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

Only $11.99/month after trial. Cancel anytime.

Quantum Mechanics
Quantum Mechanics
Quantum Mechanics
Ebook759 pages7 hours

Quantum Mechanics

Rating: 3 out of 5 stars

3/5

()

Read preview

About this ebook

Suitable for advanced undergraduates, this thorough text focuses on the role of symmetry operations and the essentially algebraic structure of quantum-mechanical theory. Based on courses in quantum mechanics taught by the authors, the treatment provides numerous problems that require applications of theory and serve to supplement the textual material.
Starting with a historical introduction to the origins of quantum theory, the book advances to discussions of the foundations of wave mechanics, wave packets and the uncertainty principle, and an examination of the Schrödinger equation that includes a selection of one-dimensional problems. Subsequent topics include operators and eigenfunctions, scattering theory, matrix mechanics, angular momentum and spin, and perturbation theory. The text concludes with a brief treatment of identical particles and a helpful Appendix.
LanguageEnglish
Release dateMay 5, 2015
ISBN9780486804781
Quantum Mechanics

Related to Quantum Mechanics

Titles in the series (100)

View More

Related ebooks

Physics For You

View More

Related articles

Reviews for Quantum Mechanics

Rating: 3.199998 out of 5 stars
3/5

5 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Quantum Mechanics - John L. Powell

    MECHANICS

    CHAPTER 1

    HISTORICAL ORIGINS OF THE QUANTUM THEORY

    1–1 Difficulties with classical models. Classical physics deals primarily with macroscopic phenomena. Most of the effects with which classical theory is concerned are either directly observable or can be made observable with relatively simple instruments. There is a close link between the world of classical physics and the world of sensory perception.

    During the first decades of the present century, physicists turned their attention to the study of atomic systems, which are inherently inaccessible to direct observation. It soon became clear that the concepts and methods of classical macroscopic physics could not be applied directly to atomic phenomena. If classical laws of physics are to be applied at all to systems of atomic size, they can be considered only in connection with models of such systems. Models are usually conceived as systems of particles which interact with one another and with electromagnetic radiation according to assumed simple laws. One attempts to construct a microscopic model which will reproduce, as nearly as possible, observed macroscopic effects.

    Many of the observed characteristics of atomic systems, however, are such that they cannot be reproduced by any model which behaves according to classical laws. The early development of atomic theory consisted of efforts to overcome these difficulties by modifying the laws of classical physics and the properties of the models to which they were applied. These efforts reached their successful conclusion in the period from 1925 to 1930, when an entirely new theoretical discipline, quantum mechanics, was developed by Schrödinger, Heisenberg, Dirac, and others. The present book is an introduction to this theory. The first chapter is a review of the history of the subject and provides a background for later work. Some of the problems which must be faced in an attempt to understand atomic phenomena will be pointed out.

    The earliest evidence of the need for revision of classical concepts came from the field of chemistry. It had long been realized that the molecules of which a pure substance is composed are chemically identical, and that they retain their identity over long periods of time. A molecule of nitrogen, for example, consists of a number of positively and negatively charged particles which are held together by electrostatic forces. However, it is stated by Earnshaw’s theorem that a system of charged particles cannot remain at rest in stable equilibrium under the influence of purely electrostatic forces. If a classical picture is adopted, these particles must, therefore, be in relative motion. Yet, this motion must be such as not to destroy the identity of the molecules, and must persist indefinitely. If the particles are confined to a restricted region of space, they must frequently or continuously change their direction of motion, i.e., be accelerated. It is, however, a well-known fact that accelerated charged particles radiate energy in the form of electromagnetic waves. Hence, the particles in a molecule should progressively change their state of motion in accordance with this loss of energy—a conclusion which does not agree with observation. This fact alone shows that the stability of molecules cannot be understood on the basis of a classical model.

    The enormous range of electrical conductivities of solid materials is an example of a property of matter which cannot be reasonably explained in terms of classical ideas. For instance, the conductivity of silver is more than 10²⁴ times larger than that of fused quartz. Electrical conduction presumably consists of a relative motion of the negatively and positively charged components of the material under the influence of an applied electric field. It is not possible to comprehend, in terms of a classical picture, how such motion occurs readily in silver, but not at all in quartz. It seems to be necessary to recognize the existence of a principle of selection which prohibits the motion in quartz, but not in silver. A similar situation is found in ferromagnetism: the magnetic susceptibility of iron is observed to be of the order of 10⁹ times larger than that of other metals. Quite generally, any attempts to understand the chemical behavior of matter, as summarized in the periodic table of the elements, must take into account the fact that not all the states of motion permitted by a classical model are accessible to the systems of particles which comprise the molecules of a chemical substance.

    1–2 Optical spectra. Light is emitted by substances which are raised to a high temperature or subjected to an electrical discharge. This light can be separated into its spectral components by means of a diffraction grating, and the wavelengths of the various components can be measured with great precision. Characteristic emission spectra, consisting of discrete frequencies or lines, are obtained in this way for each element.

    The emission of this electromagnetic radiation must be associated with the accelerated motion of the charged particles in the atoms of the emitting substance. An attempt to construct a classical model reproducing the observed frequencies naturally leads to the conclusion that these must be the same as the frequencies of the periodic motions of the particles. The spectral lines should therefore fall into groups within which the frequencies v are related by the harmonic law where ν1, ν2, ν3, . . . are the fundamental frequencies of the motions of the atomic system, and n1, n2, n3, . . . are integers denoting harmonics of these frequencies. Such a relationship does not describe the experimental facts. Rather, spectral lines occur in series, such as the Balmer series in the visible and near ultraviolet region of the hydrogen spectrum, which is reproduced in Fig. 1–1. Within a series, the lines are most widely spaced on the long wavelength side and are crowded together near the short wavelength series limit.

    The empirically observed regularities among spectral lines are expressed by the Ritz combination principle, first formulated in 1908. According to this principle, a small set of numbers or spectroscopic terms can be assigned to the atoms of a substance, such that the wave numbers[1] of the numerous spectral lines are equal to differences between terms:

    FIG. 1–1. The Balmer series in the emission spectrum of hydrogen. (From Atomic Spectra and Atomic Structure by Gerhard Herzberg, courtesy of Dover Publications, New York.)

    Classical physics does not tell how the terms arise, nor does it explain the combination principle.

    1–3 Blackbody radiation. The concept that atomic systems exchange radiant energy in discrete amounts, or quanta, rather than in a continuous way, was first enunciated by Max Planck[2] in 1901, in his theoretical development of the law that describes the frequency distribution of heat radiation. The spectrum of the radiation emitted by a hot object is continuous; experimentally, it is known to cover a wide frequency range, with maximum intensity at a wavelength which depends upon the temperature of the radiating body.

    It can be shown by thermodynamic reasoning that the spectral energy distribution of the radiation within an enclosure at constant temperature is independent of the shape and material of the enclosure (provided it is large enough), and is a function of temperature and frequency alone. The radiation within such an enclosure can be observed through a small hole in its wall, which is a perfect absorber for radiation incident from without and therefore acts as a black body.

    The electromagnetic radiation can be considered to be a superposition of harmonic waves that correspond to the various normal modes of oscillation or degrees of freedom within the enclosed space. The frequency distribution of the radiation can be obtained if it is possible to enumerate the normal modes and determine the energy associated with each.

    Consider the latter problem first. It is shown in statistical mechanics that, at temperature T, the probability of excitation of a mode whose energy lies in the range between E and E + dE is proportional to the Boltzmann factor exp (–E/kT) dE. Accordingly, the energy associated with each mode is, on the average,

    This is an expression of the classical law of equipartition of energy.

    The number of degrees of freedom can be determined by considering the enclosure as a large cubical cell with sides of length L. A closer examination of the problem shows that the shape of the enclosure is immaterial to the result. It can also be shown that the particular form of the boundary conditions at the walls does not matter. Assume, therefore, that the field is periodic with period L in each direction parallel to an edge of the cube, i.e., that an integral number of wavelengths must fit into the length L. There are two modes of oscillation (one for each independent direction of polarization) corresponding to each plane wave exp (ik·r) for which

    Here, k is the propagation vector for the wave, with components kx, ky, and kz and of magnitude k = 2π/λ, where λ is the wavelength. Hence, there are two degrees of freedom for each triplet of integers (nx, ny, nz). These integers may conveniently be exhibited as a set of lattice points in (nx, ny, nz)-space, as shown in Fig. 1–2. There is one such point per unit volume of this space.

    Then n is the radius of a sphere in (nx, ny, nz)-space, and the number of lattice points between n and n + dn is equal to the volume of a shell of radius n and thickness dn, that is, 4πn² dn.[3] It follows from the definition of n and from the conditions (1–3) that k = (2π/L)n; hence the total number of modes for which k is in the range (k, k + dk) is

    where the factor 2 has been introduced to take account of the two possible directions of polarization of the transverse electromagnetic waves. This result may be rewritten in terms of the frequency v = kc/2π and divided by the volume L³, showing that the number of degrees of freedom per unit volume with frequency between ν and ν + is 8πν² /c³.

    FIG. 1–2. Lattice points representing combinations of integers nx, ny, nz.

    Now energy kT is to be associated with each degree of freedom, and the energy per unit volume within the frequency range is therefore

    This is the Rayleigh-Jeans law, which results from the application of the classical equipartition theorem. The law agrees well with experimental results at low frequency (Fig. 1–3), but predicts a monotonic increase of the energy density with increasing frequency. The integral that represents the total energy density diverges:

    This divergence at high frequencies has been called the ultraviolet catastrophe. Attempts to secure a more satisfactory law for the frequency distribution of blackbody radiation on a classical basis have failed.

    FIG. 1–3. The spectral distribution of energy density in blackbody radiation. Comparison of observed spectrum with Rayleigh-Jeans formula. (Adapted from W. P. Allis and M. A. Herlin, Thermodynamics and Statistical Mechanics, McGraw-Hill, New York, 1952, by permission.)

    , so that the only available energy states are those for which E , . . . , etc. On this hypothesis, the average energy of an oscillator is

    Using the formulae

    and

    we obtain for the average oscillator energy

    → 0, or at very high temperature, this result reduces to kT, in agreement with the value obtained from the equipartition theorem.

    At the time of Planck’s work, Wien had already shown on thermodynamic grounds that the energy density must have the functional form ν³f(ν/T). This requirement is satisfied if one sets

    where h is a universal constant. If the expression 8πν² dv/c³ is introduced for the number of modes in the interval (ν, ν + ), Planck’s law results:

    This law is in agreement with experiment. The avoidance of the ultraviolet catastrophe on the basis of the quantum partition law is easily understood, since the available states are now widely separated in energy when ν is large, and can be reached only by the absorption of very high-energy quanta, a relatively rare occurrence.

    The constant h (Planck’s constant), as well as the Boltzmann constant k, can be evaluated by comparison with experiment. The total energy density, obtained by integrating Planck’s distribution function over the whole spectrum, must depend on the temperature in accordance with Stefan’s law:

    If the integral is evaluated, using Eq. (1–10) for the energy density, one obtains for the value of Stefan’s constant

    Furthermore, Planck’s law must yield a maximum energy density at the wavelength predicted by Wien’s law:

    The experimental values for the Stefan-Boltzmann constant σ and for Wien’s constant b then determine the value of k and of Planck’s constant,[4]

    Additional insight into the significance of the quantum hypothesis is afforded by an alternative derivation of the radiation law, due to Einstein.[5] Detailed reference to the electromagnetic radiation theory is avoided in this treatment, and the role played by statistical considerations is emphasized. Suppose that two states of energy E1 and E2 (E2 > Ε1) are available to an atom bathed in radiation of density . It is assumed that quantum jumps between these states can take place in three different ways: (1) spontaneous transitions, with the probability A12 that the atom jumps from state 2 to state 1, emitting a quantum, (2) absorption, with the probability B21that the atomic state changes from 1 to 2 with absorption of a quantum, and (3) induced emission, with probability Β12, in which the state changes from 2 to 1 under the influence of the incident radiation. In thermal equilibrium, the probabilities that states 1 and 2 are occupied are proportional to e–E1/kT and e–E2/kT, respectively. Therefore the equation

    expresses the condition of equilibrium, namely,

    Now it is assumed that the Einstein coefficients Α12, Β12, B21 are properties of the atomic states 1 and 2, and therefore independent of . Consequently, if the temperature is sufficiently high and therefore very large, it may be assumed that

    hence B21 = Β12. This is the principle of detailed balance for induced transitions, that is, the probability that a transition is induced by the action of the electromagnetic field on the charged particles in the atom is the same for the transition 1 → 2 as for 2 → 1. With this substitution in Eq. (1–14) one obtains, on rearranging,

    where E2 – E1 = . By comparison with Eq. (1–10), it is seen also that

    The above argument, which is very much in the spirit of modern theory, is based upon the idea that quantum jumps take place in a random way, and that only probabilities for their occurrence can be denned. This application of the concept of probability is to be distinguished from its application in classical statistical mechanics. It is asserted that there is a certain probability per unit time that a random event may occur. This concept will receive further elaboration in later chapters. The Einstein coefficients depend upon the detailed structure of the atom and can be computed from quantum principles (cf. Chapter 11).

    Einstein’s derivation of Eq. (1–16) is essentially nonclassical. It contains as a basic element the assumption that the atom and radiation can coexist in a stable state, and that transitions from one state to another can be predicted only statistically.

    1–4 The photoelectric effect. There is no satisfactory classical explanation for the detailed characteristics of the photoelectric effect. Apparatus for the study of this effect may consist of a vacuum chamber containing a metal plate opposite a collector (Fig. 1–4a). A retarding potential is maintained between the two electrodes. When a beam of ultraviolet light strikes the plate, electrons are emitted from the metal. Those electrons whose initial kinetic energy equals or exceeds the work that is required to overcome the retarding potential can reach the collector and contribute to the photoelectric current between the electrodes.

    The lowest retarding potential for which the photoelectric current is zero is called the stopping potential (Fig. 1–4b). It is proportional to the maximum energy with which photoelectrons are emitted from the plate. According to classical considerations, this maximum energy should increase when the intensity of the incident light is increased; the force which the light exerts on electrons in the metal surface should be proportional to the magnitude of the electric vector ε of the incident light wave, and the magnitude of this vector increases when the light is made more intense. Contrary to this expectation, experiment shows that the maximum energy of the photoelectrons is independent of the intensity of the incident light. However, the energy of the photoelectrons is found to increase with the frequency of the incident light.

    These properties of the photoelectric effect were first explained by Einstein.[6] Planck’s earlier work on blackbody radiation, described in the preceding section, had indicated that the exchange of energy between radiation and the walls of a cavity occurs in quanta of magnitude . Einstein extended this idea and suggested that, in inducing the emission of photoelectrons, light does not act like a wave, but like a stream of discrete quanta or photons. The energy of one photon can be imparted to only one electron. Conservation of energy then leads to Einstein’s photoelectric law:

    Here, E is the kinetic energy with which the photoelectron leaves the metal surface, and W is the work function, which depends on the nature of the metal. Of the energy of the incident quantum, a portion W is expended in freeing the electron from the surface, and the remainder, E, is imparted to the photoelectron as kinetic energy. Einstein’s photoelectric law agrees closely with experiment and leads to the same value for h that is obtained from Planck’s law.

    FIG. 1–4. Photoelectric effect, (a) Apparatus for the study of the photoelectric effect, (b) Typical variation of photoelectric current with potential between collector and plate.

    1–5 The Franck-Hertz experiment. Planck’s work on blackbody radiation and Einstein’s study of the photoelectric effect showed that electromagnetic radiation when interacting with matter acts like an assemblage of discrete quanta of energy rather than like a continuous wave. The idea of naturally discontinuous processes of emission and absorption was new in physics, and it is understandable that further experimental confirmation was diligently sought. In 1914, Franck and Hertz[7] reported an unusually elegant experiment which proved that mechanical energy, like electromagnetic energy, is absorbed by atoms in discrete quanta. Because of its classic simplicity and its convincing nature, this experiment deserves somewhat detailed consideration.

    The apparatus used by Franck and Hertz (Fig. 1–5) consisted of an electrically heated wire, located along the axis of a cylindrical grid, which was surrounded by a collector. The device was housed in an enclosure filled with mercury vapor. An accelerating voltage was applied between heater and grid, and a retarding voltage was maintained between grid and collector.

    From the variation of collector current with retarding voltage, the energy of electrons that had passed through the mercury vapor could be determined. It was found that electrons whose total energy was less than 4.9 ev[8] lost no detectable energy at all in collisions with mercury atoms. Any collisions between these low-energy electrons and atoms of the vapor must therefore have been perfectly elastic, imparting no energy to the atoms except for negligible recoil. However, when the accelerating potential was increased above 4.9 volts, inelastic collisions occurred near the grid, in which electrons gave up their entire kinetic energy to mercury atoms. After losing their energy in an inelastic collision, the electrons were no longer able to traverse the retarding field, and the collector current fell to a minimum.

    FIG. 1–5. Apparatus for the experiment of Franck and Hertz.

    FIG. 1–6. Collector current as a function of accelerating voltage in the Franck-Hertz experiment. [After J. Franck and G. Hertz, Verhandl. deut. physik. Ges. 16, 457 (1914).]

    A further increase in the accelerating voltage moved the region where electrons reached the critical energy of 4.9 ev closer to the central wire. After attaining the critical energy and losing it through collisions, electrons could now pick up new energy on their way to the grid, so that the collector current rose again. A second current minimum was obtained with an accelerating potential of approximately 10 volts, due to a second region of inelastic collisions in the vicinity of the grid. The graph of collector current as a function of accelerating voltage is reproduced in Fig. 1–6. The experimental results show clearly that mercury atoms absorb mechanical energy in quanta of 4.9 ev.

    The question immediately arose whether the atoms would reradiate quanta of the same energy. The frequency of the emitted radiation would then be 1.18 × 10¹⁵ sec–1, which corresponds to ultraviolet light of wavelength 2530 A. Franck and Hertz repeated their experiment in a quartz container, transparent to ultraviolet radiation, and photographed the emission spectrum from the mercury vapor. A strong line appeared at 2536 A, agreeing, within the limits of error, with the calculated value.

    1–6 The Rutherford atom. Once the quantum nature of energy exchange was recognized, theoretical physicists were confronted with the task of suggesting a mechanism for the absorption and emission of quanta, in accordance with spectroscopic evidence. The mechanism would, of course, depend closely upon a satisfactory model of the atom.

    The most important information about the structure of atoms was discovered by Rutherford in 1911. In a classic paper,[9] Rutherford analyzed experimental results of Geiger and Marsden on the scattering of alpha particles from thin metallic foils. The experiments had shown that a few incident alpha particles (about one in 20,000) were deflected through an average angle of 90° in passing through a thin (4 × 10–5 cm) gold foil. Rutherford assumed that such large deflections were produced in a single encounter of an alpha particle with an atom, since previous calculations based on multiple scattering did not give a satisfactory result. He showed that the experimental results could be explained if the atom was assumed to consist of a strong positive or negative central charge, concentrated within a distance of less than 3 × 10–12 cm and surrounded by a sphere of electrification of the opposite charge that extended throughout the remainder of the atom, i.e., to a distance of approximately 10–8 cm. The scattering could then be assumed to be due mainly to the central charge or nucleus, which would cause the alpha particle to describe a hyperbolic path with the center of the atom as one focus.

    A calculation based on classical mechanics and the Coulomb force between alpha particle and atomic nucleus led Rutherford to his well-known formula for the number n(θ) of particles which are deflected into unit solid angle in the direction θ:

    In this expression, m and υ are the mass and velocity, respectively, of the incident alpha particle, N is the number of scattering centers per unit area, n0 is the number of incident alpha particles per unit area, and Ze and Ze are, respectively, the charge of the nucleus and of the alpha particle.[10] The correctness of Rutherford’s scattering formula, and therefore of his assumption of a nuclear atom, was borne out by further experiments,[11] during which more than 100,000 scintillations produced by alpha particles impinging upon a zinc sulfide screen were counted. In this work, Geiger and Marsden confirmed the 1/sin⁴(θ/2)-dependence of the number of scattered alphas on the scattering angle, as well as the direct proportionality of the thickness of the scattering foil to the number of alphas scattered in any given direction. They found also that the scattering per atom of foils of various materials varies approximately with the square of the atomic weight, and that the scattering by a given foil is inversely proportional to the fourth power of the velocity of the incident alpha particles. Finally, Geiger and Marsden were able to calculate from their data that the charge of the atomic nucleus is approximately equal to one-half the atomic weight.

    It is interesting to note that a nuclear atom had already been considered mathematically by Nagaoka[12] in 1904, but it was Rutherford’s analysis that established this concept as an experimental fact. The problem which immediately arose, however, concerned the stability of such a system. For reasons pointed out in Section 1–1, a classical model of a nuclear atom is unstable.

    1–7 Stationary states of atoms. In 1913, Niels Bohr was able to resolve the question of the stability of the Rutherford atom and to take account of the absorption and emission of quanta. Bohr formulated a completely new theory of atomic structure, based on postulates that deviated fundamentally from the pattern of classical physics.[13] Bohr’s work constitutes one of the most brilliant advances in modern physics and was basic in the development of the quantum theory.

    In his first postulate, Bohr assumed the existence of discrete stationary states of the atom, with electrons moving about a positive nucleus in orbits which can be computed from classical theory. In the simplest case, that of the hydrogen atom, a single electron is assumed to describe a circle, or an ellipse with the nucleus at one focus. The total energy of the atom in such a stationary state remains constant; contrary to classical electrodynamics, none is radiated.

    As a second postulate, Bohr suggested that an atomic electron can make a transition from one stable orbit to another in a way which cannot be treated classically. If the energy Em of the atom in the final state is lower than its energy En in the initial state, then the energy difference is radiated as a single photon. Since the energy of a photon of frequency ν is , conservation of energy requires that

    The Bohr frequency rule follows:

    In the inverse process, a photon of the appropriate energy is absorbed by the atom, raising an electron from an orbit of lower energy to one of higher energy.

    Bohr showed also that the ionization potential of hydrogen and the frequencies of the lines of the Balmer series could be obtained if one assumed that the stationary states of the hydrogen atom were characterized by certain definite values of the angular momentum of the electron about the nucleus. According to this theory, the angular momentum is an integral multiple of Planck’s constant, divided by 2π:

    Bohr introduced his postulates of nonradiating, discrete stationary states and of transitions between states without justification other than that they explained a number of experimentally known facts. A deeper understanding of Bohr’s assumptions became possible only at a later time, with the advent of wave mechanics.

    It will be recalled that the empirical Ritz combination principle states that the frequencies of the spectral lines of an atom can be obtained as differences between pairs of term values. In the light of Bohr’s postulates, the set of characteristic term values of an atom is to be interpreted as the set of allowed energy values for the atomic system, divided by the velocity of light times Planck’s constant. With this interpretation, the Ritz combination principle and the Bohr frequency rule become identical.

    1–8 The correspondence principle. Bohr’s postulates imply that atomic systems are not entirely governed by the laws of classical physics. However, a basic condition can be stated, which can serve as a guide in the development of a more adequate theory. Whatever form it may take, the new theory must agree with classical physics in any of the very large number of situations in which classical physics has been found to provide the correct answer.

    Newtonian mechanics and classical electrodynamics are based on a wealth of thoroughly established experimental evidence. Therefore, it must be demanded that the quantum theory yield, in every instance, results that become identical with those of classical physics if the masses and dimensions of the system under consideration are made to approach the masses and dimensions of classical systems. This fundamental idea is already apparent in Bohr’s early work and was explicitly stated by him in 1923. It is known as the correspondence principle and is essential in the formulation of quantum mechanics.

    1–9 The Bohr atom. The greatest triumph of the old quantum theory was its successful interpretation of the spectrum of hydrogen. The later wave-mechanical treatment of atoms differs in many respects from the early theory. Yet, Bohr’s theory of the hydrogen atom retains sufficient interest and historical importance to justify its somewhat detailed presentation at this point.

    In accordance with Kepler’s first law, the electron orbit is, in general, an ellipse. However, the special case of a circular orbit is subject to particularly straightforward treatment and will be considered first. For further simplification, the nucleus will be assumed to be at rest. This is equivalent to assigning it infinite mass. A correction for the finite mass of the nucleus will be introduced later.

    Consider an atom in which a single electron of mass m moves with velocity υ in a circular orbit of radius a, centered at the nucleus. The charge of the electron is –e and that of the nucleus, +Ze. The centripetal force is equal to the Coulomb force between electron and nucleus:

    The kinetic energy of the electron, therefore, is

    Now according to Bohr, the stationary states of the system are characterized by definite values of the angular momentum, such that

    ("h-bar or Dirac’s h") is written for the quantity h/2π. Hence, the velocity of the electron in its nth orbit is

    When this value for the velocity is substituted in Eq. (1–24), the radius of the orbit is obtained:

    Of particular interest is the radius of the orbit in the ground state of the hydrogen atom (n = 1, Z = 1):

    This quantity is known as the Bohr radius.

    The total energy of the atom is the sum of kinetic and potential energies:

    The energy of the system in its nth state is therefore

    1–10 Spectroscopic series. The Balmer series of lines in the hydrogen spectrum is shown in Fig. 1–1. The series is named for J. J. Balmer, who in 1885 discovered the following empirical relation describing the wave numbers of the lines in this group:

    Here, RH is the Rydberg constant for hydrogen, with the numerical value[14] 109,677.58 cm–1. Substitution of the integers 3, 4, 5,. . . for n in the Balmer formula gives the wave numbers of the lines in the Balmer series. These wave numbers have been verified experimentally to the very great accuracy attainable in spectroscopic measurements.

    The Balmer formula follows from the Bohr frequency rule which, in terms of wave numbers, reads

    Substituting the energy values from Eq. (1–30), we obtain

    It is apparent that the Balmer series is produced by a group of transitions in which the electron falls from an outer orbit into the second innermost (n = 2) orbit. This is indicated in Fig. 1–7. Note that the indices m and n of the frequency rule appear in reversed order in the Balmer formula, because Em and En are negative quantities. The Rydberg constant (for infinite nuclear mass) appears as a combination of universal constants,

    The frequency rule suggests the existence of other series in the spectrum of hydrogen. In fact, other series arise when the index m in Eq. (1–33) has values different from 2. In accordance with this prediction, four other series have been observed, none of which, however, lie in the visible range.

    FIG. 1–7. (a) Transitions producing the spectral series of hydrogen (energy not to scale), (b) Scale drawing of the energy levels of the hydrogen atom.

    The five hydrogen series, named for their discoverers, are

    Representative transitions that produce lines of these series are also indicated in Fig. 1–7. The lines in each series fall closer together as the running index n is raised. In the limit of large n, the lines in each series converge toward the series limit, of wave number R/m².

    1–11 Correction for finite mass of the nucleus. The theoretical value (1–34) for the Rydberg constant was obtained under the assumption that the nucleus has infinite mass and, therefore, remains at rest. Actually, both the nucleus and the electron move about their common center of mass. When the finite mass of the nucleus is taken into account, a slightly different value for the Rydberg constant is obtained, denoted by RH. The correct energy levels do not differ greatly from those given by Eq. (1–30), since the ratio of nuclear mass to electron mass has the large value[15] 1836.13. One can account for the motion of the nucleus in the following manner: Newton’s second law, as it applies to the electron, of mass m1, moving with velocity v1 under the action of the Coulomb force F, is

    By Newton’s third law, an equal and opposite force acts on the nucleus, of mass m2, which has velocity v2:

    The relative acceleration of the electron with respect to the nucleus is

    so that the equation of motion for the electron with respect to the nucleus is

    This result, familiar from mechanics, indicates that allowance for the finite mass of the nucleus can be made by substituting the reduced mass m1m2/(m1 + m2) in place of the electronic mass m1 in the equation of motion. This substitution should be made in Eq. (1–24). Then, instead of m, the reduced mass will appear in Eq. (1–27) for the radius of the orbit and in Eq. (1–30) for the energy levels of the hydrogen atom; thus Eq. (1–30) becomes

    The Rydberg constant for finite nuclear mass is

    and the wave numbers of the lines in the hydrogen spectrum are given by

    Although the ratio of RH to R∞ is very nearly unity (approximately 1836/1837), the consequent small shift in wave numbers is within reach of spectroscopic observation. Of particular interest is the fact that Eq. (1–33), which contains R∞ and hence does not depend upon the nuclear mass, predicts that certain spectral lines from singly ionized helium should coincide exactly with lines from hydrogen.[16] Such coincidences are not observed. The discrepancy caused Bohr in 1914 to take account of the nuclear mass, as has been done above.[17]

    Another interesting consequence of the mass dependence of the Rydberg constant RH is that the spectral lines of deuterium (H²) are slightly shifted with respect to the corresponding lines of H¹. Through this shift the existence of deuterium was first proved by Urey and co-workers[18] after it had been predicted by Birge and Menzel[19]

    1–12 Quantization of the phase integral. Bohr had shown that the stationary states of the hydrogen atom can be computed if it is assumed that the angular momentum of the system can have only discrete values, equal to multiples of a universal constant. A more general postulate, which includes Bohr’s assumption and leads to the successful prediction of the allowed energy values for certain other systems as well, was discovered by W. Wilson[20] in 1915 and, independently, by A. Sommerfeld[21] the following year.

    Let the motion of a periodic system of N degrees of freedom be described by the N coordinates qi and the N canonically conjugate momenta pi (i = 1, 2, 3, . . . , N). For example, the position of a mass point in three-dimensional space can be specified by the three cartesian coordinates q1 = x, q2 = y, and q3 = z, and its velocity by the

    Enjoying the preview?
    Page 1 of 1