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

Only $11.99/month after trial. Cancel anytime.

Molecular Spectra and Molecular Structure - Vol I
Molecular Spectra and Molecular Structure - Vol I
Molecular Spectra and Molecular Structure - Vol I
Ebook1,467 pages14 hours

Molecular Spectra and Molecular Structure - Vol I

Rating: 0 out of 5 stars

()

Read preview

About this ebook

“I have endeavoured to give a presentation which is readable by the beginner in the field and also will be useful to those who do or want to do research work in this field. In order to assist the former, I have frequently made use of small type for those sections that are not necessary for an understanding of the fundamentals. For the benefit of those working in the field, numerous references to original papers have been included.
“A satisfactory presentation of molecular spectra and molecular structure is nowadays not possible without treating thoroughly, apart from the empirical results, the theoretical background also. Therefore, I have included as much of the theory of molecular spectra as is possible without going into the more difficult mathematical details. A large number of diagrams, graphical representations of eigenfunctions and potential curves, as well as energy level diagrams, serve to illustrate and to explain the theory. On the other hand, I have added numerous carefully selected spectrograms of bands and band systems (some of which have been taken specially for this purpose) in order to give an accurate idea of the experimental material that forms the basis of the developments.
LanguageEnglish
Release dateApr 18, 2013
ISBN9781446549148
Molecular Spectra and Molecular Structure - Vol I

Related to Molecular Spectra and Molecular Structure - Vol I

Related ebooks

Chemistry For You

View More

Related articles

Reviews for Molecular Spectra and Molecular Structure - Vol I

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Molecular Spectra and Molecular Structure - Vol I - Gerhard Herzberg

    TEXT BOOK

    MOLECULAR SPECTRA

    and

    MOLECULAR STRUCTURE

    I. SPECTRA OF DIATOMIC MOLECULES

    BY

    GERHARD HERZBERG, F.R.S.

    National Research Council of Canada

    With the co-operation, in the first edition, of

    J. W. T. SPINKS, F.R.S.C.

    SECOND EDITION, EIGHTH PRINTING

    D. VAN NOSTRAND COMPANY, INC.

    Dedicated

    to the Memory of

    WALTER CHARLES MURRAY

    First President

    of the

    University of Saskatchewan

    PREFACE

    Eleven years ago I published a volume entitled Molecular Spectra and Molecular Structure I. Diatomic Molecules which was followed in 1945 by a second volume Infrared and Raman Spectra of Polyatomic Molecules. The first volume has been out of print for a number of years but the demand for it seemed to justify a new edition. Although the book has been completely revised and brought up to date, its general plan has remained substantially unchanged. Concerning this plan it seems therefore appropriate to quote from the preface of the first edition:

    "I have endeavored to give a presentation which is readable by the beginner in the field and also will be useful to those who do or want to do research work in this field. In order to assist the former, I have frequently made use of small type for those sections that are not necessary for an understanding of the fundamentals. For the benefit of those working in the field, numerous references to original papers have been included.

    "A satisfactory presentation of molecular spectra and molecular structure is nowadays not possible without treating thoroughly, apart from the empirical results, the theoretical background also. Therefore I have included as much of the theory of molecular spectra as is possible without going into the more difficult mathematical details. A large number of diagrams, graphical representations of eigenfunctions and potential curves, as well as energy level diagrams, serve to illustrate and to explain the theory. On the other hand, I have added numerous carefully selected spectrograms of bands and band systems (some of which have been taken specially for this purpose) in order to give an accurate idea of the experimental material that forms the basis of the developments.

    While of course most of the material presented is not new, it seems that the actual procedure followed in analyzing a band spectrum has not previously been given as specifically in a book of this kind. The same holds for the applications of band spectra to other parts of physics, to chemistry, and to astrophysics given in the last chapter. I hope that both these features will be found useful.

    In the eleven years since the publication of the first edition the subject Spectra of Diatomic Molecules has developed vigorously even though not as rapidly as in the preceding two decades. Most of the progress made has been consolidation and slow evolution rather than revolution. Exceptions to this statement are the amazing advances made by applying the new tools of molecular beams and microwaves to diatomic molecular problems.

    Naturally I have incorporated these advances of recent years in the present new edition. In addition I have amplified and extended the theoretical sections of the book in response to suggestions of several critics of the first edition. I have added new sections on Radiofrequency (Microwave) Spectra (in Chapter II), Hyperfine Structure (in Chapter V) and Intensities of Electronic Transitions (in Chapter VI) and I have completely rewritten many other sections. In order to improve the presentation and to bring it up to date I have scrutinized every sentence and every illustration. Thus the reader familiar with the old edition will find much revision and modification in detail. Twenty-six new illustrations have been added of which some, like Fig. 57 showing the symmetric top eigenfunctions, have not previously been given in the literature of the subject.

    The table of molecular constants of the first edition included only the ground states of all diatomic molecules then known. Following the suggestions of many spectroscopists and in view of the fact that the available tables that include the excited states are now out-dated I decided to undertake the rather arduous task of preparing a comprehensive table of all known states of all known diatomic molecules. This table is presented in the Appendix (Table 39) and now covers 80 pages. I have made every effort to make this table as comprehensive and consistent as possible (see the introduction to it, p. 501). The literature up to September 1949 and in some cases up to February 1950 has been included in the table.

    The manuscript of the text matter of the book was completed in August 1948, and only in a few cases was it possible to include later developments in the proofs.

    As in the first edition the detailed subject index at the end of the book includes all the more important symbols and quantum numbers used as well as all the molecules treated. Part II of the bibliography of the first edition was taken over and a new Part III added giving all the references to work published since then or to earlier work not used in the first edition.

    G. HERZBERG

    OTTAWA, ONT.

    February, 1950

    ACKNOWLEDGMENTS

    It gives me great pleasure to acknowledge the help and cooperation I have received from many persons during the preparation of this book. My sincere thanks are due to Dean J. W. T. Spinks who translated the original version into English and to Professor R. N. II. Haslam who contributed many stylistic improvements to the first edition. I am very grateful to Professor R. S. Mulliken for suggesting several important improvements. I should like to express my thanks to Professor S. Chandrasekhar for making many valuable suggestions and for reading parts of the manuscript and the proofs. Many colleagues helped by supplying advance information on molecular constants for Table 39. Of these I should like to mention particularly Dr. R. F. Barrow (Oxford) and Professor E. Miescher (Basel). The help given by Dr. J. G. Phillips (Williams Bay) and Dr. D. Andrychuk (Ottawa) in especially taking a number of new spectrograms for the present edition is much appreciated. Thanks are also due to the following spectroscopists for supplying the originals of various spectrograms of the new and the old edition: Dr. E. Bengtsson-Knave (Stockholm), Professor R. T. Birge (Berkeley), Dr. B. A. Brice (Philadelphia), Professor W. G. Brown (Chicago), Professor F. A. Jenkins (Berkeley), Professor R. Mecke (Freiburg), Professor R. S. Mulliken (Chicago), Dr. R. W. B. Pearse (London), Professor F. Rasetti (Baltimore), the late Lord Rayleigh, and Professor R. W. Wood (Baltimore). Finally I am greatly indebted to my wife who prepared all the illustrations of the first edition and made numerous suggestions for improving the text of the new edition.

    G. HERZBERG

    CONTENTS

    INTRODUCTION

    General remarks, Physical constants, Units and conversion factors

    CHAPTER I: RÉSUMÉ OF THE ELEMENTS OF ATOMIC STRUCTURE

    1. BOHR THEORY

    Stationary energy states, Radiation, Terms, Continuous spectra

    2. WAVE MECHANICS (QUANTUM MECHANICS)

    Fundamental equations, Physical interpretation of the Ψ function, Orthogonality and normalization of eigenfunctions, Perturbation theory, Heisenberg’s uncertainly principle, Momentum and angular momentum, Space quantization, Electron spin, Emission and absorption of radiation

    3. ATOMS WITH SEVERAL ELECTRONS; VECTOR MODEL

    Quantum numbers of the individual electrons, The Pauli principle, Quantum theoretical addition of angular momentum vectors, Quantum numbers and angular momenta of the whole atom; term symbols, Influence of a magnetic or electric field, Selection rules, Nuclear spin

    CHAPTER II: OBSERVED MOLECULAR SPECTRA AND THEIR REPRESENTATION BY EMPIRICAL FORMULAE

    1. SPECTRA IN THE VISIBLE AND ULTRAVIOLET REGIONS

    Coarse structure, Fine structure, Intensity distribution

    2. SPECTRA IN THE INFRARED REGION

    Near infrared spectra, Far infrared spectra

    3. RADIOFREQUENCY (MICROWAVE) SPECTRA

    Microwave absorption, Magnetic resonance spectra

    4. RAMAN SPECTRA

    Nature of the Raman effect, Large Raman displacements, Small Raman displacements

    CHAPTER III: ROTATION AND VIBRATION OF DIATOMIC MOLE CULES; INTERPRETATION OF INFRARED AND RAMAN SPECTRA

    1. INTERPRETATION OF THE PRINCIPAL FEATURES OF INFRARED AND RAMAN SPECTRA BY MEANS OF THE MODELS OF THE RIGID ROTATOR AND OF THE HARMONIC OSCILLATOR

    (a)The Rigid Rotator

    The molecule as a rigid rotator, Energy levels, Eigenfunctions, Spectrum

    (b)The Harmonic Oscillator

    The molecule as a harmonic oscillator, Energy levels, Eigenfunctions, Spectrum

    (c)Comparison with the Observed Infrared Spectrum

    (d)The Raman Spectrum of the Rigid Rotator and of the Harmonic Oscillator

    Classical theory of light scattering and of the Raman effect, Quantum theory of the Raman effect, Vibrational Raman spectrum, Rotational Raman spectrum, General remarks

    2. INTERPRETATION OF THE FINER DETAILS OF INFRARED AND RAMAN SPECTRA

    (a) The Anharmonic Oscillator

    The molecule as an anharmonic oscillator, Classical motion, Energy levels, Eigenfunctions, Infrared spectrum, Raman spectrum, Vibrational frequency and force constant, Continuous term spectrum and dissociation, Mathematical representation of the potential curves

    (b) The Nonrigid Rotator

    Energy levels, Spectrum

    (b) The Vibrating Rotator

    Energy levels, Eigenfunctions, Infrared spectrum, Raman spectrum

    (d) The Symmetric Top

    The diatomic molecule as a symmetric top, Angular momenta, Energy levels, Eigenfunctions, Infrared spectrum, Raman spectrum

    (e) Thermal Distribution of Quantum States; Intensities in Rotation-Vibration Spectra

    Vibration, Rotation

    (f) Symmetry Properties of the Rotational Levels

    Positive and negative rotational levels, Symmetric and anti-symmetric rotational levels for homonuclear molecules, Influence of nuclear spin, Influence of nuclear statistics, Ortho and para modifications, Isotopic molecules

    (g) Isotope Effect

    Vibration, Rotation

    CHAPTER IV: ELEMENTARY DISCUSSION OF ELECTRONIC STATES AND ELECTRONIC TRANSITIONS

    1. ELECTRONIC ENERGY AND TOTAL ENERGY

    Electronic energy and potential curves; stable and unstable molecular states, Resolution of the total eigenfunction, Resolution of the total energy

    2. VIBRATIONAL STRUCTURE OF ELECTRONIC TRANSITIONS

    General formulae, Examples; graphical representation, Absorption, Excitation of single progression in emission, Sequence (diagonal groups), Vibrational analysis, Isotope effect, Applications of the isotope effect

    3. ROTATIONAL STRUCTURE OF ELECTRONIC BANDS

    General relations, The branches of a band, Band-head formation; shading (degrading) of bands, Combination relations and evaluation of the rotational constants for bands without Q branches, Combination differences and evaluation of rotational constants for bands with Q branches, Determination of the band origins (zero lines), Determination of the numbering in the branches of incompletely resolved bands, The picking out of branches, Isotope effect

    4. INTENSITIES IN ELECTRONIC BANDS

    (a) Intensity Distribution in the Vibrational Structure

    Observed intensity distribution in absorption, The Frank-Condon principle: absorption, The Frank-Condon principle: emission (Condon parabola), Wave-machanical formulation of the Frank-Condon principle, Vibrational sum rule and vibrational temperature

    (b) Intensity Distribution in the Rotational Structure

    ¹Σ—¹Σ transitions, Other transitions, Intensity alternation, Wood’s resonance series

    CHAPTER V: FINER DETAILS ABOUT ELECTRONIC STATES AND ELECTRONIC TRANSITIONS

    1. CLASSIFICATION OF ELECTRONIC STATES; MULTIPLET STRUCTURE

    Orbital angular momentum, Spin, Total angular momentum of the electrons; multiplets, Symmetry properties of the electronic eigenfunctions

    2. CCOUPLING OF ROTATION AND ELECTRONIC MOTION

    (a) Hund’s Coupling Cases

    Hund’s case (a), Hund’s case (b), Hund’s case (c), Hund’s case (d), Hund’s case (e)

    (b) Uncoupling Phenomena

    A-type doubling, Transition from case (b) to case (d) (L uncoupling), Transition from case (a) to case (b) (spin uncoupling, Transition from case (b) to case (c)

    (c) Symmetry Properties of the Rotational Levels

    Σ states, Π, Δ, . . . states

    3. TYPES OF ELECTRONIC TRANSITIONS

    (a) Selection Rules

    General selection rules, Selection rules holding for case (a) as well as case (b), Selection rules holding only in case (a), Selection rules holding only in case (b), Selection rules holding only in case (c), Selection rules holding only in case (d), More general cases

    (b) Allowed Electronic Transitions

    Notation, ¹Σ—¹Σ transitions, ²Σ—²Σ transitions, ³Σ—³Σ transitions, ¹Π—¹Σ transitions, ¹Σ—¹Π transitions, ²Π—²Σ transitions, ²Σ—²Π transitions, ³Π—³Σ transitions, ³Σ—³Π transitions, Π—Σ transitions of higher multiplicities, ¹Π—¹Π transitions, ²Π—²Π transitions, ³Π—³Π transitions, Π—Π transitions of higher multiplicity, Π—Δ transitions, Δ—Δ transitions, Band structures in Hund’s cases (c) and (d), General remarks on the technique of the analysis of multiplet bands

    (c) Forbidden Electronic Transitions

    Violation of approximate selection rules, Quadrupole and magnetic dipole radiation, Enforced dipole radiation

    4. PERTURBATIONS

    Observed phenomena, General considerations, Selection rules for perturbations, Rotational perturbations, Vibrational perturbations, Intersection of potential curves and the non-crossing rule

    5. ZEEMAN EFFECT AND STARK EFFECT

    General remarks on splitting of molecular energy levels in a magnetic field, The Zeeman splitting of ¹Σ states, Zeeman effect in Hund’s case (a), Zeeman effect in Hund’s case (b), Other cases, Polarization of resonance fluorescence, Magnetic rotation spectra, Stark effect

    6. HYPERFINE STRUCTURE

    Hyperfine structure without field, Zeeman effect of hyperfine structure

    CHAPTER VI: BUILDING-UP PRINCIPLES, ELECTRON CONFIGURATIONS, AND VALENCE

    1. DETERMINATION OF THE TERM MANIFOLD FROM THE STATE OF THE SEPARATED ATOMS

    Unlike atoms, Like atoms

    2. DETERMINATION OF THE TERM MANIFOLD FROM THE STATES OF THE UNITED ATOM

    Unlike atoms, Like atoms

    3. DETERMINATION OF THE TERM MANIFOLD FROM THE ELECTRON CONFIGURATION

    (a) Quantum Numbers of the Individual Electrons

    Single electron in an axially symmetric electric field, Several electrons

    (b) The Pauli Principle in the Molecule

    (c) Derivation of the Term Type (Species) from the Electron Configuration in Russell-Saunders Coupling

    Terms of nom-equivalent electrons, Terms of equivalent electrons, Electron configurations with equivalent and non-equivalent electrons, Like atoms

    (d) Derivation of the Term Type (Species) for Other Types of Coupling

    (ω,ω) coupling, (Ω,ω) coupling, Other types of coupling

    (e) Term Manifold of the Molecule, Examples

    General considerations, H2 and the hydrides, Molecules with nuclei of equal charge, Other molecules

    4. STABILITY OF MOLECULAR ELECTRONIC STATES; VALENCE

    (a) Homopolar Binding (Atomic Binding)

    Treatment of the H2 molecule according to Heitler and London, Generalization of the Heitler-London theory for more complicated cases, Theory of bonding and antibonding electrons for equal nuclear charges, Unlike nuclear charges

    (b) Heteropolar Binding (Ionic Binding)

    Ionic molecules, Transition cases, Ionicity and polarity

    (c) Van der Waals Binding

    5. INTENSITIES OF ELECTRONIC TRANSITIONS

    General formulae, Rydberg transitions, Charge-transfer spectra, Other sub-Rydberg transitions

    CHAPTER VII: CONTINUOUS AND DIFFUSE MOLECULAR SPECTRA: DISSOCIATION AND PREDISSOCIATION

    1. CONTINUOUS SPECTRA AND BAND CONVERGENCE LIMITS: DISSOCIATION OF DIATOMIC MOLECULES

    (a) Absorption

    Ionization continua, Dissociation continua, Upper state continuous, Lower state or both lower and upper state continuous,

    (b) Emission

    Upper state continuous (molecule formation in a two-body collision), Lower state continuous, Both upper and lower state continuous,

    2. DIFFUSE MOLECULAR SPECTRA, PREDISSOCIATION, AND RELATED TOPICS

    (a) General Discussion of Spontaneous Radiationless Decomposition Processes

    The Auger process, Passage through potential barriers,

    (b) Radiationless Decomposition Processes in the Molecule

    Diffuseness of the bands, Photochemical decomposition, Breaking-off of bands, Different types of predissociation, Preionization, Inverse predissociation, Accidental predissociation,

    (c) Selection Rules for Predissociation

    Kronig’s selection rules, Forbidden predissociations, Preionization,

    (d) The Franck-Condon Principle in Predissociation

    Case I of predissociation, Case III (predissociation by rotation); effective potential curves, Influence of rotation in case I,

    (e) Pressure Effects in Predissociation

    Induced predissociation, Suppression of breaking-off by pressure,

    (f) Other Diffuse Molecular Spectra

    3. DETERMINATION OF HEATS OF DISSOCIATION

    (a) Determination of Dissociation Limits

    (1) Band convergences, (2) Extrapolation to convergence limits, (3) Long-wave-length limit of an absorption continuum, (4) Predissociation limits, (5) Excitation of atomic fluorescence, (6) Photodissociation, (7) Chemiluminescence,

    (b) Determination of the Dissociation Products

    (1) Energy differences of dissociation limits, (2) Application of the Wigner-Witmer correlation rules, (3) Application of the non-crossing rule, (4) Observation of atomic fluorescence, (5) Use of thermochemical data,

    (c) Examples

    O2, N2,

    CHAPTER VIII: EXAMPLES, RESULTS AND APPLICATIONS

    1. ENERGY LEVEL DIAGRAMS; MOLECULAR CONSTANTS

    Empirical relations, Ionization potentials,

    2. APPLICATIONS TO OTHER FIELDS OF PHYSICS

    Nuclear physics, Paramagnetism, Collision processes, Nature of the liquid and the solid state, Determination of high temperatures, Calculation of thermodynamic quantities,

    3. APPLICATIONS TO CHEMISTRY

    Free radicals, Elementary chemical reactions in gases and Chemical equilibria, Atomic heats of formation and related topics,

    4. APPLICATIONS TO ASTROPHYSICS

    Absorption spectrum of the earth’s atmosphere, Emission spectrum of the earth’s atmosphere, Planetary atmospheres, Comets, Stellar atmospheres, Interstellar space,

    APPENDIX: VIBRATIONAL AND ROTATIONAL CONSTANTS FOR THE ELECTRONIC STATES OF ALL KNOWN DIATOMIC MOLECULES (TABLE 39)

    BIBLIOGRAPHY

    AUTHOR INDEX

    SUBJECT INDEX

    INTRODUCTION

    General remarks. In the course of the last three decades, very considerable progress has been made in the investigation and theoretical interpretation of molecular spectra. As a result the study of molecular spectra has become one of the most important, perhaps the most important, means for investigating molecular structure.

    From the spectra the various discrete energy levels of a molecule can be derived directly. From these again, we can obtain detailed information about the motion of the electrons (electronic structure) and the vibration and rotation of the nuclei in the molecule. The study of electronic motions has led to a theoretical understanding of chemical valence. From the vibrational frequencies the forces between the atoms in the molecule, as well as the heats of dissociation of molecules, can be calculated with great accuracy. From the rotational frequencies we obtain accurate information about the geometrical arrangement of the nuclei in the molecule—in particular, extremely accurate values of the internuclear distances.

    The knowledge of the various properties of the individual molecules so obtained allows us to understand many of the physical and chemical properties of the gases under consideration and, in fact, in many instances allows us to predict these properties—for example, the specific heat and the paramagnetic susceptibility. Also, on the basis of this knowledge, chemical equilibria can be predicted theoretically with great accuracy and chemical elementary processes can be elucidated.

    A further important result of the investigation of molecular spectra is that proof has been obtained of the existence of a large number of molecules which were previously unknown in chemistry or were not thought capable of free existence. Among these are CH, OH, C2, He2, Na2, CP, and many others. The structures of these molecules have also been determined.

    The investigation of the spectra of diatomic molecules is also of great significance for the physics of atomic nuclei, since certain nuclear properties influence these spectra in a characteristic manner and can therefore be determined from them, and since, furthermore, in some cases rare isotopes may be detected by means of these spectra.

    Finally, in recent years, a new field of application of the knowledge of molecular spectra has opened up in astrophysics. Not only can the presence of various molecules on fixed stars, planets, and comets, in the upper atmosphere, and in interstellar space be detected on the basis of their spectra, but also from a more detailed analysis definite conclusions can be drawn concerning the physical conditions in these objects.

    This first volume deals with the spectra of diatomic molecules and the conclusions that can be drawn from them concerning the structure of these molecules. The spectra of polyatomic molecules will be taken up in the second and third volumes of this series. The former, already published (28),¹ deals with the infrared and Raman spectra; the latter will deal with the electronic spectra of polyatomic molecules. Since a thorough treatment of molecular spectra presumes a certain knowledge of atomic spectra and atomic structure, a short survey of the fundamentals of atomic theory is given in the first chapter of this volume.

    The experimental methods for the production and investigation of spectra will not be dealt with here. Reference should be made to Sawyer (4b), Harrison, Lord, and Loofbourow (4c), Meissner (4a), Baly (2), Brode (3), Jenkins and White (6b), and Kayser-Konen (1).

    The nomenclature of the various quantities concerning the molecule—energy levels, quantum numbers, and so on—was made subject to international regulation in 1930 [Mulliken (515)]. We shall use this international nomenclature throughout.

    TABLE 1. PHYSICAL CONSTANTS

    Physical constants. The best values of various fundamental physical constants have been the subject of much discussion. We accept here the values recently recommended by DuMond and Cohen (914) which are based on and substantially identical with the values derived by Birge (809-811), but we use the defined (thermochemical) calorie recommended by the Bureau of Standards [Mueller and Rossini (1252)] instead of the 15° calorie used by Birge. In Table 1 the values most frequently used in this book are collected. These values differ only very slightly from those used in the first edition of this volume which were taken from an earlier version of Birge’s list (107–110) and which are also used in Volume II.

    Units and conversion factors. The wave lengths, λ, of the spectral lines in the infrared are measured in μ (1μ = 1/1000 mm) and in the visible and ultra violet spectral regions in Ångström units (1 Å = 10−7 mm = 10−8 cm). In theoretical discussions the frequency ν′ = c/λ is much more important than the wave length. Instead of it usually the wave number ν = ν′/c = 1/λ is used which is proportional to the frequency and is measured in cm−1 (number of waves per cm). The wave numbers can be given to the same accuracy as the wave lengths and are not limited to the accuracy with which the velocity of light is known. However, in order to have a constant proportionality factor (1/cvac) it is necessary to use the wave number in vacuo, which is the reciprocal of the wave length in vacuo. Since the wave lengths of spectral lines above 2000 Å are usually given as measured in air, they have to be converted to vacuum [by adding (n – 1)λair where n is the index of refraction of air] before the reciprocal is taken. For the region 2000-10,000 Å this calculation can be avoided by using Kayser’s Tabelle der Schwingungszahlen (45).⁴ According to Babcock (762) the use of these tables can be extended to 5 μ by inverting them and adding a very small correction. For still longer wave lengths n – 1 is very nearly constant (= 2.720 × 10−4 at 15° C and 760 mm pressure) so that the vacuum correction is easily made. In infrared work carried out before 1940 the vacuum correction has usually been neglected. This must be kept in mind when molecular constants are calculated from infrared data.

    The frequency ν′, and the wave number, ν, depend on the energy, E, according to the fundamental Planck relation E = ′ = hνc, where h is Planck’s constant. We can therefore take the wave number as a measure of the energy, as is very frequently done in spectroscopy. Using the values in Table 1, we find that 1 cm−1 corresponds to 1.98554 × 10−16 erg/molecule. If we wish to refer the energy to 1 mol instead of to a single atom or molecule, we have to multiply by the number of molecules Nch in 1 mol. Therefore, 1 cm−1 corresponds to 11.9599 × 10⁷ ergs/mol (referred to the chemical atomic-weight scale) or, converting to calories, 2.85848 cal/mol.

    Another unit used in measuring energy in atomic and molecular physics is the (absolute) electron-volt (e.v.). By V e.v. is meant an energy equal to the kinetic energy of an electron (or singly charged ion) that has been accelerated through a potential difference of V volts. This energy is eV/299.776 ergs (the factor 1/299.776 is introduced in converting to electrostatic units). Substituting the numerical value for e gives 1 e.v. equivalent to 1.601996 × 10−12 erg/molecule, or 8068.32 cm−1, or 23,063.2 cal/mol. These and some other derived conversion factors for the energy units are collected in Table 2.

    TABLE 2. CONVERSION FACTORS FOR ENERGY UNITS.

    ¹ The numbers in parentheses refer to the bibliography, p. 583.

    ¹a In an erratum received too late for inclusion in this book, DuMond and Cohen give 6.62373 × 10−27 for h. The difference is smaller than the probable error.

    ² The gas constant is referred to the standard atmosphere 1013250 dynes/cm² as adopted by the International Commission on Weights and Measures [see Rossini (50b)] not 1013246 dynes/cm² as used by Birge and DuMond.

    ³ This defined calorie should be distinguished from the calorie of the international steam tables which is used by engineers and defined by 1 I.T. steam calorie = 4.18674 abs. joulex. The 15° calorie is given by 1 cal15 = 4.1855 abs. joules.

    ⁴ In doing so it must be remembered that Kayser’s tables were calculated with values for the index of refraction of air which are now outdated. For very precise calculations the more recent values of Barrell and Sears (770) should be used [see Birge (811)].

    CHAPTER I

    RÉSUMÉ OF THE ELEMENTS OF ATOMIC STRUCTURE¹

    1. Bohr Theory

    Stationary energy states. Niels Bohr was the first to suggest that an atom or molecule cannot exist in states having any arbitrary energy, but only in certain discrete energy states called stationary states. According to the Bohr theory, these stationary states are obtained when one selects, by means of certain quantum conditions, some of the states from the continuous range of classically possible states. For example, in the case of a single electron moving in the Coulomb field of an atomic nucleus, we obtain, classically, elliptical orbits of any half-axes. According to the Bohr theory, only certain of these, with quite definite values of the axes, are possible. For these selected orbits, the major axis is proportional to the square of a whole number n, called the principal quantum number while the minor axis is proportional (with the same proportionality factor) to the product of n and a second whole number l + 1, where l is the so-called azimuthal quantum number. While the principal quantum number can take any integral value greater than 0,

    the azimuthal quantum number l, for a given value of n, can take only the values

    The energy E of the atom contains an arbitrary additive constant. If this constant is chosen so that E = 0 when the electron is completely separated from the nucleus and when at the same time there is no relative kinetic energy, the following expression for the energy of the possible elliptical orbits (Bohr orbits) in the hydrogen atom or hydrogen-like ions is obtained (neglecting a very small relativity correction):

    In this expression, Z is the atomic number (for hydrogen, Z = 1) and μ is the reduced mass, μ = mM/(m + M), where m is the mass of the electron and M that of the nucleus. Thus, neglecting relativity effects, the energy depends only on the principal quantum number n. It is negative for all stationary states. The state with lowest energy is that with n = 1, corresponding to the innermost orbit. It is the ground state of the atom or ion. With increasing n, the energy increases and eventually approaches the limit E = 0.

    In the more general case of the motion of a single electron about an atomic core which is not a point charge, the orbits are no longer simple ellipses but, to a first approximation, ellipses whose axes rotate uniformly (rosette motion). As before, they can be characterized by the two quantum numbers n and l. But now the energy is no longer independent of l (even when relativity effects are disregarded). To a first approximation it is given by

    where a, the so-called Rydberg correction, depends on the azimuthal quantum number l and rapidly approaches zero with increasing l. Z is now the charge on the core—that is, the nuclear charge (atomic number) minus the number of core electrons.

    As an example, Fig. 1 gives a diagrammatic representation of the observed energy levels of the lithium atom, which has a single electron revolving about an atomic core consisting of the nucleus and two electrons. The energy levels are indicated by horizontal lines whose vertical height corresponds to the energy of the atom as given by the energy scales to the left and right. For each value of l there is a series of energy levels (each drawn in a separate column) which correspond to the different values of n and approach one another as n increases, finally going to the limit E = 0. In agreement with the relation (I, 2) the smallest value of n for a given l is l + 1 except for l = 0 where it is 2. The reason for this exception will be explained on p. 22. The energy level diagram of the hydrogen atom is quite similar, except that, according to (I, 3), all levels with the same n have the same height, and that the level with l = 0, n = 1 does occur.

    Radiation. According to Bohr, electromagnetic radiation is not emitted while an electron moves in its orbit (as it should be according to classical electro dynamics), but only when the electron goes from one quantum orbit of energy E1 to another of energy E2. This process is called a quantum jump. The liberated energy E1 – E2 is emitted as a light quantum (pholon) of energy ′ = hcν (ν′ = frequency, ν = wave number of the light) so that

    This is Bohr’s frequency condition. It holds for absorption as well as for emission. That is to say, a light quantum is absorbed by an atom only when the energy of the light quantum is precisely equal to the energy which is necessary to bring the atom from the state E2 to the state E1.

    From (I, 5) it follows that the wave number of the emitted or absorbed light is

    If we substitute the energy values (I, 3) of the hydrogen atom or of the hydrogen like ions (He+, Li++, · · ·) into this expression, we obtain

    where R = R′/hc = 2π²μe⁴/ch³ is the Rydberg constant and n2 and n1 are the principal quantum numbers of the two states concerned. The whole spectrum of the hydrogen atom (Z = 1) and of the hydrogen-like ions is represented with great accuracy by the simple equation (I, 7). When n2 = 2 and Z = 1, the formula for the well-known Balmer series of the hydrogen atom results. A spectrogram of this series is given in Fig. 6, p. 30. As n1 increases, the lines of this series draw rapidly together and go to the limit R/4 for n1 → ∞. The other series represented by equation (I, 7) behave correspondingly.

    FIG. 1. Energy Level Diagram of the Li Atom (after Grotrian). The wave lengths of the spectral lines are written on the sloping connecting lines, which represent the transitions (see be low). The symbols S, P,· · · are explained on p. 27. The numbers in front of these symbols are the principal quantum numbers n of the outer electron.

    In the more general case, using (I, 4), we obtain

    where the first and second terms now depend on l as well as n. Therefore a greater number of transitions results. For fixed values of n2, a2, and a1 and variable n1 we obtain a series of lines similar to the Balmer series illustrated in Fig. 6. Such series are sometimes called Rydberg series. In Fig. 1, in which the transitions are indicated by sloping lines joining the levels in question, a Rydberg series corresponds to sloping lines that lead from the levels of a series, having a definite value of l, to a fixed lower-lying level.

    It can be seen from Fig. 1 that transitions do not take place between any two levels, but only such transitions are observed for which the quantum number l changes by unity. Such a limitation is called a selection rule. In the present case it is written

    Transitions for which l does not alter or alters by 2, 3,· · · units do not occur and are said to be forbidden. In the Bohr theory, this selection rule can be derived with the help of the so-called correspondence principle.

    Terms. According to (I, 6), (I, 7), and (I, 8), the wave number ν of a spectral line can be represented as the difference between two quantities called terms. More precisely, the positive quantities RZ²/n² or RZ²/(n + a)² and similar, possibly more complicated, expressions arc designated as terms. Terms are thus the negative energy values divided by hc. A series of terms of the form RZ²/(n + a)² with n = 1, 2, 3, · · · is called a Rydberg series of terms.

    The terms can be obtained empirically from the observed spectrum. The first term in (I, 8), for example, is always given by the wave number of the limit of the series of spectral lines in question, while, for the second term, a has to be chosen so that all the observed lines of the series are accurately represented.

    Often, the word term is used synonymously with energy state or quantum state and the expression term value is used synonymously with energy value. This practice is even more justifiable in molecular than in atomic spectroscopy, since here a zero of energy differing from that for atoms is used almost exclusively—namely, the ground state of the molecule. Hence term value and energy value differ only by the factor 1/hc (without the minus sign). In the following we shall employ frequently this somewhat looser usage of the concept term.

    Continuous spectra. All stable, discrete quantum states of an atom with one outer electron, according to (I, 3) or (I, 4), have negative energy values if the usual zero-point of the energy scale is used. A positive value of E corresponds to an energy greater than that of a system consisting of an electron and an ion infinitely separated and at rest. That is to say, the two particles approach or separate from each other with a kinetic energy that is not zero even at infinity. The electron then moves in a hyperbolic orbit, through which it goes, of course, only once. Such aperiodic motions are not quantized. The kinetic energy of the two particles can take any value; that is to say, all positive values of E are possible. Hence, extending from the limit E = 0 of the series of discrete energy values of an atom to higher energy values, there is a continuous region of possible energy values (continuous term spectrum).

    Corresponding to the continuous region of energy values, there is a continuous absorption or emission spectrum, which arises when the electron jumps from a discrete state into the continuous region or from the latter into the former. The continuous absorption spectrum therefore corresponds to the removal of the electron (photoeffect) with more or less kinetic energy; the continuous emission spectrum corresponds to the capture of the electron. The continuum begins at the limit of the line series having the same principal quantum number n2 of the lower state and extends toward shorter wave lengths. The series limit, which is at the same time the starting point of the continuum, corresponds to the removal or capture of the electron with zero velocity; that is to say, it gives directly the energy required to separate the electron from the atom in the lower state of the line series (ionization potential).

    One can say quite generally that continuous spectra of gases always correspond to such splitting or recombination processes, since of all forms of molecular energy only the kinetic energy of translation is not quantized.

    2. Wave Mechanics (Quantum Mechanics)

    Fundamental equations. The essential point in the Bohr theory, the idea of discrete stationary stales, is retained in wave mechanics, but the derivation of these stationary states is altered. Moreover the wave-mechanical energy values for non-hydrogen-like systems agree quantitatively with experiment, while the values calculated according to the Bohr theory do not.

    Whereas the Bohr theory starts out from the classical laws of motion, and subsequently, by means of certain quantum conditions, selects as possible only a few of the orbits so obtained, wave mechanics puts in the place of the classical laws of motion another law, which is, however, so constituted that for large masses it goes over asymptotically into the classical laws of motion. Wave mechanics is therefore more comprehensive than classical mechanics, which fails for atomic dimensions.

    The first step toward wave mechanics was De Broglie’s fundamental idea that the motion of any corpuscle of matter is associated with a wave motion of wave length

    where m is the mass and ν the velocity of the corpuscle. This idea was later on brilliantly confirmed by the discovery of the diffraction of corpuscular rays (electrons as well as atoms), so that the existence of waves of matter (De Broglie waves) is now an established fact.

    Accordingly, when we want to investigate the motions in an atomic system we must investigate the wave motions associated with them. Let Ψ be the wave function, whose physical meaning will be discussed later. Since we are dealing with a wave motion, Ψ must vary periodically with time at every point in space. We can therefore write

    Ψ = ψ sin 2πν′t    or    Ψ = ψ cos 2πν′t,

    or, combining both together,

    Here ν′ is the frequency of the vibrations, and ψ is the amplitude of the wave motion.

    In the case of the motion of a single mass point of mass m, the amplitude ψ depends only on the coordinates x, y, and z. For this case Schrödinger put forward the equation

    where V is the potential energy and E = ′ is the total energy of the masspomt He added the rather natural restriction that the solutions ψ of this equation be everywhere single-valued, finite, and continuous and vanish at infinity. The Schrödinger equation (I, 12) together with this restriction replaces for atomic systems the fundamental equations of motion of classical mechanics. For the potential energy V the same expression is used as in classical theory, that is, the laws of force are unchanged.

    The mathematical investigation of equation (I, 12) shows that under the restrictions given, it is in general soluble only for certain definite values of E, the so-called eigenvalues, just as the differential equation of a vibrating string can be solved only for certain definite frequencies, those of the fundamental and the overtones. Thus the discrete energy values of an atomic system which are experimentally observed in the spectrum appear, according to Schrödinger, as the eigenvalues of the wave equation of the system.

    For the case of several (N) particles, (I, 12) must be replaced by the equation

    where ψ now depends on the 3N co-ordinates, xk, yk, zk, and where mk is the mass of the kth particle.

    If, in (I, 12), we substitute for V the potential energy (−Ze²/r) of an electron in the field of a fixed nucleus with a charge Ze, the solution of the wave equation leads to the energy values (I, 3) with m in place of μ. [see texts on wave mechanics (11a–16)]. If we drop the assumption that the nucleus is fixed, equation (I, 13) has to be used. The resulting energy values are then exactly those of (I, 3). On the other hand, when we are not dealing with the field of a point charge but with that of an atomic core, energy values of the form (I, 4) result. In both cases it is found that in addition to the discrete negative energy values (I, 3) or (I, 4) all positive values of E are possible; that is to say, a continuous region (E > 0) adjoins the series of discrete energy levels (E < 0) just as in the old Bohr theory.

    Even for systems with several electrons,—for example, for He and Li+—the solution of the wave equation leads to energy values that are not only qualitatively but also quantitatively in agreement with the empirical values. This is in contrast to the Bohr theory, which leads in these cases to values for the ionization potential and other term values that are entirely at variance with experiment.

    Instead of starting out from the amplitude equation (I, 12) or (I, 13) one may also use, and sometimes has to use, a differential equation for the time dependent wave function Ψ. This wave equation for the non-relativistic case of several particles is, according to Schrödinger,

    By substituting for Ψ a product of a function Ψ of the coordinates alone and a function of the time alone, one is immediately led to the relation (I, 11) and, with ν′ =E/h, to the amplitude equation (I, 13) [see for example Pauling and Wilson (13)].

    In more complicated problems it is often convenient to employ a form of the Schrödinger equation that makes use of a close formal analogy between classical and quantum-theoretical quantities. In the classical treatment of a conservative system the sum H of kinetic energy T and potential energy V of a dynamical system is a constant E:

    H, expressed in terms of momenta and coordinates (or generally in terms of canonically conjugate quantities) is called the Hamiltonian function. For example, for the case of several particles, using Cartesian coordinates we have

    where Pxk, Pyk, Pzk are the components of the momentum and xk, yk, zk the coordinates of particle k. If now the momenta in H are replaced by differential operators²

    and the whole equation is multiplied by ψ (where multiplication of an operator by ψ means that the differentiations implied in the operator are carried out on ψ) the Schrödinger equation (I, 13) for the amplitude ψ is obtained.

    Symbolically we can therefore write for the Schrödinger equation

    where H is the Hamiltonian operator, obtained from H of (I, 15) by the substitutions (I, 17). The usefulness of the form (I, 18) of the Schrödinger equation rests on the fact that it allows one to set up in a simple way the Schrödinger equation in terms of coordinates other than Cartesian coordinates. This is important since as a rule the potential energy can be expressed much more simply in terms of coordinates other than Cartesian coordinates. In such cases the kinetic energy has to be expressed in terms of generalized momenta pk which are canonically conjugate to the coordinates qk used, that is, so that

    where as usual

    If in (I, 18) the operator − is substituted for the factor E the time dependent Schrödinger equation [(I, 14) in the case of Cartesian coordinates] is obtained.

    Physical interpretation of the Ψ function. The complete solution of the Schrödinger equation [(I, 12) or (I, 13)] yields not only the eigenvalues—that is, the energy values of the stationary states—but also the corresponding functions ψ (or Ψ= ψe−2πi(E/h)t), the so-called eigenfunctions. Just as in the case of vibrating strings or membranes these eigenfunctions have a certain integral number of nodes or nodal surfaces. These integral numbers are the quantum numbers that occur in the energy formulae.

    While in the case of mechanical vibrations the wave function simply gives the displacement from the equilibrium position for every point of the system the physical interpretation of Schrödinger’s Ψ function is less obvious: As was first postulated by Born ΨΨ* dτ = |Ψ|² dτ = |ψ|² (where Ψ* is the complex conjugate of Ψ) gives the probability that the particle under consideration will be found in the volume element dτ at the position given by the co-ordinates. Consequently |ψ|² is the probability density. Its dependence on the co-ordinates —that is, the probability density distribution of the electron or electrons in an atom—is the essential thing that wave mechanics has to tell us about the motions in an atom. (See A.A., Fig. 21, p. 44). However, it tells us nothing about the orbits of the electrons. The eigenfunction ψ itself is sometimes referred to as the probability amplitude.

    The electron may be observed at any point in space with a certain probability; it is, so to speak, smeared out over the whole of space. However, the ψ function decreases exponentially with increasing distance from the nucleus, and the probability of finding the electron appreciably outside the region of the old Bohr orbits is extremely small.

    On the other hand, in the states of positive energy, the eigenfunction is an outgoing or incoming spherical wave corresponding to the removal or capture of the electron. Here the decrease in |ψ|² is only inversely proportional to the square of the distance from the center and not exponential.

    Frequently there are several different eigenfunctions associated with one and the same eigenvalue. In that case we have a degenerate state. The de generacy is d-fold if there are d linearly independent eigenfunctions for a given eigenvalue. These d eigenfunctions are not uniquely determined by the wave equation since, on account of the linearity of the latter, any linear combination of the eigenfunctions belongs also to the same eigenvalue.

    The time dependent Schrödinger equation (I, 14) is fulfilled also by general solutions of the form

    where the ψn are eigenfunctions of the amplitude equation and where the cn are constant coefficients. Such a wave function according to the statistical interpretation tells us that the system under consideration may be in any of the states of energy En, and that the probability of finding it in the state En is cncn*.

    Orthogonality and normalization of eigenfunctions. It can be shown [see, for example, Pauling and Wilson (13)] that any two eigenfunctions of an atomic system belonging to different eigenvalues are orthogonal to each other, that is,

    where n and m represent the quantum numbers of two different states, and where the integration extends over the whole configuration space.

    While degenerate eigenfunctions belonging to a given eigenvalue are not necessarily orthogonal to one another it is always possible to find a set of linearly independent functions belonging to this eigenvalue which are mutually orthogonal.

    If an eigenfunction is multiplied by a constant factor the function obtained is still an eigenfunction belonging to the same eigenvalue. For many purposes it is convenient to multiply the eigenfunctions by such factors that for all of them

    The eigenfunctions are then said to be normalized to unity.

    If, in addition, the wave function Ψ of (I, 20) is normalized to unity it is easily seen that for the coefficients cn

    Perturbation theory. Only in very few cases can the wave equation be solved rigorously. Usually, particularly for molecular problems, we must start out from a wave equation in which certain interactions have been neglected. Let (I, 13) with a certain potential energy V be the approximate (unperturbed) Schrödinger equation and let

    be the Schrödinger equation of the perturbed system where is the perturbation term representing the originally neglected interactions. W may be an additional term in the potential energy or it may be an operator arising from an additional term in the kinetic energy (see p. 11).

    Considering first the case of non-degenerate states we shall let the eigenvalues and eigenfunctions of the unperturbed Schrödinger equation (I, 13) be

    and those of the perturbed equation (I, 24)

    According to wave-mechanical perturbation theory these true eigenvalues and eigenfunctions can be expressed in terms of the unperturbed eigenvalues and eigenfunctions and the perturbation function. It is found that, up to the second approximation, the eigenvalues are [see, for example, Pauling and Wilson (13)].

    In this equation

    are the so-called matrix elements of the perturbation function W. To a first approximation, perturbation theory gives for the true eigenfunctions³

    It can be seen immediately from (I, 27 and 29) (and it is also plausible) that for a sufficiently small perturbation function W the true eigenvalues and eigenfunctions lie very close to the approximate eigenvalues and eigenfunctions. The deviations depend on the eigenfunctions and eigenvalues of all the other states of the system as well as on the magnitude of the perturbation function. This state of affairs may also be expressed by saying that the perturbation causes a mutual interaction of the unperturbed energy levels: Each level produces a shift of any other and at the same; time contributes to the eigenfunction of the other; both effects increase with decreasing separation of the two unperturbed levels, and of course, the larger the perturbation function W, that is, the poorer the zero-order approximation, the larger will be these effects. It is easily seen that the shift of produced by , is opposite and equal to that of produced by , and that the direction of the shift is always such that the separation of the two levels is increased. In other words there is an apparent repulsion of any two states which is stronger the closer they are together. However, it must be realized that, this repulsion affects only the unperturbed energy levels. The actual energy levels do not, of course, affect one another.

    The formulae (I, 27) and (I, 29) hold only if the separation of and is not too small or zero. If it is zero we have degeneracy and the introduction of the perturbation will cause a splitting of this degeneracy. We consider only the case of two-fold degeneracy. Let E⁰ be the zero order energy value and and two linearly independent, mutually orthogonal, and normalized eigenfunctions. With decreasing strength of the perturbation the true eigenfunctions will not in general approach and but will approach two linear combinations of these:

    These zero-order eigenfunctions are often sufficiently good approximations to the true eigenfunctions. The coefficients cih are found to be determined by the equations

    and

    where the Wik are given as before by (I, 28) and where εa = Eb E⁰ and εb = Eb E⁰ are the shifts of the perturbed energy levels from the unperturbed position (in first-order approximation). The two equations (I, 31) give non-zero solutions for c11 and c12, and similarly (I, 32) for c21 and c22, only if

    The two roots εa and εb, of this secular equation added to E⁰ yield the first order energy values Ea and Eb which are thus immediately given as soon as the matrix elements Wik have been evaluated. Substituting εa and εb into (I, 31) and I, 32) gives relative values of the coefficients cik. The absolute values are obtained by normalizing the functions and

    Heisenberg’s uncertainty principle. The fact that, in wave mechanics, the orbits of the old Bohr theory have lost their meaning is closely connected with the Heisenberg uncertainty principle. According to it, the position and momentum (or velocity) of a particle cannot be simultaneously measured with any desired accuracy. Rather, if Δx is the uncertainty of the x co-ordinate and Δpx is the uncertainty of the x component of the momentum (mΔvx), always

    where h is Planck’s constant. Analogous relations hold for the other co ordinates, and also, for example, for energy and time:

    These uncertainty relations are a direct result of the wave properties of matter (see A.A., Chapter I) and represent a limitation in principle of the accuracy of measurements in atomic systems.

    Momentum and angular momentum. Even though in wave mechanics the orbits of the electron cannot be given—that is to say, the momentum is not defined for every point—yet a probability density distribution of the momentum can be given (see A.A., p. 47). Quite definite statements may be made regarding the angular momentum of an atom. For any mechanical system the angular momentum is given by where ρi is the perpendicular distance of the momentum mivi of the ith particle from a fixed point, usually the center of mass. In wave mechanics, as in classical mechanics, the angular momentum of an isolated system is constant.⁴ According to wave mechanics, it can take only discrete values as can the energy; in the case of one electron, it can have the values , or, approximately l(h/2π). The azimuthal quantum number l gives us therefore the orbital angular momentum of the electron in units h/2π. Consequently, the various term series of lithium, for example (see Fig. 1), differ in the angular momentum of the electron.

    The fact that even in wave mechanics the angular momentum has a perfectly definite value shows that we can still speak of an electron revolving about an atomic nucleus, provided that we are quite clear that there is no point in speaking of definite orbits. Consequently, we can in many instances use the angular momentum in the same way as in the Bohr theory; for example, under certain conditions, angular momenta can be added vectorially (see section 3).

    Throughout this book we shall use bold face letters for angular momentum vectors and thus distinguish them from the corresponding quantum numbers, printed in regular type. Thus l means an angular momentum vector of magnitude

    It should be noticed that even in the continuous states of positive energy the angular momentum l of the electron can have only the discrete values given above.

    In wave mechanics the classical dynamical variables are replaced by operators. Classically, the components of the angular momentum in the case of a single particle are

    which upon substitution of (I, 17) yields the corresponding operators

    or, introducing spherical polar coordinates, r, , φ, [see Rojansky (15b)]

    The square of the magnitude of the angular momentum is therefore represented by the operator

    In quantum mechanics the possible values of a dynamical variable are the eigenvalues of the corresponding operator; that is, the possible values of the square of the magnitude of the angular momentum are the eigenvalues a of the equation

    It is found that these eigenvalues are

    which leads to the value for the magnitude of the angular momentum |P| given above. The eigenfunctions of the Hamiltonian operator [equation (I, 18)] are also eigenfunctions of the angular momentum operator. That is why in every stationary state (characterized by a certain eigenfunction) the atomic system considered has not only a definite energy but also a definite angular momentum.

    Space quantization. According to classical theory an atom brought into a magnetic or electric field F will carry out a precession such that the angular momentum l describes a cone with the field direction as axis and with a constant component ml of the angular momentum as shown diagrammatically in Fig. 2.

    FIG. 2. Precession of the Angular Momentum l of an Electron in a Magnetic or an Electric Field F.

    FIG. 3. Space Quantization of l in a Field F for l = 3. The figure also holds generally for other angular momentum vectors (see p. 26). The length of the vector l is = 3.464 units.

    In quantum theory, too, the component of the angular momentum in the field direction is constant. However, while classically ml can have any value between + |l| and — |l| (that is, any angle between l and the field is possible) in quantum theory only the discrete values ml (h/2π) are possible where

    that is, only certain angles of l with respect to the field direction are possible. This space quantization is shown in Fig. 3 for l = 3. The number ml is called the magnetic quantum number of the electron. Since the magnitude of the vector l is which is larger than the largest value of ml(h/2π) it is clear that l can never point exactly in the field direction as it could in the old Bohr theory, a fact that is intimately connected with the Heisenberg uncertainty relation (see A.A. p. 102). The space quantization persists and the magnetic quantum number remains defined even for vanishingly small field. For zero field the 2l + 1 values of ml in (I, 42) for a given n and l correspond to 2l + 1 different modes of motion (eigenfunctions) of the same energy. We have a (2l + 1)-fold degeneracy.

    The component of the angular momentum in the field direction corresponds to the operator Pz of (I, 38). It can be shown that the eigenfunctions of the Hamiltonian operator are simultaneously eigenfunctions not only of the angular momentum operator P² but also of Pz. Instead of Pz any other component in a fixed direction may be chosen but not more than one. The eigenvalues of Pz are ml(h, 2π) with ml given by (I, 42).

    Electron spin. Investigations of the multiplet structure of spectral lines (for example, the alkali doub’ets) and of the anomalous Zeeman effect have led, at first quite independently of quantum mechanics, to the assumption (Goudsmit and Uhlenbeck) that the electron has an angular momentum of its own, called electron spin; that is to say, the electron rotates about its own axis. The corresponding quantum number is s = , and the angular momentum itself, s, has the magnitude The electron spin appears as a necessary result of Dime’s relativistic wave mechanics. [See texts on wave mechanics (1la–16).]

    Dirac showed that in a magnetic field, the spin of the electron can set itself only so that its component in the field direction has the value ms(h/2π), where m8 can be only + or − . Without a field there is a two-fold degeneracy. But whenever there is a non-zero orbital angular momentum, an internal magnetic field arises which causes a splitting of the double degeneracy even without an external magnetic field (see also below).

    Emission and absorption of radiation. If the interaction with an electro magnetic field is introduced into the Schrödinger equation of an atomic system, it is found that a non-zero probability arises of finding the system in a state En if originally it was in the state Em, and if radiation of wave number ν = (En Em)/hc is present. If En < Em, radiation of this wave number is emitted by the atomic system. Thus Bohr’s frequency condition (I, 6) holds also in wave mechanics. The frequencies of the emitted or absorbed spectral lines are determined in this simple fashion by the eigenvalues of the Schrödinger equation—that is, by the energy values of the system.

    The probabilities of the transitions under the influence of radiation are determined by the eigenfunctions of the states involved. Thus the intensities of the emitted or absorbed spectral lines can be obtained. In particular, when the eigenfunctions are known, we can calculate whether or not two states can combine with each other; that is to say, we obtain the selection rules.

    The interaction of an electromagnetic wave having an electric vector E, with an atomic system is in a first approximation the interaction with the (variable) electric dipole moment M of the system whose components are

    where the ek are the charges on the N particles of coordinates xk, yk, zk If the interaction energy M · E is introduced into the wave equation [see for example Pauling and Wilson (13)] it is found that the probability of a transition between two states n and m produced by the interaction is proportional to the square of the magnitude of certain vector quantities Rnm– the matrix elements of the electric dipole moment—whose components depend in the following manner on the eigenfunctions ψn and ψm of the two states:

    The integrals are to be taken over the whole configuration space of the 3N coordinates. If the matrix element Rnm (transition moment) differs from zero for two states n and m, the two states combine with each other with a certain probability with emission or absorption of radiation; if it is zero the transition under consideration is forbidden as a dipole transition.

    For example in the case of a one-electron system it is found that all Rnm vanish with the exception of those for which the quantum number l for the two states differs by only 1; that is to say, we have the selection rule

    This is the same selection rule as that already given above in the Bohr theory [equation (I, 9)] and found to agree with experiment.

    If one puts n = m in (1, 44) one obtains the permanent dipole moment of the system, which of course equals zero in the case of an atom, but may be different from zero for a molecule.

    A non-zero transition probability may also be produced by the interaction of an electromagnetic wave with the magnetic dipole moment, the quadrupole moment, or higher moments of the atomic system; or in the process of scattering

    Enjoying the preview?
    Page 1 of 1