency, eq. 32, we obtain

This formula clearly predicts a specific heat per molecule of κ at high temperatures, and approaching zero at very low temperatures as

It is in general agreement with experiment, except at very low temperatures (∼10°K).

The reason for the discrepancy at very low temperatures was explained by Debye.* The oscillations of each atom are actually not independent of the oscillations of the others, but are coupled to them because of the forces between molecules. The description in terms of independent oscillations is, therefore, not completely accurate.

To describe the coupled oscillations of the molecules, we may consider, for example, a one-dimensional string of particles. Suppose that each particle interacts only with its two nearest neighbors. It can then be shown† that waves are propagated through this system resembling those propagated through a chain, except that here the waves are both longitudinal and transverse, whereas the waves in a chain are transverse only. When the wavelength is large compared with the distance between particles, the propagation differs very little from that in a continuous string; but as the length of the waves approaches the mean distance between particles, the law of propagation changes. For wavelengths shorter than the mean distance between particles, propagation becomes impossible.

Problem 4: In the one-dimensional string of particles specified above, let the equilibrium distance between particles be a. Suppose the force on the nth particle to be

Here xn is the deviation of the nth particle from its equilibrium position.

Find solutions of the form xn = Aneiωt, and show that we can choose An = einα, where α is a suitable constant whose relationship to ω is obtained by solving the equations.

, where v = ω0a is the speed of sound in the system. Show also that there is a maximum possible frequency.

In three dimensions, a similar treatment can be given and, in this way, we can describe the propagation of sound waves through a crystal. As was done with the electromagnetic field, we can adopt the amplitudes of the possible sound waves as co-ordinates to describe the state of the system. Since these co-ordinates oscillate harmonically with the time, the energies of the associated oscillators must be quantized.‡ In computing the energy, however, we must take into account the fact that only a finite number of wavelengths is permissible, and also that the relation between frequency and wavelength becomes more complex as we approach wavelengths comparable with interatomic spacing. When all these factors have been taken into account, the quantum hypothesis leads to excellent general agreement with experimental specific heats at all temperatures. Thus, in addition to quanta of electromagnetic energy, we now have evidence for the existence of quanta of sound energy.

15. Summary. We may conclude that all systems which oscillate harmonically are quantized with E = nhv, whether these systems be material oscillators, sound waves, or electromagnetic waves. Since we assume that all systems can interact with each other, the quantization of any one type of harmonic oscillator requires a similar quantization of all other types. If experiments had not verified the existence of this unity, the quantum theory would have had to be abandoned, or at least fundamentally modified.

* Richtmeyer and Kennard. (See list of references on p. 2.)

† See Richtmeyer and Kennard for a derivation of this formula and also for a more complete account of blackbody radiation. The term blackbody arose because the radiation from a hole in such a cavity is identical with that coming from a perfectly black object.

* Wien did not actually introduce Planck’s constant, h, but instead the constant h/κ.

* Richtmeyer and Kennard, Chap. 2.

* ε are the only physically significant quantities connected with the electromagnetic field.

* The theoretical proof depends on the use of statistical mechanics. See, for example, R. C. Tolman, The Principles of Statistical Mechanics. Oxford, Clarendon Press, 1938.

† With these conditions, no walls are actually necessary, but the thermodynamic results are the same as for an arbitrary wall, including, for example, a perfect reflector or a perfect absorber.

* There will be, of course, the usual regularity conditions that prevent a from being infinite or discontinuous.

† The function must be piecewise continuous.

* For the origin of the term orthogonality see Chap. 16, Sec. 10; also Chap. 10, Sec. 24.

† This follows from the fact that the part of a . Such a field requires a charge distribution somewhere to produce it, i.e., at the boundaries, and since we assume that no such distribution is present, we set a0,0,0 = 0.

* See also eq. (22).

* R. C. Tolman, The Principles of Statistical Mechanics.

* See Richtmeyer and Kennard, p. 161.

* This choice involves an assumption that is justified in part by its success in accounting for the energy distribution in a blackbody. A systematic development of the theory of quantum statistics (see Tolman, The Principles of Statistical Mechanics) shows, however, that no other probability distribution would lead to thermodynamic equilibrium.

.

* See Richtmeyer and Kennard, p. 450.

† F. Seitz, The Modern Theory of Solids. New York: McGraw-Hill Book Company, Inc., 1946, pp. 121 and 125.

‡ See Secs. 10 and 13.

CHAPTER 2

Further Developments of the Early Quantum Theory

The New Concepts of the Quantum Theory

THUS FAR, quantum restrictions on allowed energies have arisen only in connection with harmonic oscillators. We shall see, however, that the results of many experiments, together with the systematic and logical development of the quantum hypothesis, lead to the conclusion that all matter is subject to quantum restrictions. This conclusion thus enables us to explain correctly a wide variety of experimental data for which the results of the classical theory are either wrong or ambiguous. As examples, we shall deal with the photoelectric effect, the Compton effect, the energy levels of material systems, and the laws governing the emission and absorption of radiation. In all these examples, we shall also study in detail how the quantum laws approach the classical limit.

1. Photoelectric Effect. We begin with a discussion of the photoelectric effect. The study of blackbody radiation would enable one to deduce indirectly that electromagnetic waves can change their energy only in units of hv; it would certainly now seem desirable to verify directly whether this statement is true by a study of the emission and absorption of radiation. The earliest experimental investigations of this problem were concerned with the photoelectric effect. These experiments showed that electrons are emitted from a metal surface* that is irradiated with light or ultraviolet rays; also, that their kinetic energy is independent of the intensity of the radiation, but depends only on the frequency in the following manner:

Here v is the frequency of the incident radiation, and W is the work function of the metal or, in other words, the energy needed to remove the electron from the interior of the metal.

Einstein was the first person to relate this result to Planck’s hypothesis (1905). Perusal of the data showed that h was a universal constant, and equal to the h appearing in Planck’s theory. This agreement is a strong confirmation of the hypothesis that the radiation field can change energy only in units of hv. If the constant had not been obtained, the theory would have been in serious difficulties.

The next important task is to try to determine why the electron absorbs energy only in quanta, independently of the intensity of the radiation. In this connection, it is worth noting that with radiation of very low intensity we simply obtain a correspondingly low rate of emission of photoelectrons.

The simplest interpretation of this phenomenon is that light consists of particles* which, because they are localized objects, can transfer all their energy to the photoelectrons during a collision. This idea is strengthened by experiments in which very low-intensity beams are directed at a photographic plate;† we obtain dark spots at random positions, with an average density proportional to the intensity of the light. In the limit of a very intense beam, the distribution of spots gets so dense that it is practically continuous.

When the beam is so intense that it seems to be continuous, it must in some way become the equivalent of what is described as a light wave in classical physics. Such a classical wave has a certain rate at which energy is incident on any surface per unit area per unit time. Let us call the rate S. Then, when many quanta are present (intense beam or low frequencies), this rate must also be equal to the mean number, N, of incident quanta per unit area per second times their energy hv. Thus, we have

If only a few quanta are present, N must be the probability that a quantum strikes a unit area per second.

Although the assumption that light is made up of localized particles enables us to explain the photoelectric effect in a very simple way, it cannot be made consistent with the extremely wide range of experiments leading to the conclusion that light is a form of wave motion. As an example of the type of experiment that calls for a wave interpretation, consider the measurement of the intensity pattern of the light that strikes a screen, after being diffracted through an arrangement of one or more slits. It very often happens that when two nearby slits are open, the intensity will be very small at certain points on the screen where either slit, separately, would produce a high intensity. This result is both qualitatively and quantitatively explained by the assumption that light is made up of waves, which can interfere either constructively or destructively so that, under some circumstances, the waves coming from each of the two slits may cancel each other.

It would be impossible, however, to explain interference if one assumed that light was made up of localized particles. Such particles would have to go through either one slit or the other, and the opening of a second slit could hardly prevent a particle from reaching a certain point to which the particle would be free to go if this second slit were closed. On the other hand, the assumption of wavelike properties for light not only explains this particular experiment, but also a whole host of other experiments involving radiation—from radio waves to x rays. It is, therefore, certainly desirable to try to understand the appearance of quanta in terms of the wave theory of light, if possible.

To do this, we now consider the classical account of what happens in the photoelectric effect. When radiation strikes an electron vibrating within an atom, it transfers energy to the electron. If the electric field oscillates at a frequency that is resonant with the frequency of the electron in the atom, the electron will absorb energy from the light wave until it is liberated. One could try to explain the photoelectric effect by assuming that the properties of the atom are such that the electron would keep on gaining energy until it had picked up an amount equal to hv, after which it would be ejected. If atoms had these properties then, with very weak light, the photoelectric effect should not be observed for a long time, since it would take a long time to store the necessary quantum of energy. Experiments were conducted, however, with metallic-dust particles and very weak light. These dust particles were so small that it would have taken many hours to store hv of energy; yet, some photoelectrons were found to appear instantaneously.

, however, it seems unlikely that electrons with so large a surplus of energy would remain indefinitely inside the metal, until their release was triggered by a light wave of exactly the right frequency. Moreover, it has been found that no matter how we try to release an electron from a metal (for example, by bombardment of the metal by protons or by other electrons), we must always supply the same minimum energy, equal to the work function W. Similarly, it has been found that electrons cannot be liberated from atoms of a gas, unless a certain minimum energy equal to the ionization potential I is supplied (see the discussion of the Franck-Hertz experiments in Sec. 15). Yet, some electrons are liberated instantaneously by very weak light from gas atoms with a kinetic energy equal to hv I. In view of all this evidence we must, therefore, rule out the possibility of explaining the photoelectric effect by assuming that some electrons initially possess nearly all the energy with which they escape.

If electrons in metals had such a range of energies, then it would be difficult to make the quantum hypothesis self-consistent, because only part of a quantum would have to be absorbed to liberate a typical electron. According to Planck’s hypothesis, however, the radiation oscillators can supply a minimum of a full quantum in each absorption process. What would then happen to the rest of the quantum if only part of it were absorbed by the electron?

These particular efforts fail to explain the photoelectric effect in terms of a process of gradual accumulation of energy, and every similar attempt that has ever been carried out has also failed. This means that the wave theory is unable to account for the sudden appearance of finite amounts of energy on a single electron. We are, therefore, in a quandary. One set of experiments suggests that light is a particle that can be localized, and the other suggests, with equal emphasis, that it is a wave. Which approach leads to the correct picture? The answer is, neither.

Before we can obtain a correct theory of the wave-particle duality of the properties of light, we shall see that it is necessary to make radical changes in some of our most fundamental concepts dealing with the properties of matter and energy. These new concepts will be developed through the remainder of this book, but primarily in Chaps. 6, 8, and 22. For the present, however, we merely state that light must be regarded as existing in the form of fundamental units, or quanta, which can, in some circumstances, act like particles and, in other circumstances, like waves. We find a strong analogy here to the fable of the seven blind men who ran into an elephant. One man felt the trunk and said that an elephant is a rope; another felt the leg and said that an elephant is obviously a tree, and so on. The question that we have to answer is: Can we find a single concept that will unify our different experiences with light, just as our concept of the elephant unifies the experiences of the seven blind men?

2. Differences between Classical and Quantum Laws of Physics. Our first step in the program of developing the new concepts needed in quantum theory will be to bring out two crucial differences between the kind of physical law obtained in classical theory and the kind suggested by experience with quantum phenomena. The first difference is that whereas classical theory always deals with continuously varying quantities, quantum theory must also deal with discontinuous or indivisible processes. The second difference is that whereas classical theory completely determines the relationship between variables at an earlier time and those at a later time (i.e., it is completely causal), quantum laws determine only probabilities of future events in terms of given conditions in the past.

3. The Indivisibility of Quantum Processes. Let us now consider some of the experimental evidence that indicates the need for introducing the concept of discontinuous or indivisible processes into the quantum theory. The first important piece of evidence comes from the photoelectric effect. We have already seen, for example, that while all efforts to explain the photoelectric effect as a process of gradual transfer of energy from radiation field to matter have failed, the assumption that the transfer of energy is a discontinuous process that takes place in jumps of size ΔE = hv is in agreement with all the experiments dealing with this phenomenon. Moreover, the same assumption is also required by Planck’s hypothesis that the energy of the radiation oscillators is restricted to discrete values. That is, if the transfer of energy took place gradually, it would be necessary to consider states in which radiation oscillators had part of a quantum and, according to Planck’s hypothesis, no such state is possible.

As we shall see later, there are many other experiments which demand the interpretation that the transfer of energy is a discontinuous process. For the present, we shall offer an experiment by Lawrence and Beams,* who tried to break up a light quantum into two parts by means of a very fast shutter, utilizing a Kerr cell that could be activated in 10−9 sec. If the light wave were continuous, as described by classical theory then, with the intensities of light used, it would have taken much longer than 10−9 sec for a full quantum of energy to come through. Thus, we should expect that the shutter would break up the quanta into smaller quanta. They found, however, that none of the quanta was ever broken up.

If we combine Planck’s hypothesis with the fact that no one has ever been able to perform an experiment in which a part of a quantum has been detected, we are led to the conclusion that a quantum is an indivisible unit of energy. We may also see, from the failure of all attempts to follow the energy gradually, that the transfer of a quantum from one system to another is an indivisible process. The indivisibility of the quantum of energy, and the indivisibility of the process of transfer go together; they are necessary for each other’s logical self-consistency. We should conclude, therefore, that in the transfer of a quantum, the system cannot be regarded as passing through a succession of intermediate states, in which the energy is exchanged in a continuous fashion. Instead, the quantum process must be regarded as discontinuous and as an indivisible unit. The transfer of a quantum is one of the basic events in the universe and cannot be described in terms of other processes. It may be called an elementary process, just as a proton or an electron is called an elementary particle, because it does not seem to be made up of other particles.

4. Probability and Incomplete Determinism in Quantum Laws. The indivisibility of quantum processes is totally at variance with classical physics, which describes all processes in a continuous fashion, each change being caused by the state of the system just before the change took place. Since classical laws presuppose the existence of continuous processes to which they apply, it is clear that discontinuous quantum jumps cannot be predicted by our classical laws. Our problem is, then, to find the new laws governing quantum transfers.

We come now to the second important difference between classical and quantum laws. It is an experimental fact, exemplified in the photoelectric effect and in a wide range of other experiments not yet studied here, that no law has been discovered which predicts exactly where and when an individual quantum will be transferred. Instead, only the probability of such a process may be predicted. For example, if only one quantum is directed at a metal surface, it is impossible to predict whether it will be absorbed and, if it is absorbed, exactly where and when. But if a beam contains many quanta, it is possible, from the intensity of the light used, to predict the mean number absorbed in any given region. Thus, in this case, quantum laws appear to control only the probability of an event and cannot predict its occurrence with certainty. We shall see that this behavior is not restricted to the photoelectric effect but is common to all quantum processes.

. Once the initial position and velocity of each particle are given, the future motion is determined exactly by the differential equations of motion. Thus, the trajectory of an electron is determined by three quantities:

(1) The position at any instant of time.

(2) The velocity at that time.

(3) The value of the force F at all times.

For an electrical particle, the force F is determined by the electric and magnetic fields. But these can be calculated exactly with the aid of Maxwell’s equations and the initial values of electric and magnetic fields everywhere. Hence, according to classical physics, the motion of a charged particle (also of any other kind of particle) can be determined precisely for all time, once certain initial conditions are known. The same can be said about changes of the electromagnetic field. Classical theory may therefore be called completely deterministic.

Applying these general ideas, one concludes from classical theory that, in a light beam of a given intensity, electrons gain energy at a continuous rate, which is calculable from the light intensity and from the initial conditions of the electrons. On the other hand, experiments show that the process of energy transfer is discontinuous and apparently not governed exactly by deterministic laws, at least not by the deterministic laws of classical mechanics. Instead, so far as we can find out from experiment, only the probability of the process is determined.

At this point, it is worthwhile to go more deeply into the connection between the appearance of probability and the indivisibility of a quantum process. First, there is the previously mentioned fact that many classical laws (including Newton’s equations of motion) which are essential for the operation of classical determinism must, by their very nature, refer to gradual and continuous processes. Hence, if only because this kind of law has no meaning in discontinuous processes, it cannot apply directly to quantum transfers. Some classical laws, however, do not require us to follow particles through a continuous path in space time, for example, conservation of energy, momentum, or angular momentum. Even in an impulsive collision in which we cannot follow the motion continuously, these laws apply for the collision as a whole. Such laws do have meaning even in discontinuous processes. It is an experimental fact that these laws can all be taken over directly into the quantum theory. For example, it has been shown experimentally that energy is always conserved in the photoelectric effect. Many other experiments also yield this result. Hence, not all classical deterministic laws must be abandoned, but only these requiring a description in terms of continuous processes.

5. Unlikelihood of Completely Deterministic Laws on a Deeper Level. One might wonder whether the appearance of probability in quantum processes is not a result of our ignorance of the correct variables to use in describing the system. In classical physics, probabilities often appear for just this reason. For example, in thermodynamics we measure the pressure, temperature, and volume of a given system. In very small regions of space, especially near the critical point, we find that these quantities no longer obey an equation of state exactly, but instead exhibit large random fluctuations about a mean value that is predicted by the equation of state. Hence, the deterministic laws of thermodynamics break down and are replaced by laws of probability. This is because the thermodynamic variables are no longer appropriate for the problem and must be replaced by the position and velocity of each molecule, which are, from the viewpoint of thermodynamics, hidden variables. The thermodynamic quantities are, then, merely averages of hidden variables that cannot be observed by thermodynamic methods alone. To find the underlying causal laws, we must accept a description in terms of the individual molecules.

The idea immediately suggests itself that probability in quantum processes arises in a similar way. Perhaps there are hidden variables that really control the exact time and place of a transfer of a quantum, and we simply haven’t found them yet. Although this possibility cannot be absolutely ruled out, we can show that this is unlikely. The first point, of course, is that no experiment has yet shown the slightest trace of such hidden variables. The second point is that there are strong theoretical arguments which make it unlikely that such hidden variables exist. These will be discussed later (Chap. 22, Sec. 19). For the present, we

Enjoying the preview?
Page 1 of 1