17471-Introduction To Atomic Spectra - Text
17471-Introduction To Atomic Spectra - Text
$<E>oC=3<
LIBRARY §
Kashmiri Gate, Delhi-110006 |
Accession No.
\oArt*> '• ■I
Book No._
DATE DUE
For each day's delay after the due date a fine of
10 Pafse per Vol. shall be charged for the first week, and
50 Paise per Vol. per day for subsequent days.
INTERNATIONAL SERIES IN
PURE AND APPLIED PHYSICS
G. P. HARNWELL, Consulting Editor
The late P. K. Richtmyer was Consulting Editor of the series from its inception in
1929 to his death in 1939. Lee A. DuBridge was Consulting Editor from 1939 to
1946, and G. P. Harnwell from 1947 to 1954.
INTRODUCTION
TO
ATOMIC SPECTRA
BY
TOKYO. JAPAN
TO MY WIFE
PREFACE
During the past few years students have frequently come to the author
with this query, “ Where can I find an elementary treatment of atomic
spectra?” Further questioning has shown, in general, that, although
these students have had at least one year of calculus and one or more years
of college physics, they have found most treatments of atomic spectra
either too brief or too highly mathematical. The desire to meet the
situation has given the author the incentive and the encouragement
to write this book.
In preparing a manuscript that will in some measure meet the desires
of beginning students, as well as those already familiar with certain
phases of atomic spectra, the author set up the following three objectives:
first, to start as nearly as possible at the beginning of each subject;
second, to develop each new concept so that the student with a working
knowledge of elementary physics and elementary calculus should have
little difficulty in following; and third, clearly to illustrate each chapter, as
far as possible, with diagrams and photographs of actual spectra.
At the outset one is confronted with the problem of describing a field
of scientific investigation that has developed, and still is developing,
at a rapid rate. On the one hand we have the Bohr-Sommerfeld theory
of an orbital atom and on the other the newer and more satisfactory
theory of quantum mechanics. Believing that one can better under¬
stand the principles of quantum mechanics by first becoming well
acquainted with the observed spectra and with the vector model, the
orbital model of the atom is first treated in some detail. Once developed
the orbital model then furnishes a very easy step to the quantum-
mechanical atom as given by Schrodinger’s wave equation or by the
Dirac electron. At frequent intervals throughout the* book both the old
and new models, of orbits and probability density functions, are compared
with each other.
The treatment as it is given here starts with the historical background
of spectroscopy (Chap. I) and develops in the first ten chapters the old
and the new quantum theories of one-valence-electron atoms. The
second half of the book (Chaps. XI through XXI) deals with complex
atomic systems of two or more valence electrons and includes a brief
account of x-ray spectra. Believing that frequent comparison with the
observed spectral lines, by giving actual reproductions of spectrograms,
is all too often neglected, the author has spent a great deal of time in
photographing many of the spectra used as illustrations. That photo
vii
viii PREFACE
CHAPTER I
Early Historical Developments in Atomic Spectra. 1
1.1. Kirchhoff's Law. 1.2. A New Era. 1.3. Balmer's Law. 1.4. Ryd¬
berg's Contributions. 1.5. The Rydberg-Schuster Law. 1.6. Series Nota¬
tion. 1.7. Satellites and Fine Structure. 1.8. The Lyman, Balmer,
Paschen, Brackett, and Pfund Series of Hydrogen. 1.9. The Ritz Com¬
bination Principle. 1.10. The Ritz Formula. 1.11. The Hicks Formula.
1.12. Series Formulas Applied to the Alkali Metals. 1.13. Neon with 130
Series. 1.14. Normal and Abnormal Series. 1.15. Hydrogen and the
Pickering Series. 1.16. Enhanced Lines.
CHAPTER II
Introduction to the Quantum Theory and the Bohr Atom.23
2.1. Planck's Simple Harmonic Oscillator. 2.2. The Bohr Atom. 2.3.
Bohr's First Assumption. 2.4. Bohr’s Second Assumption. 2.5. Bohr’s
Third Assumption. 2.6. Characteristics of the Bohr Circular Orbits. 2.7.
Bohr Orbits and the Hydrogen Series. 2.8. Series of Ionized Helium, He II
(IIe+). 2.9. Series of Doubly Ionized Lithium, Li III (Li++) and Triply
Ionized Beryllium, Be IV (Be+++). 2.10. Energy Level Diagrams. 2.11.
Unquantized States and Continuous Spectra. 2.12. The So-called “Reduced
Mass” of the Electron. 2.13. Variation of the Rydberg Constant. 2.14.
Bohr’s Correspondence Theorem.
CHAPTER III
Sommerfeld’s Elliptic Orbits and Space Quantization.42
3.1. Two Degrees of Freedom. 3.2. The Radial Quantum Number. 3.3.
The Total Energy W. 3.4. General Characteristics of Sommerfeld’s
Elliptic Orbits. 3.5. Space Quantization. 3.6. Larmor’s Theorem. 3.7.
Magnetic Moment and the Bohr Magneton.
CHAPTER IV
Quantum Mechanics and the SchrOdinoer Wave Equation.54
4.1. De Broglie's Corpuscular Wave Equation. 4.2. The Schrodinger Wave
Equation. 4.3. Schrodinger's Wave Equation Applied to Hydrogen.
4.4. Eigenfunctions. 4.5. The <p Factor <t»m of the Eigenfunction if/. 4.6.
The 0 Factor 6m,i of the Eigenfunction \J/. 4.7. Correlation of and
0»»,i9w,i with the Bohr-Sommerfeld Orbits. 4.8. The Radial Factor Rnti of
the Eigenfunction 4.9. Correlation of Rn,iRt,i with the Bohr-Sommerfeld
Orbits. 4.10. A General Interpretation of the Eigenfunction 4.11. Use¬
ful Atomic Models. 4.12. Spherical Symmetry.
CHAPTER V
The Alkali Metals and the Periodic Table.77
5.1. Energy Level Diagrams. 5.2. The Bohr-Stoner Scheme of the Building
Up of the Elements. 5.3. The First Period. 5.4. The Second Period. 5.5.
LX
X CONTENTS
Pagi
The Third Period. 5.6. The Fourth Period or First Long Period. 5.7. The
Fifth Period or Second Long Period. 5.8. The Sixth Period or Third Long
Period. 5.9. The Seventh and Last Period. 5.10. Energy Levels of the
Alkali Metals. 5.11. The Effective Quantum Number and the Quantum
Defect. 5.12. The Selection Principle.
CHAPTER VI
Excitation Potentials, Ionization Potentials, and the Spectra op Ionized
Atoms. .92
6.1. Critical Potentials. 6.2. The Spectra of Ionized Atoms.
CHAPTER VII
Penetrating and Nonpenetrating Orbits in the Alkali Metals.100
7.1. The Quantum-mechanical Model of the Alkali Metals. 7.2. Penetrat¬
ing and Nonpenetrating Orbits. 7.3. Nonpenetrating Orbits. 7.4. Pene¬
trating Orbits on the Classical Model. 7.5. Quantum-mechanical Model
for Penetrating Orbits.
CHAPTER VIII
Doublet Fine Structure and the Spinning Electron.114
8.1. Observed Doublet Fine Structure in the Alkali Metals and the Boron
Group of Elements. 8.2. Selection Rules for Doublets. 8.3. Intensity
Rules for Fine-structure Doublets. 8.4. The Spinning Electron and the
Vector Model. 8.5. The Normal Order of Fine-structure Doublets. 8.6.
Electron Spin-orbit Interaction. 8.7. Spin-orbit Interaction for Non¬
penetrating Orbits. 8.8. Spin-orbit Interaction for Penetrating Orbits.
CHAPTER IX
Hydrogen Fine Structure and the Dirac Electron.132
9.1. Sommerfeld Relativity Correction. 9.2. Fine Structure and the Spin¬
ning Electron. 9.3. Observed Hydrogen Fine Structure. 9.4. Fine Struc¬
ture of the Ionized Helium Line X4686. 9.5. The Dirac Electron and the
Hydrogen Atom. 9.6. The Angular Distribution of the Probability Den¬
sity P$. 9.7. The Radial Distribution of the Probability Density Pr. 9.8.
The Probability Density Distribution W*. 9.9. The Sommerfeld Formula
from Dirac’s Theory.
CHAPTER X
The Zeeman Eppect and the Paschen-back Effect.149
10.1. Early Discoveries and Developments. 10.2 The Vector Model of a
One-electron System in a Weak Magnetic Field. 10.3. The Magnetic
Moment of a Bound Electron. 10.4. Magnetic Interaction Energy. 10.5
Selection Rules. 10.6. Intensity Rules. 10.7. The Paschen-Back Effect.
10.8. Paschen-Back Effect of a Principal-series Doublet. 10.9. Selection
Rules forthe Paschen-Back Effect. 10.10. The Zeeman Effect, and Paschen-
Back Effect, of Hydrogen. 10.11. A Quantum-mechanical Model of the
Atom in a Strong Magnetic Field.
CHAPTER XI
Singlet and Triplet Series op Two-valence-electron Systems.171
11.1. General Series Relations. 11.2. Triplet Fine Structure. 11.3. The
Quantum Numbers n and l of Both Valence Electrons. 11.4. Penetrating
and Nonpenetrating Electrons for Two-electron Systems. 11.5. The Excita¬
tion of Both Valence Electrons.
CONTENTS xi
Pag*
CHAPTER XII
The Atom Model fob Two Valence Electrons.184
12.1. U-coupling. 12.2. Spin-spin-, or 88-coupling. 12.3. L&-, or Russell-
Saunders Coupling. 12.4. The Pauli Exclusion Principle. 12.5. Triplet
Multiplets in Ionized Scandium, Sc II. 12.6. Coupling Schemes for Two
Electrons. 12.7. T Factors for L>S-coupling. 12.8. The Land6 Interval
Rule. 12.9. ^/-coupling. 12.10. ^-coupling in the Carbon Group of
Elements. 12.11. Term Series and Limits in Two-electron Systems. 12.12.
The Great Calcium Triad. 12.13. The Branching Rule. 12.14. Selection
Rules. 12.15. Intensity Relations. 12 16. Relative Intensities of Related
Multiplets. 12.17. Helium and Helium-like Atoms. 12.18. Quantum
mechanical Model of Helium. 12.19. Fine Structure of Helium-like Atoms.
CHAPTER XIII
Zeeman Effect, Paschen-Back Effect, and the Pauli Exclusion Principle,
for Two Electrons.215
13.1. The Magnetic Moment of the Atom. 13.2. The Zeeman Effect.
13.3. Intensity Rules for the Zeeman Effect. 13.4. The Calculation of
Zeeman Patterns. 13.5. LS- and jj-coupling and the g Sum Rule. 13.6.
Paschen-Back Effect. 13.7. //^-coupling and the Paschen-Back Effect.
13.8. jj-coupling and the Paschen-Back Effect. 13.9. Complete Paschen-
Back Effect. 13.10 Breit’s Scheme for the Derivation of Spectral Terms
from Magnetic Quantum Numbers. 13.11. The Pauli Exclusion Principle.
13 12. Pauli’s g Permanence Rule. 13.13. Pauli’s g Sum Rule for All Field
Strengths. 13.14. Land6’s T Permanence Rule. 13.15. Goudsmit’s T
Sum Rule. 13.16. The T Permanence and V Sum Rules Applied to Two
Equivalent Electrons.
CHAPTER XIV
Complex Spectra.248
14 1. The Displacement Law. 14.2. Alternation Law of Multiplicities.
14 3. The Vector Model for Three or More Valence Electrons. 14.4. Terms
Arising from Three or More Equivalent Electrons. 14.5. The Land6
Interval Rule. 14.6. Inverted Terms. 14.7. Hund’s Rule. 14.8. The
Nitrogen Atom. 14.9. The Scandium Atom. 14.10. The Oxygen Atom.
14.11. The Titanium Atom. 14.12. The Manganese Atom. 14.13. The
Rare-gas Atoms, Neon, Argon, Krypton, and Xenon. 14.14. The Normal
States of the Elements in the First and Second Long Periods. 14.15.
Houston’s Treatment of One s Electron and One Arbitrary Electron. 14.16.
Slater’s Multiplet Relations. 14.17. Multiplet Relations of Goudsmit and
Inglis. 14.18. Relative Intensities of Multiplet Lines.
CHAPTER XV
The Zeeman Effect and Magnetic Quantum Numbers in Complex Spectra 286
15.1. Magnetic Energy and the Land6 g Factor. 15.2. The Calculation of
Zeeman Patterns. 15.3. Intensity Rules and Zeeman Patterns for Quartets,
Quintets, and Sextets. 15.4. Paschen-Back Effect in Complex Spectra.
15.5. Derivation of Spectral Terms by Use of Magnetic Quantum Numbere.
15.6. Equivalent Electrons and the Pauli Exclusion Principle.
CHAPTER XVI
X-Ray Spectra.299
16.1. The Nature of X-rays. 16.2. X-ray Emission Spectra and the Moseley
Law. 16.3. AbsorDtion Spectra. 16.4. Energy Levels. 16.5. Seieeti<»*
xii CONTENTS
P\Q*
and Intensity Rules. 16.6. Fine Structure of X-rays. 16.7. Spin-relativ¬
ity Doublets (Regular Doublets). 16.8. The Regular-doublet Law. 16.9.
Screening Doublets and the Irregular-doublet Law. 16.10. A Predicted
Structure in X-rays. 16.11. X-ray Satellites. 16.12. Explanation of X-ray
Absorption Spectra.
CHAPTER XVII
ISOELECTRONIC SEQUENCES.331
17.1. Isoelectronic Sequences of Atoms Containing One Valence Electron.
17.2. Optical Doublets and the Irregular-doublet Law. 17.3. Optical
Doublets and the Regular-doublet Law. 17.4. Isoelectronic Sequences of
Atoms Containing Two Valence Electrons. 17.5. Isoelectronic Sequences
Containing Three or More Valence Electrons. 17.6. The Irregular-
doublet Law in Complex Spectra. 17.7. Energy Relations for the Same
Atom in Different Stages of Ionization. 17.8. Centroid Diagrams.
CHAPTER XVIII
Hyperfine Structure. .C.r2
18.1. Introduction. 18.2. Hyperfine Structure and the Land6 Interval
Rule. 18.3. Nuclear Interaction with One Valence Electron. 18.4.
Nuclear Interaction with a Penetrating Electron. 18.5. Classical Explana¬
tion of Normal and Inverted Hfs. 18.6. Hyperfine Structure in Atoms with
Two or more Valence Electrons. 18.7. Hyperfine Structure in Complex
Spectra. 18.8. Nuclear gt Factors. 18.9. Zeeman Effect in Hyperfine
Structure. 18.10. Back-Goudsmit Effect in Hyperfine Structure 18.11.
Isotope Structure. 18.12. Isotope Structure and Hyperfine Structure
Combined.
CHAPTER XIX
Series Perturbations and Autoionization. 386
19.1. Observed Abnormal Series. 19.2. Energy Level Perturbations. 19.3
The Nature of and Conditions for Term Perturbations. 19.4. The Anoma¬
lous Diffuse Series of Calcium. 19.5. The Anomalous Principal Series in
Copper. 19.6. The Inverted Alkali Doublets. 19.7. Autoionization. 19.8.
Autoionization in Copper. 19.9. Autoionization in Calcium, Strontium,
and Barium. 19.10. Hyperfine-structure Term Perturbations.
CHAPTER XX
The Stark Effect. .401
20.1. Discovery of the Stark Effect. 20.2. The Stark Effect of Hydrogen.
20.3. Early Orbital Model of Hydrogen in an Electric Field. 20.4. Weak-
field Stark Effect in Hydrogen. 20.5. Strong-field Stark Effect in Hydrogen.
20.6. Second-order Stark Effect in Hydrogen. 20 7 Stark Effect for More
than One Electron. 20.8. The Stark Effect in Helium.
CHAPTER XXI
The Breadth of Spectrum Lines.418
21.1. The Doppler Effect. 21.2. Natural Breadths from Classical Theory.
21.3. Natural Breadths and the Quantum Mechanics. 21.4. Observed
Natural Breadths and Doppler Broadening. 21.5. Collision Damping.
21.6. Asymmetry and Pressure Shift. 21.7 Stark Broadening.
Appendix.437
Index .443
INTRODUCTION
TO ATOMIC SPECTRA
CHAPTER I
ful arc was first allowed to pass through a sodium flame just in front
of the slit of a spectroscope, two black lines appeared in exactly the same
position of the spectrum as the two D lines of the sun’s spectrum. Not
many years passed before evidence of this kind proved beyond doubt
that many of the elements found on the earth were to be found also
in the sun. Kirchhoff1 was not long in coming forward with the theory
that the sun is surrounded by layers of gases acting as absorbing screens
for the bright lines emitted from the hot surfaces beneath.
Fraunhofer Lines
* D
In the year 1859 Kirchhoff gave, in papers read before the Berlin
Academy of Sciences, a mathematical and experimental proof of the
following law: The ratio between the powers of emission and the powers of
absorption for rays of the same wave-length is constant for all bodies at the
same temperature. To this law, which goes under Kirchhoff’s name, the
following corollaries are to be added: (1) The rays emitted by a substance
excited in some way or another depend upon the substance and the tem¬
perature; and (2) every substance has a power of absorption which is a maxi¬
mum for the rays it tends to emit. The impetus KirchhofFs work gave to the
field of spectroscopy was soon felt, for it brought many investigators
into the field.
In 1868 Angstrom2 set about making accurate measurements of the
solar lines and published an elaborate map of the sun’s spectrum. Ang¬
strom’s map, covering the visible region of the spectrum, stood for a
number of years as a standard source of wave-lengths. Every line to be
used as standard was given to ten-millionths of a millimeter.
1.2. A New Era.—The year 1882 marks the beginning of a new era in
the analysis of spectra. Realizing that a good grating is essential to
accurate wave-length measurements, Rowland constructed a ruling
engine and began ruling good gratings. So successful was Rowland3 in
this undertaking that within a few years he published a photographic
map of the solar spectrum some fifty feet in length. Reproductions
from two sections of this map are given in Fig. 1.2, showing the sodium
1 Kirchhoff, G., Monatsber. Berl. Akad. Wiss., 1859, p. 662; Pogg. Ann.j 109,
148, 275, 1860; Ann. chim. et phys., 58, 254, 1860; 59, 124, I860.
* Angstrom, A. J., Uppsala, W. Schultz, 1868.
* Rowland, H. A., Johns Hopkins Univ. Circ. 17, 1882; Phil. Mag., 18,469,1882;
Nature, 26, 211, 1882.
Sec. 1.2] EARLY HISTORICAL DEVELOPMENTS 3
D lines, the iron E lines, and the ionized calcium H lines. With a wave¬
length scale above, the lines, as can be seen in the figure, were given to
ten millionths of a millimeter, a convenient unit of length introduced by
Angstrom and now called the Angstrom unit The Angstrom unit
Hydrogen
-1---1-1_i_■
00004 0 0005 Q0006 Millimeters
4000 5000 6000 Angstroms
Fig. 1.3.—The Balmer series of hydrogen.
hydrogen, pointed out that the first, second, and fourth lines were the
twentieth, twenty-seventh, and thirty-second harmonics of a fundamental
vibration#whose wave-length in vacuo is 131, 274.14 A. Ten years later
Schuster2 discredited this hypothesis by showing that such a coincidence
is no more than would be expected by chance.
1 Stoney, G. J., Phil. Mag.} 41, 291, 1871.
* Schuster, A., Proc. Roy. Soc. Itmdon, 31, 337, 1881.
4 INTRODUCTION TO ATOMIC SPECTRA [Chap. I
■■»SHH■■HB
Not
Fiq. 1.4.—Schematic representation of the sodium and potassium series. (After Liveing
and Dewar.)
Dewar say that while the wave-lengths of the fifth, seventh, and eleventh
doublets of sodium were very nearly as ^ the whole series cannot
be represented as harmonics of one fundamental. Somewhat similar
harmonic relations were found in potassium but again no more than would
be expected by chance.
Four years later Hartley2 discovered that the components of a
doublet or triplet series have the same separations when measured
in terms of frequencies instead of wave-lengths. This is now known as
Hartley’s law. This same year Liveing and Dewar3 announced their
discovery of series in thallium, zinc, and aluminum.4
1.3. Balmer’s Law.—By 1885 the hydrogen series, as observed in the
spectra of certain types of stars, had been extended to 14 lines. Photo¬
graphs of the hydrogen spectrum are given in Fig. 1.5. This year is
significant in the history of spectrum analysis for at this early date
Balmer announced the law of the entire hydrogen series. He showed
that, within the limits of experimental error, each line of the series is
given by the simple relation
(1.1)
where h = 3645.6 k and ni and n2 are small integers. The best agree¬
ment for the whole series was obtained by making nx = 2 throughout
1 Liveing, G. D., and J. Dewar, Proc. Roy. Soc. London, 29, 398, 1879.
* Hartley, W. N., Jour. Chem. Soc., 43, 390, 1883.
* Liveing, G. D., and J. Dewar, Phil. Trans. Roy. Soc.f 174, 187, 1883.
4 A more complete and general account of the early history of spectroscopy is to
be found in Kayser, “Handbuch der Spektroscopie, Vol. I, pp. 3-128, 1900.
Sec. 1.4] EARLY HISTORICAL DEVELOPMENTS 5
and n2 ® 3,4,5,6 • • • for the first, second, third, fourth, • • • members
of the series. The agreement between the calculated and observed
values of the first four lines is shown in the following table:
Table 1.1.—Balmer's Law for the Hydrogen Series
Angstrom's
Calculated wave-lengths Difference
observed values
Ha = P= 6562.08 A 6562.10 A +0 02 A
Hp = iJA = 4860 80 4860.74 -0.06
Hy = |*fc = 4340 00 4340.10 4-0 10
Hi = |p = 4101 30 4101.20 -0 10
Fig. 1.5.—Stellar and solar spectrograms showing the Balmer series of hydrogen.
accuracy with which lines were then known is seen from the following
measurements of five different investigators for the first member of the
hydrogen series:
Table 1.2.—Early Measures of the Hydrogen Line H«
6565 6 . Van der Willigen
6562 10 . Angstrom
6561.62... MendenhaU
6560 7. Mascart
6559 5 . Ditscheiner
separation of 1574 A, one line lying in the green at X5350 and the other
in the near ultrarviolet at X3776. This discovery proved to be of con¬
siderable importance, as it suggested the possibility that all series arising
from the same element were in some simple way connected with each
other.
Using Liveing* and Dewar’s data on the sodium and potassium
series, Rydberg made for the first time the distinction between what is
called a sharp series and a diffuse series. The Na and K series given in
Fig. 1.4 were each shown to be in reality two series, one a series of sharp
doublets (designated by s) and the other a series of diffuse doublets (des¬
ignated by d).
A third type of series found in many spectra is the so-called principal
series, involving as its first member the resonance or persistence line of the
entire spectrum. Resonance or persistence lines are those relatively
strong lines most easily excited. Such lines usually appear strong in a
Bunsen flame. The yellow D lines are a good example of this.
Still a fourth type of series was discovered by Bergmann. This type
of series is usually observed in the near infra-red region and is sometimes
called the Bergmann series. Since Bergmann did not discover all such
series, Hicks called them the fundamental series. Although this name is
perhaps less appropriate than the other and is really misleading, it has
after 20 years become attached to the series. New series were dis¬
covered so rapidly about this time that many different names and
systems of notation arose. Three of these systems commonly used are
given in the following table.
where rn is the wave number of the given line n, and C, and m are
constants. In this equation, approaches v . as a limit, as n approaches
infinity. While this formula did not give the desired accuracy for an
entire series, Rydberg was of the opinion that the form of the equation
could not be far from correct.
The next equation investigated by Rydberg was of the form
where, as before, JV and ^ are constants, v«, is the limit of the series, and
n is the ordinal number of the line in the series. This formula proved
Vn = V. - Kn = 2, 3, 4, 5, • • • ». (1.4)
If we now let v%, vt, and v{ represent vn for the sharp, principal,
diffuse, and fundamental series, respectively, then the four general
series may be represented by
Sharp series:
R
where n = 2, 3, 4, • • • (1.5)
(n + S)2’
8 INTRODUCTION TO ATOMIC SPECTRA [Chap. I
Principal series:
R
K = "£ (n + py> where n - 1, 2, 3, • •1 • CO . (1-6)
Diffuse series:
R
< = vt - (n + Dy> where n - 2, 3, 4, • - • 00. (1.7)
Fundamental series:
R
K - vl - ^ + Fy< where n - 3, 4, 5, • • • oo. (1.8)
Here S, P, D, and F represent the values of and v l, v&, vt, and the
limits of the different series. In applying the above equations to the
three chief series of lithium, Rydberg obtained the following expressions:
109721.6
K - 28601.6 - (1.9)
(ti + 0.5951) 2‘
109721.6
K = 43487.7 - (1.10)
(n + 0.9596)2
109721.6
28598.5 (1.11)
(n + 0.9974)2
Hence, from Eqs. (1.7) and (1.18), Eq. 1.8 can be written
, _ R R
(1.18a)
'* (2 + DY (n + FY’
and this equation added to the group (1.14), (1.15), and (1.16) shows
that the frequency limit of every one of the four series has now been
expressed in terms of the constants of some other series. For the diffuse
series, n usually starts with two in place of unity.
The Rydberg-Schuster law as well as the* Runge law is shown in
Fig. 1.7 by plotting the spectral frequencies, in wave numbers v, against
the order number n, for the four chief series of sodium. To prevent
confusion among the lines belonging to the different series each one is
plotted separately. In order to show that the first member of the
principal series becomes the first member of the sharp series, when taken
with negative value, the scale has been extended to negative wave
numbers. Unfortunately the frequency scale is too small to show the
doublet nature of each line. The Rydberg-Schuster law is indicated
by the intervals X\ of the figure and the Runge law by the intervals Xz.
A study of the singlet and triplet series discovered in a number of
elements shows that similar laws are to be found. This may be illus-
1 Rydberg, J. R., Astrophye. Jour., 4, 91, 1896.
* Schuster, A., Nature, 55, 180, 200, 223, 1896.
10 INTRODUCTION TO ATOMIC SPECTRA [Chap. I
Fla. 1.7.—Schematic plot of the four chief series of sodium doublets showing the Rydberg-
Schuster and Runge laws
= IP — nSf v* =* IP — nD.
yp = IS — nP, j/£ = 2D — nP. (1.19)
Similar relations are found to exist among the triplet series of the
alkaline earths. The first four members of the four chief series of
calcium are shown schematically in Fig. 1.10. From the diagram it is
seen, first, that the common limit of sharp and diffuse series is a triple
limit with separations equal to the first member of the principal series;
second, that the limit of the fundamental series is a triple limit with
separations equal to the separations of the strongest line and satellites
of the first member of the diffuse series; and third, that the principal
series has a single limit. It is to be noted in the doublet series (Figs.
1.9 and 1.7) that, while the first principal doublet becomes the first
sharp doublet when inverted, the reverse is true for the triplet series of
calcium. By inverting the first sharp-series member (Fig. 1.10), the
lines fall in with the principal series in order of separations and intensities.
Sue. 1.7] EARLY HISTORICAL DEVELOPMENTS 13
In the development of Rydberg’s formula, each member of a series
was assumed to be a single line. In the case of a series where each
member is made up of two or more components, the constants vK and p
of Eq. (1.4) must be calculated for each component. Rydberg’s formulas
for the sharp series of triplets, for example, would be written, in the
accepted notation,
R R
i13r
P.s _ n
n3.Cf,
o i,
= a + sp»)2 ” (n + *Si)2
R R 13p,
iii _ n3Q,
n (1.20)
“a + ‘Pi)1 “ (n + ‘Si)1
R R
AFPn
X 0 — n3Si
AJlj
=a + SPo)J + ‘Si)1
where 3$i, 8Po, 3Pi, 3Pzf occurring in the denominators, are small con¬
stants. Symbolically 13P2 stands for the term R/{ 1 + 8P2)2 which is
one of the three limits of the sharp series. The subscripts 0, 1, and 2,
used here to distinguish between limits, are in accord with, and are part
of, the internationally adopted notation and are of importance in the
theory of atomic structure.
A spectral line is seen to be given by the difference between two
terms and a series of lines by the difference between one fixed term and a
series of running terms. The various components of the diffuse-triplet
series with three main lines and three satellites are designated
1SP2 — n3D3f first strong line; l3Pi — n3P2, second strong line.
13P2 — n8D2, satellite; l3Pi — n*Di, satellite. (1.21)
13P2 — n8£>i, satellite; l3Po ~ n3Z>i, third strong line.
Internationally
Series Fowler Paschen
adopted
later that each letter and number has a definite meaning, in the light of
present-day theories of atomic structure.
1.8. The Lyman, Balmer, Paschen, Brackett, and Pfund Series of
Hydrogen.—It is readily shown that Balmer's formula given by Eq. (1.1)
is obtained directly from Rydberg's more general formula
R_R
(1.22)
Vn (ni + mi) 2 (n2 + M2)2
(1.24)
Lyman series:
» where n2 = 2, 3, 4, • • • . (1.25)
Balmer series:
» where n2 = 5, 6, 7, • • • . (1.28)
Pfund series:
Knowing the value of R from the well-known Balmer series the positions
of the lines in the other series are predicted with considerable accuracy.
The first series was discovered by Lyman in the extreme ultra-violet
region of the spectrum. This series has therefore become known as the
Lyman series. The third, fourth, and fifth series have been discovered
Sec. 1.9] EARLY HISTORICAL DEVELOPMENTS 15
where they were predicted in the infra-red region of the spectrum by
Paschen, Brackett, and Pfund, respectively. The first line of these five
series appears at X1215, X6563, X18751, X40500, and X74000 A, respec¬
tively. It is to be seen from the formulas that the fixed terms of the
second, third, fourth, and fifth equations are the first, second, third, and
fourth running terms of the Lyman series. Similarly, the fixed terms of
the third, fourth, and fifth equations are the first, second, and third
running terms of the Balmer series; etc. This is known as the Ritz
combination principle as it applies to hydrogen.
1.9. The Ritz Combination Principle.—Predictions by the Ritz
combination principle of new series in elements other than hydrogen
have been verified in many spectra. If the sharp and principal series of
the alkali metals are represented, in the abbreviated notation, by
the series predicted by Ritz are obtained by changing the fixed terms
12P and VS to 22P, 32P and 22£, 32£, etc. The resultant formulas are
of the following form:
Combination sharp series:
’•-«(?- ?) <i-33>
and realizing that p and q must be functions involving the order number
n, Ritz obtained, from theoretical considerations, p and q in the form of
infinite series:
, , b\ , Ci , di ,
V = n1 + a1 + ^ + ^ + ^+ -- ‘
(1.34)
. . &2 , C2 . di
Q = tt2 + a2 + -2 + -^ + -6 + •
Using only the first two terms of p and q, Ritz’s equation becomes
identical with Rydberg’s general formula which is now to be considered
only a close approximation. In some cases, the first three terms of the
expansion for p and q are sufficient to represent a series of spectrum lines
to within the limits of experimental error.
1.11. The Hicks Formula.—The admirable work of Hicks1 in develop¬
ing an accurate formula to represent spectral series is worthy of mention
at this point. Like Ritz, Hicks starts with the assumption that Ryd¬
berg’s formula is fundamental in that it not only represents each series
separately but also gives the relations existing between the different
series. Quite independent of Ritz, Hicks expanded the denominator of
Rydberg’s Eq. (1.3) into a series of terms
M +
1
+
1
_a +. _
&
+
, c +
,
(1.35)
1 Hicks, W. M., Phil. Tram. Roy. Soc., A, 210, 57, 1910; 212, 33, 1912; 213, 323,
1914;217, 361, 1917; 220, 335, 1919.
* Wood, R. W., Astrophys. Jour., 29, 97, 1909.
Sec. 1.12] EARLY HISTORICAL DEVELOPMENTS 17
Table 1.7.—The Ritz Formula Applied to the Sodium Series Observed by Wood
(Calculations after Birge)'
Fio. 1.12.—Four diffuse series in neon showing normal and abnormal progression of the
residual m- (After Paachen.)
neon, a Ritz formula with at least one added term is necessary to ade¬
quately represent the series. If Tn represents the running term of a
Ritz formula, and Ti the fixed term,
R
vn Tl — Tn = Ti (1.37)
(n + a +
between the two satellites / and c and the interval between the satellite
c and the chief line a that follows Hartley’s law of equal separations in
both series. In cadmium it is seen that the main lines and satellites
converge toward the three series limits very early in the series. In
calcium, on the gther hand, the lines first converge in a normal fashion,
then spread out anomalously and converge a second time toward the
three series limits. These irregularities now have a very beautiful
explanation which will be given in detail in Chap. XIX.
.0 —
a d e f a be d e f
u -
_J_ 1 i , 1 J_Li_1 i ■
i_L_L_ 2 l L_ k
1 l II 3 I 1. Ii.
I l L 4 1 1 k
_I_ I. II 5 . 1 It It
-1 ii_Lu_ 6 i l . k
7 i L l
L 8 i L l
. I L Ii. 9 i k l
. i L 1 10 _i__l__1_
i l l 11 _l_ i
| Limits j Limits
l_ oo L_
Abnormal Series of Calcium Normal Series of Cadmium
Triplets Triplets
The Balmer and Pickering series are both shown schematically in Fig.
1.14. So good was the agreement between calculated and observed
wave-lengths that the Pickering series was soon attributed to some new
form of hydrogen found in the stars but not on the earth.
Sue. 1.16] EARLY HISTORICAL DEVELOPMENTS 21
All of the lines of this predicted series, except the first, are in the ultra¬
violet region of the spectrum. With the appearance of a line in the
spectrum of f-Puppis at X4688, the position of the first line of the pre¬
dicted series, Rydberg’s assumption was verified and the existence of a
new form of hydrogen was (erroneously) established.
He Hi Hr H* Ha
E9
First
Period PERIODIC TABLE
■IHI
m
n i m ■fl H El
Second
Period
n HEln H Hi
Third 18
Period Nil A
zSw ‘So
12gllrtllfSlgig 1
m
►
ii|iI I ft Ip El
Fourth 25 26 27 28 29 30 31 32 33
Period 2
Fe Co Ni Cu Zn Go Ac
*D4 4FVi ’So % *P#
T$i IIE053II
'5(i 4Stt
.37 38 39 40 41 42 43 48 49 50 51
Fifth
Period 1lb Sr
S* ’S.
155 56
y
57 72
Zr Cb
4 *Fi
73
Ii
*Dvk
Mo Mo
74 75 76 77 78 79
Cd
•Sb
80
In
81
Sn
%
82
Sb
4S*
83 84 85
m
86
Sixth
4 Ba
JeL £&
m1III I■■ ■■■■ 1■■■ 1 | I
Period ' Lfl “*rHf To W Jr Pt Au Pb Bi Pb Rn
s*
*svi lD *h. •Sn 5Pi
% _!5b_
Seventh i
Period J
HIEHll
60 61 62 63 64 65 66 67 70 71
Hd 11
at
•L iyjt
Sm
rHU
Eu
8Hyi
Tb
®HfW ft
Ho
i
Yb
H*
Lu
*Dyi
He Ne A Kr Xe Rn
Z = 2(12 + 22 + 22 + 32 + 32 + 42) (5.1)
2 10 18 36 54 86
N = 2n2.
Table 5.2
K L M N
n - 1 n = 2 n = 3 n = 4
Subshell System
Is 28 2p 3# 3p 3d 4s 4p 4d 4f
tions and the corresponding states for normal H and He atoms aae there¬
fore written:
H, Is, *Si
He, Is2, XS0
6.4* The Second Period.—With the K shell completed in helium the
third electron bound to the third element lithium, in its normal S state
(see Fig. 5.1), is a 2s electron. The fourth electron bound to the fourth
element beryllium, Z — 4, is also a 2s electron to complete the 2s sub-
shell. The normal state of the fifth element boron, Z = 5, is known
from spectroscopic data to be a P state. On the Bohr-Stoner scheme
the first electron added should be one with quantum numbers n = 2 and
l — 1. Such an electron is called a 2p electron. This, as well as other
examples to follow, indicates the origin of our present electron nomen¬
clature s, p, d, etc. The known spectra of the remaining elements
in the second period, carbon, nitrogen, oxygen, fluorine, and neon, show
that 2p electrons are added one after the other to complete not only the
2p subshell but also the main L shell, n = 2. The electron configurations
and the corresponding states for elements in the second period are
designated as follows:
Li, ls22$ , *Sh N , ls22$22p8,4S,
Be, ls22s2 , O , ls22s22p4, 8P*
B , ls22s22p , 2P* F , ls22s22p*, 2P,
C , ls22s22p2, 3P0 Ne, ls22s22p6, %
6.6. The Third Period.—The building up of the eight elements in the
third period of the periodic table is similar to that of the eight elements
in the second period. The normal level of the well-known sodium
diagram (see Fig. 5.1) is an S state, so that the eleventh electron to
be bound to sodium must be a 3s electron. Following sodium the
elements magnesium, aluminum, silicon, phosphorus, sulphur, chlorine,
and argon are built up by adding successively one more 3s electron and
then six 3p electrons. The normal states of the atoms and the corre¬
sponding configurations of the electrons are;
Na, ls22s22p63s , 2£* P , ls22s22p«3s23p*, 4£,
Mg, ls22s22p63s2 , lS0 S , ls22s22p63s23p4 8P2
A1 , ls22s22p63s23p , 2P* Cl, ls22s22p83s23p6 2P,
Si , ls22s22p63s23p2, 3P0 A, ls22s22p63s23p* X/S0
For brevity the first 10 electrons of these configurations are often
omitted, e.g., the configuration for A1 would be written 3s23p. The chief
reason for this is that only the last bound electrons are responsible for
the so-called optical spectra. In dealing with x-rays, however, the inner
electrons are involved in the process of emission or absorption and should
when necessary be given (see Chap. XVI).
82 INTRODUCTION TO ATOMIC SPECTRA [Chap. V
S.6. The Fourth Period or First Long Period.—The first long period
of elements, starting with potassium Z = 19, and ending with krypton
Z = 36, contains 18 elements. Since the subshells 3s and 3p have
been filled in argon Z = 18, and the 3d subshell is next in line, one might
well expect the normal state of potassium, Z = 19, to be a D state, i.e.,
one given by a 3d electron, n * 3, l = 2. The well-known spectrum of
potassium, on the other hand, is like that of sodium with an S state
at the bottom of the energy level diagram. It is therefore evident that
an 8 electron, and necessarily a 4s electron, must have been added in
place of the expected 3d electron, and this in turn must be due to the
fact that the 4s electron is more tightly bound than the 3d electron.
Table 5.3.—The Electron Configurations of the 92 Elements in Their Normal
Atomic States
K L M N Ncnnal
Shell
n- 1 »-2 ft — 3 »—4 state
Subehell 1-0 1-0 1-1 1-0 1-1 1-2 1-0 1-1 1=2 1-3
1 H la
2 He la* •So
3 Li la* 2a
4 Be 2a* •So
6 B la* 2a* 2p
■■
'P\
6 C “ 2p* *Po
7 N “ 2p» **3
8 0 14 2 p* *P2
0
10
F
Ne
M
1
“
“
2p*
2p* •So
11 N» la* 3a
12 Mg 1 3a* •So
Is* Is*
H H«
PERIODIC TABLE
2s* ~ 2s1*
Li Be NORMAL STATE ELECTRON
CONFIGURATIONS
F FTk
3s' * 3s*
Ha Mg 7m
86 INTRODUCTION TO ATOMIC SPECTRA [Chap. V
to radon 86, completes the third long period. The normal S state of
radon is given by the complete electron configuration
Rn, l«*2s22p63s23p83d104s24p64d104/145s25p65d106s26p6, ^So.
6.9. The Seventh and Last Period.—The seventh period, like all but
the first period, starts with an alkali metal, cornellium, followed by
radium, actinium, thorium, uranium X, and uranium. Though the
spectroscopic analysis for these elements is very meager, this period
very probably starts in just the same way as the first and second long
periods.
A very convenient and informing way of representing the addition
of the various electrons to the elements of the periodic table is shown
in Table 5.4. This arrangement not only retains most of the spectro¬
scopic and chemical relations between elements and groups of elements
but also places corresponding electron configurations and subshells
together in individual columns.
6.10. Energy Levels of the Alkali Metals.—Many attempts have
been made to calculate the energy levels arising from other than hydro¬
gen-like atoms. While calculations of the energy levels for the two and
three body problems, hydrogen, and helium have been reasonably
successful, many body problems still remain unsolved. In a few special
cases, however, some headway has been made with the more complex
atoms.
Next to hydrogen the alkali metals Li, Na, K, Rb, and Cs have
the simplest of the known spectra. Each alkali atom contains, in
addition to the completed subshells of electrons, one single electron in
an outer shell (see Table 5.3). The complete configurations are written
out in Table 5.5.
L M N 0 P
Li Is* 2s
Na Is* 2s* 2 p* 3s
K Is* 2s* 2p* 3s* 3p* 4s
Rb Is* 2s* 2p* 3s* 3p* 3d10 4s* 4p* 6s
C8 Is* 2s* 2p* 3s* 3p* 3d10 4s* 4p* 4d10 5s* 5p6 6s
(5.2)
Sec. 5.10] THE ALKALI METALS AND THE PERIODIC TABLE 87
r
Rh is the Rydberg constant for hydrogen, and ni and n2 are integers.
The success of Bohr’s theory of the hydrogen atom lay in the fact that
he was able for the first time to obtain from theoretical considerations
the fundamental Rydberg constant and to show that ni and n* are the
necessary quantum numbers to be associated with each quantized
state of the atom. Neglecting small relativity and spin corrections, to
be treated in Chaps. VIII and IX, there is no difference in energy between
the Bohr circular orbits of hydrogen and the Sommerfeld elliptic orbits
with the same total quantum number n.
L. I I K I Rb 1 Cs I H
2S2P2D2F 2S2P2D2F 7S2P'D1? *S 2D 2F ^^^‘F *SPDF
Fiq. 5.2.—Energy level diagrams of the alkali metals, lithium, sodium, potassium, rub**
dium, and caesium, and hydrogen.
Term Electron n = 2 n - 3 n = 4 n — 5 n = 6 n - 7
Term Electron n = 3 n = 4 n — 5 n = 6 w - 7 n = 8
R R
V (ni - mi)2 (n, - M2)2 { }
and a spectral term by
7 ‘ a <5-5)
where n is now the total quantum number. The factor n — m is still
called the Rydberg denominator, m is the quantum defect, and nerr is
the so-called effective quantum number.
The bracketing together in Fig. 5.1 of the energy levels having
the same total quantum number reveals a number of interesting and
important relations. One important result is that in each group of
levels the S level lies deepest and is followed by P} D, and F levels in
this order. A second relation indicates that in going from the more
elliptic s orbits (S states) to the more nearly circular p, d, and / orbits
(P, D, and F states) the term values approach those of hydrogen. Using
the Eq. (5.5), the effective quantum numbers n«// for the term values of
lithium and sodium (see Tables 5.6 and 5.7) are calculated and given in
Tables 5.8 and 5.9.
Term Electron n — 2 n *= 4 n = 5 n - 7
s
II
..
Table 5 9 —Values of the Effective Quantum Number n^t for Sodium
(An* * 109734 cm"1)
Term Electron n - 3 n = 4 n = 5 n - 6 n - 7 n « 8
which it may return to the normal state: first, by the direct transition
4 2P to 3 2S; second, by the three transitions 4 2P to 3 2D to 3 2P to 32S; or
third, by the three transitions 42P to 42£ to 32P to 32S. There are some
exceptions to the selection rule for the quantum number J, but the
forbidden transitions occur so seldom that the resultant spectrum lines
are exceedingly faint. These exceptions are found experimentally and
should occur theoretically when high external fields are present (see
Chaps. X and XIII).
It should be pointed out in passing that the quantum numbers
assigned to any electron really belong to the atomic system as a whole.
It is only because we can approximate a complex atom by a set of inde¬
pendent one-electron systems with given values for the quantum numbers
n and Z, that we can associate the latter with the individual electrons.
Problems
1. If the valence electron in sodium is excited to the 42D state, what are the
different routes open for the electron in returning to the normal state ?
2. Calculate the Rydberg denominators for the first five term values of the sharp,
principal, diffuse, and fundamental series of potassium (see Fowler, “ Series in Line
Spectra”). Where the fine structure is known, use the larger of the two values,
R ~ 109736 cm"*1.
3. From the values of the effective quantum number calculated in Prob. 2, compute
the quantum defect ix.
4. The first five members of a certain series have wave-lengths as follows: X*ir -
7800.29, 4201.82, 3587.08, 3348.72, and 3228.05 A. Using Eq. (5.4), compute at
least four values of ^ and from their average determine m and the term value of the
fixed term. With the fixed term known, determine the five running terms. Assume
R - 109737 cm-1, and ni = 5 and n2 = 5, 6, 7, 8, and 9.
.
5 It will be found very useful to memorize the order of the elements in the
periodic table and the corresponding electron configurations of the normal state. This
task is greatly facilitated by taking one period at a time and writing them in the form
shown in Table 5.4.
CHAPTER VI
where e is the charge on the electron, m is its mass, and V is the acceler¬
ating potential in electrostatic units. Expressing V in volts Vi the
energy becomes1
me ~ *“*• <6-2>
In order for an electron to collide inelastically with an atom, it must
give up sufficient energy $mv2 to raise the valence electron from its
normal state to the first excited state. In the case of a sodium atom (see
Fig. 5.1) this amounts to raising the valence electron from the 32/S to the
3 2P state, an energy given by
Wd *P — W3*s = hv = hvc, (6.3)
1 For a treatment of the early work on critical potentials see K. T. Compton and
F. L. Mohler, “Critical Potentials,” Nat. Research Council. Bull., Vol. 9, Pt. I, No. 48,
1924.
94 INTRODUCTION TO ATOMIC SPECTRA [Chap. VI
c
ft)
fc
3
u
L
ft)
*
E
o
n
§
§
0 2 4 6 8
Accelerating Potential, Vol+3
Fig. 6.2.—Critical potential curve of sodium. (After Tate and Foote.)
the grid increase and the plate current rises. When the velocity has
increased sufficiently to excite sodium atoms, inelastic collisions occur.
With a further increase in Vi more electrons reach the critical velocity,
are stopped by inelastic collision, and not being able to reach the plate
cause a temporary drop in galvanometer current. This drop in current
continues until the critical speed is attained far enough in front of the
grid to collide inelastically and again reach the plate P. The current
therefore rises again and continues to rise until the electrons after one
inelastic collision attain the critical speed and make a second inelastic
collision. Except for the doubling of the third and fourth maxima, this
process and explanation repeats itself with each major rise and fall of the
curve.
The double peaks in Fig. 6.2 show that not only have collisions
occurred in which the valence electron has been excited but also collisions
in which the valence electron has been completely removed from the
atom, i.e., the atom has been ionized. Complete ejection of the electron
from sodium would, from Eq. (6.5), require an equivalent electron
velocity of 1.2336 X 41444.9 = 5.1 volts. In view of the initial electron
1 Tate, J. T., and P. D. Foote, Jour. Wash. Acad. Sci., 7, 517,1917.
Sec. 6.1] EXCITATION P0TENTIAL8} IONIZATION POTENTIALS 96
velocity at the filament the first ionization should occur at about Vi » 4.5
volts. With ionization taking place at this potential, it is to be expected
that the electrons starting from rest at this point will collide inelastically
at 2.1 volts further on, as observed.
Improved experimental technique on critical potentials has made
it possible to obtain not only one excitation potential for a given atom
Element
Computed Computed
Observed Observed
from spectra from spectra
Li 3 1 84 5 37
Na 11 2.12 2.09 5.13 5.12
K 19 1.55 1.60 4.1 4.32
Rb 37 1.6 1.55 4.1 4.16
Cs 66 1.48 1.38 3.9 3.88
Be I Mg I Cal Sr I Ba I
2s2 3s2 4s2 5s2 6s2
Li I Na I K I Rb I Cs I
Bell Mg II Call Sr II Ball
28 Ss 4s 5s 6s
visible and near ultra-violet region. In potassium, for example, the first
member of the principal series is in the far red between 7000 and 8000 A,
whereas the first line of the principal series of ionized calcium is in the
n2 n2
n = 1, 2, 3, • • • (see Chap. II) (6.6)
By analogy with the formulas of hydrogen and ionized helium it follow s
that for ionized atoms like those under discussion the expression for
term values may be written
4R 4R
T = (6.7)
nlti in - m)2>
where n is the total quantum number of the electron and m the quantum
defect. The well-known enhanced series of the alkaline-earth elements
are represented therefore [see Sec. 1.16 and Eq. (1.41)] by the formula
4R _4 R
(6.8)
(ni — m)2 (n2 — M2)2
Fig. 6.5.—Energy level diagrams of singly ionized beryllium, magnesium, calcium, stron¬
tium, and barium, and hydrogen.
hydrogen scale shown at the right. With the relatively large energy
values for the different states, electron transitions between states will
in general be large. The difference between each term value and the
hydrogen term values of the same n, shown by horizontal lines, is a
measure of the quantum defect m. A comparison of these diagrams with
those of Fig. 5.2 brings out a number of interesting relations. The
relatively large quantum defect of s electrons in Li I and Be II, of a
Sac. 6.2] EXCITATION POTENTIALS, IONIZATION POTENTIALS 99
0 12 3 4 5 6 7a,
.K
Fig. 7.2.—Probability-density distribution curves for the neutral alkali atoms, lithium,
sodium, potassium, and rubidium. In each case the core is shown by one curve and the
valence electron by another.
to say, however, that the resultant electric field obtained for any atom
is such that the solutions of Schrddinger’s wave equation for all of the
core electrons in this field give a distribution of electrons which reproduces
the field.
1 Habtbxb, ,
D. B., Proe. Camb. Phil. Soc.t 24 89, 111, 1928.
102 INTRODUCTION TO ATOMIC SPECTRA [Chap. VII
W = P. E. = --eV, (7.1)
r
where r is in centimeters. Expressing V in volts and r in Angstroms,
T/ _ 300e _ 300 X 4.77 X 10"10 _ 14.31 n 0,
r X 10“8 r X 10“8 r ^ '
Expressing r in units of ai (a2 = 0.528 A),
_ 27\1_27\1_(73)
rnu* V ionization potential in volts ^ * *
From the ionization potentials of the alkali metals given in Table 6.1
the following values of the orbital extremities are obtained:
Li Na K Rb Cs
r— = 5.0ai 5.3ai 6.3ai 6.5ai 7.0ai
1 Paulino, L., Proe. Roy. Soc., A, 114, 181, 1927.
* Thomas, L. H., Proc. Canib. Phil. Soc., 28, 542, 1927; see also Gaunt, Proc. Camb.
Phil. Soc., 24, 328, 1928; and Fermi, ZeiU. f. Phys., 48, 73, 1928.
Ssc. 7.2] PENETRATING AND NONPENETRATING ORBITS 103
SODIUM
Penetrating Non* Penetrating
I.
3d
8 ? 10 II 12 13 14ci|
sponding 3s, 3p, and 3d classical orbits based on model b (see Fig. 4.8),
are shown in the lower part of the figure. A comparison of these orbits
with the corresponding hydrogen orbits shows that, due to penetration
into the core, 3s and 3p are greatly reduced in size radially. The 3d
orbit, on the other hand, remains well outside the main part of the core
and is hydrogen-like. Corresponding to the penetration of the 3s and 3p
orbits the probability-density-distribution curves (above) have small
loops close to the nucleus.
104 INTRODUCTION TO ATOMIC SPECTRA [Chap. VII
greatly from a Coulomb field, the orbit will be a Kepler ellipse precessing
slowly (due to small deviations from a Coulomb field) about the atom
center. In the remaining figures increased penetration is shown accom¬
panied by an increase in the precession at each turn of the orbit. As
the electron goes from aphelion (r^) to perihelion (rmln), it leaves
behind it more and more of the core charge. With the steady increase
in force field the electron is drawn from its original path into a more
and more eccentric path, with the result that at its closest approach
to the nucleus the electron has turned through somewhat more than
180 deg. Upon reaching again, there has been an advance, i.e., a
precession, of the aphelion. The increased force of attraction between
nucleus and electron at penetration increases the binding energy, the
kinetic energy, and the term values but decreases the total energy of the
atomic system [see Eq. (2.15)].
Bsc. 7.3] PBNBTRA TINQ AND NONPBNBTRA TINO ORBITS 105
valence electron the atom core is pushed away and the nucleus is pulled
toward the electron by virtue of the repulsion and attraction of like and
unlike charges, respectively. The effect of this polarization is to decrease
the total energy of the system. On an energy diagram this means a
lowering of the level, i.e,f an increase in the term value. Theoretical values
of the polarization energy calculated for the alkali metals with the aid
of the quantum mechanics are found to account for the major part of
these very small deviations from hydrogen-like terms.1
7.4. Penetrating Orbits on the Classical Model.—Although no sharp
line of demarkation can be drawn between penetrating and nonpene¬
trating orbits, the former may be defined as those orbits for which the
term values are appreciably different from those of hydrogen. Referring
to Fig. 5.2, the s orbits of Li, the s and p orbits of Na and K, and the
s, p, and d orbits of Rb and Cs come under this rough classification. Cer¬
tainly on the quantum-mechanical model all orbits are penetrating.
To effect a calculation of the term values for penetrating orbits on the
classical theory, one is led by necessity to simplify somewhat the atom-core
model given in Fig. 7.4. A suitable idealized model was first put forward
by Schrodinger2 in which the core elec¬
trons were thought of as being distrib¬
uted uniformly over the surface of one or
more concentric spheres. This same
model has been treated by Wentzel,3
Sommerfeld,4 * Van Urk,6 Pauling and
Goudsmit,® and others. Since the
classical treatment of penetrating orbits
is so closely analogous to the quantum-
mechanical treatment of the same
Fio 7.6 -Valence electron penetrat- orbitg to be taken up in the next
mg an ideal core where the core electrons * ^
ue distributed uniformly over the sur- section, Schrodinger’s simplified model
face of a sphere. will be considered here in some detail.
Consider the very simplest model in which the core electrons are
distributed uniformly over the surface of a sphere of radius p (see Fig. 7.6).
Let Z»e represent the effective nuclear charge inside the charge shell and
Z<# the effective nuclear charge outside the shell. Usually Z0 is 1 for the
alkali metals, 2 for the alkaline earths, etc. The potential energy
of the valence electron when outside the thin 'spherical shell of charge
(,Zi — ZQ)e will be
y_
(7.4)
Vt r
while inside it becomes
_Zte* Z<e2 _ Zof'
Vi (7.6)
r p p '
The total energy for an elliptic orbit, in polar coordinates r and <p, is
given by Eq. (3.26) as
That part of the electron path which is outside the shell is a segment
of an ellipse determined by the azimuthal quantum number k and the
radial quantum number r0, whereas the path inside the shell is a segment
of an ellipse determined by the same azimuthal k (pv = constant) but a
different radial quantum number, rx. Substituting successively the
potential energies of Eqs. (7.4) and (7.5) in the total energy [Eq. (7.6)]
snd solving for pr) the radial quantum conditions can be written down as
The total quantum numbers to be associated with r0 and r< will, as usual,
be given by
n0 = k + r0 and n< == k + r<. (7.10)
Since the electron does not complete either of the two ellipses in
one cycle, the integrals of Eqs. (7.8) and (7.9) are not to be evaluated
over a complete cycle as indicated but over only that part of the ellipse
actually traversed. The radial quantum number r for the actual path
traversed is therefore given by the sum of the two integrals
108 INTRODUCTION TO ATOMIC SPECTRA [Chap. VII
f
•fouteide
Rodr + f
•/inside
R4r = rh. (7.11)
The total energy in the outside region by Eqs. (2.14), (2.15), (2.30), (2.33),
and (7.4) is
f = -r = - = = Zl* (7.12)
2 2r 2 o,n*
zy = (Zt - Z0)e2
(7.14)
2a in2 2ainf p
Consider now the special case shown in Fig. 7.7, in which the two
partial Kepler ellipses are almost complete.1 If the outside orbit were
Penetrating Orbit
a complete Kepler ellipse the electron
would never penetrate the shell,
whereas if the inner orbit were com¬
plete the electron would always remain
inside. As the outer ellipse is made
less and less penetrating, the two
ellipses become more and more com¬
plete and the integrals of Eq. (7.11)
approach those of Eqs. (7.8) and (7.9).
Expressing this in terms of the radial
quantum numbers,
Flo. 7.7.—Special case where inner r = r0 + 7». (7.15)
and outer ellipses are almost complete.
Core charge distributed uniformly over In a similar fashion the perihelion
the surface of a sphere. distance of the outer orbit a0( 1 — Co)
approaches the aphelion of the inner ellipse a<( 1 + c<), both approaching
at the same time the radius of the spherical shell p. We write, therefore,
n — n0 = ni — k = p. (7.17)
from which
m = pZx(2paiZi - a2fc2)~*. (7.20)
This expresses the experimental result, well known before the quantum
mechanics, that for a given atom the quantum defect n is a function
of the azimuthal quantum number and is independent of the total
quantum number (see Table 7.2).
Na s
* s, l = 0 1 37 1.36 1 35 1.35
2p p, l — 1 0.88 0 87 0 86 0.86
1
1 A\ A2 A3
* = A1 A2 A31 (7.22)
V6 Bi Bz
where the A's and A*s are Is hydrogen-like wave functions for the
two K electrons and the IPs are 2s wave functions for the L electron.
From Eq. (4.55) the solutions of the wave equation for Is and 2s
electrons are
A - +u = Nitr*% (7.23)
/w Zr Zr\
B « - NsjfnT * - e~^y (7.24)
1 Wilson, E. B., Jour. Chem. Phys., 1, 210, 1933. For other references see this
paper.
* Slater, J. C., Phys. Rev., 34, 1293, 1928.
* Slater, J. C., Phys. Rev., 42, 33, 1932.
Sec. 7.5] PENETRATING AND NONPENETRATING ORBITS 111
1
Ai At At
* A,
\/6 Bi + bAi
At At > (7.25)
Bt + bAt Bz *+• bAt
where the A1s and Bfs now represent the simple hydrogen-like wave
functions with modified values for Z.
When determinants involving hydrogen-like functions with Z put
equal to the atomic number are used to calculate the energy, the result
Ai = ils = (7.26)
/y — (Z — m)rt — (Z—vi)r% — (Z—ai)rA
where <ri = 0.31, <r2 = 1.67, 0-3 = 0, and N1 and Nt are normalizing
factors. The probability-density-distribution curve obtained by plotting
4?rr2^2 against the electron-nuclear distance r is shown by the heavy
curve in Fig. 7.8. The dotted curve for the hydrogen 2s state does
not rightly belong here but is shown as a comparison with the 2s state of
lithium. The latter curve is obtained by plotting 4irrVi««
The pulling in of the inner loop of the Li 2s curve over the inner
loop of the hydrogen 2s curve is due to the lack of screening by the Is
electrons of the core and is to be compared with the deeper penetration
112 INTRODUCTION TO ATOMIC SPECTRA [Chap. VII
and speeding up of the electron in the inner part of the classical orbit.
The 2s electron, since it is most of the time well outside the core, is
screened from the nucleus by the two core electrons. This average
screening is well represented by the screening constant <r* = 1.67. The
screening of each Is electron from the nucleus by the other Is electron
should lie between 0 and 1. The value <ri = 0.31 is in good agreement
with this. When the 2s electron is inside the core (i.e., the smaller
loop), the screening by the outer Is electrons is practically negligible.
The value <r* = 0 is in good agreement with this. This same analogy
between the quantum-mechanical model and the orbital model should
extend to all elements.
The accuracy with which Eq. (7.25) represents the normal states of
lithium and singly ionized beryllium (see Figs. 5.2 and 6.5) is shown by
the following values:
Spectroscopic Calculated
tration is not greatly different for all states of the same series. The,
first seven of a series of d states, l * 2, for example, are shown in Fig.
7.9. The orbits given above are drawn according to model a* and
beneath them are drawn the hydrogen probability-density-distribution
curves of the same quantum numbers.
On the orbital model the perihelion distances are very nearly the
same. On the quantum-mechanical model the lengths of the first
Fig. 7.9.—Series of d orbits illustrating nearly equal penetration for all orbits with the same
l on either the classical or the quantum-mechanical model.
loop (indicated by the first nodal points), with the exception of 3d, are
nearly the same. Figure 7.9 brings out better than any other, perhaps,
the close analogies that may be drawn between the orbits of the early
quantum theory and the probability-density-distribution curves of the
newer quantum mechanics. This is one of the reasons why the terms
orbit, penetrating orbit, nonpenetrating orbit, etc., are still used in discussing
quantum-mechanical processes.
Problems
1. Assuming that the s and p orbits in potassium penetrate only the shell of eight
M electrons, compute the value of pM from Eq. (7.21). Compare these values of p with
the density-distribution curve in Fig. 7.2. If the values of the quantum defect are no*
known from Prob. 2, Chap. V, they are readily calculated from the term values
directly.
.2 Compute, on the classical theory, model a, the maximum electron-nuclear
distances attained by the valence electron of sodium in the first 10 8 orbits. Compare
these values with the corresponding s orbits of hydrogen by plotting a graph. Plot
Turn* against n for hydrogen and for sodium.
CHAPTER VIII
states shown in Fig. 8.1. In boron, for example, the first excited 2S,
state finds the electron in a 38 orbit, the Is and 2s subshell being already
filled. There being no d electrons in boron and no virtual d orbits
with a total quantum number n lower than 3, the excitation of the valence
electron to the first 2D state places the electron into a 3d orbit.
Neglecting for the moment the doublet nature of the different levels,
it should be noted that the 2F terms in all five elements, just as in the
alkali metals, are nearly hydrogen-like, indicating nonpenetrating
In
2S 2P 2d 2f 2s 2p 2d V 2s 2p 2d 2f £s 2p 2d 2f h 2p 2d 2f
t {Hi51
Table 8.1.—Doublet Separations for the First Members of the Sharp Series
»PrVJ, >pr*5j AX Ap
X A cm~l
structure intervals for the first member of the principal series in each of
the alkalies and ionized alkaline earths are observed as follows:
Li I Na I K I Rb I Cs I
0.338 17.2 57.9 237.7 554.0 cm-1
Bell Mg II Ca II Sr II Ba II
6.61 91.5 223.0 800.0 1691.0 cm-1
In lithium the 22S-22P interval (see Fig. 5.1) is over 4000 times the
fine-structure interval 0.338 cm-1, whereas in caesium the corresponding
interval is only 20 times larger.
Principal Series
ill
n: i _ Na
■■■MM ill
— 1 1 7 .. Rb
mu_
-4-- Js =L= i 4-4-
Cs
A general survey of the energy level diagrams (Figs. 5.1, 5.2, 6.4, and
6.5) will enable certain general conclusions concerning fine-structure
intervals to be drawn: First, corresponding doublet separations increase
with atomic number. Second, doublet separations in the ionized
alkaline earths are larger than the corresponding doublets in the alkali
metals. Third, within each element doublet separations decrease in
going to higher members of a series. Fourth, within each element
Sac. 8.2] DOUBLET FINE STRUCTURE 117
2P— 2d . — ■■■ 2f
*P ?D =-2F
aT ;%L_ AF
AP *AP ’off
$1 AP AP
AD
Fvr
± JL JL
AP AP AP aq-p—vaDap ADaf AFad
Fia. 8.4.—Illustrating selection and intensity rules for doublet combinations.
apply to all spectra in general, they will be stated here for doublets
only, (a) The sum of the intensities of those lines of a doublet winch
come from a common initial level is proportional to the quantum weight
of that level. (6) The sum of the intensities of those lines of a doublet
which end on a common level is proportional to the quantum weight
of that level. The quantum weight of a level is given by 2j + 1 .
This,
it will be seen in Chap. X, is the number of Zeeman levels into which a
level j is split when the atom is placed in a magnetic field.
In applying these intensity rules, consider again the simple case
of a principal-series doublet. Here there are two lines starting from
the upper levels 2P\ and 2P\ and ending on the common lower level
2S$. The quantum weights of the 2P levels are 2(f) + 1 and 2(£) + 1,
giving as the intensity ratio 2:1. The same ratio results when the 2S
level is above and the 2P level below.
120 INTRODUCTION TO ATOMIC SPECTRA {Chap. VIII
2P,
4 2
2D,6 X 0
2D,4 Y Z
The numbers directly below and to the right of the term symbols are
the quantum weights 2j + 1. Let X, F, and Z represent the unknown
intensities of the three allowed transitions and zero the forbidden transi¬
tion. From the summation rules (a) and (b) the following relations
are set up: The sum of the lines starting from 2Z>S is to the sum starting
X
from 2Z)j as 6 to 4, i.e., and, similarly, the sum of the lines
Y +Z
X + Y 4
ending on 2Pj is to the sum ending on 2P* as 4 is to 2, i.e.,
Z 2
The smallest whole numbers which satisfy these equations are X = 9,
F*l, and Z = 5. If the 2Z> terms are very close together so that the
observed lines do not resolve the satellite from the main line, as is usually
the case, the two lines observed will have the intensity ratio 9+1:5
or 2:1, the same as the principal-series or sharp-series doublets. Intensity
measurements of the diffuse series of the alkali metals by Dorgelo8 confirm
this.
A favorable spectrum in which the satellite of a diffuse-series doublet
can be easily resolved, with ordinary instruments, is that of caesium.
The first three members of this series are in the infra-red and are not
readily accessible to photography. The fourth member of the series,
composed of the three lines XX 6213, 6011, and 6218 has been observed
and l but as an inherent and fixed property of the electron. The total
angular momentum contributed to any atom by a single valence electron
is therefore made up of two parts: one due to the motion of the center
Fig. 8.5.—Spin and orbital motion of the electron on the classical theory.
of mass of the electron around the nucleus in an orbit, and the other
due to the spin motion of the electron about an axis through its center of
mass (see Fig. 8.5). Disregarding nuclear spin the atom core, as we shall
see later, contributes nothing to the total angular momentum of the atom.
By analogy with the quantum-mechanical developments in Chap. IV,
we return now to the orbital models a, by c, and d (Fig. 4.8) to find a
suitable method for combining these two angular momenta. For this
Vector Vector
Vector diagrams for the composition of orbit and spin, on models a and 6,
are given in Fig. 8.6 for the two possible states of the d electron.
On model b the spin angular momentum 8 • h/2ir is added vectorially to
the orbital angular momentum Z • h/2r to form the resultant j • h/ 2vi where
j = l ± 8. On model a the spin angular momentum s* • h/2oc is added
vectorially to the orbital angular momentum Z* • h/2ir to form the resultant
j* • h/2ir, where s* = V*(* + 1), Z* = y/UJi + 1), j* = VJif+Tj, and
j = Z + 8. It should be noted that two is the maximum number of j
values, differing from each other by unity, that are possible on either
model and that for s electrons there is but one possibility.1 For s, p,
/, . . . orbits Z = 0, 1, 2, 3, • • • . The quantum conditions are that
j shall take all possible half-integral values only, i.e.9j = i, h %, * * • .
For a d electron Z = 2, s = j, l* = \/6, and s* — £\/3- The only
possible orientations for Z and s, or for l* and s*, are such that j = %
and j, and j* = and For a p electron l = 1, s = |,
Z* = \/2, and s* = |\/3. The only possible orientations for Z and s, or
for Z* and s*, are such that j = f and J, and j* — |\/15 and For
an s electron Z = 0, s = i, l* = 0, and s* = %\/3. The only possible
value for j is and j* = i\/3.
..
8 6 The Normal Order of Fine-structure Doublets.—In the doublet
energy levels of atomic systems containing but one valence electron
it is generally, but not always, observed that the fine-structure level
j — l — \ lies deeper than the corresponding level j — Z + J. For
example, in the case of p and d electrons, 2Pj lies deeper than 2P| and
2D} lies deeper than 2Dj. This result is to be expected on the classical
theory of a spinning electron and on the quantum mechanics. Classically
we may think of the electron as having an orbital angular momentum
and a spin angular momentum. Due to the charge on the electron
each of these two motions produces magnetic fields.
Due to the orbital motion of the electron, of charge 6, in the radial
electric field of the nucleus Ef there will be a magnetic field H at the
electron normal to the plane of the orbit.2 That this field is in the direc¬
tion of the orbital mechanical moment Z may readily be seen by imagining
the electron at rest and the positively charged nucleus moving in an
orbit around it. In this field the more stable state of a given doublet
will then be the one in which the spinning electron, thought of as a
small magnet of moment lines up in the direction of H. In Fig.
8.7 the electron spin moment is seen to be parallel to H in the state
j = Z — J, and antiparallel to H in the state j = Z + J. Of the two
1 The small letter 8 used for electron spin must not be confused with the small
letter for 8 electrons.
1 This field is not apparent to an observer at rest with the nucleus but would be
experienced by an observer on the electron.
124 INTRODUCTION TO ATOMIC SPECTRA [Chap. VIII
A j-l + s B j*l-s
Fig. 8.7.—Illustrating the mechanical and magnetic moments of the spinning electron
for the two fine-structure states j ** Z -f \ and j — l — J. The vectors are drawn accord¬
ing to the classical model (b).
the 2P terms are normal, and the 2D, and very probably the 2F, terms
are inverted. For a possible explanation of the inversions see Sec. 19.6.
It can be shown quantum mechanically that, neglecting disturbing
influences, doublet levels arising from a single valence electron will be
normal. Where resolved, all of the observed doublets of the boron
group of elements, the ionized alkaline earths, and the more highly
ionized atoms of the same type are in agreement with this.
..
8 6 Electron Spin-orbit Interaction.—The problem which next
presents itfeelf is that of calculating the magnitude of the doublet separa¬
tions. Experimentally we have seen that doublet-term separations, in
general, vary from element to element, from series to series, and from
member to member. Any expression for these separations will there¬
fore involve the atomic number Z, the quantum number l9 and the
quantum number n.
A calculation of the interaction energy due to the addition of an
electron spin to the atom model has been made on the quantum mechanics
by Pauli,1 Darwin,2 Dirac,8 Gordon,4 and others. By use of the vector
1*^k = mr X v (8.1)
where l* = \/l(l + 1), m is the mass of the electron, v its velocity, and
r the radius vector. According to classical electromagnetic theory, a
charge Ze on the nucleus gives rise to an electric field E at the electron
given by
Ze
E = -yr. (8.2)
r3
h-EXv- (8.3)
c
From these two equations,
H = —t X v. (8.4)
crs
2rmi X v = Z* h, (8.5)
the field becomes
H = A.££.^. (8.6)
2t me r3
In this field the spinning electron, like a small magnetic top, under¬
goes a Larmor precession around the field direction. From Larmor's
theorem [Eq. (3.58)] the angular velocity of this precession should be
given by the product of the field strength H and the ratio between the
magnetic and mechanical moment of the spinning electron:1
e Ze . 1 . 2_e_
H * 2 (8.7)
2 me me r3 2mc
This is just twice the ordinary Larmor precession f given in Eq. (3.58)
1 The ratio between the magnetic and mechanical moment of a Bpinning electron
is just twice the corresponding ratio for the electron's orbital motion. This is in
agreement with results obtained on the quantum mechanics and accounts for the
anomalous Zeeman effect to be treated in Chap. X.
126 INTRODUCTION TO ATOMIC SPECTRA [Chap. VIII
(8.10)
In this equation for the interaction energy the last two factors are
still to be evaluated. In general the electron-nuclear distance r is a
function of Z, n, and l and changes continually in any given state.
Because the interaction energy is small compared with the total energy
of the electron's motion the average energy AWi,a may be calculated by
means of perturbation theory. In doing this, only the average value
(A & (8.11)
W a\nH{l + m+ 1)'
= h2 (8.12)
4w2me2
For the last factor of Eq. (8.10) we turn to the vector model of the
atom. In calculating the precessional frequency of s* around the field
produced by the orbital motion the vector l* was assumed fixed in space.
One might equally well have calculated the precession of the orbit in tjie
field of the spinning electron. It is easily shown that this frequency is
just equal to the ordinary Larmor precession and to the precession of the
electron around Z*. In field-free space both orbit and spin are free to
move so that l* and s* will precess around their
mechanical resultant j*. By the law of conserva¬ .*
J
tion of angular momentum this resultant j* and
hence the angle between l* and s* must remain
invariant. The vector model therefore takes the
form shown in Fig. 8.8. With the angle fixed the
cosine does not need to be averaged and l*s* cos
(1*8*) is calculated by the use of the cosme law
j* = Z*2 + s*2 + 2Z*s* cos (1*8*) (8.13)
from which
j*2 - Z*2 - 8*
1*8* COS (1*8*) = (8.14)
where
RaW
cm- (8.19)
n*l(l + M+ 1)
Measured from the series limit down, the term value of any fine-
structure level will be given by
T = To - r, (8.20)
128 INTRODUCTION TO ATOMIC SPECTRA [Chap. VIII
r
/
\
\
\
\
l Z
Ra2ZA Ra2ZA
Av cm-1. (8.21)
nH{l + \){l + 1) nH(l + 1)
(IU2)
For hydrogen-like systems the effective nuclear charge is given
simply by the atomic number Z.
Sic. 8.8) DOUBLET FINE STRUCTURE * 129
In the previous chapters we have seen hoW for other atomic systems
the deviations of the term values from those of hydrogen-like atoms are
attributed to a polarization of the atomic core or generally to a quantum
defect [see Eq. (5.4)],
T - HZ2 _ (8.23)
(n - aO* nht ’
where Z * 1 for neutral atoms, 2 for singly ionized atoms, etc. Instead
of attributing the increased binding of the electron to a defect in the
quantum number n, one may argue that it should be attributed to a
screening of the valence electron from the nucleus by the intervening
core of electrons, and that Z should be replaced by Z.ff where ZM = Z — <r,
Z is the atomic number, and a is a screening constant:
RZ2M = R{Z — a)2
(8.24)
n2 n2
Most of the doublets to which this formula applies are known only for
singly and multiply ionized atoms. Although its general application
will be left to Chap. XVII on Isoelectronic Sequences, it should be
remarked here that this formula gives doublet intervals in remarkably
good agreement with experimental observations.
8.8. Spin-orbit Interaction for Penetrating Orbits.—In the preceding
chapter on penetrating and nonpenetrating orbits we have seen how
penetrating orbits may be considered as made up of two parts, an inside
segment of an ellipse and an outside segment. In attempting to apply
Eq. (8.25) to the doublets of penetrating orbits, much better agreement
with the observed values is obtained, especially for the heavier elements,
by again considering separately the inner and the outer part of the orbit
(see Fig. 7.6). Whatever atomic model is formulated, the electron in a
deeply penetrating orbit is by far the greater part of the time in an outer
region where the field is nearly hydrogen-like. If the electron remained
in an outer orbit like the outer segment, the doublet formula [Eq. (8.21)]
would be
R**zi
A r. (8.26)
n*l(l + 1)
To bring these two formulas together for the actual orbit, the motion
in each of these segments is weighted according to the time spent in each.
Now the time t required to traverse the whole path is so nearly equal to
the time ta required to traverse a complete outer ellipse that we may
write, to a first approximation,
f = njh8
(8.28)
" 4r2me4Zf
This equation for the period of an electron in a Kepler ellipse was left
as an exercise at the end of Chap. III. The time required to traverse a
completed inner ellipse is
n\h3
(8.29)
4ar2meAZ\
Now the resultant frequency separation Av for the actual orbit will
be A Vo times the fractional time tQ/t spent in the outer segment, plus
A Vi times the fractional time tx/t spent in the inner segment:
Av = + A,f (8.30)
Ra2Zl
Av —
<1(1 + 1) (Z\ + Zl) cm" (8.31)
Av =
n<x2zm (8.32)
nll(! + 1).
This equation was derived from the quantum theory and used by Land6l
before the advent of the spinning electron and the newer quantum
mechanics. In calculating Z, for a number of atoms Land£ showed that
the penetration in many cases is almost complete, being almost equal
to the atomic number Z (see Table 17.4A). It is to be noted that nQ is
the effective quantum number.
Inserting screening constants for each of the Z’s, in Eq. (8.32),
Problems
1. Compute doublet-term separations for the nonpenetrating 2p states of lithium
and singly ionized beryllium. Assume complete screening by the core electrons.
Compare the calculated values with the observed values given in Sec. 8.1.
2. Determine theoretical intensity ratios for the doublet transitions %Ff-*Gj.
8. Construct vector-model diagrams for JFj, *Gj and *Gj states based on model
a, Fig. 8.6.
4. Determine the electron spin-orbit precession frequency «/2t for a 4f state in
potassium. Assume complete screening by the 18 core electrons.
6. Compute a theoretical doublet separation for the 6*P state in caesium. Assume
complete penetration when the electron is inside the core, t.e., «,* ** 0, and perfect
screening when it is outside, t.c., 8P = Z — 1. The effective quantum number n»
can be determined from the observed term values (use the center of gravity of the
doublet). All other factors remaining the same, what value of *,• will give the observed
doublet separation?
CHAPTER IX
Fig. 9.1. Photographs of the Ha line of both of the hydrogen isotopes H1 and H*. (After
Lewis and Spedding.)
energy as the Bohr circular orbit with the same n. This energy in waive
numbers is
l- m
where R is the Rydberg constant
2t2W64
ch*( i +
h is Planck's constant, c the velocity of light, m and e the mass and charge
of the electron, and M the mass of the nucleus with charge Ze.
Bohr pointed out in his earliest papers that the relativistic change in
mass of the orbital electron should be taken into account in computing
the energy levels. Introducing elliptic orbits, Sommerfeld applied the
special theory of relativity to the
electron mass. Due to the different
velocity of the electron in orbits of
the same n but differing azimuthal
quantum number, the mass of the
electron and hence the resultant
energy levels are all different. If the
rest mass of the electron is m0, its
mass when moving with velocity v is
given by the special theory of rela¬
tivity as
= ro{1-^)1
\ / Fig. 9.2.—Schematic representation of
the precession of an electron orbit due to
As a result of this change in mass, the relativity change in mass of the elec-
, . , • , , , , tron with velocity. {After Sommerfeld.)
which is greatest at perihelion and
greatest for the most elliptic orbits, there is an advance of the perihelion,
or a precession of the electron orbit, similar to that of a penetrating orbit
in the alkali metals (see Fig. 7.4), or to that of the planet Mercury moving
about the sun. This precession is shown schematically in Fig. 9.2.
While the derivation of Sommerfeld's equation for the change in energy
due to this precession is out of place here, we shall find use for it in making
comparisons with the quantum-mechanical results.1 According to the
Sommerfeld theory the term values of hydrogen-like atoms are given by
t= = maJ i a2Z2_
—~
Ac
-|-
h\l + (n - k + y/k2 — a2Z2)2 }"+£
1 For a derivation of Sommerfeld’s relativistic fine-structure formula see “Atomic
Structure and Spectral Lines,“ p. 467,1023; also A. E. Ruark and H. C. Urey, “ Atoms,
Molecules and Quanta/’ p. 132.
134 INTRODUCTION TO ATOMIC SPECTRA [Chap. IX
and
Mm
** * M + m
+ • • • • (9.6)
The first term of this expansion is the same as that derived by Bohr
for circular orbits, neglecting relativity, and gives the major part of the
energy. With n = 1, 2, 3, • • • , and with Z — 1 for hydrogen, Z = 2
for ionized helium, and Z = 3 for doubly ionized lithium, this term gives
the following values:
The corrections to be added to each of the above given terms are there¬
fore given by
<U>
Fiq. 9.3.—Fine structure of the hydrogen energy levels. A TV and AT*i,« represent
the relativity and the spin-orbit corrections respectively. The dashed linea represent
Sommerfeld’s relativity corrections.
of hydrogen given by Table 9.1. The shifted levels for each value of
n and k are shown by the dotted lines with the term value increasing
downward. The left-hand side of each diagram has to do with the
spinning-electron picture of the atom and will be taken up in the following
section. For ionized helium and doubly ionized lithium the intervals
given in Fig. 9.3 must be multiplied by 16 and 81, respectively.
9.2. Fine Structure and the Spinning Electron.—With the introduc¬
tion of the spinning electron and the quantum mechanics another account
of the hydrogen fine structure has been given. Heisenberg and Jordan1
A 6563 A4861
Fig. 9.4.—Schematic diagrams of the lines Ha and H/g, in the Balmer series of hydrogen.
AT =
Ra2Z V 1_ r\ (9.14)
n3 \j + J 4nj
the two known hydrogen isotopes. Using Fabry and Perot ^talons,
photographs similar to the one shown in the center of Fig. 9.5 have been
obtained. For this photograph the first order of a 30-ft. grating mount¬
ing (of the Littrow type) was used as the auxiliary dispersion instrument.
Microphotometer curves of both H1 and H2 are reproduced above and
below each pattern. It is to be noted that the components of H2 are
considerably sharper than Hi, and that a third component is beginning to
show up. The broadening is due to the Doppler effect and should be
greater for the lighter isotope.
Fio. 9.5.—Fine structure of Hi and Hi from the Balmer series of the two hydrogen
isotopes. Microphotometer curves above and below were made from the interference
patterns in the center. (After Spedding, Shane, and Grace.)
1 SoMMXSBniLD, A., and A. Unsold, Zeite. f. Phys., 36, 259, 1926; 38, 237, 1926;
a@e also ScrbOdxnoxr, E., Ann. d. Phyt., 80, 437, 1926.
Sec. 9.5] HYDROGEN FINE STRUCTURE 139
Z = 2 the fine-structure separations should bd 16 times as great as in
hydrogen [see Eqs. (9.13) and (9.14)]. The predicted fine structure
shown above in Fig. 9.6 was first given by Sommerfeld and Unsold. At
least four and possibly five of the predicted
components may be said to have been resolved
by Paschen.1 The appearance of certain com¬
ponent lines in this pattern, which are not
allowed on Sommerfeld’s original theory of
hydrogen fine structure, are strong points in
favor of the newer theory of the coincidence
of levels having the same j values.
9.5. The Dirac Electron and the Hydrogen
Atom.—On Diracs2 theory a single electron in
a central force field is specified by a set of
four wave functions ^i, ^3, and ^4, in place
of just one as in the case of the Schrodinger
theory. Each of these functions is a solution
of a wave equation. Although the setting up
of the equations is out of place here we shall
accept the solutions arrived at by Darwin
and Gordon3 and show in what way they 4686 1 0 9 8 7 6 .5 4 3
Fig. 9.6.—Diagram of the
correspond to the earlier theories and, at the
fine structure of the ionized
same time, get some picture of the new atom helium Une X4686. (After Som-
merfeld, Unsold, and Paschen.)
With each wave function \j/i, ^2, ^3, and ^4 properly normalized, the
probability density, just as in the case of Schrodinger's theory (see
Chap. IV), is given by
P = W, (9.15)
where
W* = Mi* + Mi + Mi + Ml (9.16)
For given values of the azimuthal quantum number l and the mag¬
netic quantum number m (m = u + J), there are two sets of solutions
corresponding to j = l + \ and j = l — $, respectively (j equals inner
quantum number).
j = 1+\
\fri = —iMePf+i • MrFi
^2 = —iMeP^ll • MrFi 17x
J-1-*
\j/i = —i(l + u) * MePi-i' MrF-i-1.
*2 = i(l -u- 1) • MaP?^1 * MJP-u.1.
(9.18)
= ilfflP1/ • MrG-i-1.
*4 = JI^P’/*1 • MrG-l-1.
P, - (-!)■<! + .)!
where «i is the radius of the Bohr first circular orbit and a is the fine-
structure constant. The functions iPi are of the form of series
for two 'electrons with the same n, j, and m, and l * j ± J is the same.
As an example of these two theorems, the angular charge distributions,
as they are called by Hartree, for the magnetic states m =s* ± f of a *Pj
term are not only the same but are also identical with the two magnetic
states m =* ±4 of the 2Dj term. It should be pointed out, however, that
the radial charge distributions of the 2P and 2D terms are different.
With this simplification of the problem the charged distributions need
be determined for j = l + \ and positive m only.
S9
The first part of the solution is a function of the angles ip and 0 alone.
From Eqs. (9.19) and (9.29) it may be seen that the angle tp occurs
in each polynomial in the form of an exponential eim* and is always
given in the last column of Table 9.2 and graphically by the shaded
m— -fj
Energy states i m Pg X Pe
*Sh or *P* i \ 1 1
| sin* 0
*P| or 3 J(3 cos* 0 + 1) 2
V1 sin4 0
8
*£>, or V, 3 i | sin* 0(15 cos* 0 + 1) 3
1 h f(5 cos4 0 — 2 cos* 0 + 1)
i sin6 0
1 t5# sin4 0(35 cos* 0 + 1)
*Fj or *Crj I ! [8 sin* 0(21 cos4 0 — 6 cos* 0 + 1) 4
i 15(175 cos6 0 - 165 cos4 0+45
cos* 0 + 9)
areas under the straight lines in Fig. 9.7. Since the probability density
for the negative m states is the same as that for the corresponding
positive m states, the sum of the probability densities for all negative m
states is also a constant. This means that four electrons in 2P* states
or two electrons in 2P\ states will form a spherically symmetrical charge
distribution.2 Another and similar consequence of the Dirac theory
is that a single electron in a 2P\ state is two electrons in a 2Pj state
present spherical symmetry. Not only are all S states, formed from a
_ (N + p + nrWW+'V&p + nr + 1)_ ,Q ^
r nr\2N[Y(2p + l)]2[(Ar + Z + l)2 + nr(nr + 2p)f
which from tables of the gamma function are readily evaluated. The
radial function Fj, as compared with G], is extremely small and for
hydrogen is of the order of magnitude of the square of the fine-struckire
constant. If a is set equal to zero throughout the radial equations
the gamma functions become simple factorials, F\ vanishes, and Pr
reduces to the radial factor of the Schrodinger theory (Rn,i)2 (see Chap.
IV)
Since the radial densities on the Schrodinger and Dirac theories
are so nearly identical, it is difficult to show their differences graphically
Sac. 9.7] HYDROGEN FINE STRUCTURE 145
However, by splitting up Mf(F\ + Gf) into tfro parts M*Ff and MfGf,
curves may be given for each on different scales. In Fig. 9.9, for example,
curves for the 4p, 2P? state are drawn. The factors MfG\ and M*F\
are shown by the dotted lines in the top and middle figures, respectively.
Multiplying by 4xr2 the density-distribution curves (shaded areas) are
obtained. The main reason for representing the density distribution
in this way is to show that the zero points of each of the two upper
Fiq. 9.9.—The radial factor Pr of the probability density for the state 4p,*P|.
The spin correction (middle figure) added to the Schrodinger distribution (upper figure)
gives the Dirac distribution (lower figure).
. i.it
m-+ior-£ 22P^ m-+j or-J [32D3/^] m•+ Jorj> 2gP% [3gDfo] m-+Jor-£
Fio. 9.10.—Photographs representing the Dirac electron cloud for a number of the
simpler hydrogen-like states made by means of photographing a mechanical model. The <f>
(magnetic) axis is vertical and in the plane of the paper. The scale for each state is given
below each symbol in AngstrSm units.
so that the <p (magnetic) axis is vertical and in the plane of the paper.
While these photographs were made to represent the electron cloud
and not simply cross sections, the latitude lost in photograph copying
and printing makes them appear as cross sections. The states repre¬
sented are given beneath each figure. Since states with negative m
values are identical with the corresponding positive m states, only
one picture is given for both +m and — m. The electron cloud for each
state, given in brackets, is so nearly like the figure given for the unbrack¬
eted state that the difference is not distinguishable in such a photograph.
The scale in Angstrom units for each state is given beneath each symbol.
Graphical comparisons of the classical electron orbits, without
spin, have shown (see Chap. IV) that the electron path closely follows
i White, H. £., Phy*. RmH 88, 518, 1931.
Sec. 9.0] HYDROGEN FINE STRUCTURE 147
0» - J
a2Z2
\ + VP a2Z*)
r
where j' is a new quantum number given for the various terips by the
+
nc
(9.34)
Since f occurs as the square only, the minus sign can be neglected.
Due to this fact f2 can be replaced by (j + $)2, and Eq. (9.34) becomes
T = -^jl +
a2Z2
+ Vj+hy-ct'Z*)2 r+r
(9.35)
The first term gives the Balmer terms and the second the corrections
given by Eq. (9.14).
Problems
1. Calculate the fine-structure pattern for the electron transition n * 3 to n ■ 2
for ionized helium. Determine the wave-length at which this line is to be found, and
compare the fine-structure intervals in Angstroms with those of Ha in hydrogen.
2. Construct radial-density-distribution curves for a 4d,*D| and states
(see Fig. 9.9).
CHAPTER X
If the field is normal to and up from this page of the book, then
electrons moving in a counterclockwise direction in the plane of the
paper are speeded up by an amount Av and those moving in a clockwise
direction are slowed down by the same amount. It will now be shown
how these modified motions have been employed in giving a classical
explanation of the normal Zeeman effect.
In the following explanation we are concerned with an assembly
of electrons moving in orbits oriented at random in space. We start
by selecting one of these orbits and resolve the motion into three com¬
ponents along three mutually perpendicular axes (see Fig. 10.1a). The
1 The electrons are here moving at right angles to the field, and they therefore
experience a force at right angles to their motion the direction of which depends upon
the direction of the field and the direction of the motion.
Sec. 10.1] ZEEMAN EFFECT AND THE PASCHEN-BACK EFFECT 151
motions are now combined, as shown in Fig. 10. Id, to form plus an<J
minus resultants. Thus the motion of a single electron in a magnetic
field is represented by a linear motion along the field direction with
unchanged frequency vo and two circular motions at right angles to
this, one with the frequency v0 + Av and the other with frequency
vo — Av. On summing up such motions for all of the electrons, the
result will be the same as if one-third of the electrons are moving with
unchanged frequency along the z axis, one-third moving with a counter¬
clockwise circular motion normal to z of frequency vo + Av, and the other
third moving with clockwise circular motion normal to z of frequency
vo — Av (see Fig. 10.le).
We are now interested in the nature of the light that should classically
be radiated from these motions. When viewed in the direction of the
field, only the circular motions are observed and these as right- and
left-handed circularly polarized light (Fig. 10.IB). Since light is a
transverse wave motion, the z motions will not emit light in the field
direction. When viewed perpendicular to the field, the z motions are
observed as plane-polarized light with the electric vector vibrating
parallel to the field, and the circular motions, seen edge on, are observed
as plane-polarized light with the electric vector at right angles to the
field. A spectrum line viewed normal to H should therefore reveal
three plane-polarized components (see Fig. 10.1A), a center unshifted
line and two other lines equally displaced one on either side. This is
called a normal triplet. The abbreviation p stands for a vibration
parallel to the field and s (senkrecht) stands for a vibration normal to
the field. The experimental agreement with the direction of rotation
of the circularly polarized components is proof that the radiation is due
to moving negative electric charges.
In Zeeman’s early investigations he was not able to split any lines into
doublets or triplets, but he did find that they were widened and that
their outside edges were polarized as predicted. He was later able to
photograph the two outer components of lines in a number of the ele¬
ments, Zn, Cu, Cd, and Sn, by cutting out the p components with a
Nicol prism. Preston1 using greater dispersion and resolving power
was able to show not only that certain lines were split up into triplets
when viewed perpendicular to the field, but that others were split into
as many as four and even six components (see Fig. 10.2a). He also
pointed out that the pattern of all lines (usually called Zeeman patterns)
belonging to the same series of spectrum lines was the same and was
characteristic of that series. This is now known as Preston9s law. With
Preston’s law firmly established, the Zeeman effect has been, and still
is, a powerful tool in spectrum analysis.
,
1 Preston, T., Phil. Maa.< 46 326, 1898.
152 INTRODUCTION TO ATOMIC SPECTRA [Chap. X
m mm
mi ini wig ~~
1-1 J
Normal Triplet
I 1
Anomalous
1 I—L-J
Patterns
No field
Weak field
I—l—I » ■ ■ U-L-J
Anomalous Patterns
Fig. 10.26.—Normal and anomalous Zeeman effect. Viewed perpendicular to the magnetic
field.
man patterns showing just three lines with exactly these separations
are called normal triplets. All other line groups, as, for example, the
complex patterns observed in the chromium spectrum (see Fig. 10.2a),
Sac. 10.2] ZEEMAN EFFECT AND THE PASCHEN-BACK EFFECT 153
are said to exhibit the anomalous Zeeman effect. One of the niost impor¬
tant of the early investigations of the anomalous Zeeman effect was
carried out by Paschen and Runge.1 Each member of the principal
series of sodium, copper, and silver was observed to have 10 components
as shown in Fig. 10.26. The sharp-series triplets in mercury are still
more complicated, the strongest line in each triplet having nine com¬
ponents, the middle line six, and the weakest line only three. This
last line does not form a normal triplet since it has twice the norma)
separation.
came the development of the LaruU vector model of the atom and the
calculation of the famous Landi g factor. The accuracy with which
this model, with its empirical rules, accounted for all observed Zeeman
patterns, and predicted others which were later verified, is one of the
No Field Weak Field .
-i
-f
Fia. 10.4.—Splitting up of an energy level in a weak magnetic field. This figure is
drawn for the case where j — f.
the field direction H (see Fig. 10.5). The quantum conditions impose,d
upon this motion (see Sec. 9.6) are that the projection of the angular
momentum j*h/2r on the field direction H will take only those values
given by mh/2t, where m ±J, ±f, ±f • • • , ±j. In other words the
projection of j* on H takes half-integral values from +j to —j only.
The discrete orientations of the atom in space, and the small change
in energy due to the precession, give rise to the various discrete Zeeman
levels. While the number of these levels is determined by the mechanical
moment j*h/2iry the magnitude of j*
the separations is determined by the
field strength H and the magnetic // N\
moment /*. In field-free space an energy / N\
level is defined by the three quantum CTJ' *
numbers n, l, and j. In a weak magne- y7 ^ /
tic field an additional or fourth quantum * — -~jl^
number m is necessary to define the s -
state. \\ //'
10.3. The Magnetic Moment of a \\ //
Bound Electron.—To determine the ^ ^
magnitude of the separations between ///i\w
///t\\\^
Zeeman levels, it is essential that we //f✓V'l'W
/;jiv\ j^\\\
first determine the total magnetic / / / 1 \ \ \
moment of the atom. In the simplest A/L j \\\ _J\ \ M
case to be considered here the atom
core and nucleus will be assumed to -^_I___
have zero magnetic and mechanical 7 \]
' s
/ \i
I
\
I
\ / /
/
(10.3)
156 INTRODUCTION TO ATOMIC SPECTRA [Chap. X
This result has also been derived theoretically on the quantum mechanics
(see Chap. IX).
A schematic vector diagram of the magnetic and mechanical moments
is shown in Fig. 10.6. Here it is seen that the resultant magnetic
moment is not in line with the resultant mechanical moment j*h/2t.
Since the resultant mechanical moment is invariant, Z*, $*, m, and
M,g precess around j*. As a result of this precession, only the component
of parallel to j* contributes to the magnetic moment of the atom.
This may be seen by resolving mt9 into two components, one parallel to
j* and the other perpendicular. The perpendicular component, owing
to the continual change in direction, will average out to zero. The
parallel component may be evaluated as follows:
By Eqs. (10.2) and (10.3) m and are given as
i*h
e ergs , 0 * h e ergs
Ml = -and m* = 2 • «V • n— (10.4)
27r 2me gauss ( 27r 2mc gauss'
Making use of the vector model and the cosine law that
S*2 = 1*2 + j*2 _ 21*j* cos (Z*j*), (10.8)
we obtain
<1*2 4_ 7*2 _ 0*2
l* COS (IT) = 1 2j* (10.9)
Similarly,
n*2 7*2 _L o*2
8* COS (s*j*) = 3 2j* (10.10)
_ , , id + i) + «(• + i) - id + l)
(10.12)
* = 1 +-MTT)-
The importance of this g factor cannot be overemphasized, for it
gives directly the relative separations of the Zeeman levels for the
different terms.1 We shall now see how this comes about.
10.4. Magnetic Interaction Energy.—By Eqs. (10.6) and (10.7) the
ratio between the total magnetic and mechanical moments of the atom,
y, and p„ is just
M/ __
(10.13)
p, 9 2me1
where p, = j*h/2ir.2 The precession of j* around H is the result of a
torque acting on both l* and s*. Due to the electron's anomalous spin
magnetic moment, s* tends to precess twice as fast around H as does Z*.
If the field is not too strong, the coupling between Z* and s* is sufficiently
strong to maintain a constant j*, so that this resultant precesses with a
compromise angular velocity, by Larmor's theorem [Eq. (3.58)], given
by g times the orbital precession angular velocity
■- - <ioi4>
The total energy of the precession is given by the precessional angular
velocity wL times the component of the resultant mechanical moment
j*h/2ir on the axis of rotation H :3 * * * * *
1 The values of g given by Eq. (10.12) are exactly the same as those given by
Land6’s model.
* In any experiment like the Stern-Gerlach experiment (Zeits. /. Phys8, 110,
1922), performed for the purpose of determining the magnetic and mechanical moment
of the atom, the moments and pj are oriented at some angle with the field just as
in the Zeeman effect (see Fig. 10.7). What one measures in this experiment is the
component y of the resultant magnetic moment along H. By theory we say the
component of /xy will be y *= m, cos (,j*H), and the component of j*h/2x along H will
be mJt/2v, where m takes values differing from each other by unity from m = +j
tom = — j.
9 The magnetic energy can be considered as the energy of a permanent magnet of
moment y, at an angle 9, in the field H, or as the added kinetic energy of the electron's
orbital motion. If, in the case of a circular orbit normal to the field, E represents
the kinetic energy before the field is applied, and E' = £/(<■> + «l)* the kinetic energy
after, then the change in energy is just AE ** E' — E = J/wJ, 4- /««£. Since the
added field does not change the size of the orbit, I remains constant. With <•> > >
the first term iB negligibly small and the energy change is given by the product of the
mechanical moment, /« « j*h/2*} and «l.
158 INTRODUCTION TO ATOMIC SPECTRA [Chap. X
AW Krp He
(10.17)
Table 10.1.—The Land£ g Factors and the Splitting Factors mg for Doublet
Terms
1 Term 9 mg
0 ! ±1
'Pi I ±i
1 'Pi I ±i, ±*
1 ±i, ±i
2 t ±1, ±1, ±¥
V, f ±4, ±1, ±¥
3 »F, 4 ±4, ±¥, ±¥, ±V
1 ±4, ±¥, ±¥. ±V
4 •0| ¥ ±t, ±¥. ±¥> ±¥. ±¥
1 Here m, the magnetic quantum number in the numerator, must not be confused
with m, the mass of the electron in the denominator.
Sec. 10.6] ZEEMAN EFFECT AND THE PASCHEN-BACK EFFECT 159
just derived from the Zeeman splitting, the complete term value T of
any magnetic level may be written
T = To - T - mg • L. (10.19)
Am = 0, ±1. (10.20)
For the stronger of the two field-free lines there are six allowed transi¬
tions and two forbidden transitions. For the other line there are four
160 INTRODUCTION TO ATOMIC SPECTRA [Chap. X
8 2
%
r~ -
V r- % —2—
w ■%
a % -%
V - - - ~ - - -
4
-4—1 -1
4
-
-'/j - ^
■a
r- o
- - - - ~
and Omstein (see Sec. 8.3) are readily shown to follow directly from
the intensity rules for the same levels in a weak magnetic field. In
short these rules may be stated as follows:
The sum of all the transitions starting from any initial Zeeman level
is equal to the sum of all transitions leaving any other level having the same
n and l values. The sum of all transitions arriving at any Zeeman level
is equal to the sum of all transitions arriving at any other level having the
same n and l values.
For any given field-free spectrum line these rules are better expressed
in terms of formulas which have been derived from the classical1 as well
1 For a classical derivation of the intensity rules, based on Bohr’s correspondence
principle, see J. H. Van Vleck, “Quantum Principles and Line Spectra,” Nat. Research
Council, Bull., Vol. 10, 1926.
Sec. 10.6] ZEEMAN EFFECT AND THE PASCHEN-BACK EFFECT 161
A and B are constants that need not be determined for relative intensities
within each Zeeman pattern. These formulas take into account the
. i
1 1
_ Zpeman Patterns _
For Doublets
\ Vi '1
Yi
•
1 «
1, ,
l1M
2c 2j\
fi ri r Vi 11 r,ir i •
Vi
i ii 1
i r * ,
[^
ft
V f | Vi
n 11
Fia. 10.9.— Zeeman patterns for all of the commonly observed doublet transitions. The
dots represent normal triplet separations.
m “ \ i i ~~2 ~ £
mg initial state ^ 1 $ $ —| —$ —
mg final state
\JX1XIX!^
£ f —f —£
fields the internal motions are greatly perturbed and the atom gives
rise to the so-called Paschen-Back effect.
Just as the doublet fine-structure separations are a measure of the
classical frequency with which l* and 8* precess around their resultant
j* (see Sec. 10.4), so the Zeeman separations of the same energy states
in a weak magnetic field are a measure of the frequency with which j*
precesses around H. In calculating
the Zeeman separations in Sec. 10.4, it
was tacitly assumed that the precession _is*
of l* and s* around j* was much faster
than that of j* around H. This was
necessary in order that the components
of l* and s* normal to j* average to
zero and do not appreciably perturb
the other precession. If now the field
H is increased until the two precessions
are of the same order of magnitude,
then the Zeeman levels of the doublet P fiddu
will begin to overlap, there will be no so strong that 1+ and**precessindepend-
averaging to zero, and Eqs. (10.17) ently around the field direction H.
and (10.19) will not hold. Under these conditions the coupling between
l* and s* will be partially broken down, the classical motions of l* and
8* will become complicated, and j* will no longer be fixed in magnitude.
As the field H is still further increased, l* and s* will soon become
Since the ratio between the magnetic and mechanical moment for the
spin of the electron is twice the orbital ratio, s* should, on the classical
picture, precess twice as fast as l*. Multiplying each of these angular
velocities by the projection of the angular momentum on H [see Eq.
(10.15)], one gets the first two terms of the energy:
He
-ATh = (mt + 2m,)4-^ cm"1, (10.27)
To this magnetic energy the small correction term due to the inter¬
action between l* and s* must be added. Although these two vectors
precess independently around H, each motion still produces a magnetic
field at the electron which perturbs the motion of the other. This
interaction energy, though small as compared with that due to the
external field, is of the same order of magnitude as the fine-structure
Sac. 10.8] ZEEMAN EFFECT AND THE PASCHEN-BACK EFFECT . 105
We may now write down a general relation for the term value of any
strong field level,
Tem-1 = To - {mt + 2m.)L — amtm9i (10.35)
m =
Term m 0 mg mi m. ' mi -f- 2m, am\m$
mi 4- rn.
i
+! +» +1 +* +» +2 -ha/2
-hi 0 +1 +* +1 0
Vi t
-i -i -1 +i -i 0 -a/2
-i -i +1 -i 0 -a/2
+1 +* 0 -i -i -1 0
—a i
-i -i -1 -i -! -2 -ha/2
m +i
2
+1 0 +i -hi +1 0
M -i -1 0 -i -i -1 0
fields, therefore, arises, and one asks, how does each weak-field level
go over to a corresponding strong-field level? Darwin’s treatment
of this problem, which will not be given here, answers this question in a
very simple manner.1 According to the classical law of the conservation
of angular momentum, the sum of the projections on H of the various
angular-momentum vectors must remain the same for all field strengths.
Since in weak field this sum is given by m and in strong field by mi + ra„
we may write, as part of the correlation rule, m = mi + m,. This
alone is not sufficient to correlate all weak- and strong-field levels, since
in most instances there will be more than one level with the same m
value. The more specific rule, in keeping with the quantum mechanics,
may be stated as follows: Levels with the same m never cross.
An ingenious method for obtaining the same correlation has been
given by Breit.2 An array of weak- and strong-field quantum numbers
Fiq. 10.12.—Energy levels for a principal-series doublet starting with no field at the
left and ending with a strong field (Paschen-Back effect) at the right. Allowed transitions
are shown below.
mi and m,. These sums are the weak-field quantum numbers, divided
into two parts by the dotted lines. Each weak-field level m is to be
correlated with the strong-field level given by the value of mi directly
above, and the value of mt directly to the right of the m value. The
2Pj, m = $ state, for example, goes to the state mi = 1, and m, = $.
P d
nr>| » 1 0 -1 nrije 2 1 0-1-2
m» 1 ± -± * m_-
l
m» m- f i-i -!]-!
* -j: -i
ms ms
*•» i%
Fia. 10.13.—Correlation of weak- and ig-field quantum numbers and energy levels.
Brett.)
It is obvious that there are two ways of drawing the L-shaped dotted
line. Of the two ways only the one shown will give the correct correla¬
tion for doublets from a single electron.
168 INTRODUCTION TO ATOMIC SPECTRA [Chap. X
six times as wide as the field-free line. A photo- published paper of Pas-
graphic reproduction of this triplet is given in Back. ()>) Same
Fig. 10.15. Normal triplets have also been observed paper moved parSiei to
by Paschen and Back for Hp and H7. lines during enlargement.
10.11. A Quantum-mechanical Model of the Atom in a Strong Mag¬
netic Field.—In this chapter the Zeeman effect, as well as the Paschen-
Back effect, has been treated chiefly from the standpoint of the semi-
classical vector model. Quantum-mechanical treatments of the same
problems have been made by different investigators and found to lead
to exactly the same formula. Chronologically, the more accurate
quantum mechanics led the way to a simpler formulation of Landd’s
vector model. We have seen in Chaps. IV and IX how, in field-free
space, this model is surprisingly similar to the quantum-mechanical
model of probability-density distributions for the electron. As would be
expected from energy relations, the weak- and strong-field distributions
for the electron on any model should be little different from each other
1 Paschen, F., and E. Back, Ann. d. Phys., 39, 897, 1912. The photograph in
Fig. 10.15 is a copy of the photograph given in Plate VIII of Ann. d. Phys., Vol. 39,
1912.
170 INTRODUCTION TO ATOMIC SPECTRA IChap. X
Problems
1. Compute the Zeeman pattern (separations and intensities) for the doublet
transition
.
2 Find the total width in wave numbers of the Zeeman pattern of Prob. 1 in a
weak field of 5000 gauss.
3. Compute the weak- and strong-field energies for a diffuse-series doublet, and
tabulate them as in Table 10.2. Plot the initial and final states, as shown in Fig. 10.12,
and indicate the allowed transitions by arrows.
.
4 Plot, as in Fig. 10.14, the field-free lines, the weak-field lines, and the strong-
field lines of the above example. [Note:—Certain components of the forbidden
transition appear in strong fields and should be indicated (see Fig. 13.14)].
5. What field strength would be required to carry the first member of the principal
series of sodium over to the Paschen-Back effect where the separation of the resultant
normal triplet (see Fig. 10.14) is four times the fine-structure separation of the field-
free doublet?
CHAPTER XI
composed of three lines with the same separations and approach a triple
limit; (2) all members of the principal series are composed of three lines
with decreasing separations and approach a single limit; (3) all members
of the diffuse and fundamental series contain six lines, three strong lines
and three satellites, and approach triple limits.
Fig. 11.1.—Energy level diagram for the four chief singlet and triplet series of the neutral
calcium atom. For complete diagram, see Fig 11.9.
As in the doublet series of the alkali metals nearly all of the observed
singlet or triplet series are closely represented by the general series
formula
R_R
(ni - juj)2 (n2 — m)2’
(11.1)
(11.2)
Such a diagram is given for calcium in Fig. 11.1. / This same energy level
diagram may be arrived at in another way.
If all of the limits of the eight chief series of calcium (Fig. 1.8) are
brought together into one common limit, the spectrum lines, including
3P ===== 3D 1 *F -
3p
3D 3F
3P
3P 3D
Fig. 11.2.—Schematic representation of the triplet fine structure of the first four members
of the three chief 3P, 3Z), and 3F term series.
the ones plotted as negative, become the energy levels of Fig. 11.1.
Taking into account the fine structure of the triplet series, shown
in Fig. 1.10, this process transforms the triplet sharp-series lines into
single levels and the three remaining triplet series into series of con-
Electron
Terms n = 3 71—4 71—5 n — 6 n = 7 n = 8
configuration
verging triple levels (see Fig. 11.2). Although the so-called *S levels
are single, their triplet nature will be brought out when the atom is
placed in a magnetic field (see Chap. XIII). The assignment of the
total quantum numbers n to the valence electrons in the different levels
7195 7392 7905
Fia. 11.3.—Photographs of triplets taken from original spectrograms kindly loaned to
the author by Dr. A S. King
1
Be 4 2 36 0 69 Unresolved
Mg 12 40 92 19.89 Unresolved
Zn 30 388.90 189 8 5 5 3.4
Cd 48 1171.00 541 8 18 2 11 7
Hg 80 4630.60 1767.3 35 1 60.0
Ca 20 105 9 52 30 21 7 13 9
Sr 38 394.4 187.1 100.3 60 0
Ba 56 878.2 370.6 381.1 181.5
A brief study of the diffuse and fundamental triplets will show the
reader that certain general selection rules are in operation. These
rules are most easily stated in terms of the quantum numbers or quantum
values of the different energy states. For various types of levels the
capital letter L is almost universally used and assigned the following
quantum values:
L = 0, 1, 2, 3, 4, 5,
for S, P, D, F, G, H, . . . terms.
The subscripts to the right of each term are the inner quantum numbers
J and always take the following values:
'So 'Pi 'D2 'F 3 'Ga 'Hi
The allowed transitions for a number of triplets are shown in Fig. 11.4.
Relative separations and intensities are shown directly below the arrows
by the respective separations and heights of the lines. Qualitatively
the following rules for the intensities are found to hold experimentally:
The stronger lines arise where AL changes in the same direction as AJ.
These are shown by the heavier arrows. Of these the strongest line
arises from the transition involving the largest values of L and J. These
two well-known rules should be memorized, for they are quite generally
applicable in complex spectra. It should be pointed out in passing that
small letters s, Z, and j are used for the quantum values of individual
electrons, whereas capital letters S, L, and J are used to designate terms
in all atomic systems where there are one or more valence electrons.
11.3. The Quantum Numbers n and l of Both Valence Electrons.—In
order to correlate energy levels with the various types of electron orbits
we return to the Bohr-Stoner scheme of the building up of the elements
and the normal states of the atom. According to Table 5.2, the complete
normal electron configurations for the neutral atoms of Groups IIA and
[IB (Table 5.1) are as follows:
T = 7 ff)2, (11.8)
n'1
where Z is the atomic number and a is a screening constant, and on the
ether hand the Rydberg formula
T
zi (11.4)
(n — m)*
Sec. 11.5] SINGLET AND TRIPLET SERIES 179
Fiu 11.6.—Energy level diagrams of beryllium, magnesium, zinc, cadmium, and mercury.
(Hydrogen comparison.)
4* nd lDi
*Dt
Hydrogen.. 1.0 0.0 1.0 0.0 1.0 0.0 1.0 0.0 1.0 0.0
from three triplet terms which have no place in the ordinary chief series
of terms and end on the lowest 3D 1,2,3 of the diffuse series. The 3Z>
intervals, common to all three multiplets, are shown by brackets in the
frequency plot. It is to be noted that each multiplet is composed of
Vv _3pO
,du/3f2°34 3DUjf D1°2.3 UU.5 r0,U
CD'V O) O lO CM
<7> 0> GO vo «£>
ID LO if) CM CM CM
ID ID lO if) if) UO
GREEN GREEN
Fig. 11.7.—Photographs and frequency plot of the great calcium triad of triplet multiplets
one in the red region of the spectrum and the other two in the green.
three relatively strong lines and three or four fainter lines. The observed
intensities indicate that the inner quantum numbers of the upper triad of
Sac. 11.5] SINGLET AND TRIPLET SERIES 181
triplet levels are those of sP0li.*, and */Ys,4 terms. The wave-'
lengths, frequencies, and estimated relative intensities are given in
multiplet form in Table 11.5.
Table 11.5.—Relative Intensities, Wave-lengths, and Frequencies of thb
Great Calcium Triad
(Given in multiplet form)
150
*f< 6439.09
16626.87
78.15
40 125
6471.66 6462.58
16447.72 15469.44
88.28
1 30 80
6508.84 6499.65 6493.79
15859.40 15381.20 15395.08
80 25
5588.74 5581.97
17888.16 17909.85
40.01
30 60 20
*Di 5601.28 5594.46 5590.11
17848.11 17869.87 17883.79
26.73
i 20 50
3/>i 5602.83 5598.48
17843 18 17867.03
60 20 2
5270.27 5264.24 5260.38
18969.08 18990.89 19004.77
4.80
40 20
'Pi 5265.56 5261.70
18986.07 18999.99
1.94
26
*Po 5262.24
18998.05
Vp'dVp 3d
1 The suggestion that two electrons are excited to higher energy states was first
made by Bohr and Wentzel, Phys. Zeits., 24,106,1923, to account for wide separations
in the anomalous spectral terms in calcium and strontium.
Sec. 11.5] SINGLET AND TRIPLET SERIES 183
one of the two valence electrons is in a 3d orbit, while the other occupies
successively the states 3d, 4d, 5d, 6d, etc. With one of the electrons tak¬
ing on larger and larger n values (n—> »), the atom will finally be left
ionized with one electron in a 3d state. The series limit of the anomalous
ZP terms therefore becomes an excited state of the first spark spectrum,
Ca II, and lies above the common limit of the chief series.
Returning to the energy level diagram of Ca II (Fig. 6.4), the normal
state is seen to be in the new notation 4s2S, followed by 3d2D and
4p2P. The first excited state (the metastable state 2D) lies 13711 cm""1
above the normal state 2S. The series limit of the anomalous SP terms
is calculated by Russell and Saunders to be 13961 cm""1 above 2S. The
agreement is so remarkable that there is little doubt that the two are
one and the same.
The question now arises as to how two electrons in d orbits give
rise to 3P terms, or how one electron in a 4p orbit and the other in a 3d
orbit give rise to a triad of triplets like the ones given above. The
answer to this question will serve as a starting point for the next chapter.
Problems
1. With the term values of the chief senes in strontium given, plot an energy level
diagram similar to the one shown in Fig. 11.1. Indicate the hydrogen levels with
dotted lines. For term values see Bacher and Goudsmit, “Atomic Energy States.”
2. Compute values of Z — a and the quantum defect m for the first four members
of the chief term series in strontium. Show which terms indicate penetrating and
nonpenetrating orbits (see Table 11.4).
CHAPTER XII
l — 0, 1, 2, 3, 4, 5, 6, * * *
for s, p, d, ft g, h, i, . . . electrons, respectively,3
Fia. 12.1.—Vector diagrams of the space quantizat of the orbital motions of two
valence electrons.
Table 12.1.
S-8, S ps, P ds, D /*. F
sp, P P’Pf SPD dp, PDF fp, DFG
sd, D Pdy PDF dd, SPDFG fd, PDFGH
*/> F Pfl DFG df, PDFGH //, SPDFGHI
The dot between two electrons of the same type indicates that the
electrons have different total quantum numbers. If they are alike
certain of the terms shown are not allowed. (Such special cases will be
treated in Sec. 13.11.) Where l2 < h the roles of the two electrons are
interchanged.
Returning at this point to the discussion of calcium in Sec. 11.5, we
observe from Table 12.1 how it comes about that a P term may arise
from two electrons in d orbits. It is also observed how with-only two
valence electrons, one in a p orbit and the other in a d orbit, the atom
may be in a 3P, a 3D, or a SF state. In
SPIN-SPIN-COUPLING
calcium these three terms, combining with
another 3D term arising from the con¬
S,* S?
s‘ /•*
r7 figuration 4s3d, give rise to the three great
triplet multiplets of Fig. 11.7.
12.2. Spin-spin-, or ss-coupling.—
With two electrons, each having a spin
angular momentum of s*h/2tt, where
s* = \/s(s + 1) and s = J, there are two
S,«*2 s*-'/2 S|*k s2-J$
ways in which a spin resultant S*h/2ir may
s 0 S-l
Fig. 12.2. -Vector diagrams of
be formed. Let s* and s* represent the
the space quantisation of the spins of respective spin vectors of the two electrons,
two valence electrons. Quantizing these (see Fig. 12.2), we find,
with sf = £\/3 and sj = two resultants, one with S* = 0, and the
other with S* = y/2. These give the resultant quantum values S = 0
and S = 1. The resultant S* = 0 will now be shown to give rise to
singlet terms and S* = y/2 to triplet terms.
12.3. LS-, or Russell-Saunders Coupling.—With the orbital motions
of two electrons coupled together to give a resultant L*f and the spins
of the same electrons coupled together to form S*, both L* and S*
will in turn be coupled together to form «/*, which is a vector representing
J*h/2v the total angular momentum of the atom. The quantum con¬
ditions imposed upon this coupling are that J* = y/J(J + 1), and that
J take non-negative integral values. Consider, as a specific example, the
case of one electron in a p orbit and the other in a d orbit. From Secs.
12.1 and 12.2 we have the following possible values to work with; L -- 1
Sec. 12.3] THE ATOM MODEL FOR TWO VALENCE ELECTRONS 187
/
With these term symbols it may now be pointed out that the super¬
script, which we have already used, expressing the multiplicity of the
fine-structure system to which the term belongs, is always 2S + 1.
o
L-1S-0 L-2 S-0 L-3 S-0 L"1 S’l L-2 S*1 L-3 S-l
J-l J-2 J-3 J "0,1,2 J“ 1,2,3 J-2,3.4
Fiq. 12.3.—Vector diagrams of two valence electrons in LS- (Russell-Saunders) coupling.
be used for the rapid calculation of J values, when they are not remem¬
bered. With given values of L and S} (1) where L ^ S, all integral
values of J between L — S and L -f S are allowed, or, (2) where S ^ L,
all integral values of J between S — L and S + L are allowed. All
allowed terms arising from some of the more common electron con¬
figurations are given in Table 12.2.
Terms for which the sum h + U is odd are called odd terms, all others
are called even terms. Odd terms are distinguished by the small super¬
script “°” to the right of the term symbol S, P, D . . . . Thus we
have completed the explanation and meaning of the modem spectro-
188 INTRODUCTION TO ATOMIC SPECTRA [Chap. XII
Initial states 3d4d ^So lPi lD2 XF3 lGi 3£i 3Po,i,2 3Z>i,2.3 3F2f3.4 ^3.4.5
The quantum values, i.e., the term types assigned each level are
determined by selection rules for complex spectra (see Sec. 12.14).
The upper set is composed of a pentad of singlet and triplet levels and the
lower set of a triad of singlet and triplet levels. It should be noted
that the selection rule for electrons is obeyed in that one electron goes
from a 4d to a 4p orbit, Al = 1, while the other remains in a 3d orbit.
12.6. Coupling Schemes for Two Electrons.—Attempts to calculate
the fine-structure separations of the various energy levels arising from
any given electron configuration have been made by many investigators.
Prior to the development of the quantum mechanics, the vector model,
qualitatively, accounted for the fine structure of all analyzed spectra.
Although quantitative agreement was also found in a few cases, it
remained for the quantum mechanics to bring about a more general
agreement. It should be pointed out that, even at this time, there is
much to be desired especially in the case of the more complex atomic
systems. We shall first consider, in detail, the vector model as it applies
to atoms containing two electrons and then formulate a quantum-
mechanical picture of the same atomic states.
In hydrogen and the alkali metals, the simplest of all atomic systems,
the interaction between the electron spin $* and the orbit l* of the
190 INTRODUCTION TO ATOMIC SPECTRA [Chap. XII
In the case of two valence electrons there are four angular momenta
If, If, sf, and s% which give rise to the following six possible interactions:
Applying Eq. (12.1) to these six interactions, there will be six energy
relations:
ri = aisfsf cos (sfs*)! r3 = ajfsf cos (Ifsf), r& = abZfs? cos (Ifsf),
(12.4)
r2 = adflf cos (Iflf), r4 = ad*2sf cos (l*s%), r6 = a6l*s* cos (Ifsf).
J* J*
Fia. 12.5.—Ideal vector models for (A) LS-coupling, and (B) jy-coupling. Examples
drawn with lx * 1, U *= 2, / = 3.
r -
jf
T>, mmmm
i'/o*
SINGLET r -*Pi SINGLETS-
I
1
■31
-2A
Sp—+T0- pd-^T0-7^ “H
•* A
°A ♦A Ay #. —---A1-
Element Element
Obs. Obs.
Configura¬ 3Po - *Pi »Pl - *P! Configura¬ - *D* *D, - *D,
ratio ratio
tion tion
Ca, 3d3d 13.5 26.9 2.0 Ca, 3d48 13.6 21.7 1.6
Ca, 4s4p 52.3 105.9 2.0 Ca, 3d4p 26.7 40.0 1.6
Sr, 5«5p 187.0 394.6 2.1 Zn, 4s4d 3.4 4.6 1.4
Mg, 3s3p 20 0 40.9 2.0 Cd, 5s5d 11.7 18.2 1.6
Zn, 4s4p 190.0 389.0 2.0 Ca, 4s4d 3.8 5.6 1.6
taken from one group of elements in the periodic table are given in Table
12.4.
The agreement with theory clearly justifies the use of the vector
model and identifies the electron coupling as Russell-Saunders.
A graphical representation of the Land6 interval rule is shown in
Fig. 12.7 for a 3D term. With
S = landL = 2, S* is shown quan¬
tized with respect to L* in the three
allowed positions J = 1, 2, and 3.
The J vectors are not shown.
Projecting S* on L*, we obtain S*
cos (L *S *). The resultant intervals
have the ratio 2:3, and the central
dotted line is at the center of
gravity.
The chief characteristics by
which LS-coupling is readily recog¬
nized are: (1) The singlet-triplet
intervals are large compared with
the triplet fine structure. (2) The Fiq. 12.7.—Graphical representation of the
triplet intervals follow the Land6 Land6 interval rule for a *D term.
interval rule. Both of these conditions are observed when the coupling
is LS.
12.9. jj-Coupling.—We now turn to the ideal case of ^'-coupling for
two electrons. Before calculating general expressions for the inter¬
action energies on such a scheme, let us first determine all of the various
196 INTRODUCTION TO ATOMIC SPECTRA [Chap. XII
Ti * ojsfsjf cos (s?sj) = aisfs* cos (sfjf) cos (j*jt) cos (jjs?) (12.19)
and, similarly,
Ti = ajffi cos (1*1*) = OiZfZJ cos (Jfr'f) cos (j*j*) cos (jtl*). (12.20)
Each of these cosines is constant and readily evaluated with the cosine
law:
Ti + r2 = i(ai/3i + atEt){J*2 - jV - jV\, (12.21)
where
R ay + jv - iy a? + j? - iy .
(12.22)
Pl 2jV 2jt*
and
_ i? + jy - «p is* + jy - s*t\ (12.23)
Pi 2j? 2jV
Writing
A = Qifii + (I2P2, (12.24)
Eq. (12.21) may be written
shown by the formulas, that the large interaction is due to r8, the spin-
orbit interaction of the p electron; the s electron, with h = 0, contributes
nothing (ri = 0). That this splitting is due to the p electron is evidenced
by the 5p,2P|,j separation, observed in the ionized spectrum where the
6s electron is absent. It is seen that the effect of the addition of a 6s
electron to Sn II, 6p, to form Sn I, 5p6s, is to split each doublet level
into two levels as calculated. This smaller interaction is due to the
spin-spin coupling IY If this latter splitting were due to the interaction
between h and s8 [see Eq. (12.4)], the two upper terms would be inverted.
All of these electrons, with the exception of the last two similar p electrons,
form completed subshells. Two similar p electrons (see Table 12.3)
give rise to the terms ^So, *P<u,2, lDt. Applying Hund's rule, which is
always valid for normal electron configurations, the 3P0,i,2 terms should
lie deepest, followed by xDi and then %, as observed.
Exciting the atom now by raising one electron to the lowest possible
available state 5s, changes the configuration to 4p5s. From Table 12.2
Sue. 12.10] THE ATOM MODEL FOR TWO VALENCE ELECTRONS 199
these two electrons give rise to lPi and 3Po,i,s terms. On examining
the fine-structure intervals, it is observed that the coupling is not LS
but very closely jj. These four terms constitute the first member of a
series of terms arising from the series of configurations 4pns, where
likewise form series, which have as a limit the 4p,*P*.| state of the
ionized atom. Subsequent excitation of the ionized atom will raise the
one remaining electron to other doublet states, which as series have a
single limit, lSo. When this limit is reached, the atom is doubly ionized
and 30 electrons still remain in completed subshells.
In taking up each one of the carbon group of elements in detail,
good examples of intermediate coupling between LS- and ^-coupling
are found. Consider the first member of the 4pns series mentioned
above. In carbon this first member, due to the configuration 2p3s,
i,
shows good LjS-coupling; the 8P0 — 8P 3Pi — 3Pi intervals are 20 and 40
cm"1, with the theoretical ratio 1:2 and the lPi — 3P2 interval 1589 cm"1
Carbon Silicon Germanium Tin Lead Carbon Silicon Germanium Tin Lead
Jj
CoudmS
4p5s 5p6s \\ 6p7s
Fig. 12.12.—LS- to .//-coupling as shown by the normal and first excited states in the
carbon group of elements.
is correspondingly large. These levels are shown at the left of the right-
hand figure in Fig. 12.12. In lead the same four levels have gone over
to ^-coupling with two groups of two levels rather far apart. The
intermediate elements (Si, Ge, and Sn) furnish transition stages between
the two extreme cases. While jj-coupling for similar electrons has
not been treated above, the terms making up the normal states of these
atoms show LS-coupling in C and relatively good jj-coupling in Pb.
12.11. Term Series and Limits in Two-electron Systems.—The
fine structure of two series of singlet and triplet terms is shown graphically
in Fig. 12.13. Fine-structure intervals are plotted vertically, and the
term values, on a very much smaller scale, horizontally. The cadmium
series, it is observed, shows fairly good LS-coupling as far as the series
is observed. The silicon series, on the other hand, shows fairly good
L/S-coupling for the first series member and goes over to good ^-coupling
as the series approaches the double limit.
In the cadmium series 5snp, the 5s electron remains in the Same
tightly bound state while the p electron takes on one of the series of
more and more loosely bound states 5p, 6p, 7p, . . . The limit
Sec. 12.121 THE ATOM MODEL FOR TWO VALENCE ELECTRONS 201
CADMIUM SILICON
Fig. 12.13.—Two types of term series and series limits.
along with the jj-coupling scheme developed in Fig. 12.8, will show what
levels go to each limit.
One interesting result brought out in Fig. 12.13 is that the over¬
all separation of each of the 3P terms in silicon is approximately the
same as that of the 2P limit. This quite frequent observation occurs
when one of the electrons is an s electron. For a ps configuration in
L£-coupiing (see Fig. 12.6) the 3P0 — interval is calculated to be
3A, or 3a3/2. In ^/-coupling the same interval (see Fig. 12.10) is again
3a3/2. For the ionized atom the 2P\ — 2P% interval is given by Eq.
(8.18) (see also Fig. 8.9) as 3a/2. In cases where the $ electron does not
appreciably change the effective nuclear charge for the p electron, the
a’s in each case should be nearly the same. These conditions are met
in a number of other atoms, and the same phenomenon is observed in
the following series, in C I, N II, and O III, 2pns, *P, where n = 3, 4,
5, . . . oo; in Si I and P II, 3pns, 3P, where n = 4, 5, 6, • • • oo; in
Sc II, Ti III, V, IV, and Cr V, 3dns, 3D, where n = 4, 5, 6, • • • oo.
12.12. The Great Calcium Triad.—Sufficient preparation has now
been made to enable us to return to the energy levels of the well-known
calcium triad and to calculate the width of each triplet in the upper set
1 Hund, F., Zeits. f. Phys., 52, 601, 1928, and previous papers.
,Shbnstonei A. G., Nature, 121, 619, 1928; 122, 727, 1928.
202 INTRODUCTION TO ATOMIC SPECTRA {Chap. XII
of levels. The relative positions of the 8P, *Df 8F, lP, lD, and ]F terms
arising from the electron configuration 3d4p may be seen in the energy
level diagram of calcium (Fig. 11.9). Assuming LS-coupling for this
configuration the triplet intervals may be calculated with a reasonable
degree of accuracy. With the treatment of LS-coupling in Sec. 12.7, we
have only to determine the values of two coefficients, <z8 and a4, in the
A’s at the bottom of Fig. 12.6:
—, 3a4#
*P, A = 4 1- 4 »
CLl . 5o4 #
*£, A' (12.26)
12 + T2;
*F, A" = ?i _L. —
6 3*
Since a8 and o4 are both due to the interaction between the spin
of each electron and its own orbit, they may be evaluated directly from
the electron configurations 4p4s and 3d4s, respectively. Here with one
electron in an 8 orbit the observed triplet separations 4p4s,3P and 3d4s,
lD are due to the 4p and 3d electrons, respectively. The following table
gives the observed intervals and the calculated coefficients [see Eq.
(12.9)].
Table 12.5.—Coefficients for 4p and 3d Electrons in Calcium
3449 14 21 2A 3A aa = 14
II
Using these values of 105 cm-1 and 14 cm-1, for a3 and a4, in the
preceding Eq. (12.26), the following values for the 8P, 3D, and 3F triad are
obtained:
Observed.
Calculated.
166 cm"1
156 HH
Although the intervals throughout this configuration do not indicate
ideal LS-coupling, the agreement between observed and calculated
intervals is quite good.
12.18. The Branching Rule.—The branching rule is a very simple
rule frequently used in the construction of energy level diagrams.
Suppose, for example, that the atom of germanium has been doubly
ionised. This state of the atom is represented in Fig. 12.11 by the
Sec. 12.14] THE ATOM MODEL FOR TWO VALENCE ELECTRONS 203
term at the top of the diagram. Now allow one electron to return to
the atom and end up in a 4p orbit. This state, 4p,2P, of the atom is
represented in the middle of the diagram as the lowest doublet state
of the singly ionized atom and at the same time the doublet limit of
the various series of the neutral atom. Suppose now that the second
electron is allowed to return to the atom in a 5s, or a 5p, or a 4d orbit.
The various possible states of the atom corresponding to these configure
tions may be written down as follows:
4p 4p 4p
Ionized atom
2P ip 2P
A
1P 3P 'S'P'D 3S3P3D lPhDlF 3P3D3F
Neutral atom
4p5s 4 p5p 4p4d
3d 3d 3d
Ionized atom
2D 2D 2D
A
lD 3D
^/\
ipiD'F 3P3D3F 'S'P'D'F'G *S*P*D*F*G
Neutral atom
3dis 3dip 3did
The 4p2P and 3d2D terms in these two atoms are known as parent
terms to the ones below them. If the L values of a parent term is equal
to or greater than the l of the added electron, the allowed terms go from
L — l to L + 1. If l > L, they go from l — L to l + L (see Sec. 12.1
and Fig. 12.1).
12.14. Selection Rules.—An extension of the selection rules of
hydrogen and the alkali metals to two-electron systems introduces
new sets of rules, of which those for one-electron spectra may be thought
of as special cases. With two electrons taking part in producing
204 INTRODUCTION TO ATOMIC SPECTRA [Chap. XII
the various types of terms, transitions may occur in which two electrons
jump simultaneously with the emission of a single radiated frequency.
Selection rules for two-electron systems in general may be written as
If a single electron jumps, the l value of one changes by unity, and the
other by zero. If a double electron jump occurs, the l value of one
changes by unity and the other by zero or by two. There are no restric¬
tions on the total quantum number n of either electron. For the various
types of terms arising from all possible electron configurations the
further restrictions are divided into two parts.
A. For L/S-coupling the further restrictions are:
AS = 0
AL = 0, ±1 (12.28)
AJ = 0, ± 1 (0 —> 0 excluded)
The singlet and triplet transitions shown for ionized scandium in Fig.
12.4 will serve as good examples of these rules.
B. For .^/-coupling the further restrictions are:
Aji = 0 >
A . „ , i Vor vice versa (12.29)
Aj 2 = , ± )
0 1
AJ = 0, ± 1 (0—0 excluded)
According to these rules transitions are allowed which, under the rules
for LS-coupling, are not allowed. Intercombination lines, although
weak, are good indications that the coupling in either the initial or final
state is not good Uncoupling.
A more convenient way of expressing the selection rules for the
individual electrons [Eq. (12.27)] has been brought out by the quantum
mechanics. The rules are expressed in terms of the oddness or evenness
of the electron configurations and terms. All spectrum terms arising
from an electron configuration for which the sum of the l values is even
are called even terms. All terms for which the sum is odd are called odd
terms. Odd terms are designated by small exponents “ °.” For example
we write:
The selection rules for l values are that only even terms can combine with
odd terms, and only odd terms with even terms. This rule1 was dis¬
covered empirically by Laporte. The even terms in the above example
1 For a quantum-mechanical proof of this rule see H. Weyl, “The Theory of Groups
and Quantum Mechanics,” p. 201, 1931.
Sec. 12.151 THE ATOM MODEL FOR TWO VALENCE ELECTRONS 205
can combine with the lower set in what is called a double electron jump,'
Ad to 4p (AZ = 1), and 3d to 4s (AZ = 2), or 3d to 4p (Al * 1), and
4<Z to 4s (AZ = 2). In calcium the 3dnd,*P terms combining with
4s4p,3P° terms give rise to all but the first member of the anomalous
series observed by Russell and Saunders,1 (see Sec. 11.5). In jy-coupling
odd terms will be designated in analogous fashion, e.g., (? f)J.
12.15. Intensity Relations.—In calculating relative intensities of
spectrum lines in genera) it is not always sufficient to use the summation
rules of Ornstein, Burger, and Dorgelo (see Secs. 8.3 and 10.6). In
thermal excitation, for example, the intensity of a line will depend upon
the temperature of the source. At low temperature there will be very
few atoms in excited states. With an increase in temperature the numbei
in higher states will increase, and hence the possibility of jumping back
with the emission of radiation becomes more probable. Just as in the
classical theory of the emission of energy from an electric oscillator the
intensity of a spectrum line will also depend upon the frequency. The
ratio between the intensities of two emitted spectrum lines may be
written as
hv,n
1 The first member of the series is now ascribed by Russell to the transition 4p* to
4*4p.
2 The quantum weight (2/ + 1) gives the number of magnetic levels into which a
given fine-structure level J will split when the atom is placed in a magnetic field.
3 Kronig, R. de L., Zeits. f. Phys., 31, 885, 1925; 33, 261, 1925.
4 Russell, H. N., Proc. Nat. Acad. Sci., 11, 314, 322, 1925; Nature, 115, 835, 1925.
6 Sommerfeld, A., and H. H6nl, Sitz-ber., Berl. AkacL Wiss., 9, 141, 1925.
206 INTRODUCTION TO ATOMIC SPECTRA [Chap. XII
(/ + D
(12.31)
For transitions L —► L
A(L + ^ + S + 1)(L + / - S)(L - / + S + 1)(L - / - S)
J-l- J» / J
A[L(L + 1) + J(J 1) - S(S ± 1)]«(2J ± 1)
ij-;, /
J(J + 1)
4-
168 30 2 200 5 *P* 100 17.9 1.2 *P, 100 17.5 1.6
90 30 53.6 17.9
120 t »Pi aPi 54.0 17.5
»Po 40 40 1
aP» 23.8 *P0 22.2
168 120 72
7:6:3
According to the sum rules, the sums of the intensities in the columns
have the ratios given by the integers 7:5:3 below and the sums of
the intensities in the rows have the ratios given by the integers 5:3:1
at the right in the array. The three strongest lines are those for which
L and J change in the same direction and they are called the diagonal
1 Dirac, P. A. M., Proc. Roy. Soc.t A, 111, 281, 1926.
Sac. 12.15] THE ATOM MODEL FOR TWO VALENCE ELECTRONS . 207
/ t
lines. Of these lines the one involving the largest L and J is the strong¬
est. The three weaker lines are called off-diagonal lines. At times
it is convenient to compare the intensities of the lines of each multiplet
with the strongest line in the multiplet designated as 100. These
reduced intensities are given in the center in Table 12.6. Relative
intensities expressed in this way are given in the Appendix for all triplet
combinations up to 3J. The relative intensities of the calcium *P — lD
multiplet (see Fig. 11.7), as measured by Burger and Dorgelo, are given
in the third array. Wherever LS-coupling is revealed by the fine
structure, the calculated intensities are in good agreement with those
observed. Where deviations occur in one they occur as expected in
the other. It should be noted that the sum rules make it possible to
interchange the r61es of the initial and final states in the above given
formulas.
For the relative intensities of lines arising from ^-coupling schemes,
Eqs. (12.31) and (12.32) may be employed by replacing S by ji and L
by j2. Here ji is taken as the quantum value which, in any given transi¬
tion, does not change. Applying the modified formulas to the intensities
of the lines arising from the configurations sp —> s-s, the values shown
in the first two arrays in Table 12.7 are obtained:
Table 12.7
sp, (ii) (if) sp
/ = 0 1 J = 1 2 J — 0 1 1 2
J * 0 s 1 J = 0 24 1 J - 0 12 24 1
*, (ii) 1 - * • «, (ii) •• 8 • 8 ••
/ = 1 3 6 3 J - 1 12 60 3 J - 1 12 24 12 60 3
1 :3 3 : 5 1 : 3 3 : 5
Table 12.8
\s. 36 1
••
’Si 12 36 60 S
3 1 : 3 : 6
Table 12.9
*P 3F HI
3p° 24 64 42 3
w° 18 70 112 5
3jfp° 8 66 216 7
<! 1:3:5 : 7 : 9\
of 100 for the strongest multiplet, the calculated and observed valuesf
are as follows:
Table 12.10
*S *P »D *F *G *P «F *Q
tpo
11.1 25.0 19.4 apo 24.0 18.3 14.1
*D° 8.3 32.4 51.8 »D° 4.2 41.0 35.2
»p° 3.7 25.9 100 apo 9.9 22.5 100
calculated observed
fine structure of the very narrow orthohelium series was triplets and
not doublets was first predicted by Slater1 and verified by Heisenberg.2
Similar series have been observed also in ionized lithium. The two sets
of series in each of the elements may be arranged into two apparently
noncombining sets of energy levels which, in the case of helium, have
come to be known as parhelium and orthohelium levels.
1 Slater, J. C., Proc. Nat. Acad. Sci11, 732, 1925.
* Heisenberg, *W., ZeiU. f. Phys., 39, 499, 1926.
210 INTRODUCTION TO ATOMIC SPECTRA [CHAJr. XII
Energy level diagrams of helium and singly ionizod lithium are
shown in Fig. 12.14. The enormous energy shift from the normal
l«VSo state to the first excited states 1s2s,1iS0 and 3£i is characteristic
of all the inert gases and similar atomic systems.
Many attempts to calculate the energy levels of helium, a three-body
problem, from a purely theoretical standpoint have been made, and with
reasonable success. The normal state of helium, for example, has been
calculated by Hylleraas1 to be 24.470 volts below the series limit, a
value differing only 0.003 volt from the spectroscopic value of 24.467
volts. For Li+ he obtained the value 75.272 volts, in excellent agreement
with the observed value 75.279 volts.
Heisenberg has shown that the enormous energy difference between
the first two or three triplet levels and their associated singlet levels
is due not to the magnetic interactions of the two electrons but to an
electrostatic resonance interaction. This phenomenon of resonance,
which so frequently arises in the quantum mechanics, is analogous
to the resonance between two mechanical oscillators. The analogy
has been drawn of two similar pendulums, connected to the same support,
and set swinging. If the support is not too rigid the energy of oscillation
is found to shift back and forth between one pendulum and the other.
At one instant one pendulum is at rest and the second vibrates with a
maximum amplitude. Some time later the second pendulum comes
to rest and the first vibrates with the maximum amplitude. It may be
said that the two pendulums have changed places.
For helium-like atoms this resonance phenomenon has been shown
to yield fair agreement with experiment by Heisenberg, Dirac, Slater,
Gaunt, Unsold, Breit, Hylleraas, and others. The first attempts to
solve the three-body problem of helium on the quantum mechanics led
to a wave equation, quite like Schrodinger’s equation for hydrogen, and
of the form,
where the subscripts refer to the two electrons respectively. The first
term in parentheses represents the total energy of the system; the second
and third terms represent the potential energies of the two electrons,
due to their attraction by the nucleus; the last term represents the
potential energy due to the mutual repulsion of the two electrons. If
it were not for this last term, which accounts for the fine structure we
wish to measure, the equation could be solved by analytical methods.
Neglecting this term, the solution is of the form of the product of two
hydrogen-like functions,
i = la, m«; n, dx, fa) • tH(nb, lb) mb; rlf 6l} fa), (12.34)
1 Hylleraas, E. A., ZeiU. f. Phy54, 347, 1929; 65, 209, 1930.
Sec. 12.18] THE ATOM MODEL FOR TWO VALENCE ELECTRONS 211
where n, l, and m are the quantum numbers of the two electrons and.
r, 6, and <t> their polar coordinates. This solution is degenerate in that
the energy of any state of the atom is not altered by interchanging the
two electrons, i.e., interchanging the coordinates r, 6, and fa
V = ^«(n0, la, fna; r2, 02, fa) • (n&, lb, m&; r\, $i, fa). (12.35)
1 For an elementary treatment of this problem the reader is referred to that given
by E. U. Condon and P. M. Morse, “Quantum Mechanics,” Chap. IV.
212 INTRODUCTION TO ATOMIC SPECTRA [Chap. XII
are oppositely directed for the singlets, S = 0, and more nearly in the
same direction for the triplets, S — 1. It follows that the interaction
energy is positive when the angle between $* and s* is large, and negative
when the angle is small (see Sec. 12.7).
According to Heisenberg’s results the interaction energy between
S and L for the configuration ls2p is given by
(12.36)
These intervals are shown in Fig. 12.16. The dash-dot line repre¬
sents the center of gravity of each triplet. As Z becomes large, the
second and third terms of Eq. (12.36) become negligibly small compared
with the first. Dropping them we obtain
This is the Land6 interval rule which for sPo,i,a gives the interval ratio
1:2. The relative separations of the first two elements, He I and Li II,
have been observed to have, quite accurately, the intervals given by
3a Z-1 5 4 5 6 X i —»oo
Fia. 12.16.—Relative separations for the l«2p, 8Po,i,j. states of helium-like atoms. (After
Heisenberg.)
Problems
1. Derive all of the terms arising from the electron configuration fg.
2. Graphically represent the Land6 interval rule for a 9P term as in Fig. 12.
214 INTRODUCTION TO ATOMIC SPECTRA [Chap. XII
.
8 Show for a ds electron configuration that the total lD separation is the same in
both LS- and ^-coupling. Show also that the d electron alone gives the same %D
separation (see Sec. 12.11).
4. Using Eqs. (12.31) and (12.32), calculate the relative intensities in each of the
three triplets arising from the transition pp —► ps. Assume L5-coupling.
5* Calculate the relative multiplet intensities for the triad of triplets arising in
Prob. 4 (see Sec. 12.16).
6. Repeat Prob. 4 assuming jj-coupling between the electrons. Compare these
intensities, as in Tables 12.7 and 12.8, by applying the sum rules.
CHAPTER XIII
ZEEMAN EFFECT, PASCHEN-BACK EFFECT, AND THE PAULI
EXCLUSION PRINCIPLE FOR TWO ELECTRONS
e ergs e ergs
2me gauss* 2me gauss (13.1)
Similarly the two Z’s are coupled together and quantized to form a
resultant L*. Projecting each orbital moment on L* and adding,
215
216 INTRODUCTION TO ATOMIC SPECTRA [Chap. XIII
e ergs
Mi * K cos (J?L*) + £ cos (ZJL*)}A . • „ (13.3)
11 u ' n v3 /j2t 2mc 2t 2me gauss
Just as the spin and orbit of a single electron are coupled together to
form a resultant moment j*, so L* and S* are coupled together to form
a resultant moment «/* (see Sec. 10.3 and Fig. 10.6). Projecting ns
and nL on J* and adding, we get for the total magnetic moment of the
atom,
The terms within the brackets give hj in units of the Bohr magneton,
he/fame. Replacing these terms in the brackets by g times J*f i.e.,
Since the angles between L*, S*, and J* are fixed, in LiS-coupling,
the cosine terms in Eq. (13.5) are readily evaluated.
Substituting these two cosines in Eq. (13.5), we get for the g factor,
This is the Lande g factor in exactly the same form as Eq. (10.11)
for a single electron, except that l* is here replaced by L*, s* by S*f and
j* by J*. Values of the g factor calculated from this equation are given
in the Appendix.
In ^'-coupling the s* and Z* of each electron are coupled together
and quantized to form their own resultant j*. The magnetic moment
of a single electron has already been given in Eqs. (10.6) and (10.7).
With two electrons we may write for each one separately,
w = (13-9)
With j* and j* in turn coupled together and quantized to form a
resultant J*, these moments are projected on J* and added to give the
total magnetic moment of the atom,
Sac. 13.2] ZEEMAN EFFECT, PASCHEN-BACK EFFECT 217
LS-Coupling jj-Coup!ing
Fio. 13.1.—The vector model for LS- and /./-coupling in a weak magnetic field (Zeeman
effect).
<1316>
AW = H ■ g • ^ J* • £ cos (13.17)
Dividing by he, the energy in wave numbers (see Sec. 10.4) is given by
He
—AT\t = M • g^~~c2 cm"1 = M • g • L cm’1. (13.19)
Fig. 13.2.—Graphical representation of the splitting of a *D8 level in a weak magnetic field.
the g factors for the initial states 3P0, 3Pi, and 3JP2, and for the final
state ZS\, are &, f, f, and 2, respectively. The resultant magnetic
levels are shown schematically in Fig. 13.4. At the bottom of the
figure the observed patterns are shown for a zinc triplet. With a g factor
of two for the lower state 3Si, the pattern 3Si- 3Po has twice the normal-
triplet separation.
13.3. Intensity Rules for the Zeeman Effect.—Intensity rules for the
Zeeman effect of atomic systems containing more than one valence
electron are independent of the type of coupling and depend only on
the quantum numbers M and J. These rules, first discovered empirically
by Ornstein and Burger,1 may be derived from the sum rules. The
1 Ornstein, L. S., and H. C. Burger, Zeits. f. Phys., 28, 135, 1924; 29, 241, 1924.
Sec. 13.4] ZEEMAN EFFECT, PASCHEN-BACK EFFECT 221
sum of the intensities of all transitions starting from any initial Zeeman
level is equal to the sum starting from any other level. Similarly, the sum
of the intensities of all the transitions arriving at any final Zeeman level is
equal to the sum arriving at any other level. From the classical model
and these rules the following equations have been derived,
Transition J -+ J:
Transition J —» J + 1: (13.21)
In deriving these equations account has been taken of the fact that
when observed perpendicular to the field only half the intensity of
the s components is observed. Observations parallel to the field give
the other half of the s components. In any direction the total light
emitted is unpolarized just as in field-free space. The constants A and B
in the above equations are proportionality constants and are not needed
for any given Zeeman pattern. Applying these equations to the principal
triplet of Fig. 13.4, the values shown at the arrow tips are obtained.
In order that the sum rules hold for all of the levels in the figure the
values derived for the first pattern *Si — 3Po have all been multiplied
by two and those of the second pattern 3Si — 3P± by three. Doubling
the s components the sum of the intensities of all transitions starting
from any upper level is 24 and the sum ending on any lower level is 72.
13.4. The Calculation of Zeeman Patterns.—A scheme for the rapid
calculation of Zeeman patterns has been given by Sommerfeld (see
Sec. 10.6). Consider, for example, the complex pattern arising from the
transition 3D3 — 3P2. Here the J values are 3 and 2, and the respective
g factors, from Eq. (13.8), are $ and f. The separation factors Mg
for both the initial and final states are first written down in two rows
with equal values of M directly below and above each other as follows:
M = 3 2 1 0 -1 -2 -3
Mg for initial state:
ponents ± j, ±J, ±$, ±§, and ±V-. These may be abbreviated to give
the separations (see Fig. 13.5 and Sec. 10.6):
(0), (1), (2), 6, 7, 8, 9, 10
-ATm = ± L cm”"1. (13.22)
6
The common denominator is called the Runge denominator and is
the least common multiple of the two denominators of the Mg factors.
This pattern is plotted in Fig. 13.5. The p components are plotted
V? Vi
i■
Anomalous Zeeman
’ 1 1 ‘1 _ Patterns
for
Triplets
r'i 1
®i& |
II' | l Ii.Ml| ’V+1 H |
5d,3ds 3d,5d, |
r 1 I 1 .1, ,1.
w L_ ill
i r %\
ir i l li, ,|iil ■ .'
3r 5C
'5 rl
r 1 1 f jJTr i 11
1 Y 1r
TP ''iii
«il yj 1L_ w
r i T | '1 W ■ "W
nili' 1 fi 1 1 i
Fig. 13.5.—Anomalous Zeeman patterns for triplet combinations. Dots show normal
triplet intervals (LS-coupling).
above the line in the usual way, and the s components are plotted below.
The heights of the lines represent the relative intensities as calculated
from the equations in Sec. 13.3.
13.5. LS- and jj-coupling and the g Sum Rule.—In classifying a
given spectrum line for the first time it is not always known to just
what type of coupling scheme the corresponding levels belong. Since
the g factors are in many instances different for ^/-coupling than they
are for L£-coupling, the Zeeman patterns will also be different. For
this reason it is sometimes difficult to assign a given pattern to a definite
transition. This uncertainty is often alleviated by Pauli's1 so-called
g sum rule. This rule states that out of all the states arising from a
given electron configuration the sum of the g factors for levels with the
same J value is a constant independent of the coupling scheme. As
an example, consider the four terms 8Po,i.2 and IPi, arising from the
configuration ps. The following table gives the g values of these four
>p. 17-1
LS
»Po.l,« 0 i i
I**?* r*C%
0 4
II II
II II
1 jj
**■»
0 a
1
When only one term occurs for a certain J value the g sum rule
slates that the g factor is the same in all coupling schemes. Observed
Zeeman patterns illustrating the g sum rule are to be found for many
< leinents. In Table 13.2 zinc, tin, and lead are given as examples.
g = 1.000
Zn LS
0,1,2 0 1 500 1.500
lPl 1.125
Sn* Intermediate
•Po.1.2 0 1 375 1.500
'Px 1.150
Pbt jj
0 1.350 1 500
Sp 0 1.500 1 500
jj-coupling, is broken down. This will occur when the interaction energy,
given by Eq. (13.19), exceeds the L*S* interaction energy given by Eq.
(12.13) or the jfjj interaction energy given by Eq. (12.25). In other
words when the Zeeman levels of the different levels of a multiplet begin
to overlap each other, Eq. (13.19) will no longer hold. As the field
t S S P p 5 S
Fio. 13.6.—Observed Zeeman patterns for the transition 3Si — 3P1 in Zn, Sn, and Pb.
H Strong field H
L S-Coupling jj-Coupling
Fig. 13.7.—Classical vector model for LS- and //-coupling schemes in a strong magnetic
field (the Paschen-Back effect).
pendently around H, their resultant «/* ceases to have any meaning. The
two special cases of LS- and ^'-coupling will now be treated separately.
13.7. LS-coupling and the Paschen-Back Effect.—The method used
in deriving an expression for the interaction energy when the Paschen-
Back effect sets in, for original LS-coupling, is exactly the same as that
Sec. 13.7] ZEEMAN EFFECT, PASCHEN-BACK EFFECT 225
used for a single electron in Sec. 10.7. With L*' and S* independently ,
quantized with the field H, the quantum conditions are (1) that the
projection of L*h/2x on H is equal to MJi/2v, where Ml = 0, ±1,
±2, • • • ±L; and (2) that the projection of S*h/2ir on H is equal to
Msh/2x, where Ms = 0, ±1, ±2, • • • ±S. For singlets Ms = 0, and
for triplets Ms = +1, 0, and —1. On the classical model the Larmor
precessions of L* and S* will be given by H times the ratio between
the magnetic and mechanical moments,
“d "■-h'2'25V (13-23)
The sum of these two energies gives the main energy shift of each mag¬
netic level from the field-free level from which it sprang:
where A is given by Eq. (12.12). Adding this term to Eq. (13.26), the
total energy shift becomes
Next come the weak-field Zeeman levels symmetrical about the field-
free levels followed by the Paschen-Back levels at the right. The
Fig. 13.8.—Strong-field energy levels for >5, 3P, and lD terms (Paschen-Back effect.
I/<S-coupling).
equally spaced dots in the strong-field levels represent the two inter¬
action energies due to the external field. Adding the L*S* interaction
term divides the 3Z>3 levels into three groups of five equally spaced
levels.
Table 13.3.—Computation of Magnetic Energies for LaS-coupling in Weak
and Strong Magnetic Fields
1 strong
Term rWMk M Mg M Ms Ml 2Ms 2MS + Ml
A ■ MlMs
+2 +1 +2 +1 +1 +2 +3 +A
+1 +! +1 +1 0 +2 +2 0
3P2 +A 0 o 0 -1 +2
+1 +1 —A
~1 -2
+1 +1 +1
-2 -3
0 0 0
+1 +1
-1 -1 -1
*Pi -A 0 0
0 +1 -2 -1 -A
-1 0 0
-* I -1 “2 -2
*Po -2 A o -2 -1 -2 —3 +A
M = Ml 4- Ms. (13.31)
where gi and p2 are the g factors for the two electrons, respectively.
Multiplying each of these angular velocities by the corresponding pro¬
jections of the angular momenta on H (see footnote 3 on page 157) gives
the two energies,
He
-ATh = (gimn + QWn)^-^ cm_1 = + ^m,,) - Lem"1. (13.37)
To this energy must be added the interaction energy due to the coupling
between and j*. By analogy with LS-coupling this energy is given
by [see Eq. (13.28)]
where A is given by Eq. (12.24). Adding this small term to Eq. (13.36),
the total magnetic energy becomes
Fio. 13.10.—Strong-field energy levels for the electron configurations sp and a-s (Paschen-
Back effect, //-coupling).
+2 +i +2 +i +! +1 +l +34/4
+1 +i +1 +1 +1 +! +4/4
(i 1)2 +34/4 0 0 +i -i +1 -! -4/4
-1 -! -1 +* -i +1 -i -34/4
_o
— U -t
+1 -i +1 -1 +1 -34/4
+1 +i 0 -J +1 -1 +! -4/4
(1 l)i -54/4 0 0 -1 -i -1 -1 -1 +4/4
-1 -I -* -! -1 -1 +3A /4
230 INTRODUCTION TO ATOMIC SPECTRA [Chap. XIII
strong-field levels for the two terms at the left are calculated by tabulating
the various interaction energies as in Table 13.4.
It is to be noted that for each weak-field level Af, there is a strong-
field level with the same M value. Just as in the case of LS-coupling
the conservation of angular momentum requires that the sum of the
\projections of the mechanical moments on H does not change with changing
field. As an equation,
M — mn + mJt. (13.41)
Since there are in general several levels with the same M value,
the further restriction that no two levels with the same M cross each other
G+z ^ ^ ms
electron transition sp
These rules are in agreement with the polarization rules that the
polarization of a line is the same in all field strengths. In the pattern
it is seen that lines tending to go from p to s or from s to p components
fade out and should not be observed in a strong field. Unlike the
Paschen-Back patterns for L£-coupling the lines do not resemble a
normal triplet but possess an anomalous pattern. Were it not for the
fine structure, the pattern shown in Fig. 13.11 would be the same as
the pattern for a principal-series doublet line, — 2P§ (see Fig. 10.9).
Sxc. 13.0] ZEEMAN EFFECT, PASCHEN-BACK EFFECT 231
The Larmor precessions of each spin and orbit around H [see Eq.
(10.4)] are given by
(13.44)
Table 13.5.—Very Strong Field Energy Levels tor the Electron Configura¬
tion sp
(Complete Paschen-Back effect)
M=m.l+mn
Sec. 13.9] ZEEMAN EFFECT, PASCHEN-BACK EFFECT 233
These average cosines are evaluated just as in Sec. 10.7 [see Eq. (10.33)]
to give
Fia. 13.13.—Very strong field energy levels for an sp electron configuration (complete
Paschen-Back effect).
each very strong field level will be given by the addition of Eqs. (13.46)
and (13.48):
no two levels with the same M values cross. It should also be pointed
out that the coefficient Oi, due to the spin-spin interaction, is negative and
Normal triplet
Fio. 13.15.—Complete Paschen-Back effect for the transition ep —♦ (//-coupling).
resultant patterns for all allpwed transitions in weak, strong, and vety
strong fields are shown in Figs. 13.14 and 13.15. The initial levels giving
rise to these patterns are the ones shown in Fig. 13.13. The selection
rules for very strong fields may be stated as follows:
Am* = Am* = 0
Amh =
or with subscripts
i 0 for p components/ (13.50)
Am,, = interchanged
[±1 ior s components)
Were it not for the fine structure arising from the T factors, each
very strong field pattern would be a normal triplet. This result is
arrived at with all electron configurations. It should be noted that
the two strong field patterns for ^‘-coupling, in Fig. 13.15, closely resemble
the anomalous patterns of a principal-series doublet in a weak field,
f P s
rrv 1 0 -1
msrl \ -J4
"V
t 0 “1
_ 20
i
■_
mj2
M» 2 1 0 1 Ml= 1 0 -i
M =2 1 \ M *1 0 %
M= i o']-1 0 M ■ t 0 -i _0_ \
M ■f 1 0 "m" B0l
-4
-I
M= *0] -1 j-2 -1 *P| MS
0 j -1 -<>!_-
\ -HX —KM
lP0'3P, 3P2 MS I??
m
a row and column as in the array in Fig. 13.16. To the left of raZl = 0
and below mlt = 1 the sum Mh = 1 is written. This process continued
gives the strong-field quantum numbers for a P term. Combining
mtl and m9t in the same way, the array shown at the top of the figure
is constructed. The values Ma — 0 and Ma = 0 and ±1 divided by
the L-shaped line are just the Ms quantum numbers for singlets and
triplets, respectively. Combining these two sets of Ms with the set of
Ml, the two arrays in the lower left of the figure are obtained. The
quantum sums divided by the dotted lines are just the weak-field quan¬
tum numbers for the 3P0, ®Pi, %P*, lP\ terms. In the case of ^‘-coupling,
shown at the right in the figure, the values of ma and mz are first combined
for each electron separately to give the strong-field quantum numbers
m,. These in turn are combined to form the array shown at the bottom
of the figure. These give the same runs of weak-field quantum numbers
and correspond to the same four levels obtained for L£-coupling.
13.11. The Pauli Exclusion Principle.—In 1925 Pauli put forward a
new principle which came to play an important role in the development
of complex spectra (see Sec. 12.4). In its simplest form the principle
may be stated as follows: No two electrons in the same atom can have all
of their quantum numbers the same. This principle is well known as the
Pauli exclusion principle. In order to assign quantum numbers to each
electron we go to very strong magnetic fields where the coupling between
the electrons is completely broken down. In such a field the four
quantum numbers for any electron are n, l, mi, and mt.
Let us now calculate the allowed spectral terms arising from two
equivalent electrons. The term equivalent refers to any two electrons
having the same n and l values. As a first example, consider two equiva¬
lent p electrons. First all possible combinations of mt and mi for a
single p electron are written down in two rows as follows:
™. = $ \ i -i -i -i
mi = 1 0 -1 1 0-1
(a) (b) (c) (d) (e) (f)
It is seen that there are six possible states, (a), (6), (c), (d), (e), and
(/), in which a single p electron may exist in an atom. With two equiva¬
lent p electrons Pauli's exclusion principle says that one of these quantum
values m, or mi must be different. We therefore obtain all possible
states for two electrons by writing down all combinations of the above
states taken two at a time, with no two alike. They are
ab
ac be
ad bd cd
ac be ce de
af bf cf df ef
Sec. 13.11] ZEEMAN EFFECT, PASCHEN-BACK EFFECT 237
Ms « 1 1 0 0 0 1 0 0 0 0 0 0 -1 -1 -1
ML = 1 0 2 1 0-1 1 0-10-1-2 1 0-1
Ms = 0 0 0 0 0\
Ml = 2 1 0-1-2/
Taking these out we find just enough numbers left to form a 8P and a
l8:
Ms — i 1 1 0 0 0 -1 -1 -n Ms = o\w
Ml = 1 0-1 1 0-1 1 0-1/ Ml = 0/ ’
electrons give rise to 1S, lP, lD, *S, 3P, and 3Z) terms. The Pauli exclu¬
sion principle thus excludes IP, 3Sy and 3Z>, when the two electrons have
equal n values.
The same terms are readily calculated from Breit’s scheme given
in the preceding section. Consider, for example, two d electrons in
LiS-coupling. We first write down all possible combinations of mtl
and mti in one array and those of mZl and mZa in another (see Fig. 13.17).
If the total quantum numbers are different, all possible combinations
of the magnetic quantum numbers are allowed. The runs ML = 4 to
—4, 3 to —3, 2 to —2, 1 to —1, and 0 correspond to G, F, D, P, and S
terms, respectively; and the two sets of Ms = 1, 0, —1, and 0 correspond
to triplets and singlets. These are just the terms given in Table 12.2
for d-d.
238 INTRODUCTION TO ATOMIC SPECTRA [Chap. XIII
Ms = 1, Ml- 3 2 1 0 -1 -2 -3
Ms = -1, Ml = 3 2 1 0 -1 -2 -3
Ms = 1, Ml = 1 0 -1
Ms = -1, Ml = 1 0 -1
If, on the other hand, we permit the ML values to be alike, the values
of M8 — 1, or —1, are forbidden. Since the two Ms = 0 values are
identical, one of these must also be excluded. The remaining com¬
binations are tabulated as follows:
Ms = 0, Ml = 4 3 2 1 0 -1 -2 -3 —4
Ms = 0, Ml = 3 2 1 0 -1 -2 -3
Ms = 0, Ml = 2 1 0 -1 -2
Ms « 0, Ml = 1 0 -1
Ms = 0, Ml - 0
The second and fourth rows go with the preceding tabulation to com¬
plete the quantum numbers for *F and 3P terms. The remaining rows
correspond to Hj, 1D, and lS terms.
Consider, as a third example, the calculation of terms arising from
two equivalent p electrons in jj-coupling. The first step is to write
down all possible combinations of m, and mi for each electron separately.
This forms the two arrays at the top of Fig. 13.18. Each of the two
runs of mn combined with each of the two runs of m3l will give all possible
combinations. Of the four resultant arrays there will be two like the
lower right one in the figure. Since the electrons are equivalent, one
of these must be entirely excluded. In the lower left-hand array, the
diagonal and either the upper right or lower left half of the array must
be excluded. The same is true for the lower center array. The remain¬
ing combinations are just sufficient to form the five terms (f |)0, (4 I)2,
(i £)o, (4 !)i> and (4 4)2. These correspond to the five terms bS0,
*Po, *Pi, and 8P2 in L£-coupling, derived above.
1 Each value of Ml in the lower left half of the array in Fig. 13.17 is identical with
a corresponding value in the upper right half. It is observed that each pair of values
arises from the same mi values. Although the mi subscripts are just interchanged in
each of the two combinations of a pair, the equivalence of the two electrons makes the
two states identical in all respects.
Sec. 13.12] ZEEMAN EFFECT, PASCHEN-BACK EFFECT 239
13.12. Pauli’s g Permanence Rule.—In 1923 Pauli1 proposed a new
rule for the g factors of any given multiple term in weak and strong
fields. This rule states that the sum of the g factors for a given M is the
same in all field strengths. In a weak field the g factor of any state is
given by the ratio between the magnetic moment of the atom in Bohr
"Vi o -i "Vt o -i
1 m .S1 i .i 1
mj = \ 2 “2 2 J2 2 2 2 2
~r n 3 i
77}! "2l 2 "2 '2
mS2
ms,
Coupling
3 1.13
mjis2 2 2 2 _ mjfH_ - "vf ?4"i
i o l M* 2 1 0 4 T
2 X 0 |-l 4 o\\ 1 0 -1 }-2 "2i
1 o'l^K |-2 -1 mJz
-V 1 \
0 |-l |-2 |-3, -f
(fi)0 cii\ 1 mj2 (H), (41),
Fig. 13.18.—Magnetic quantum numbers for two similar p electrons 07-coupling).
Qwmk
M£. (13.51)
P/
Here
"-sfe “d <13®>
In weak or strong field the g factor is given by the ratio of the total
projection of the magnetic moments on the field direction to the total
projection of the mechanical moments.2
For L/S-coupling,
_ ML + 2MS
0*m" - Ml + Mb' (13.54)
For ^‘-coupling,
_ giinH + gtmh
{/strong — t (13.55)
mh + ^;S
For a given multiple term it often happens that there are several
magnetic levels with the same M value. Pauli’s g permanence rule
states that for given L, S, and M, or for given jh jt, and M the sum
of the weak-field g factors is equal to the sum of the strong-field g factors:
2^we*k ~ trons* (13.56)
Values of weak- and strong-field g factors are given in Table 13.6
for a 8P term.
Table 13.6.—Pauli’s g Permanence Rule for L£-coupling
s
Term M = 1 M = 0 M = -1
II
1
wm
•P* ! ! 1 ! 1
•Pi ! ! ! Weak field
»Po 8
•Pa 1 2 1 i
•Pi 1 2 Strong field
•Po
i 3 3 !
8
Similarly for the two upper states arising from the sp configuration
in ^-coupling we obtain the values given in Table 13.7.
Table 13.7.—Pauli’s g Permanence Rule for jj-coupling
<N
Term M 2 M - 1 M = 0 M = -1
II
=
1
(H)i i l
Strong field
(1 i). i i i §
2* ! i (1) 1 i
Darwin has shown that the g permanence rule holds for all inter¬
mediate field strengths.1
13.13. Pauli’s g Sum Rule for All Field Strengths.—In a very strong
magnetic field the individual electron couplings are completely broken
down and each spin and orbit becomes quantized independently with
the field. Pauli showed that even under these extreme conditions
the sum of the g factors for a given M, of a given electron configuration is
constant.2 Starting with LS- or ^‘-coupling the g factor for any term
in a very strong magnetic field is given by the sum of the projections on
H of the magnetic moments in Bohr magnetons divided by the sum of the
projections of the mechanical moments in units h/2ir.
/2 m9l + 2 m,t + mh + mt\
(13.57)
y m81 + mtz + mh + mi% )
field g factors can be determined from the g permanence rule and hence
the weak-field g factors can be evaluated (see the previous section).
13.14. Land€’s r Permanence Rule.—Analogous to the g permanence
rule we have what is known as the r permanence rule discovered by
LandA The rule may be stated as follows: For a given multiple term,
i.e.y given S and L, or given ji and the sum of all the T factors for terms
with the same magnetic quantum number M is a constant independent
of the field strength.
Table 13.9.—Pauli’s g Sum Rule por the Electron Configuration sp in
./[/-coupling
In going from weak fields to strong fields the r factors always change/
if at all, so that ST * constant for each M. It is to be noted that the
weak-field T's can be determined from the strong-field T’s. The Land6
interval rule is thus obtained without the use of the quantum-mechanical
or classical model of the atom [Eq. (13.58)].
o
SI
SI
M = —1 M = -2
(M
Term M « 1
II
II
3P2 A A A A A
Weak field
3Pi -A -A -A
T factors
3Po -2 A
3P2 A 0 -A 0 A
Strong field
•Px 0 0 0
T factors
3Po —A
xr = A 0 -2 A 0 A
Table 13.11.—T Permanence Rule for the Two Terms (£ |)i and ($ §)*, Arising
from the Configuration sp (if-couPLiNG)
Term M = 2 M = 1 M = 0 M « -1 M = —2
MB -A/4
Tt< rfi
Term M - 2 M = 1 M = 0 M = —1 M - -2
lPx 0 0 0
5P* a/2 o/2 a/2 a/2 a/2 Weak field
*Pi -o/2 -a/2 -a/2 T factors
*Po —a
*Pi 0
aP2 a/2 -a/2 a/2 Strong field
aPi | 0 r factors
aPo -a/2
1 Goudsmit, S., Phys. Rev., 31, 946, 1928; see also Paulino, L., and 8. Goudsmit,
“Structure of Line Spectra.”
Sue. 13.16] ZEEMAN EFFECT, PASCHEN-BACK EFFECT 246
be pointed out that the r sum rule may also be applied and shown to'
hold for Ti and IV
13.16. The r Permanence and r Sum Rules Applied to Two Equiva¬
lent Electrons.—Consider the permanence and sum rules as they apply to
the case of two equivalent d electrons, i.e., to two d electrons having
the same total quantum number n. In this case the Pauli exclusion
principle is in operation and the allowed terms (see Sec. 13.11) are
X/S, 8P, 1DJ 8P, and Hr. In a very strong field there are 10 possible states
for each d electron:
m. = i i i \ i
mi — 2 1 0 —1 —2 2 } o -I -2 («•«»
Following the first scheme developed in Sec. 13.11 for the calculation
of terms arising from two equivalent electrons we combine these columns
two at a time, taking no two alike. The resultant combinations should
be tabulated by the reader as follows: (1) Make four columns of all
possible combinations of, and giving each value of, mtl, mtt, milf and
(2) Using the formulas for r3 and r4, Eq. (12.9), multiply each m,
by its own mt times a to form the fifth and sixth columns under r3 and r4.
(3) Since the a coefficients are identical in the two electrons, add each
r8 and r4 to form the seventh column T^ry .tron«. (4) Add the first
four columns to form the last and eighth column of total magnetic
quantum numbers M = mtl + + mh + m^. We now collect all
values of Tvery „tron, and tabulate them under corresponding values of M
as shown at the top of Table 13.13.
The strong-field r sums may now be used to calculate the weak-field
T factors, provided the coupling of the two electrons is specified. Sup¬
pose first that the coupling is LS. In LS-coupling the T factors for
singlet terms are always zero. We therefore write zero in all columns
for the 1S, 1Z>, and lG terms, as shown. Since 3P4 is the only remaining
term with M = 4, and the r sum for if = 4 is 3a/2, we write T = 3a/2
for all 8P4, M’s. This leaves —a/2 for 3P3, M = 3, which is next written in
all 8P3, M’s. With — 3a/2 for 3P4, the r permanence rule, which also
gives the Land6 interval rule, gives T = —a/2 and —2a for SP3 and *F2)
respectively. This in turn leaves only a/2 for 8P2, then —a/2 for 8Pi
and —a for 8P0.
If the coupling of the two equivalent d electrons is ^-coupling, the
individual j’s and the field-free terms are first determined by the scheme
given in Sec. 13.11. The resultant values are given in the lower left
hand column of Table 13.13. For each d electron, separately, ji = |
and |, for which the T factors are a and —3a/2, respectively. The
T factors for both electrons, taken together, will be just the sum of the two
individual T's. These inserted in the lower third of Table 13.13 agree
with Goudsmit’s T sum rule.
246 INTRODUCTION TO ATOMIC SPECTRA [Chap. XIII
Tablb 13.13.—r Sum Rule for Two Equivalent d Electrons in Weak and in
Very Strong Magnetic Fields
M* ilf- M* M= M— M-
Term
4 3 2 1 0 -1 -2 -3 -4
K?4 0
3o/2 3a f 2 3a/2 3a/2 3a/2 3a/2 3a/2 3a/2 3a/2
—a/2 -a/2 -a/2 -a/2 -a/2 -a/2 —a/2
'Ft -2a -2a -2a -2a -2a Weak field
'Dt 0 0 0 0 0 T factors
'P* a/2 a/2 a/2 a/2 a/2 L$-coupling
*Pl -a/2 -a/2 -a/2
*Po —a
'So 0
III
ill
-a/2 -a/2
-a/2
-a/2
-a/2
— a/2
-a/2
-a/2
-a/2
—a/2
-a/2
-a/2 -a/2 -a/2
-a/2 -a/2
mim -a/2 -a/2
-a/2
-a/2
—a/2
-a/2
-a/2
-a/2
Weak field
(iDt
(H)« 2a 2a 2a 2a 2a 2a 2a 2a 2a T factors
(II). 2a 2a 2a 2a 2a //-coupling
(I f). 2a
(1 !). —3a -3a -3a -3a -3a
(ID. -3a
Interval Interval
Element *^4, — *Fj — 8F j *p. - «Pi - *p. a cm"1
ratio ratio
Problems
1. Calculate the relative separations and intensities for the Zeeman pattern arising
from the transition ZFA — *GA.
2. Calculate the Zeeman patterns for the transition 3Pi — 3D2 (sp-sd) in LS- and
in jy-coupling.
3. Show that Pauli's g sum rule holds for the terms arising from a pp configura¬
tion in LS- and in ^/-coupling.
4. Find the relative intensities of the lines comprising the Paschen-Back pattern
at the bottom of Fig. 13.9. See Fig. 13.8 for the levels, and use the sum rules.
6. Compute the terms arising from two similar / electrons.
6. Show that the T sum rule holds in weak, strong, and very strong fields for a pd
electron configuration (a) in L£-coupling and (6) in ^-coupling.
CHAPTER XIV
COMPLEX SPECTRA
All but singlets in Cr and Fe, and doublets in Mn, have been observed.
Photographs of multiplets illustrating term multiplicities for these
elements are given in Fig. 14.1. The multiplicity and term designation
1 Kossel, W., and A. Sommerfeld, Verh. d. Devisch. Phys. Ges., 21, 240, 1919.
250 INTRODUCTION TO ATOMIC SPECTRA [Chap. XIV
*$ -*p POTASSIUM
ALTERNATION LAW
MULTIPLICITIES
Fig. 14.1.—Illustrating the alternation law of multiplicities for the elements in the
first long period. Photographs taken by the author with a 15-ft. Rowland grating and
mounting.
Sl
+
sa
+
S*
+
*4
252 INTRODUCTION TO ATOMIC SPECTRA [Chap. XIV
p *P *P
p *S *P !D *5 SP 2jD 4S 4P 4Z>
*P'P ii i ti i * M Hi 14 nil
These correspond exactly with the 18 terms obtained for the same
resultant electron configuration in L/S-coupling: 2£j, 2P*j, 2D|j, 2S^,
*M. 'DU> M..» 4*>M.I.I-
The processing vector models for LS- and //-coupling for complex
atomic systems are represented in Fig. 14.3. Just as for two-electron
systems, L% and I*, and and $*, precess rapidly around their respective
resultants L* and S* which in turn precess more slowly around their
resultant J* In //-coupling the spin and orbit of the added electron
are coupled together and precess rapidly around.their resultant j*
which in turn precesses more slowly with JP of the parent term around
J*. It should be pointed out that in //-coupling the parent term
254 INTRODUCTION TO ATOMIC SPECTRA [Chap. XIV
s
LS-Coupling Jj-Coupling
Fig. 14.3.—Processing vector models for complex atomic systems.
by the method outlined above. Here the Pauli exclusion principle must
be taken into account and the strong-field or very strong field quantum
numbers resorted to (see Sec. 13.11). Although this necessitates the
postponement of the calculation of terms until the Zeeman and Paschen-
Back effects have been treated in the next chapter, the results may be
given here.
V\ aP
p2, 'S 1D
p3, 2P 2D AS
pA, >£ lD 3P
p*, 2P
p®, 'S
dl, *D
d\ 'SlD*G 3P3F
*D s I APAF
dA, WZW *P3F WD'FKPI 'P*D'F*G'H
d*, 2£) J PS*D*F*G*I4PAF *AD*0 *S
d\ 'S'D'G *P*F WWW *P*D*F*G*H *D
<P, *D I APAF
W2W *P*F
<*,
d*
It is important that we note here the lSo term arising from six equiva¬
lent p electrons or from ten equivalent d electrons. Omitting these
configurations in the tables, the upper half of each table is identical
with the lower half.
Sec. 14.5] COMPLEX SPECTRA 255
14.5. The Land! Interval Rule.—The Landd interval rule, as already
shown in Sec. 12.8, applies to the special case of Russell-Saunders coupling
and gives the interval rule for the fine structure of each multiple level.
Fig. 14.4a.—Graphical representation of the Land6 interval rule for 4P, 4D, and 4F terms.
If, for example, the spins of three or more electrons are coupled together
to form a resultant S and the Vs coupled together to form a resultant L,
the derivation of the interval rule becomes identical with the derivation
cor two electrons. The interval rule is
written, therefore,
F = AL*S* cos (L*S*) =*
±A(J*2 -L*2 - S*2), (14.1)
where A is a constant for a given
multiple term and is given by 6A
i.e., by the larger of the two J*s. From Fig. 14.4a it may be seen that
the interval ratios for 4P and *D terms are 3:5 and 3:5:7, respectively.
For a hF term, shown in Fig. 14.46 with J = 1, 2, 3, 4, and 5, the interval
ratios are 2:3:4:5.
Of the many examples revealing good LS-coupling and the Land6
interval rule in complex spectra, only a few will be given as illustrations
(see Table 14.4).
Sc I
and, second, the coupling between and the orbit L*, which is given by
_ ,S^_+j*2 - S*2
r = a‘ (J*2 - L*2 - S*2), (14.5)
4 S*2
„ _ _,„S” + sp - s**
r" = a - L*2 ~ S*2). (14.5a)
Adding these,
^r - i +, rr,„ -_ ^
r /-S*2 +igss
•** - S*2 ,S*2 + S*2 - s*2
+ o'
4 S*2 ].
CJ*2 - L*2 - S*2). (14.6)
Since for a given multiplet the values in the bracket are constant,
the fine structure will follow the Land6 interval rule. Substituting the
values of S, s, SP, L, and J in this formula, the following values are
obtained:
t4, r = +K + V«"; 6^3, r = -|a' + V-a";
Ts, r = -ia' - ia"; 6P2, r = +ia' - |al#;
7P2, r = -4a' - Va"; 5Pi, r = +«af -
Since a! and a" are magnetic in character, and therefore positive, the
7P term can only be normal. If, however, a' is greater than 7a", the
6P term will be inverted. This may be seen from the diagrams of
Fig. 14.5, for in the 5P diagram the component of s* along S* adds to
S* and in 5P it subtracts. If, therefore, the s*L* interaction is sufficiently
strong, the energy will be greatest when the component of 8* is most
nearly parallel to L*. This state corresponds to 7P4 on the one hand
and 6Pi on the other. A good example of this is observed in the chromium
258 INTRODUCTION TO ATOMIC SPECTRA [Chap. XIV
spectrum where the two terms arising from 3d54p follow the Land6
interval rule and have the separations 112.5 and 81.4 cm-1 for 7P2.3.4
and —5.7 and —8.8 cm"1 for 6P3,2,i for which the coupling coefficients
a' = 104.2 cm"1 and a" = 12.4 cm"1.
The most common cause for the inversion of spectral terms is to
be attributed to more than half-filled, but incomplete, subshells of
electrons. Although the derivation of spectral terms arising from
more than two equivalent electrons will not be given until the next
Following the scheme already described in Sec. 13.15 these give for
the two triplet terms arising from d8,
Sec. 14.8] COMPLEX SPECTRA 259
The last three equivalent p electrons (in heavy type) represent the
three valence electrons of each atom and are responsible for the optical
spectra. As a typical example of these spectra we shall consider nitrogen
in some detail.
N II, 2p2, lS •P lD
N I, 2p23s, 2S 2P 4P 2D
NII, 2p2, 1S SP 1D
The parent terms are again the low levels of the ion N II. Of the
resultant terms all but the 2S and 2F have been observed in combination
with those of 2p23s.
Similarly the excitation of an electron to a 3d state gives rise to the
terms:
N II, 2p2, l8 8P lD
Of these terms only those built upon *P of the ion are observed. In
this connection it is to be noted that no term above the 8P series limit
262 INTRODUCTION TO ATOMIC SPECTRA [Chap. XIV
has been observed in nitrogen. The longest observed series of terms are
the 4P terms of 2p2nd, where n = 3, 4, 5, 6, and 7.
The observed spectrum lines of the neutral atom of nitrogen lie in
two widely separated regions of the spectrum: one group of lines is
in the extreme ultra-violet and the other is in the visible and near infra¬
red. This is due to the enormous amount of energy required to raise
a 2p electron to a 3s state, as compared with the difference between the
3s and other excited states (see Fig. 14.6). A schematic diagram of
With all but one electron forming completed subshells, the norma1
state, as observed, is given by the one 3d electron as 2D. Observation
shows that when the scandium atom is excited, it is easier to raise one
1 Bowen, I. S.f Asirophys. Jour., 67, 1, 1928.
Sec. 14.9J COMPLEX SPECTRA 263
3d4s2
Fiq. 14.8.—Energy level diagram of scandium.
Sc III, 3d,
Sc II, 3d np,
Sc III, 3d, 2D
Sc II, 3d nd,
These form five series of singlets and five series of triplets approaching
3d,2D as a limit. The second members of these series are the levels
shown in Fig. 12.4 in combination with the triads of 3d4p. The first
member of the configuration series, viz., 3d3d, finds both electrons
in the same orbits so that the Pauli exclusion principle comes in to
exclude three triplets and two singlets. All predicted terms and no
others are observed for 3d2. The higher series members shown in Fig.
14.8 in brackets are predicted but not observed.
Sec. 14.9] COMPLEX SPECTRA 265
In turning to the neutral scandium atom we/find that the main energy
level diagram is built upon two parent configurations 3d48 and 3d* of
the ion. Starting at the lower left in the figure we have the terms arising
from the addition of a 4d electron to 3d4s. These are obtained as
follows:
Sc II, 3d4s, 1D
Sc I, 3d4«4d,
A
2S2P2D2F2G 2S2P2D2F2GA SAPADAF*G
Of these terms five doublets and three quartets have been found.
The second group of terms shown are those attributed to 3d4s4p.
All predicted terms are observed for this first member of the series
but none of those for 5p, 6p, etc.
In the third set of series 3d4sns the first member 3d4s4$ comprises
the normal state of the neutral atom.
The fourth group of series 3d2ns is built upon 3d2 of Sc II. These
terms arise as follows:
Sc II, 3d2, 8P lD 8F Hi
Sc I, 3d2ns, 2S 2P 4P 2D 2P 4F *G
With the exception of 2S the first member of each of these series has been
identified.
Similarly the fifth group of series 3d2np are built upon 3d2 as follows:
Sc II, 3d2, 1S 8P 1D 8P
Sc I, 3d2np,
IAAAA
*P 2S2P2D 4£4P4D 2P2D2F% W'FKi ADAF*G 2FH}2H
Nearly all of these have been observed for the lowest configuration
3d*4p. The remaining terms, 8P and 8F, arise from the configuration
3d*.
Schematic diagrams of the fine structure of two quartet multiplets
are shown in Fig. 14.7. Photographs of the same two multiplets are
reproduced at the top in Fig. 17.11. Altogether about 160 lines have
been identified as belonging to the Sc II spectrum, and about 360 lines
have been identified as belonging to Sc I.1
1 Russell, H. N., and W. F. Meggers, U. S. Bur. Standards, Sei. Papers, 22,329,
1927.
266 INTRODUCTION TO ATOMIC SPECTRA [Chap. XIV
OII, 2p\ 4S 2D 2P
O I, 2p33s, 3S 5£ lD *D lP 3P
& •,% '?%% 'dU 'DVD -fTy 'gVg ’h’h’h 4ft, Vtt
• 3d*4s4s
• 3d*4s5s
• 3d*4s6s
• 3d*4t4p
- 3d*4i4d
- 3d*4»5d
tions are given by the legend at the lower right in the figure. The
normal state of the atom is given by the abbreviated electron configura¬
tion 4s23d2. All of the predicted terms 3Pf lDy 3F, and lG are observed.
In agreement with Hund’s rule the 3F lies deepest. Not far above the
normal state lies a bF term arising from the configuration 3d34$. This
multiple level as well as all levels lying below the first odd bG are metas¬
table levels. Nearly all of the predicted terms arising from 3d34s
are found. As shown by the branching rule, these terms are:
Ti II, 3d3, 2D
Til, 3d34s, lD ZD *P 3P JD 3D l¥ 3F
combine with even terms or even terms with odd. Expressed in terms
of the electron configurations:
Odd Odd
3d24s4p 3<234p
Even Even
The transitions 3d34p to 3d24s2 give rise to some of the strongest lines
in the titanium spectrum and constitute double electron jumps 4p to
4s (Al = 1) and 3d to 4s (Al = 2). The titanium triplet shown in Fig.
14.1 is a photograph of the double electron jump 3d24s2,3F — 3d34p,3F°.
A very plausible explanation of the great strength of such lines has
been given by Condon.1 By a quantum-mechanical treatment Condon
has shown that the terms of two configurations having the same sign
(odd or even), and lying close together or overlapping each other, belong
in part to both configurations. Because of a sort of resonance or periodic
interchange of states, the atom may be thought of as jumping back
and forth between two states having the same quantum numbers,
without radiation. In this sense Condon has shown that double-electron
transitions may be reduced to single-electron transitions (see Sec. 19.3).
Intercombinations between singlet and triplet levels and between
triplet and quintet levels are also commonly observed. A schematic
diagram of an intercombination multiplet is shown in Fig. 14.11, along
with two quintet-quintet combination multiplets. A photograph of the
*F-*G° multiplet is reproduced at the top of Fig. 17.12 and at the left
1 Condon, E. U.t Phys. Rev36, 1121, 1930.
270 INTRODUCTION TO ATOMIC SPECTRA [Chap. XIV
3d6ns _3d*np_
Even Odd
3d54sns 3d54sn ♦po 4qo 4p0 6po 6qo SpO
3d54s HWs
MANGANESE
3d54s4s*
Fig. 14.12.—Energy level diagram of manganese.
-61
-103
Mn II Mn I
9473 cm"1 6199 cm-1
7/2 36 cm”1
i S/2 -
3/2 ■
25
■*/*
151
7‘
117
jl/2
Observed Calculated
Fia. 14.13.—Schematic representation of the branching rule for the p terms in manganese.
Configuration 3d64a 4p.
4700 cm”1 and the lower terms by about 6100 cm""1. The observed
order 8P, #P, 4P, ®P indicates that the interaction between the spins
of 4* and 4p is considerably greater than that between the spin of 4p
and the spin of 3d5. We shall return to the calculation of the fine structure
shown at the right in Fig. 14.13.
If one of the 4s electrons from the normal state of manganese is
excited to a 3d orbit, the most probable among the many terms predicted
for the new configuration 3d64s are the 6D and 4D terms observed. Being
built upon the inverted 3d6,6Z> term of Mn II, both of these D terms are
inverted.
11/2
9/2
6 DV2
V2
3/2
9/2
8r
7/2
3/2
JL Ui JJL ill
•p0-‘D d5sp-d5sd ‘D-6D° d‘s-d‘p ‘D-6F°
J = i j- i-
Xe II, 5p*» *Pi •Pi
Xel, 5p66p,
/A\
(§ i)o.i (i 4)1.2
/A\
(1 2) 1*2 (f ■§■)(),1,2,3
Fia. 14.17.—Fine structure of the « and p terms in neon, argon, krypton, and xenon.
KI CaI ScI Til VI CrI MnI FeI Col Nil Cul Rat Sal YtI ZrI Cal Mol MaI RuI RhI Pol AgI
Fig. 14.18.—Deep lying electron configurations in the elements of the first and second long
periods, illustrating breaks in the building up of the s and d shells of electrons.
To show the general change taking place in the term fine structure
as we go from Ne to A to Kr and to Xe, this structure for 8 and p series
has been plotted to scale in Fig. 14.17. At the left the coupling is neither
LS nor Jj but something between the two.
278 INTRODUCTION TO ATOMIC SPECTRA [Chap. XIV
14.14. The Normal States of the Elements in the First and Second
Long Periods.—It was pointed out in Chap. V that certain irregularities
in the Bohr-Stoner scheme of the building up of the elements occur in
the first and second long periods. Energy level diagrams of all but one
of the known elements in these two periods have made possible the
construction of Fig. 14.18.1 Here the deepest level in each of four of
the deepest lying configurations dn~2s2, dn~ls, dn~lp, and dn has been
plotted with respect to dn_1s of the same element as zero. It is observed
that, while there is a regular progression of the configurations dn~lp
across each period, there is an apparent breaking up of the dn~2s2 curve
into two similar parts. This is only apparent, for when the dn~2s2 curve
is used as the origin in plotting, the other two curves for dn_1s and dn~lp
appear as the broken curves. The curves plotted either way show,
however, a gradual increase in the binding of 3d or 4d electrons over
48 and 5s electrons as we progress from element to element. With the
addition of one, two, and three d electrons there is a strong tendency
to complete half a subshell of electrons so that with the addition of a
fourth d electron to form the normal state of Cr, or Mo, one of the s
electrons also goes over into a d state to yield 3d64s, or 4d55s. With
the shell half filled, the next electron added to form the next element
goes into a 4s or 5s orbit to again fill the s subshell. At the approach
of a complete subshell of d electrons we observe that additional d electrons
are again added at the expense of the s subshells. No such anomalies
are found to exist in the sequences of elements where p electrons are
added.
14.15. Houston’s Treatment of One s Electron and One Arbitrary
Electron.—A quantum-mechanical treatment of the fine structure due
to the interaction of one s electron with one arbitrary electron has
been made by Houston,2 by applying the Darwin-Pauli treatment of the
electron to the Schrodinger wave equation for two electrons. By
setting up suitable wave functions, based upon Schrodinger’s theory,
solutions for the spin-spin resonance interaction between the two electrons
and the spin-orbit interaction for the arbitrary electron are obtained.
The treatment is one involving perturbation methods in which only
the first-order perturbation to the energy is evaluated. Houston’s
results may be given most easily by his resultant equations;
1 Graps, R. C., and H. E. White, Nat. Acad. Sci., 14, 559, 1928.
* Houston, W. V., Phys. Rev., 33, 297, 1929.
Sue. 14.15] COMPLEX SPECTRA 279
and the triplet fine structure following the interval rule with the ratio
2:1. If X << A we have ^‘-coupling with the same outside triplet
separation. This is just the result obtained from the classical theory
given in Secs. 12.7, 12.8, and 12.9 (see also Fig. 12.13). For negative
values of X the middle triplet level becomes the singlet, and the singlet
level becomes the triplet. Such cases of the singlet lying below the
triplet are observed in the diffuse series of Cd and Hg.
Observed triplet and singlet intervals for Si, Hg, Ge, Sn, and Pb
are seen (in Fig. 14.19) to fit the theoretical curves to a remarkable
degree of accuracy. In plotting the observed terms, the value of A
is first determined by dividing the observed intervals 3P0 — 5P* by three.
This gives a value of A in wave numbers which, with the observed inter¬
vals of *Pi and 3Pi substituted in Eqs. (14.7a) and (14.7c), respectively,
gives two values of X. These are averaged and the terms plotted at
that X. The configurations plotted are: Se, 3p4s; Hg, 6p6s; Ge, 4p5«;
Sn, 5p6s; Pb, 6p7s.
Many other configurations are found to fit these curves at large
values of X (not shown in the figure), where the singlet lies far above
280 INTRODUCTION TO ATOMIC SPECTRA [Chap. XIV
the triplet and the triplets have the ratio 2:1. This figure is also to be
compared with the observed silicon series in Fig. 12.13.
In addition to the interaction energies Houston’s formulas include
the Landd g factors for the weak-field Zeeman effect. They are:
_X±1_
Singlet lgL
1 + 2L(L + 1) 2L(L + 1)[(X + 1)’ + 4L(L + 1)]»
(14.8a)
L + 2
(14.8b)
L+ 1
X + 1
Triplet ^ 2L(L + 1)
+ 2L(L + 1)[(X + 1)* + 4L(L + 1)]*
(14.8c)
L - 1
(14.8eJ)
L
For the outside triplet terms these formulas give the same g factors
as before and are independent of the coupling scheme. According to
the g sum rule, the sum of the g’s for the two terms with the same J
value is the same for all couplings and all field strengths. This is given
by the sum of Eqs. (14.8a) and (14.8c):
it is inverted. This latter result means only that the above given A
factor is negative instead of positive. In order that the singlet be above
the inverted triplet, as observed, X must also be negative.
Substituting —X' for X and —A' for A in Eqs. (14.7a, b, c, and d),
formulas for the configurations p5s, d?s, and are obtained. Curves
plotted from these new equations are given in Fig. 14.20, along with
the observed intervals for a number of elements. It will be noted
that, even for highly ionized atoms like In IV and Sn V, the observed
intervals are in excellent agreement with theory. The configurations
Fig. 14.20.—Fine-structure curves for the configurations ph8 and d*a. Observed terms are
shown to agree with the theoretical curves. (After Laporte and InQiis.)
Value »F - - HI W _ id *F - 1S
Table 14.5, for example, the following experimental values have been
obtained by Harrison.1
Fig. 14.21.—Fine-8tructure curves for the configurations ps and p4. (After Goudsmit.)
Observed terms are shown to agree with the theoretical curves.
155)
*i!
-+-LS | Coupling
Fig. 14.22.—Fine-structure curves for the configuration p3. (After Inglis.) Observed
terms are shown to agree with the theoretical curves.
It has also been shown that the relative intensities between related
multiplets, taken as a whole, obey the sum rules (see Sec. 12.16). For
a titanium triad of quintet multiplets, for example, the following values
have been obtained by Harrison and Engwicht:1
90 70 50 calculated
81 70 49 measured (without and T corrections)
89 70 48 measured (with vK i
The measures taken from many multiplets show that better agreement
between theory and experiment is obtained when v4 and temperature
corrections are made.
Problems
1. The following list of observed lines form a multiplet. Plot the lines to a
frequency scale and, looking for equal frequency differences by marking them on a
1 Harrison, G. R.. and H. Engwicht, Jour. Opt. Soc. Amer., 18, 287, 1929.
Bsc. 14.18] COMPLEX SPECTRA 285
strip of paper, find the term intervals. Tabulate the lines in multiplet form, as in
Table 14.3. Applying the Land6 interval rule to the observed terms, determine the
J1S, and L values.
2. Using the values of X as read from Fig. 14.19 for Si, Ge, and Hg, compute the
g factors for each of the four terms, respectively [see Eqs. (14.8)]. Show that the
g sum rule holds.
.
8 Derive all the terms arising from the electron configurations rf1*, d*pt p*df p4s,
and d7p.
4. Carry out the calculation of the F, factors for d* as outlined in the middle of
Sec. 14.6, and tabulate them. From this table make another tabulation like that
given in the top third of Table 13.13. From these r sums determine the T values of
the terms as shown in the middle third of Table 13.13.
CHAPTER XV
Fio. 15.1.—Classical models for the motion of complex atoms in a weak magnetic field.
Should the coupling on the other hand be Jj (i.e., the spin and orbit
of the last bound electron coupled together to form jf, and jf in turn
coupled together with JJ of the parent term of the ion to form J*),
the g factor is obtained in exactly the same manner as it is for two
electrons in ./^-coupling [see Eq. (13.14)]. JJ of the parent term is
now treated like j2f the spin-orbit resultant of a single electron to be
coupled with the j* of the other electron, so that
where gi is the g factor for the single added electron derived from Eq.
(10.11), and gP is the g factor for the parent term derived from Eq. (13.8)
or (13.14). It should Be pointed out that the parent term may arise
from an LS- or jj-coupling scheme even though the added electron j*
is coupled to it finally in Jj-coupling. Tables of g factors for LS-coupling
are given in the Appendix. For Jj-coupling there are so many possi
bilities that Eq. (15.3) must be used.
A graphical representation of the magnetic energy corresponding
to the vector model, given by Eq. (15.1), is shown in Fig. 15.2 for a 4Dj
term. The vector J* is quantized with respect to the field direction H
so that its projection M = f, $, i, —i, —— $, —J. Multiplying
J* by the g factor ^ and projecting on H we obtain the values of Mg.
These are proportional to the magnetic energy. The result is eight
288 INTRODUCTION TO ATOMIC SPECTRA [Chap. XV
Fig. 15.2.—Schematic representation of Zeeman splitting for a 4Z)j term in a weak magnetic
field.
Fig. 15.3.—Zeeman effect, theoretical and observed patterns, for a rSt — 7P*,«,4 multiplet.
(Observed patterns are from original spectrograms by H. D. Babcock.)
Fig. 13.4. The intervals between the three lines indicate L&-coupling
with a ratio approximately 4:3. The q factors for 7/Ss, 7Pj, TPs, and 7Pi
(see Appendix) are 2, !•}, and £, respectively. A graphical construe-
Sec. 15.3] ZEEMAN EFFECT, MAGNETIC QUANTUM NUMBERS 289
tion, such as that shown in Fig. 15.2, gives th6 intervals shown in Fig.
15.3.
With the selection rule that in any transition the magnetic quantum
number M changes by +1, 0, or —1 only, the Zeeman patterns shown
below in the same figure are obtained. Another method is to write
down the Mg values for both the initial and final states of a given line
with equal values of M directly above and below each other, as follows:
M =3 2 1 0 -1 —2 -3
Similarly the patterns for the two other lines are calculated as
Fig. 15.5.—Zeeman patterns observed and calculated, for typical quintet multiplet lines
in chromium. {Observed patterns are from original spectrograms by H. D. Babcock.)
292 INTRODUCTION TO ATOMIC SPECTRA [Chap. XV
are just the same as those for two-electron systems. Each spin and
each orbit has a component along H. The sum of these components
-f Smz, gives the magnetic quantum number M of the entire
atomic system in that state.
When considering transitions between levels in weak, strong, and
very strong fields, each multiplet in L£-coupling in a strong field will
go over to a pattern resembling a normal triplet (see Fig. 13.9). In
a very strong field all the lines arising from the configuration of electrons
will spread out into a pattern again resembling a normal triplet (see
Fig. 13.14). Patterns similar to those of Figs. 13.11 and 13.15 will
result in the case of /./-coupling. We shall now make use of the weak-
and strong-field and very strong field quantum numbers to calculate
the allowed spectral terms arising from given electron configurations.
"V V* -/2 _ ^
= 1 0-1 ESS 0
Ms = 1 0 V4 Ms = 3/2 '/z -yz a ms = ’/2 Vz
; -Vi Ms= K -V4jj-H -Ji J/2 -Ji
Sing.'TripJ Doubt 'Quar. ms3 Doub.1 mS3
D
d :| 2 1 0 -1 -2 |
m, s 2 1 0-1-2 _ Ml; 3 2 10 -1 T
*2
2 1 0-1-2 0 S Ml = 2 1 6 -i ! -2 0 P
Mtp=
D mlt Ml. 1 0 -1 ;-2 -3 -i
■
P | D F
Fia. 16.6.—Combination of very strong field quantum numbers for the electron configura¬
tion dap.
Parent terms
Final terms
The first and fourth arrays give the first row; and the second, third,
and fifth arrays give the last row. If now each final run of Ms values
is combined with each run of ML values, all of the weak-field quantum
numbers of the above final terms are obtained. This has been done
F
ml= 3 2 10-1-2-3
M = % 7/2 5/2 34 1/, -1/2-3/2 %
M« i % % \ 4 -%l-%
M= % % 14 J/2
M = % V2 -Vfe-%1 -%!-%!-%
4p |4p 4p ,4p Ms
F3/! F5/l F%\ F%
Fig. 16.7.—Combination of the strong-field quantum numbers of a AF term showing the
formation of resultant weak-field quantum numbers.
for the 4F term in Fig. 15.7. The others will be left as exercises for the
reader.
15.6. Equivalent Electrons and the Pauli Exclusion Principle.—In
the calculation of spectral terms arising from two or more equivalent
electrons we must take the Pauli exclusion principle into account and
start with the very strong field quantum numbers (see Sec. 13.11).
Consider as an example three equivalent p electrons. We first write
down the six possible states for a single p electron in a very strong field.
They are:
m.= i i i ~i
mi = 1 0 -1 1 0-1 (15.4)
(o) (b) (c) (d) (e) (/)
Since the exclusion principle requires that no two electrons have
all quantum numbers n, l, m, or mi alike we collect all possible com-
294 INTRODUCTION TO ATOMIC SPECTRA IChap. XV
binations of the above states three at a time, with no two alike. They
are:
Table 15.1
Writing Ms for 2m, and ML for 2m/, the results are tabulated as follows:
For Ms — h Ml = 0
For Ms = i = 0, 2, 1, 0, — 1, —2, 1, 0,-1
For Ms = -i ML = 0, 2, 1, 0,-1,-2, 1, 0,-1
For Ms - - Ml = 0 (15.5)
These are just the strong-field values for *S, 2D, and 2P terms (see Table
14.3).
In a similar calculation for four equivalent p electrons we write
down the same six possible states for one p electron [Eq. (15.4)] and
take all possible combinations of four states at a time with no two alike.
They are:
Table 15.2
abed abde aede adef beef
abce abdf aedf bede bdef
abef abef acef bedf edef
When m, and mi values are summed for each combination and tabulated
one finds:
For Ms = 1, Ml = 1, 0,-1
For Ms « i, Ml = 1, 0,-1
For Ms = -i, Ml - 1, 0,-1
These are just the same values given in Eq. (15.4) for one electron and
give rise to a 2P term.
For six equivalent p electrons there is but one combination of the
very strong field quantum numbers, taking six states at a time, with no
Sec. 15.6] ZEEMAN EFFECT, MAGNETIC QUANTUM NUMBERS . 295
i- t t ! t i i
m, = | \ -4 -f 4 -4
(a) (b) (c) (d) (e) (f)
Since the Pauli exclusion principle requires that no two electrons have
all the quantum numbers n, l, j, and m, alike, we collect all possible
combinations of the above given states three at a time, with no two alike.
These will be just the combinations given in Table 15.1. Summing up
the values of m, for each of the 20 combinations the results may be
tabulated as follows:
For ji = i) J 2= |= % j wij==I A) 2) 4
Forji555^, ji j 3 2j 4* 2) 1) 4»4» 4> 4
Forji = |, j2 = i, j 3 = i m, = |,4,-4,“f
These correspond to the five terms (j f |)j, (f | i)4, (f f 4)*, ({ |J)t,
and (| 4 i).], which go over in LS-coupling to 2P3, 2P3, 2PS, 2D3, and
4£3 (see Fig. 14.22).
We shall now turn to the more complicated cases of equivalent d
electrons. Two equivalent d electrons have already been treated in
Sec. 13.11 and shown to give rise to lS9 3P, lD, 8P, and *G terms. For
three or more electrons we continue the same scheme by first writing
down the 10 possible states for one d electron:
m. = 4 4 4 4 4 —* —* —* —* —*
mi = 2 1 0-1-2 2 1 0-1-2 (15.6)
(a) (6) (c) (d) (e) (f) (g) (A) (i) (j)
For d8 we take these three at a time with no two alike and find 120
combinations that have to be collected and segregated. The resultant
sums will correspond exactly to the terms given in Table 14.3. For
d4, taking four at a time with no two alike, there are 210 possible Com¬
binations to be evaluated. Such calculations become tedious and
296 INTRODUCTION TO ATOMIC SPECTRA [Chap. XV
*P
A *D 2F *P 2D 2F HI 2H
Assume, as another example, that the correct terms have been derived
for d, d*, d*, and d4 and we wish to calculate the terms for d6. Referring
1 Gibbs, R. C., D. T. Wilber, and H. E. White, Phys. Rev., 29, 790, 1927.
1 Russell, H. N., Phys. Rev., 29, 782, 1927.
Sac. 15.6J ZEEMAN EFFECT, MAGNETIC QUANTUM NUMBERS 297
to Eq. (15.6) we take only those combinations five at a time where all
five m, values are plus, or all five minus. These are abode and fghij
Summing these we have for the plus values
Ms = 4, Ml = 0
and for the minus values
Ms - Ml - 0. (15.8)
These are just one-third the values needed to form a 6S term.
There will now be a set of combinations from Eq. (15.6) in which
four spins will be plus and the fifth one minus, or vice versa. That
part of the combinations in which four of the spins add gives quintets.
From Table 14.1 we find for d4 one quintet only, 8Z>. The fifth electron
with opposite sign corresponds to 2D. Combining these in all possible
ways we get
6Z>
Striking out the S term to go with Eq. (15.8), we have left parts
of 4P, 4Z>, 4F, and HI terms. It should be pointed out here that we now
have those parts of the *S term for which Ms = f, — f, and — f, and
those parts of each-quartet term for which Ms = $ and —f.
Again there will be a set of combinations from Eq. (15.6) in which
three spins will be plus and the two others minus, or vice versa. The
combinations of d3 have already been shown to give 4P and AF terms,
and the combinations of d2 to give 3P and 3F terms. Combining these
in all possible ways to form doublets,
4p 4p ip 4p
2s 2P 2Z> 2D 2F HI 2D 2F HI 2S 2P 2D 2F HI 2H 2I
Striking out one S term to complete the *S term above, and one
each of P, D, P, and G, to complete the 4P, 4D, 4P, and HI terms, respec¬
tively, we have the remaining terms as doublets. The resultant terms
arising from five equivalent d electrons are, therefore,
d6, 2Z>, 2P, 2Z>, 2P, 2G, 2H} 2S, 2D, 2P, 2G, 2/, 4P, 4P, 4Z>, HI, *S.
Since a subshell of d electrons lacking n electrons to complete it
will give rise to exactly the same terms as a configuration of n equivalent
298 INTRODUCTION TO ATOMIC SPECTRA [Chap. XV
electrons, written symbolically d10”* = dn, the lower half of Table 14.3
is symmetrical with the upper half.
A continuation of this process for equivalent f electrons leads to
the terms given in the tables in the Appendix. The number of possible
combinations for each configuration of equivalent electrons is given in
parentheses at the beginning of each row. These numbers are computed
from the well-known combination-theory formula for p things taken q
at a time.
Problems
1. Compute and plot the Zeeman patterns (intervals and relative intensities)
for the transitions ’^G^ — 7F6, 7U6 — 7Fe, 7Cr6 — 7F6, and *Pg — 8aSj.
2. Starting with the very strong field quantum* numbers, derive the terms arising
from the electron configurations 3d4p5p and 3dHp (see Secs. 15.5 and 13.11).
3. Using the shorthand method outlined in Sec. 15.6, calculate the spectral terms
arising from (o) four equivalent d electrons, (6) two equivalent/electrons.
CHAPTER XVI
X-RAY SPECTRA
and the equal angles of incidence and reflection 6, the x-ray wave-length
X is given by the well-known Bragg equation
nX = 2d sin 6. (16.1)
1 Walter, B., Ann. d. Phys., 74, 661, 1924; 75, 189, 1924.
2 Backlin, E., Dissertation, Uppsala Univ. Arsskr., 1928.
* Larsson, A., Dissertation, Uppsala Univ. Arsskr., 1929.
4 Kellstrom, G., Dissertation, Uppsala Univ. Arsskr., 1932.
Sec. 16.1] X-RAY SPECTRA 301
the reflection of x-rays at a grazing angle from plane polished mirrors'
(Lloyd mirror experiment) have also been photographed by Linnik1
and Kellstrom (see Fig. 16.3a).
Bergen Davis and Slack,2 * and Siegbahn,8 have been able to refract
x-rays by passing them through prisms of various materials, and Comp¬
ton and Doan,4 * Siegbahn, and others have diffracted x-rays from
mechanically ruled gratings. It is with ruled gratings that Osgood,*
Thibaud,6 and Siegbahn and Magnusson7 and others have crossed
the last gap in the electromagnetic wave chart by photographing and
1 Linnik, W., Zeiis. f. Phys., 66, 107, 1930. For other photographs of single-slit
diffraction patterns Bee M. Siegbahn, “Spektroskopie der Rontgenstrahlen,” 2d ed.
1931.
* Davis Bergen, and C. M. Slack, Phys. Rev., 26, 881, 1926.
* Siegbahn, M., Jour. Phys., 6, 228, 1925.
4 Compton, A. H., and R. L. Doan, Proc. Nat. Acad. Set., 11, 598, 1925.
* Osgood, T. H., Nature, 119, 817, 1927; Phys. Rev., 30, 567, 1927.
* Thibaud, J., Jour, de phys. et radium, 8, 13, 1927; 8, 447, 1927; Phys. Zeiis. 29,
241, 1928.
1 Siegbahn, M., and T. Magnusson, Zeiis. f. Phys., 62, 435, 1930.
* Barkla, C. G., and C. A. Sadler, Phil. Mag., 17, 739, 1909.
302 INTRODUCTION TO ATOMIC SPECTRA [Chap. XVI
16.2. X-ray Emission Spectra and the Moseley Law.—In 1913
Moseley,1 from a systematic study of the K radiation of Ca 20, Sc
21, Ti 22, V 23, Cr 24, Mn 25, Fe 26, Co 27, Ni 28, Cu 29, and Zn 30,
i
Angstrom Units
0 1000 2000 3QG0,
X-Umts
Fig. 16.4.—Wave-lengths of K-series x-ray lines. (After Moseley.)
announced what is known as the Moseley law. In brief this law may be
stated as follows: The frequency of each corresponding x-ray line is
approximately proportional to the square of the atomic number of the
10 20 30 40 5j) ^60 70 60 90
change in separation of, and the stepwise shift of, corresponding wave¬
lengths in going from one element to the next are unmistakable in their
meaning. For the first time one could say with certainty that the order
of the chemical elements is as shown in Fig. 16.4, and also that the
elements in the periodic table are constructed with extreme regularity.
Prior to Moseley's discovery, cobalt and nickel, because of their atomic
weights 58.9 and 58.7, respectively, were listed in the order nickel, cobalt.
Moseley's x-ray photographs show without doubt that so far as structure
is concerned, the order is cobalt, nickel. Optical spectra are also in com¬
plete agreement with this result. In Moseley's work the x-rays were
photographed in the second and third orders as diffracted from a crystal
of potassium ferrocyanide (grating constant d = 8.408 A).
A continuation of Moseley's early work by other investigators
has shown that in the heavier elements each of the Ka and K$ lines
(see Fig. 16.4) is in itself a close doublet. These four lines are usually
designated Ka, KatJ Kffi, and K$z. X-ray spectrograms of all available
elements now show that Moseley's law continues throughout the periodic
304 INTRODUCTION TO ATOMIC SPECTRA [Chap. XVI
table. From a graph similar to the one shown in Fig. 16.4a Moseley
arrived at the expression
Pcm-I = KR(Z - a)\ (16.2)
where R is the Rydberg constant = 109737 cm”1, and <r and K are
constants. For Ka, a = 1 and K = $. By analogy with hydrogen-like
atoms and Balmer’s formula [see Eq. (1.24)], Moseley wrote
L line Cb 41 Mo 42 Ru 44 Rh 45 Pd 46 Ag 47 Cd 48 In 49 Sn 60
1 Siegbahn, M., Verh. d. Deutsch. Phys. Ges., 18, 278, 1916; Compt. rend., 162,
787, 1916.
Zeits. f. Phys., 10, 129, 236, 1922.
1 Dolejsek, V.,
’These photographs are reproduced from original negatives kindly loaned by
F. R. Hirsch.
Sec. 16.3] X-RAY SPECTRA 307
A search for absorption edges over the large range of x-ray wave-'
lengths reveals in the heavier elements, in particular, several different
edges for each element. For a given element the absorption edge
occurring at the shortest wave-length is called the K absorption limit.
Beyond the K limit to longer wave-lengths three relatively close absorp¬
tion edges have been photographed for the elements starting at about
Rb 37. These three edges are called the L limits, Ltf Ln, and Lm, in
the order of their wave-lengths.1
Ag
Ag
Co
In
Sn
Sb
Te
I Ag
Fig. 16.9.—X-ray absorption edges of silver, cadmium, indium, tin, antimony, tellurium,
and iodine. (After Erode and Burmann.)
Shells K L M N 0
V 2 3 p* 4pa
d 3d10 4d10
4s, 4p, 4d; and 5s, respectively. Since these binding energies are meas¬
ured by the absorption edges, discussed in Sec. 16.3, an energy level
diagram is obtained by plotting just the frequencies of the absorption
limits.
At the bottom of Fig. 16.11 the normal state of the cadmium atom
with all of its electrons is represented by a state, where S *= 0, L = 0,
and J = 0. The removal from the atom of one of the 5s electrons
raises the atom to the lowest state, where £ = J, L = 0, and J —
This ionized state is the single limit of the regular optical series of singlets
and triplets (see Fig. 11.6). The notation 5s_1 means that a 5s electron
has been removed from the neutral atom. This leaves the atom with an
incompleted subshell, a 5s electron.
The removal of a Is electron from the neutral atom, on the other
hand, raises the atom to the highest energy state K shown at the top of
the figure. The removed electron is represented by Is"1. With but
one electron remaining in the Is subshell this energy state is represented
by 2Si, in the notation used in optical spectra. The term values given
at the left in the figure show that the binding energy of a Is electron is
about 3000 times that of a 5s valence electron.
310 INTRODUCTION TO ATOMIC SPECTRA [Chap. XVI
2.16-10* K M- 2 1
l S^- s - ts-V
[2 5510*1
K'Series
L-Series
63 106 M,
51 10°
. (Li J. 3s
3p?
33 10* My'
fe.81107*
IS
M-Seriei
3d9
1 Any subshell of electrons lacking but one electron to complete it gives rise to an
inverted doublet level having an L value equal to the l value of the missing electron.
Sec. 16.5] X-RAY SPECTRA 311
as in optical spectra are always single; other levels P, D, F, . . . are
double.
Attempts oy Moseley and his successors to find a simple law governing
the terms, or, what is the same thing, the absorption limits v, as a function
of the atomic number Z, has led to the so-called Moseley diagram shown
An arbitrary,
AL = +1 or -1, (16.5)
AJ = +1, 0, or —1.
1 Einstein, A., Verh. d. Deutsch. Phys. Ges., 18, 318, 1916; Phys. Zeits., 18, 121,
1917; see also Dirac, Proc. Roy. Soc., 114, 243, 1927. There is now some question
whether the ** factor belongs in the intensity formulae for spectrum lines.
1 Coster, D., Phil. Mag., 48, 1070, 1922; see also Kaufman, S., Phys. Rev., 40,
116, 1932.
Sbc. 16.5J X-RAY SPECTRA 313
Table 16.3.—Observed Relative Intensities for the K Series Lines
{After Meyer)1
’Pi 'Pi
9 0 My
!0i 1 5 Miy
Lni Ln
The observed intensities for eight elements are given in Table 16.4.
. ., , ,
» JONsaoN, A., Zeits /. Phye 36 426, 1926; 46 383, 1928.
. .,
* Allison, S. K, Phyv Rev 80, 245, 1927; 81, 1, 1928.
Peculiarly enough when vx corrections are made for these lines they
are far from being in agreement with theory. Bethe1 explains this on the
basis of the fact that the measured intensity is not equal to the transition
probability but to the product of the life time of the initial state and the
• transition probability. The measurements thus indicate simply that the
life time of Li (due, e.g.y to a greater probability of autoionization) is
1 Bethe, H., Handbuch der Physik., Vol. 24, No. 1, p. 468, 1933.
314 INTRODUCTION TO ATOMIC SPECTRA [Chap. XVI
shorter than that of Lu,m. It may be seen that the Lm lines in Fig. 16.5,
where resolved, are quite faint as compared with the strongest line Lai.
The Lp lines are taken from different plates and may not justly be
compared.
16.6. Fine Structure of X-ray Levels.—The three L levels, the five
M levels, the seven N levels, etc., may be referred to as the fine structure
of x-rays. Since a fairly satisfactory account of this fine structure may
be given, a treatment of the splitting of each main energy level into
2n — 1 sublevels is made the object of this and the three succeeding
sections.
It has already been pointed out that the L electrons move in an approx¬
imately Coulomb field, the two K electrons acting very much as if they
had coalesced with the nucleus. This screening of the nucleus by the
K electrons reduces the nuclear charge by about two, so that, neglecting
for the moment the fine structure, the energy should be given by the
hydrogen-like formula
T = ^ 7 g)-> (16.6)
n2 n2
formula for hydrogen [see Eq. (9.6)] to his x-ray formula. Replacing
Z by Z — a, or Z — 8,
in Eq. (16.7), for with his values of k we shall obtain the quantum-
mechanical term values (see Sec. 9.9).
Table 16.5
The effect of the screening constants <r and s in shifting the energy
levels from those of hydrogen-like atoms to the observed x-ray levels
is illustrated in Fig. 16.13. At the left the Bohr term values are given
r= 7---' (i6.9)
where <r is different for the different
sublevels s,p,d, ... . Adding finally
Niv — Ny, Nvt — Nyiif Ou — Om, Oiv — Ov, etc., which prior to the
quantum mechanics were called regular doublets. The intervals between
each of these doublets in the various elements in the periodic table are
known accurately from observed differences between a number of x-ray
spectrum lines (see Figs. 16.5 and 16.7).
The substitution in Sommerfeld’s formula of the known quantum
numDers n and k for a given doublet leads to the so-called regular doublet
law. Substituting in Eq. (16.7), one obtains two equations for each
doublet. For the L doublet (22P* — 22P8),
+ ^|>-*)4 + (16.11)
}•
Now if Ay represents the difference between tnese two equations,
then
Although only the first term of the analogous Eq. (9.6) is required
in describing the hydrogen fine structure, Sommerfeld1 has shown that
the first term of this equation does not suffice to determine s and that
higher terms are necessary.
From the observed x-ray lines, shown in Fig. 16.5 and tabulated in
Table 16.1, the following screening constants (Table 16.6) are computed:
Element Z Av/R Z - s 8
Cb 41 ■m 37.53 3.47
Mo 42 38.50 3.50
Ru 44 9.49 40.54 3.46
Rh 45 10.48 41.53 3.47
Pd 46 11.57 42.55 3.45
Ag 47 12.69 43.52 3.48
Cd 48 13.96 44.52 3.48
In 49 15.29 45.50 3.50
Sn 50 16.72 46.51 3.49
2 *Si 2- 2aPj 3aS* 3aPj - 3aPj 3’Dj - 3*Z>, 4’S, 4aPj - 4aPj
Li Lu — Lm Mi Mn M in M iv My Ni Nu Nm
2 0 3 50 6 8 8.5 13 14 17
Now the values of Ay for some 70 elements are known quite accurately
from the average of several observed x-ray doublet intervals Kai — Kat,
Lfa — La„ L„ — L\, Lfit — L«, and the absorption limits Ln — Lm.
The average values of Ay are plotted in Fig. 16.14 with four different sets
of ordinates y/Av/R, y/Av/R, yfAv/R, and y/Av/R. The y/Av/R
curve, with an intercept at 3.5, shows that the fourth power law [Eq.
(16.16)] gives the best agreement.
16.9. Screening Doublets and the Irregular-doublet Law.—Making a
systematic study of L series lines, Hertz1 in 1920 announced what is known
as the irregular-doublet law. This law applies to those pairs of energy
levels having the same n, S, and J values but different L values and
states that the difference between the square roots of the term values of a
given doublet is a constant independent of the atomic number Z. This
law found by Hertz to hold for the Li and Ln levels of the six elements
Cs, Ba, La, Ce, Pr, and Nd is now known to hold for all doublets la — Ln,
Mi — Mu, Mm — Miv, Ni — Nn9 Nin — Nw, and Ny — Nyi (see
Fig. 16.12). That the law is only approximately true is shown by the
values of Ay/T/R given in the last column of Table 16.9. A general
widening of the interval with increasing Z is observed.
In formulating the irregular-doublet law Hertz began with the
equation
1 Hertz, G., Zette. /. Phys.} S, 19, 1920.
320 INTRODUCTION TO ATOMIC SPECTRA [Chap. XVI
r _*(*-*)».
(16.17)
n2
Transposing J? and taking the square root, this equation becomes
(16.18)
If now <r i and <r2 are the screening constants for the two levels of a doublet,
the difference is written
Element
y/T/R Vt/r A VrjR
(M (in)
evaluated. Dropping all but the first two terms in Eq. (16.7), and
substituting the value of T", we obtain
T, _ T _ R«*(Z - aWn _ 3\
(16.21)
1 “ 1 n4 \k i)
Sic. 16.0] X-RAY SPECTRA 321
J
Here T represents the observed term values given directly by the absorp¬
tion limits.
Cs 55 La 57 Ce 58 Nd 60 Sa 62 Gd 64 Dy 66 Er 68
a^/W/r
2*S - 2*P 0.59 0 60 0.59 0 59 0 59 0.58 0.58 0 58
Ay/W/R
3‘S - 3*P 0.54 0.54 0.51 0.54 0.56 0 51 0.51 0.53 0.54 0.53
Ay/TjR
3*P - 3 *D 1.21 1.19 1.20 1.19 1.19 1.19 1.20 1.19 1.18 1.19
322 INTRODUCTION TO ATOMIC SPECTRA [Chap. XVI
Although the values shown here are quite constant, some of the values,
for elements not given in the table, deviate slightly more from the
average value. It should be pointed out that there are no spin-relativity
doublets from which to calculate s and a for the 2S levels. Wentzel,1
however, making the assumption that the irregular-doublet law is valid
for the reduced terms, has adjusted the screening constants $ for these
levels until the curves for 22S, 3 2S,
and 42S (see Fig. 16.15) are parallel
NffNJv to the center curves as shown. These
NfNm values of S have been included in
Table 16.8.
In Fig. 16.15 the dashed lines
through the origin represent hydrogen-
Ml like atoms for which there is no
screening. The difference between
these ideal curves and the corres-
j-n Lm
Ll ‘ ponding reduced curves
LS-coupling JJ-coupling
AS = 0 AJ i = 0
|or vice versa
AL = +1, 0, -1 AJ2 = +1, 0, -1
AJ = +1, 0, -1 AJ = +1, 0, -1
subject to the further restriction that odd terms combine only with even
terms.
Since the splitting of x-ray energy levels, due to a resultant angular
momentum of the valence electrons, is in general very small, a number
of the radiated lines will fall together as a single line. For this reason,
and by virtue of the fact that the sum rules are valid for multiplets, the
selection and intensity rules given in Sec. 16.5 will, when applied to the
one-electron jumping in the inner part of the atom, describe adequately
the observed lines.
16.11. X-ray Satellites.—X-ray satellites are those relatively weak
lines often observed close to and on the high-frequency side of the chief
x-ray diagram lines. These satellites, sometimes called non-diagram
lines, were first discovered by Siegbahn and Stenstrom3 on the high-
frequency side of the K series lines of the elements Na 11 through Zn 30.
Since their discovery an extensive investigation of the L series lines by
Richtmyer and Richtmyer4 and others has shown that a number of
satellites appear on the high-frequency side of the chief L lines of most of
the elements from Cu 29 through U 92. Similarly, Hirsch5 and others have
shown that M series satellites are to be found in the elements Yb 70
through U 92. The regions of a number of these satellites are marked
in Figs. 16.5 and 16.7 and may in some of the reproductions be traced
from one element to the next. Semi-Moseley plots of the satellites of
Fia. 1C. 18.—Semi-Moseley diagram of x-ray satellites. (After Richtmyer, Hirach, and
Ford.)
1 Wentzel, G., Ann. d. Phys., 66, 437, 1921; Zeits. /. Phys., 31, 445, 1925.
s Richtmyer, F. K., Phil. Mag., 6, 64, 1928; Jour. Franklin Inst., 203, 325, 1929.
* Druyvesteyn, M. J., ZeiU. f. Phys., 43, 707, 1927; Dissertation, Groningen,
1928.
326 INTRODUCTION TO ATOMIC SPECTRA [Chap. XVI
Calculated
Element Observed
Z A v/R
I'll.hi Li
A1 13 1.79 2 23 1.82 2 34
Si 14 2 07 2 74 2.11 2.63
P 15 2 38 2.97 2.40 2.92
S 16 2.97 2.70 3.22
a 17 3.27 2.97 3 49
K 19 3.59 3 55 4 07
Ca 20 3.83 3 84 4 36
Sc 21 3.77 3.53 4 05
Ti 22 3! 72 3 73 4 25
V 23 3.92 3 95 4 47
Cr 24 4.00 4 16 4.68
Mn 26 4.34 4.38 4.90
Fe 26 4.69 4 58 5.10
When one takes the fine structure of the L levels and neglects the
relatively smaller fine structure of the M levels, Eq. (16.26) predicts a
1The L differences being always greater than the M differences, Eq. (16.25)
indicates that the satellites will appear on the high-frequency side of the parent
x-ray line (see Fig. 16.5).
Sbc. 16.11] X-RAY SPECTRA 327
fine structure for Kp"’. Since the Ln — Lm interval is small as compared
with Li — Ln, the observed fine structure should first reveal itself as a
doublet. This splitting is observed in the first three elements, Al, Si, P.
It should be pointed out that, if the fine structure of both the initial
and final states is considered in the above example, many different
transitions become possible. The Kp satellites should therefore be
composed of a number of lines. Consider, for example, the possible
ways in which satellite transitions KL — LM may take place. All
of the possible terms and transitions are shown in the following scheme
(Table 16.12).
Table 16.12
Even Odd
K Li K Lii.ni
1«-1 28"1 l*"1 2p-'
lS *S xp ip
S' I
iP 3P ip »p ipiDlF*P*D*F 'S *S *D 'S'P'D'S'P'D
2s"1 3p~l 2p~l 3s*1 2p“l 3d"1 2s-1 3s"1 2s"1 3d"1 2p-» 3?-1
Li Mii.hi Lnjii Mi Lii'in AfIV,v Li Mi Li Afiv.v Ln,m AfntIXI
Odd Even
The configurations are given in terms of the missing electrons and the
levels in LS-coupling. Regardless of the type of coupling to be expected
between two adjacent incompleted subshells, the number of terms will be
the same as that given in the table above. Of the many allowed transi¬
tions it should be noted that those given by KLi — Ln,mMiv,y involve
the double electron jump 2p3d to 152s.1 All double electron jumps are
shown by dotted lines. In this diagram we see that both single and
double electron jumps are possible. Although the double jumps are
possible, they should give rise to weaker lines.2
Langer3 and Wolfe4 have suggested that the five well-known satellites
of Ka (see Fig. 16.18) are to be attributed to the transitions KL—>LL.
Following Wolfe’s assignments, based upon quantitative calculations
of the electron-interaction energies in a Hartree field for the potassium
atom (see Fig. 7.2), a schematic energy-level diagram has been con¬
structed in Fig. 16.19. The parent K and L levels are shown at the left
with the unresolved Kaiit doublet below. The removal of a second
electron from the atom splits these levels (neglecting valence electrons if
1 The selection rules for double electron jumps are Ah * 1, AU * 0, 2.
* A treatment of double electron jumps in x-ray spectra has been given by E. G.
Ramberg, Phys. Rev., 46, 389, 1934.
* Langer, R M., Phys. Rev., 37, 467, 1931.
4 Wolfe, H. C., Phys. Rev., 48, 221, 1933.
328 INTRODUCTION TO ATOMIC SPECTRA [Chap. XVI
here are any in unclosed subshells) into singlets and triplets as shown.
As pointed out by Wolfe, LS-coupling is to be expected in these levels
since the spin-orbit interaction energy (shown by the Ln — Lm interval)
is small compared with the interaction between the two electrons (shown
by the satellite separations). Five of the six allowed transitions account
for the five observed satellites. The sixth allowed transition, shown
as a dotted line, is a double electron transition and should give rise to a
relatively weak line on the low-frequency side of KThe results show
Via. 16.19.—Splitting up of x-ray levels due to the removal of a second electron from the
atom. Illustrating a possible explanation of x-ray Ka satellites.
time of the excited state is far too small to account for the observed*
satellites.
Although few quantitative data are to be found on < the excita¬
tion potentials of x-ray satellites, observation supports the theory
of double ionization by a single impact. Druyvesteyn has shown for
example that while the parent x-ray lines Kaui in vanadium appear at
5.45 kv.,the satellites K*tA appear only at potentials above 6.09 ± 0.1 kv.
This is not in agreement with WentzePs early suggestion that two
K electrons have been removed from the atom. Assuming Wolfe’s
assignments are correct for Kaiti (see Fig. 16.19), the removal of one
K and one Ln,m electron from vanadium requires about the same
energy as that necessary to remove a K (Is) electron from vanadium,
Z = 23, and an Lu,m (2p) electron from chromium, Z = 24. The
sum of these two ionization potentials is 5.450 + 0.586 = 6.036 kv.,
in excellent agreement with observation. In a similar manner the Lp,
line (due to the transition Lm — N\) in Ag 47 has an excitation potential
of 3.34 kv., and the Lp, satellites should, if attributed to the allowed
transitions LmMx — MxNv, appear at about 20 per cent higher voltages.
i.e., 4.00 kv. This likewise is in good agreement with experiment.
16.12. Explanation of X-ray Absorption Spectra.—In photographing
x-ray absorption edges, it often happens that absorption lines are found
close to and on both sides of the absorption edge. Under high dispersion
and resolving power the lines on the low-frequency sides appear as
absorption lines, whereas those on the high-frequency side appear as
subsidiary absorption edges. In 1920 Kossel1 made the suggestion
that an inner electron, as the result of the absorption of a photon hv, need
not be completely removed from the atom but instead be simply excited
to one of the outer unoccupied states. The lowest possible frequency
to be absorbed therefore would be given by the energy necessary to
carry an inner electron to the most tightly bound but unoccupied electron
orbit. These transitions are governed by the same selection rules valid
in emission spectra and lie within only a few volts of the absorption
edges2 which would correspond to complete removal of the electron.
A detailed study by Kievit and Lindsey3 of the subsidiary absorption
edges, observed on the high-frequency side of the K edges of the elements
Ca 20 to Zn 30, appears to confirm the theory already postulated, that
they are due to the removal of more than one electron from the atom.
Kronig4 has, however, been able to explain this extended structure
more satisfactorily on the basis of the existence of prohibited zones
in the velocity spectrum of free electrons in crystals.
ISOELECTRONIC SEQUENCES
From the theory of penetrating orbits on the other hand they may
also be represented by (see Sec. 7.4)
r <17-2>
1 Millikan, R. A., and I. S. Bowen, Phys. Rev., 24, 209, 1924; 26, 296, 1926;
26, 160, 1926; 27, 144, 1926.
2 Gibbs, R. C., and H. E. White, Phys. Rev., 29,426,1927; 29,666,1927; Proc. Nat.
Acad. Set.. 13. 626. 1927.
Sec. 17.1] ISOELECTRONIC SEQUENCES 333
In the first formula, Z is the atomic number, n is the principal quan-,
turn number and a is a screening constant. In the second formula
Z0 is the effective nuclear charge outside the atom and fi is the quantum
defect. (Z0 takes the values, unity for neutral atoms, two for singly
ionized atoms, three for doubly ionized atoms, etc.) The differences
between the term values of all atoms as given by the first formula and
those of hydrogen-like atoms as given by
RZ2
T = (17.3)
n2
are accounted for by the screening constant cr, whereas in the second
formula the differences are attributed to the quantum defect /*. We
shall now see how both of these formulas apply separately to the doublets
of an isoelectronic sequence. Transposing R in the first formula [Eq.
(17.1)] and taking the square root,
(17.4)
In Table 17.1 and in Fig. 17.2 the observed term values for the potassium¬
like doublets are given, as is customary in x-rays, in terms of y/T/R.
It is observed from the difference columns in Table 17.1 and from
the Moseley diagram that levels with the same principal quantum
number (n = 4) run nearly parallel. It is to be noted further that
the slopes of the curves, as shown by the dotted lines, are in agreement
with Eq. (17.4), i.e.t they are very nearly equal to 1/n. If the dotted
lines are drawn through the zero-zero origin of coordinates, as they
334 INTRODUCTION TO ATOMIC SPECTRA [Chap. XVII
Term KI Diff. VV
1 Term values T for a Moseley diagram are always computed with respect to the aeries limit as sero.
constants have been calculated and are given in Table 17.2. Since
the doublet fine-structure separations of 2P, 2D, and 2F are small as
compared with 2D — 2F or 2P — 2D separations, they are neglected
in the figures and in the tables.
Electron
Term KI Ca II Sc III Ti IV V V
configuration
(4) that the 2F curves start at approximately the same value for all.
sequences and toward heavier elements curve slightly upward, (5)
that the 2S, 2P, and 2D curves for heavier elements curve downward,
Fig. 17.3.—Moseley diagrams of the doublet levels in five periods of the periodic table.
(After Gibbs and White.)
and (6) that the 3 2D curves change their slope abruptly between the
Na and K sequences.
Although little is known concerning the energy levels of the rare-
earth elements, it is suspected that the 2F curve in the last sequence
should have a much steeper slope than that indicated in the figure.
This seems necessary from the standpoint of the building on of / electrons
in the 14 rare-earth elements starting with cerium. The figure as it is
336 INTRODUCTION TO ATOMIC SPECTRA [Chap. XVII
KI Can Sc 1 Tin VT H
and a Moseley diagram combined (see Fig. 17.4). The observed transi¬
tions may now be shown in the usual manner by arrows pointing down.
Let us now turn to the second formula [Eq. (17.2)] derived for pene¬
trating orbits and see how it applies to the energy levels of a typical
isoelectronic sequence. Transposing, Eq. (17.2) may be written
T_ R
(17.5)
Z\ (n - m)2
constant <r [Eq. (17.1)] or a quantum defect n [Eq. (17.2)] depends upon
the purpose for which these constants are to be used. In identifying'
higher members of a spectral series and in evaluating terms and series
limits, the second formula is to be preferred. From the theoretical
standpoint, on the other hand, the first formula is to be preferred in that
Electron
Term K I Ca II Sc III Ti IV V V.
configuration
quantum numbers should take integral values only. Moreover the x-ray
laws have been used so successfully in the analysis of the spectra of
highly ionized atoms and in the extension of isoelectronic sequences in
general that the screening formulas frequently can be used to greater
advantage.
17.2. Optical Doublets and the Irregular-doublet Law.—The irregu¬
lar-doublet law, extended from x-ray to isoelectronic sequences in optical
Ra\Z - $)*
Avom-i — (17.6)
tj.3
where Av represents the fine-structure doublet separations in wave
numbers, Z the atomic number, n the principal quantum number,
R the Rydberg constant = 109737 cm-1, a2 = 5.3 X 10~6, and
1 = 1
(17.7)
k\ &2 ki fci *i" 1 k\(k\ -{- 1)
Ra2(Z - s)4
Av cm”1. (17.8)
nH(l + 1)
This is the same formula as that derived in Chap. VIII from a different
treatment [Eq. (8.25)]. From the observed separations in the lithium¬
like sequence of nonpenetrating 2p orbits, screening constants s are
calculated and given in the last column of Table 17.4.
Assuming a complete screening by the two core electrons, i.e., s = 2,
Z - « - 1, 2, 3, 4, 5, and 6 for Li I, Be II, B III, C IV, N V, O VI, the
separations given in Col. 3 have been calculated. Still better calculated
values are obtained by first computing values of Z — v from the observed
Sec. 17.31 ISOELECTRONIC SEQUENCES 339
term values and Eq. (17.1), and substituting <r for 8 in Eq. (17.8). The
values computed in this way are given in Col. 4.
Table 17.4.—Doublet Separations for Lithium-like Atoms, 2*P$ — 2*P§
z -r, Z — st
Ay Ay Ay
z Atom from term from doublet 8
(obs.) (calc.) (calc.)
values separations
_ ... _
Atom Ay (obs.) Z Z0 So zx sx
Na I 17 18 11 1 10 7.66 3.34
Mb II 91 55 12 2 10 9.56 2 44
A1 III 234 0 13 3 10 10.91 2.09 Third period
Si IV 461.8 14 4 10 12.08 1.92 3*Pj - 3*P|
P V 794 8 15 5 10 13.15 1.85
s VI 1267 1 16 6 10 14 23 1.77
Cl VIII 1889 5 17 7 10 15 29 1 71
Rb I 237 6 37 1 36 31 23 5.77
Sr II 801 5 38 2 36 34.81 3.19 Fifth period
Yt III 1553 7 39 3 36 36 29 2.71 6*P| - 6*P,
Zr IV 2484 9 40 4 36 37.10 2.90
Ra'ZlZf
Av (17.9)
nll(l + 1)'
Here n0 is the effective quantum number. Values of Z* and s< have been
calculated for the first principal doublets of four isoelectronic sequences
in Table 17.5.
The large values of Z* indicate very deep penetration of the p orbits.
Regardless of what physical significance can be attached to these screen-
X6494 6462 6439 55989188
ing constants sif Eq. (17.9) serves as an excellent guide in predicting new
doublet separations for spectra not yet photographed or analyzed.
17.4. Isoelectronic Sequences of Atoms Containing Two Valence
Electrons.—In the previous sections of this chapter we have seen how
the x-ray doublet laws apply to the optical doublets of atoms containing
one valence electron. In this and the following sections we are con¬
cerned with the extension of these laws to the complex spectra of atoms
containing a number of valence electrons. As a first step in this direc¬
tion, let us consider the isoelectronic sequence of semicomplex atoms,
viz,, neutral calcium (Ca I), singly ionized scandium (Sc II), doubly
ionized titanium (Ti III), triply ionized vanadium (V IV), etc. In this
intermediate case each atom contains, in addition to the closed subshells
Sue. 17.4] ISOELECTRONIC SEQUENCES 341
of 18 elections (Is2 2s2 2pe 3s2 3p6), two valence electrons. With two,
electrons in incomplete subshells the energy level diagram of each atom
is composed of a system of singlet and triplet levels and the spectrum
of each atom is made up of groups of lines called singlets, triplets, and
triplet muUiplets.
Photographs of two sets of these triplet multiplets are shown in Fig.
17.7 for the first four elements of the calcium-like sequence. A schematic
representation of the triplet terms and intervals is given below each
photograph. Upper states are shown on the left and lower states on the
right.
in going from one element to the next are displaced to higher and higher
frequencies by very nearly equal frequency intervals. In the calcium-like
sequence, two such sets of frequencies have been observed, first the set
of lines arising from the electron transition 3d4d to 3d4p and second the
set of lines arising from the transition 3d4p to SeMs.1 The allowed terms
for these transitions are as follows:
None of the terms arising from the electron configuration 3d4d hatf
as yet been determined for Ca I; however, an extrapolation by means
of the second and third columns of Table 17.6 locates the *F\ — 3(j6
transition not far from 13974 wave numbers.
The almost linear displacement of frequency with atomic number,
for all lines due to the transition 3d4p to 3d4$, is shown graphically
in Fig. 17.9. The frequency scale for each element starts 11000 wave
numbers higher than the preceding element. This type of diagram
shows not only the successive displacement of each singlet and triplet
group of lines toward the violet by about 11000 cm~l but also magnifies
1 The many spectrum lines arising from the transition 3<i4<i to 3d4p are
shown schematically in Fig. 12 4 for singly ionized scandium (Sc II).
Sec. 17.5J ISOELECTRONIC SEQUENCES 343
Fig. 17.9.—Illustrating the irregular-doublet law for triplet multipleta arising from
the configuration transitions 3d4a — 3d4p. The almost linear displacement of frequency
with atomic number holds only for An = 0. (After White.)
1 Gibbs, R. C., and H. E. White, Phys. Rev., 29, 655, 1927; White, H. E., Phys.
Rev., S3, 672, 1929; 33, 914, 1929.
344 INTRODUCTION TO ATOMIC SPECTRA [Chap. XVII
(Is* 2s* 2p6 3s* 3p8), three valence electrons. As seen by the energy level
diagram of neutral scandium in the lower part of Fig. 14.8, each atom
in this sequence will give rise to complex systems of doublet and quartet
levels.
Along with this group of elements we may also consider the iso-
electronic sequence starting with neutral titanium, Z = 22. Each atom
in this second sequence Ti I, Y II, Cr III, Mn IV, Fe V, . . . contains,
in addition to the same closed subshells of 18 electrons, four valence
Fig. 17.10.—Modified Moseley diagram for doublets, triplets, quartets, and quintets
in the first long period. (After White.)
Since each configuration gives rise to many levels, the deepest lying
level is chosen as representative of that configuration on the diagrams.
In accord with Hund's rule, these are the levels of largest S and L
values shown schematically in Fig. 17.10. The corresponding terms
Sec. 17 6] ISOELECTRONIC SEQUENCES 345
of the one- and two-electron sequences of Figs.' 17.4 and 17.8 are also,
given for comparison. It should be mentioned that if all possible
levels for each configuration are known, the center of gravity of the
configuration can be plotted in place of the lowest term. (Each level
is assigned its quantum weight 2J + 1.) In cases where the termB are
all known, the resultant diagrams are so little different from the type
here shown that the essential features of the diagram as a whole remain
triad of doublets and the AD°t AF°, *G° terms, combining with AF, give
rise to a strong triad of quartet multiplets. These two triads have been
Fig. 17.14.—Showing the contrast between the irregular-doublet law (b) and (c), where
An *= 0, and the transitions (a) where An ** 1.
Fio. 17.15.—Modified Moseley diagram and the irregular-doublet law extended to the
energy levels and radiated frequencies of the same element, vanadium, in different stages
of ionization. (After White.)
constructing Fig. 17.15 there are many terms And spectrum lines fron*
which to choose. From any one given electron configuration, for
example, there may be as many as 100 theoretically possible levels.
Since ail possible levels in complex spectra are seldom observed, the
deepest lying term is conveniently chosen as representative of that
configuration. This is quite general practice in predicting new terms
and in making an analysis of a new spectrum.
A V I Multiplet
all points along the spark. Lines from higher stages of ionization, V III
and V IV, appear strong at the electrodes and very weak, if at all, in the
middle. These are but a few of the criteria that are often employed
in the unraveling of an unanalyzed spectrum.
17.8. Centroid Diagrams.—The centroid diagram, introduced into
atomic spectra by Mack, Laporte, and Lang,1 is often drawn for the
purpose of showing just how the energy levels of a given electron con¬
figuration change in going from one element to the next. As illustrations
Fio. 17.17.—Centroid diagrams for two different electron configurations in the isoelectronic
sequence Pd I, Ag II, Cd III, In IV, and Sn V. {After J. E. Mack.)
of this, two centroid diagrams for the levels arising from the electron
configurations 4dB5s and 4d95p in the sequence Pd I, Ag II, Cd III,
In IV, and Sn V are given, following Mack,2 in Fig. 17.17. Term
separations are plotted vertically against atomic number Z horizontally.
The straight horizontal line serves as an origin and represents the center
of gravity of each system of levels, each level being assigned the usual
weight 2J + 1. Since the intervals increase rapidly from element
to element the 4dB5«, iDi — *D8 interval is made the same for all elements,
and all levels of the same element are plotted to that scale. Since the
observed intervals show that the coupling is somewhere between LS-
and jj-coupling (see Fig. 14.20), the terms are given L£-coupling designa¬
tions on the left and ^‘-coupling designations on the right.
1 Mack, J. E., O. Laporte, and R. J. Lang, Phys. Rev., 31, 748, 1928.
,
* Mack J. E., Phye. Rev.t 34 17, 1929.
Sec. 17.8] ISOELECTROS 1C SEQUENCES 351
HYPERFINE STRUCTURE
ft ‘"Till iEMI
elements known to have but one isotope this hypothesis \\as abandoned,
as is always done in science, for a newer and better theory.
Apparently the first suggestion that hfs is to be attributed to a small
magnetic moment associated with the nucleus was made by Pauli,1 and
later, but independently, by Russell.1 Experimentalists were not long
in showing that this hypothesis, in most cases, is apparently correct.
Since the time that this theory was accepted generally, an isotope effect
has been observed in some spectra so that now there are two types of hfs
to be distinguished from each other. First, there is a hfs due to a nuclear
magnetic and mechanical moment and, second, a hfs due to the different
isotopes of the same chemical element. The first of these will be referred
to as hfs and the second as isotope structure. In some elements hfs
alone is observed, in others isotope structure alone is observed, and in
still others both are observed.
1 Pauli, W., Naturwissenschaflen, 12, 741, 1924. The suggestion that hfs is due
probably to a spinning proton was suggested by H. N. Russell to W. F. Meggers and
K. Burns, Jour. Opt. Soc. Amer.f 14, 449, 1927.
354 INTRODUCTION TO ATOMIC SPECTRA [Chap. XVIII
Flo. 18.3.—Orbital and vector models illustrating the coupling of a nuclear moment I *
with an electron moment/* to form a resultant F.*
for the first time that a new quantum vector should be added to the
atom model, and that the Land6 interval rule for fine structure also
applies to hfs. The model proposed is essentially the one shown in
Fig. 18.3, where the quantum vector/* of one electron, or J* representing
the total mechanical moment of all the extranuclear electrons, is coupled
with the quantum vector /*, representing the total mechanical moment
of the nucleus (I*h/2w) to form a resultant F*. It is this resultant F*
that now represents F*h/2ic, the total mechanical moment of the atom,
in place of J* as previously stated.
Goudsmit and Back have shown that just as the interaction energy
between L* and S* is proportional to the cosine of the angle between
them [see (Eq. 8.18)], so the interaction energy between the nuclear
moment J* and the electron moment J* is given by
where A'I*J* is constant for each given fine-structure level «/, and
A' is a measure of the strength of coupling between /* and J*. Just
as in fine structure the starred quantities are given by
Fig. 18.4.—Vector diagram illustrating graphically the interval rule for hyperfine struc¬
ture. Specific example of a level, where J is taken as $.
AF = 0, ± 1. (18.5)
1 For a derivation of hfs intensity rules see E. L. Hill, Proc. Nat. Acad. Sci., 15 ,
779, 1929. Intensity tables are given in the Appendix.
Sec.'18.2] HYPERFINE STRUCTURE 357
figure by means of the microphotometer curve, for comparison with the
calculated intensities given just above.
The second hfs pattern in Fig. 18.5 is drawn for one of the two reson¬
ance lines of manganese, an intercombination line, X5394. With an
I value of 4, each level J = f and 4 is split into six components as shown.
Applying the relative intensity formulas to the allowed transitions, the
theoretical pattern shown near the bottom is obtained. For want of
greater resolving power this very narrow pattern is only partially resolved
as shown by the microphotometer curve at the bottom of the figure.
In making an analysis of the hfs of an uncharted spectrum, patterns
such as we have shown here are all one has to work with. Where all
6p27S,(H!)l ” .830
l A
4
iii i 1 1
3 J-
eps. Gil) 3 I.-1 - =;||
V
III 11
Similarly the magnetic field at the nucleus due to the orbital motion
of the electron is
H = E X
c
Substituting Bohr's relation for the angular momentum
2xmr X v — l*h,
the magnetic field becomes
H - <18«
where m and e are the mass and charge on the electron, and r is the
electron-nuclear distance. Since r is not constant in any orbit, either
on the classical theory or on the quantum mechanics, (1/r3) must be
averaged. Now the nucleus with a mechanical moment I*h/2ir and a
magnetic moment /*/ tends to carry out a Larmor precession around this
field with an angular velocity wL given (see Secs. 3.7 and 8.6) by the
1 Fermi, E., Zeits. f. Phys., 60, 320, 1930.
* H. B. G. Casimir, quoted in L. Pauling and S. A. Goudsmit, “The Structure of
Line Spectra/1 p. 208.
* Hargreaves, J., Proc. Roy. Soc.f A, 127, 141, 1930.
4 Breit, G., Phys. Rev., 37, 51, 1931.
‘Goudsmit, S. A., Phys. Rev., 37, 663, 1931; see also Pauling, L., and S. A
Goudsmit, “The Structure of Line Spectra," p. 204.
Sec. 18.31 HYPERFINE STRUCTURE 35§
product of the field strength H and the ratio between the magnetic and
the mechanical moment. For this latter ratio we write
Mi _ „ e
(18.7)
I*h ^I2mc
2r
where c/2me is the classical ratio and gt is an unknown factor called the
nuclear g factor. If the nuclear moment were due to a proton in an orbit,
gi would be expected to be that of an orbital electron. If it is due
to a spinning proton, gi should be between and 7 times this.1
The precessional angular velocity becomes therefore
e2 l*h(l\ „oox
Ul 9'2mV 2jtVV 18'8)
The interaction energy is now given by the product of ^ and the
projection of the nuclear mechanical moment I*h/2ir on l*f
e2 l*hf 1
(18.9)
th( 1
(18.11)
-W cos ( wcos
Since both of these angles are fixed, the cosines are readily evaluated in
terms of the quantum vectors s*, l*} j*, I*, and F*. This will be done
after we obtain the interaction energy for the electron spin s* and the
nuclear spin 7*.
According to classical electromagnetic theory the mutual energy
of two magnetic dipoles with moments hi and and at a distance r apart,
is equal to
The nuclear magnetic moment fit is given by Eq. (18.7), and the
magnetic moment of the spinning electron by
-2;(_3* (18.13)
2me 2t
The mean value of the term in braces, involving the electron-nuclear
distance r, is readily shown by the use of direction cosines to be1
—J cos cos (j*s*) — 3 cos (j*l*) cos (18.14)
Inserting these results in Eq. (18.12),
a' = fir/(
cos (l*r) + w* 008 (s*r)
Qi . Ra2Z3
cm- (18.21)
1838 n\l + i)j(j + 1)
This af is the A' of Eqs. (18.1) and (18.3) for hydrogen-like atoms. If
gr is positive, a' is positive and the hfs terms will be normal. If gj is
negative, a' is negative and the hfs terms will be inverted. If the nuclear
magnetic moment is due to a spinning proton, the gt factor would be
expected to be between 4| and 7.1
As already pointed out, the above equations have been derived
for hydrogen-like atoms. Since hfs has not been observed in these
simplest of atomic systems we must look to the more complex spectra
and the heavier elements for a comparison of theory with experiment.
18.4. Nuclear Interaction with a Penetrating Electron.—In a classical
treatment of fine structure Landd showed, by the use of an idealized
model, that Z4 in the resultant equation should be replaced by Z\ times
Z2. The alkali doublet separations, for example, are given by Eq. (8.32),
where Zx is the effective nuclear charge inside the core of closed electron
shells, and ZQ the effective nuclear charge outside. By analogy with
this equation and its derivation, Z3 in Eq. (18.21) is broken up into
two parts so that the interaction energy of a single valence electron,
screened from the nucleus by a shell of electrons, is given by [see Eq. (18.3)]
Tf = W(F*2 - I*2 - J*2), (18.23)
where gj Ra2ZtZ2
(18.24)
1838 ' nl(l + $)j(j + 1)'
1 In any experiment performed for the purpose of determining the magnetic and
mechanical moments of the nucleus, m and I*h/2ir are so oriented with the field, as
in the Zeeman effect, that what one measures is the component of m along the field H.
From the experimental value of this component the moments and I*h/2ir are calcu¬
lated from theory. In Frisch, Stem, and Estermann’s experiments on the hydrogen
molecule a magnetic moment of approximately 21 ± 10 per cent nuclear magnetons
has been measured for the proton. By a different method Rabi, Kellogg, and
Zacharias find a value of 31 ± 10 per cent nuclear magnetons. The agreement
between these two values must be regarded as more important than their difference.
Dividing by / = J gives a value of qi from 41 to 7 for the proton. This comes from
Eq. (18.7), for, if the ratio jJfcjfiic g*ves ^ie 9* factor, then the ratio of their com¬
ponents along any line will also give the same ratio.
See Frisch, R., and O. Stern, Zeits. f. Phys., 86, 4, 1933; Estermann, I., and
O. Stern, Zeits. f. Phys., 86, 16, 1933. Rabi, I. I., J. M. B. Kellogg, and J. R.
ZXcharias, Phys. Rev., 46, 157, 1934.
362 INTRODUCTION TO ATOMIC SPECTRA [Chap. XVIII
Breit1 and Racah2 have shown that relativity corrections are differ¬
ent for the two levels of a doublet and must be taken into account. The
relativity corrections to be added to Eq. (18.25) are obtained by multi¬
plying a' by k/X, where
, _ Qi AvlQ, + 1)k
(18.28)
0 1838 ’ Z,(l + \)j(j + 1)X
1G. Breit, see Goudsmit, 8. A., Phys. Rev., 43, 636, 1933.
* Racah, G., Zeits. f. Phys., 71, 431, 1931.
Sec. 18.5J HYPERFINE STRUCTURE 363
As stated above, the equations thus far developed should not apply
to s orbits. McLennan, McLay, and Crawford1 have shown, however,
that the equations do seem to be approximately correct even for this
special case. Goudsmit has suggested that we write for s electrons,
Term Av a' K X zx z* gi
energy, will be the one for which the nuclear magnetic moment is most
nearly parallel to the field of the electron. If this magnetic moment
is due to something like an orbital or spinning proton, this will be the
state for which the mechanical moments I* and J* are oppositely
directed. These are the hfs states fP3 and §P* in Fig. 18.6.
If we now think of the spinning electron and the nucleus as small
magnets acting on each other at some distance from each other, the
most stable position of the two will be that for which their magnetic
moments are in opposite directions and the mechanical moments are
in the same direction. These correspond to the |P3 and \P$ states
in Fig. 18.6. Since the orbital interaction with the nucleus is greater
than the spin interaction, the hfs of both levels is normal. Although
the greater splitting is observed in the 7p,2P3 level in T1 III (see Table
18.1) the o' coefficient, which measures the total interaction with the
nucleus, is greatest for the 7p,2P* level.
Fermi has shown from a quantum-mechanical treatment of $ electrons
that the interaction with the nucleus is opposite in sign to that of the
spin of an arbitrary electron, i.e., it has the same sign as an orbital
interaction. Classically, therefore, an s electron may be thought of
as a cloud of negative charge, distributed as shown in Figs. 4.6 and 4.7
and rotating about the nucleus. A positive charge rotating at the
center of this spinning negative cloud is most stable when the magnetic
moments are parallel and the mechanical moments are opposite. This
shows that s electrons also give rise to normal hfsf as observed. Because
of the deep penetration of an s electron on the classical theory, or the
large probability density near the nucleus on the quantum-mechanical
model, the interaction between s* and I* and hence the hfs for 2&j states
should be very large.
18.6. Hyperfine Structure in Atoms with Two or More Valence
Electrons.—In 1927 Meggers and Burns1 published an account of the hfs
in the lanthanum spectrum, at which time they pointed out that the
largest structures were always observed for configurations involving a
single unbalanced s electron. They also pointed out at the same time
that, while certain levels revealed wide structures, others of the same
multiplet were either single or else very narrow and unresolved. In
the 685d,3Di,2,3 terms of La II, for example, the two outer levels 3Di
and 8Dj show a wide splitting, whereas the middle level 3Z)2 shows
practically none. An explanation of this anomaly in terms of the vector
model has been given by White,2 and equations for the splitting have
been given by Goudsmit and Bacher.3
1 Meggers, W. F., and K. Burns, Jour. Opt. Soc. Amer., 14, 449, 1927.
* White, H. E., Phys. Rev., 34, 1288, 1929; 34, 1397, 1404, 1929; 36, 441, 1930;
Proc. Nat. Acad. Set., 16, 68, 1930.
•Goudsmit, S. A., and R. F. Bacher, Phys. Rev., 34, 1499, 1601, 1929.
Sec. 18.6] HYPERFINE STRUCTURE 365
In Fig. 18.7 we picture, on both the classical model and on the
quantum-mechanical model, an atom with two valence electrons, one in
a 'penetrating s orbit and the other in a nonpenetrating d orbit. A vector
diagram for the same atom in LS-coupling is drawn schematically at
the left. From this picture it is easy to see how the penetrating electron
has an opportunity of coupling strongly with the outer electron when
it is outside the core and also with the nucleus when it is inside the core.
Thus the s electron may greatly strengthen the coupling between the
electron resultant J* and the inner nuclear resultant I*. Due to the
screening of the nucleus by the core electrons, the interaction between
the nucleus and the d electron is small compared with that between
sj and I*. For strong coupling with the nucleus the electron must
not only be deeply penetrating but also tightly hound. All s orbits, for
Fio. 18.7.—Vector and orbital model for the interaction of a penetrating s electron and
a nonpenetrating d electron with a nuclear moment.
example, will be deeply penetrating, but only those that spend a con¬
siderable amount of their time near the nucleus will couple strongly with
7* and give wide hfs.
Due to the large spin-spin interaction of the electrons in LS-coupling,
and si precess rapidly around their mechanical resultant S*. Due
to the somewhat weaker coupling between the spin sj and the orbit
ll of the d electron, S* and ll precess more slowly around their mechanical
resultant J*. Finally J* and I* are coupled weakly together and
precess slowly around their mechanical resultant F*. The relative
frequencies of these precessions are given by, and follow directly from,
the observed energy differences, i.e., by the singlet-triplet intervals, the
triplet fine-structure intervals, and the hfs intervals, respectively.
The well-known general relation that the energy of interaction
between any two magnetic moments is proportional to the cosine of the
angle between them (see Sec. 8.6) enables one to write
Since each of the angles is constant, all three cosine terms are readily
evaluated in terms of the mechanical moment vectors to give for LS-
coupling
Tf = f A'(F*2 - 7*2 - J*2), (18.34)
,J*2 + S*2 - Ip S*2 + sl2-s*2
where (18.35)
A a 2 J*2 2£*2
One interesting result regarding this equation is that A', as shown by
Goudsmit and Bacher, may be simplified by expressing it in terms of the
electron g factor. Substituting f for Si and for s2 and the value of g as
given by Eq. (13.8), the A' factor for L/S-coupling reduces to
A' = W(9 - 1) (18.36)
If the coupling between the s electron and the outer electron is
jjf the average cosine is obtained in a similar fashion by projecting
I* on J*, then J* on sj, to give
This again leads to Eqs. (18.32) and (18.34), where, for jj-coupling,
,J*2 + sf2 - jV
A' => a'yj cos (J*sf) = a (18.38)
2J*2
Fig. 18.8."- -Calculated and observed hfs intervals for an sd configuration *D term in
thallium.
those of the configuration 6s7d in singly ionized thallium (T1 II). The
observation of normal splitting in 3D3 and of inverted splitting in 3Z) i
substantiates here, as it does in many other cases, the supposition that
the penetrating s electron is responsible for the strong coupling of J*
with the nucleus.
Table 18.2.—Coupling Coefficient o' for a 6s Electron in Tl II; Calculated
from the Observed Separations
Configura¬ Configura¬
Terra Ai/(0b. ) A' a' Terra Av(.obm.) A' a'
tion tion
+3 45 +0 99 +5 94
6;8 d 3Z>2 + 1 66 +0 66 +7 92
3/>i -2 25 -1 50 +6 00
EB
/
Configuration Term
mem
Av
A comparison of the three values 5.90 for 6s, 0.72 for 7s, and 0.12
for 9s indicates the rapid decrease in the coupling with the nucleus
with increasing total quantum number. The conditions for wide hfs
therefore specify that at least one electron be not only deeply penetrating,
as all s orbits are, but that it also be tightly bound to the atom.
18.7. Hyperfine Structure in Complex Spectra.—Attempts to corre¬
late observed hfs in complex spectra with theory have in some respects
been more successful than in the simpler cases of one-electron systems.
As shown in the last section this is due largely to the general observation
that nearly all wide structures and patterns arise from configurations in
which there is a single tightly bound s electron. The classical and
quantum-mechanical model for a complex atom may be represented by
Fig. 18.7. This figure is generalized by letting the sy ly and j of the
nonpenetrating electron represent the resultant Sy Ly and J of several
nonpenetrating electrons. In such a system the penetrating $ electron
couples strongly with the outer electrons when it is outside the core
and with the nucleus when it is inside the core. The treatment for several
electrons will therefore be essentially the same as that given in the preced¬
ing section for two valence electrons. Replacing ZJ, sj, and j* in
Eqs. (18.35) and (18 38) is L£, S*t and J*y the interaction energy for
complex spectra becomes
TV = JA'(F*2 - I*2 - J*2)y (18.40)
where, for L$-coupling,
,J*2 + S*2 - L*2 S*2 + s*2 - S*2
(18.41)
A a 2 J*- 2/S*2
and, for .//-coupling,
j*2 + s*2 - j?2.
(18.42)
A 2J*2
Here SPy LPy and JP represent the resultants of the nonpenetrating
electrons, Si the spin of the penetrating s electron, and S and J the
total resultants of all the electrons together.
Let us now apply these equations to a specific example where hfs
separations have been observed. In the neutral spectrum of lanthanum
the first excited state is a AF term arising from 5eZ26s. Three of the
four levels in this term have quite wide hfs whereas the fourth is rela¬
tively narrow. The observed intervals are shown at the right in Fig. 18.9.
To calculate relative splittings for each level, we first evaluate A' coeffi¬
cients for each. The observed fine-structure intervals of 341.8, 484.6,
and 627.0 cm-1 have the ratios f: \ revealing good L£-coupling between
electrons. From the configuration and term 5d26s,4F we have the quan¬
tum values S = f, si » \y SP = 1, LP = 3, and J * f, $, J, and f.
Substituted in Eq. (18.41), we obtain A' * \a’y t^o', fya', and
370 INTRODUCTION TO ATOMIC SPECTRA [Chap. XVIII
for 4Fg, 4Fj, 4Fj, and 4Fj, respectively. Substituting these in Eq. (18.40)
the over-all separations —a', ^a', 50aa', and ^a', shown at the left
in Fig. (18.9), are obtained. The calculated intervals in the first column
are obtained by computing a constant a' from 4F% and using this for all
four levels. Better agreement between calculated and observed intervals
is obtained by taking the orbital motions of the 5d electrons into account.1
Assuming that the interaction will be given to a first approximation by
a"L*I* cos (L*/*), the energy will be given by Eq. (18.40), where
Af = a"(J*2 + L*2 — S*2)/2J*2. Substituting the AF quantum num¬
bers, the energy contributions ^a", -Mpa", -^a", and ^a" are obtained.
The intervals in the middle column are obtained by calculating constants
o' and a" from 4F# and 4F? and using these for the other two levels.
Fig. 18.9.—Calculated and observed hfs intervals for d2s, 4F, in lanthanum.
Element Element
/
and atomic Isotope I Qi M/ and atomic Isotope / ffi
number number
Just how far the above equations can be expected to give correct gt
factors is a question that will be answered when we know more con¬
cerning the structure of the nucleus. From the recent experiments of
Sec. 18.9J HYPERFINE STRUCTURE 373
Frisch, Stern, and Estermann as well as Rabi, Kellogg, and Zacharias1
there is some evidence that the magnetic moment of the neutron is
negative. This should play an important rdle in any new theory of
nuclear structure.
18.9. Zeeman Effect in Hyperfine Structure.—We now turn to the
Zeeman effect of hfs, a subject which had its beginning with a study
of the bismuth spectrum by Back and Goudsmit. The paper published
by them in 1928 is considered a classic in the annals of spectroscopy.2
At this early stage in the development of hfs, in general, it was shown
by Back and Goudsmit that all of the simple classical derivations and
formulas for the anomalous Zeeman effect and the Paschen-Back effect
could be carried over to hfs by making the following substitutions:
I, J, F, gh gJy gF] Mh Mj, and MF
for S, L, J, gs, gl, gj, Ms, ML) and Mj, respectively.
Unfortunately the so-called Zeeman effect of hfs has not yet been
observed, chiefly because of the lack of sufficient resolving power in
the optical instruments thus far employed. What has been observed
by Back and Goudsmit, however, is an effect similar to the so-called
Paschen-Back effect in fine structure. Since there is good reason to
believe that the hfs Zeeman effect will be observed and studied in the
near future, a simple treatment of the subject
will be given here. H
In a very weak magnetic field the atom
may be thought of as precessing as a whole
around the field direction H subject to the
quantum conditions that the projection of
the mechanical resultant F*h/2w on H is equal
to MFh/2ir. The magnetic quantum number
Mf takes values differing from each other by
unity ranging from MF = —F to MF = +F.
A classical vector model for very weak field
is shown in Fig. 18.10. To calculate the mag¬
netic energy it is necessary that we first com¬ Fio. 18.10.—Vector model
pute the magnetic moment of the atom as of an atom with a nuclear
spin in a very weak magnetic
a whole. This moment will be made up of field (hfs Zeeman effect).
two parts or components: iij, the magnetic
moment of the extranuclear electrons, and /zz, the magnetic moment
of the nucleus. These moments [see Eqs. (13.1) and (13.6)] should be
given by
1Frisch, R., and O. Stern, Zeits. f. Phys., ,
85 4, 1933; Estermann, I., and O.
. ,
Stern, Zeits. f Phys., 85 17, 1933; Rabi, I. I., J. M. B. Kellogg, and J. R.
Zacharias, , ,
Phys. Rev., 45 769, 1934; 46 157, 1934.
* Back, E., and S. A. Goudsmit, ,
Zeits. f. Phys., 47 174, 1928.
374 INTRODUCTION TO ATOMIC SPECTRA [Chap. XVI11
^ = 9fF*A^Tc (18-50)
Referring to the vector model again, we may now evaluate the cosines
in terms of the mechanical vectors and obtain, for Eq. (18.49),
F*2 _|_ J*2 _ 1*2 f*2 _|_ J*2 __ J*2
Because of the general observation that hfs separations are many times
smaller than the fine structure the gi factor for the nucleus must also
be very small compared with the gj factor for the extranuclear electrons.
If we make the assumption that gi << gJ) we may write
F*2 -b J*2 - 7*2
gF _ gj-2F*2-■ (18.52)
If the nuclear spin is due to one or more particles with the mass
of protons, the nuclear gi factors should be in the neighborhood of one
or two thousandths of those for electrons. By Eq. (18.50) the ratio
between the magnetic and mechanical moments of the entire atom is
therefore closely represented by
The energy of this precession is given by oof times the component of the
mechanical resultant F*h/2w along H:
<18M)
Dividing by Ac, the energy in wave numbers becomes
Tip
— ATf = gF • Mft-cm-1 = gF - Mf * L cm-1. (18.57)
47rmc
AW ir = + gsMJH-- (18.61)
Let us now apply this equation to the specific example of one of the
resonance lines of thallium, X * 5349, 6p,2Pi — 7s,2Sj (see Fig. 8.1).
378 INTRODUCTION TO ATOMIC SPECTRA [Chap. XVIII
Table 18.4.—Magnetic-energy Factors for the Terms 2Pj and 2S±, in an Atom
Where the Nuclear Shin I ~ \
2 2 9, 6 IA"
\ a
j 1 1 \ •2 a \A"
+A]" 0 1 0 _1 2 -\A"
ip* \ 2 j
-1 -1 1 _a \ 0
2 a 3 -IA"
-2 -2 3 fi ~4 A"
-i J
— 1 2. -J A "
2 2 J
1 7 _1 _2 \A"
a 2 J
0
s
0 —2 —2 _0 IA"
\Pi -M" a A
_ 6
-1
1 1 1 1 1 \A'
i 2
+ \A' 0 1 o 1 1 -1
ISj 2 2 -M#
-1 -1 1 2 1 -\A'
2 2
1 _ J2 _1 1 A'
1 1 4/l
JSi -\A' 0
0
0
0 0
1 0 -i
Fig. 18.14.—Schematic diagram of the change in the T1 hfs pattern X5349 from no
field to very weak field to weak field Back-Goudsmit effect). (Observed pattern after
Wulff). In reproducing the microphotometer curve the field-free lines have been left
out as shown by the break at the center of the figure.
the weak field where the coupling between I and J is broken down, the
lines have crossed each other to form a pattern resembling closely the
Zeeman pattern of the same transition in an atom where the nuclear
380 INTRODUCTION TO ATOMIC SPECTRA [Chap. XVIII
moment is zero (see Fig. 10.8). A comparison of Fig. 18.14 with Fig.
13.11 for the Paschen-Back effect of the fine structure in ^'-coupling
is of interest. With I = % it is to be noted that each line of the Zeeman
pattern should theoretically be double, and that the separations should
be of the same order of magnitude as the field-free hfs. The micro¬
photometer curve at the bottom of the figure has been obtained in a
field of 43350 gauss by Wulff.1
A similar calculation for any spectral line where 7 has an integral
or half-integral value will lead to a pattern resembling closely the Zeeman
pattern for the same line if 7 were zero. Due
to the interaction energy term A'MMj in Eq.
(18.68) each of the Zeeman terms will be split
up into 27 + 1 equally spaced levels, and
each line in the observed Zeeman pattern will
be split into 27 + 1 equally spaced lines. This
may be seen to better advantage from the
following description of an atom where 7 is
large. In weak field, 7* and J* precess inde¬
pendently around H. For each orientation of
J*, as shown in Fig. 18.15, there will be
27 + 1 orientations of the nuclear moment 7*.
In Fig. 18.15 the 7 value, as in bismuth, is
taken as -f. Since the coupling of 7* with
H, as measured by the energy factor giMIy
Fio. 18.15.—Vector dia¬
gram showing the 21 + 1
is negligibly small, 7* is coupled to, and pre-
orientations of the nuclear cesses around, the magnetic component of J*
moment/* for a fixed orienta- along JJ Like the ordinary
tion Of the electron moment —& “• Like the ordinary Zeeman effect
J*. Drawn for the special this energy of interaction is proportional to Mj.
GoudsmiTeffect).= f (Back‘ Since the field for the precession of I* is fur-
nished by J*, and its magnitude by Mj, the
energy of the splitting will be proportional to MiMj, or equal to A'MlMJ,
as already shown.
As an example of this effect a diagram of a weak-field pattern observed
by Back and Goudsmit in bismuth is reproduced in Fig. 18.16. Each
of the four middle groups has been resolved by Back and Goudsmit and
found to have exactly 10 components. The two center groups of lines
marked a have been reproduced and published by Zeeman, Back, and
Goudsmit.2 Figure 18.17 has been enlarged and reproduced from the
halftone reproduction in their published paper. To obtain the lower half
of the figure, the photographic paper was moved parallel to the lines
1 Wulfp, J., Zeits. /. Phys., 69, 74, 1931; see also Green, J. B., and J. Wulff,
Phys. Rev., 88, 2176, 1931; 38, 2186, 1931.
* Zeeman, P., E. Back, and S. A. Goudsmit, Zeits. f. Phys., 66, 11, 1930.
Sec. 18.10] HYPERFINE STRUCTURE 381
during exposure thus enabling the 10 components to be easily distin- 4
guished. The effect is much the same as a microphotometer trace.
That there are 10 components in each group furnishes conclusive evidence
that, for the Bi nucleus, J * f
Fia. 18.16.—Calculated pattern for the bismuth line X4722 in a weak field (Back-Goudsmxi
effect).
Fig. 18.17.—Observed patterns, by Back and Goudsmit, of the two center components
marked (a) in Fig. 18.16. (A) Enlarged from the halftone reproduction in the published
paper of Zeeman, Back, and Goudsmit. (B) Photographic paper moved parallel to the
lines during enlargement.
Intensity rules for hfs in weak fields are the same as those for the
Zeeman effect in weak fields and are given by Eq. (13.21). From the
assumption that Mi does not change in any transition the intensities
will be independent of the value of Mt. This means that for each
component of an ordinary Zeeman pattern the intensity will be divided
equally between the 2/ + 1 components due to the nuclear moment.
382 INTRODUCTION TO ATOMIC SPECTRA [CkAP. XVIII
This means that Eq. (13.21) should be valid and that Af, as given there,
is the Mj as given here. This is in fair agreement with the observations in
Tl. In Bi the appearance of so-called forbidden transitions (AAf j — ±1)
shows that the field is not strong enough for these simple rules to hold.
Calculated intervals and intensities have been shown by Goudsmit and
Bacher,1 and by Back and Wulff2 to be in excellent agreement with
observed patterns in Tl and Bi in different intermediate field strengths.
18.11. Isotope Structure.—It is now known with certainty that the
hfs observed in the spectrum lines of some elements is due to small
differences in the energy levels of two or more isotopes and not to a
Fig. 18.18.—Isotope structure observed in the spectrum lines of tungsten. {After Grace,
More, and White.)
mercury isotopes,
Hfs terms for two energy levels in mercury are shown in Fig. 18.19.
The heavy lines near the center of both the upper and lower levels
represent the even isotopes 196, 198, 200, 202, and 204. The very light
lines represent the levels of the odd isotope 201, and the remaining levels
belong to 199. The observed pattern is shown at the bottom, and the
corresponding transitions are given directly above. The relative
abundance of the isotopes as given by Aston’s mass-spectrograph meas¬
urements is in agreement with the observed hfs in that the center line
contains about 70 per cent of the total intensity. If there were an
1 Since we know relatively little about nuclear structure at the present time, it is
difficult to say whether one, or both, of the nuclear moments, magnetic or mechanical,
is zero. If a nucleus has a large magnetic moment, but no mechanical moment,
there can be no space quantization with J * and hence no hfs. If a mechanical moment
exists, but no magnetic moment, there will be quantization but no splitting. Such
cases of one or the other moment equal to zero are known to exist in extranuclear
electron structure.
Sue. 18.12] HYPERFINE STRUCTURE 385
appreciably different isotope shift in the initial dr final states, as there ,
is in the resonance line — zPi at X2536, this line would not be single
as observed but would reveal five components. The relative positions
of the levels belonging to different isotopes are arbitrary, for, if one system
of levels is moved up or down, the radiated lines remain unchanged.
Returning to Eqs. (18.34) and (18.35), the observed hfs in mercury
may be used to calculate the coupling coefficient for the deeply penetrat¬
ing and tightly bound 6s electron. For the ZP2 state the calculated
values are as follows:
Hg 199 Hg 201
A' = +0.303 Af = —0.114
o' « +1.212 a! = -0.456
Since a! is negative for Hg 201, the gt factor [by Eq. (18.24)] is negative.
For Hg 199, the factor is positive and by the ratio of the o' coefficients
is about 2.66 times as large. If the same calculations are made for the
upper term zDi, a smaller ratio is obtained. This may be attributed
to perturbations of certain of the hfs terms by those of a 1D2 term, only
three wave numbers away. When the perturbation corrections shown
in Fig. 19.15 are made, the ratio of the a’s, and hence the ratio of the
gi factors, comes out to be 2.79, in very good agreement with similar
calculations for six or more other terms.
In addition to mercury, line patterns of a number of other elements
have been sufficiently analyzed to show the presence of both isotope
structure and hfs. In Cd, for example, there are six isotopes of mass
110, 111, 112, 113, 114, and 116. Like mercury the even isotopes are
observed to have a nuclear moment of zero, and both of the odd isotopes
to have the same mechanical moment Z = i and the same magnetic
moment gt = —1.0.
Problems
1. Show that the interaction energy of two s electrons with a nuclear moment I
is given by Eq. (18.38).
.
2 Construct a diagram similar to Fig. 18.9 for the term d8s,6F. Assume a '—
2 cm-1 and 7 = f-
3. Construct a diagram similar to Fig. 18.13 for the Bi levels given in Fig. 18.5.
Assume gj factors of 2 and f for the upper and lower levels, respectively. Calcula¬
tions are most easily made by tabulating as in Table 18.4. Note that, for the lower
level, A1 is negative.
.
4 From the levels derived in Prob. 3, construct a diagram similrr to Fig. 18.14.
CHAPTER XIX
Term values
Fig. 19.3.—Anomalous copper series. Fig. 19.4.—Anomalous aluminum series.
of four series of terms are shown for cadmium, calcium, copper, and
aluminum, respectively. In each figure the fine structure is plotted
vertically and the term values horizontally. The dotted center lines
represent the center of gravity of each series member, each level being
assigned the quantum weight 2J + 1. The cadmium series in the first
figure represents a series of triplets approaching in a regular fashion
the series limit.
The 3F2,3,4 series in singly ionized aluminum, like the 3Di,2,s series
in calcium, expands abnormally and contracts again as it approaches
the single limit. The copper series discovered by Shenstone1 is of
particular interest: the first term is right side up (2P3 above 2Pj), the second
term is very narrow (both terms falling practically together), the third
term is inverted, the fourth term is right side up and very wide, and the
fifth term is inverted. The beautiful way in which each of these anoma¬
lous series has been explained by Shenstone and Russell2 is the subject
of the following sections of this chapter.
19.2. Energy Level Perturbations.—In the early chapters of this book
we have seen that the term values of hydrogen-like atoms are given by
the simple Bohr formula
Fio. 19.6.—Aluminum series showing per- Fig. 19.7.—Calcium series showing per¬
turbations. turbations.
resemblance between these curves and anomalous dispersion curves is
quite striking.
A generalized form of the Rydberg-Ritz formula, derived from
perturbation theory and the quantum mechanics, has been given by
Langer.2
RZl
Tn (19.4)
m _ _Ml_ (19.5)
(^.)i
where the effective quantum number is
n* = n — n acTn + (19-6)
With the exception of the first series member, the agreement between
the observed and calculated term values is very good. The origin of
the foreign term is discussed in Sec. 19.5.
19.3. The Nature of and Conditions for Term Perturbations.—A
first-order perturbation treatment and calculation of atomic energy
levels have been given by Slater1 and by Condon.2 Including electron
spin in the calculations, Condon shows that under certain conditions
terms belonging to different electron configurations will perturb each
other. If two levels have the same characteristic quantum numbers,
and lie close together, second-order terms of the perturbation correction
become large and cannot be neglected in the calculation of the energy cf
either level. A large second-order correction implies that the eigen¬
functions for each of the respective levels contain components of the
eigenfunction of the other level. Thus both levels belong in part to
both electron configurations. The assignment of a given level to a
definite electron configuration is therefore only approximate and becomes
more exact when the levels are far apart.
The nature of the perturbation between the two levels in question
reveals itself essentially as a repulsion. This apparent repulsion may
be thought of as being due to a sort of resonance phenomenon, there
being a certain probability that the atom will jump back and forth
between the two characteristic states without the radiation of energy.
The frequency of this interchange, and hence the repulsion, increases
as the interval between the levels decreases.
The quantum conditions for term perturbations may be stated as
follows: Two levels belonging to the same or to different electron configurations
will perturb each other when (1) both electron configurations are of the same
parity, i.e., both even or both odd,3 and (2) both levels have the same J value.
In addition to these necessary conditions, observation shows that the
greatest effect is to be expected when the two levels have the same
L and S values in L&-coupling, or the same ji and values in ^’-coup¬
ling. For intermediate coupling schemes the more nearly alike the
coupling schemes the greater will be the effect.
Consider as a simple example two neighboring sets of 3Di,2,3 terms
arising from different electron configurations of different parity, i.e.,
one configuration odd and the other even. In the ideal case of LS-coup*
ling these levels will be represented as shown at the left in Fig. 19.8.
If now the configurations are of the same parity, i.e., both even or both
odd, a sort of resonance will be set up tending to repel corresponding
levels. These are shown at the right in the figure. The nearer two
1 Slater, J. C., Phys. Rev., 34, 1293, 1929.
* Condon, E. U., Phys. Rev., 36, 1121, 1930.
3 An electron configuration or term is said to be odd when the sum of the l values
of the electrons is odd, and even when the sum is even.
Sec. 19.5] SERIES PERTURBATIONS AND AUTOIONIZATION 391
corresponding levels are to each other, the greater is the repelling effect'
(see Figs. 19.6 and 19.7). In Fig.
19.8 the 3Z)3 terms, being the nearest
to each other, undergo the greatest
displacement, with the result that the
narrow triplet is widened and the wide
triplet is narrowed. This same result
is to be expected when the unperturbed
wide triplet is above the narrow triplet.
19.4. The Anomalous Diffuse
Series of Calcium.—We now return to
the anomalous 3D (4snd) series of
neutral calcium discussed in the
preceding section. Referring to the
calcium energy level diagram in Fig.
11.9, it is seen that the first member
of the diffuse series 4snd is also the
first member of another 3Z> series
3dns. Now the seventh member of
the anomalous series is without doubt Fig. 19.8.—Schematic representation of
term perturbations.
an extra member and should be
called the second member of the 3dns series. If this assignment is
correct the first two terms of the latter series should have as a limit the
first excited 2D state (3d) of the ionized atom. Term values taken with
respect to 2D as a limit give Rydberg denominators 1.60 and 2.64, in good
agreement with the assignment.
A more detailed study of the fine structure of the two 3D series is to
be seen in Fig. 19.9. At the extreme left and right in this figure the two
series are shown as they would occur without perturbation. In the
center the observed series is shown. It should be pointed out that the
unperturbed 3dns triplets should have a total separation of approxi¬
mately 60 wave numbers, the same as that of the series limit 2D (see
Sec. 11.5), and that the higher members of the 4snd series should be very
narrow. The effect of the perturbation is to widen the triplets on the
left and to narrow those on the right, as observed. T^ie third member
of the right-hand series, shown by dotted lines, because it lies above the
first series limit, should not be observed. This will be discussed in
Sec. 19.9.
19.6. The Anomalous Principal Series in Copper.—The fine structure
of the anomalous 2P series in copper is shown in Fig. 19.3. This very
irregular series, probably the most distorted series known, was discovered
by Shenstone and now has a very beautiful and simple explanation.
Although the first term of the series, 3dl04p,2Pj.|, in combination with
3dl04s,2£$, gives rise to the strongest lines in the copper arc spectrum,
392 INTRODUCTION TO ATOMIC SPECTRA [Chap. XIX
it was not until the spectrum had been so thoroughly analyzed and every-’
thing else was accounted for, that the higher members of the series were
definitely identified.
It is now known that the irregularities in this series are due to the
perturbing influences of a foreign, widely spaced, inverted 2P term.
As shown in Fig. 19.10 these two fine-structure terms, 3952 cm-1 and
5973 cm”1, respectively, are due to the odd electron configuration 3d®4s4p,
and the perturbed series of terms are due to the odd configurations
3dlonp. Plotting the quantum defect n — n* against the term values,
Fig. 19.11 results. It is observed that, as the series approaches the
perturbing terms from either side, the mutual repulsions between cor¬
responding terms (i.e., terms with the same J values) increase slowly at
first, then more rapidly.
19.6. The Inverted Alkali Doublets.—It may be shown from the
quantum mechanics or from the classical model of the atom that the
doublet levels of the alkali metals should be normal, i.e., of the two levels
j = l + i and j = l — \ the level j = l — $ should lie deeper. While
this is observed to be generally true, it is well known that the 2D and 2F
terms of some of the alkalies are inverted. In caesium, for example, the
first 2P and 2D terms are normal and
the second 2F term is inverted.1 In
rubidium the 2P terms are normal and
the 2D and 2F terms are inverted.2
In potassium the 2P terms are normal
and the 2D terms are inverted.3 In
view of the anomalous turning over
of the normal copper doublets, dis¬
cussed in the preceding section, it
seems highly probable that the inver¬
sion of the narrow alkali doublets
might be due to the perturbing influ¬
ence of widely spaced inverted doub¬
let terms lying high above the series
limit.4 This inverting effect is shown
schematically in Fig. 19.12.
Inverted doublet terms arise in Fia. 19.12.—Possible explanation of the
inversion of alkali doublets.
general from electron shells lacking
but one electron to complete them. Consequently the excitation
of an electron from one of the completed electron shells to an outer
orbit may give rise to several of the desired doublets. Consider, for
that these hypothetical levels are not discrete in the usual sense but are
spread out over the continuum.
In some instances the quantum condition for resonance between a
given discrete level and continuum are only partly satisfied (z.e., there is
a weak resonance coupling between the discrete state and the continuum,
see Sec. 19.3) so that the process of either spectral emission or autoionizar
tion may take place. This is the subject of the following sections.
19.8. Autoionization in Copper.—It has long been known that the
copper arc spectrum, unlike most spectra, consists of many spectrum
Even Odd Even
lines half of which are sharp and the other half unusually broad and
diffuse. A thorough investigation of these lines by Allen1 has shown
that the diffuseness is not due to a pressure or temperature effect but
to a natural breadth of the energy levels. Such broad levels immediately
suggest that we are dealing with short mean lives and with the process
of autoionization.
A thorough analysis of the copper arc spectrum by Shenstone2 reveals,
in addition to the well-known series of normal doublets, a system of
inverted doublet and quartet levels. From the energy level diagram
of copper in Fig. 19.13 it is observed that several of the inverted doublet
1 Allen, C. W., Phys. Rev., 39, 42, 55, 1932.
* Shenstone, A. G. Phy*. Urn.. 28. 449 1926: 34, 1623, 1929.
396 INTRODUCTION TO ATOMIC SPECTRA [Chap. XIX
and quartet levels lie above the series limit of the normal doublets
and in the continua of even 2S and 2Z> and odd *P and *F terms. Of
these high levels the inverted 4Z>i,t,f,i terms (3d94«5s) are of particular
interest. In combination with a number of lower levels it is observed
that all lines with AD\ or AD\ as initial levels are narrow sharp lines
whereas those with or 4Z>| as initial levels are very broad and diffuse.
Photographs of the combinations 4Fhit^ — 4Z>|,|,|,* (3d94$4p — 3d94s5s)
are shown in Fig. 19.14 and the wave-lengths are given in Table 19.2.
3d3d, 3d4d, 3d5d, and 3d6d. All but the first member of this series
(see Fig. 11.9) lie above the first series limit 4 2S of the four chief series of
singlets and triplets. The continuum above these four chief series
corresponds to even S and D terms and to odd P and F terms. Russell’s
ZP terms (3dmd), being characteristically even, enable the electron in any
one of these states to have a mean life sufficiently great to combine
normally with odd lower terms and give rise to the respectably sharp
1 Russell, H. N., and F. A. Saunders, A atrophys. Jour., 61, 38, 1926.
* Russell, H. N., A atrophy a. Jour., 66, 13, 1927.
398 INTRODUCTION TO ATOMIC SPECTRA [Chap. XIX
spectrum lines observed. The even *S term Sd4d combining with *P 0,1,2,
4s4p in a double electron jump gives rise to three hazy lines at X 2757.40,
X 2749.34, and X 2745.49. The broadness of these lines is due to the
broad 3&i level and indicates that (1) the quantum conditions for auto¬
ionization are not far from satisfied (with respect to the zSi continuum),
(2) the mean life of the atom is long enough to permit some atoms to
return to lower levels giving rise to radiation, and (3) that some of
the atoms starting toward this state go over into the continuum and
self-ionize.
In strontium a ZF° term 4d6p [the corresponding term in calcium
(see Fig. 11.9) is ZF°, 3d5p] lies just above the 52S limit and, as in calcium,
is in a continuum of *F° terms 5smf. These negative 3F°2,3.4 terms
4d6p combine with lower 3D 1,2.3 terms 5s4d, giving rise to five very diffuse
lines (see Table 19.3).
Table 19.3
Strontium, hazy lines (short mean life) Barium, sharp lines (long mean life1)
,
» White, H. E., Fhya. Rev., 38 2016, 1931.
In barium these same 3F°2(3>4 terms 5d7p lie just below the series
limit 62*S so that autoionization is not possible. Combining with lower
3Di,2,3 terms 6s5d, six relatively sharp lines are expected and observed
(see Table 19.3). Where the mean life in a given state is long the lines
will have a very narrow natural breadth, and where the mean life is
relatively small the lines will have wide natural breadth. This question
of the breadths of spectral lines will be taken up in Chap. XXL
19.10. Hyperfine-structure Term Perturbations.—In the early
development of hfs and nuclear spin it was observed that the Land6
interval rule was always valid. This observation is in complete agree¬
ment with the present-day atomic models, and, were it not for one
fairly well-established case in mercury, found by Schuller and Jones,1
the rule might be laid down as infallible. Since the one known departure
from the interval rule is attributed by them to the same type of pertur¬
bation phenomena as those discussed in the first sections of this chapter,
a discussion of this single known example is not out of place here.
1 SchOlbr, H. and E. G. Jones, Zeits f Phys., 74, 631, 1932; 77, 801, 1932.
Sac. 19.10] SERIES PERTURBATIONS AND AUTOIONIZATION
I- W 1-0 l-Yz
Hg 199 Hg 198.200,202,204 Hg 201
1D2; their respective hfs terms perceptibly perturb each other. These
perturbations are shown schematically in Fig. 19.15.
Due to the relatively large number of mercury isotopes (198, 199,
200, 201, 202, and 204) the hfs patterns of a majority of the spectrum
lines are veiy complex. As stated in Sec. 18.12 the very interesting and
praiseworthy analysis of many of these patterns by Schuler and Jones has
led to the assignment of different nuclear spins to the different isotopes:
Starting with the even isotopes with zero spin in the center column
of Fig. 19.15, the two terms *Di and lDt will, for the 199 isotope, each
be split into two hyperfine levels. The same two terms for the 201 iso¬
tope will be split into three and four hyperfine levels, respectively. The
dotted lines represent the centers of gravity of each set of levels (each
level is assigned the quantum weight 2F + 1), and the arrows indicate
the levels that repel each other. It is assumed that there is no isotope
displacement.
By analogy with the perturbation rules given in Sec. 19.3 the quantum
conditions for hyperfine term perturbations may be postulated as follows:
Two levels belonging to the same or to different electron configurations
will perturb each other (1) when both electron configurations are of the
same sign, and (2) when both levels have the same F value. The I value
being always the same for every energy level of each atom, the greatest
effect is to be expected when the two levels have the same J value.
(In the case of mercury the J's are different.) Schuler and Jones point
out that the repulsion is not only a function of the distance between the
levels but also a function of the F and I values. Though the F = $
levels are farther apart than the F = £ levels, the shift is greater. Fur¬
thermore the shift of the F = $ levels for I = \ is twice as great as for
7 = |. A possible explanation lies in the relative magnitudes of the
nuclear factors.
It is to be noted that the levels for 7 = f are inverted with respect
to the levels for I = J. Whatever the coupling scheme may be for the
6s6d configuration, the *7>i term can only be L/S-coupling with S = 1,
L * 2, and J = 1. With a positive g factor and a strong coupling
between the 6s electron and the nucleus, the hfs terms of 3Di will be
large and inverted as observed for I = With a negative nuclear
g factor they will be normal as observed for I = f. The magnitude of
the hyperfine separations shows that the positive g factor is 2.7 times the
negative g factor.
CHAPTER XX
1 Stark, J., Berl Akad. Wiss., 40, 932, 1913; Ann. d. Phys., 43, 965,1919.
1 Lo Stjrdo, A., Accad. Lincei Alii, 22, 665, 1913; 23, 83, 256, 326, 717, 1914.
* A rather complete bibliography and account of the Stark effect is given by R.
Minkowski in H. Geiger and K. Scheel, Handbuch der Physik, Vol. 21, p. 389, 1929.
4 Epstein, P. S., Ann. d. Phys., 50, 489, 1916; also Phys. Zeits., 17,148, 1916.
1 Schwarzschild, K., Sitz-ber. Berl. Akad. Wiss., 1916, p. 548.
4 SchrOdinger, E., Ann. d. Phys., 80, 437, 1926; Epstein, P. S., Phys. Rev., 28,
695, 1926.
7 Schlapp, R., Proc. Roy. Soc., A, 119, 313, 1928.
401
402 INTRODUCTION TO ATOMIC SPECTRA [Chap. XX
We shall not attempt to derive any of the equations for the Stark
effect but shall begin by first writing down the general energy relations
for a hydrogen atom in an electric field and then proceed to interpret
them in terms of atomic models and the observed spectrum lines. The
interaction energy of a hydrogen-like atom in an electric field is given by
AT = AF + BF2 + CF* + • • • (20.1)
Here AT represents the change in the term value of the atom in
wave numbers, i.e.s the shift in the energy levels from the field-free states
to the states in the electric field, and F is the strength of the field in
electrostatic units. The coefficients A B, and C, in this equation, have
been calculated from classical and quantum-mechanical considerations by
Epstein, Wentzel, Waller, Van Vleck, S.Doi, Schrodinger, and others,
and are given by
A = JLn(n* -ni)’
Here n is the usual total quantum number, and m, n2) and mt are electric
quantum numbers, subject to the condition
mi = n — m — 7ii — 1. (20.3)
n = 1, 2, 3, • • • «, n\ = 0, 1, 2, 3, • • • n — 1,
(20.4)
mi — 0, ±1, ±2, • ■ • ± (n — 1), n2 = 0, 1, 2, 3, • • • n — 1.
further, viz., that the first-order Stark effect is observed first in hydrogen¬
like atoms, whereas the second-order Stark effect predominates in most
others. The latter will be considered more in detail in Sec. 20.8.
20.3. Early Orbital Model of Hydrogen in an Electric Field.1—The
early treatment by Epstein and Schwarzschild of the hydrogen atom in
an electric field is shown schematically in Fig. 20.1. We have seen in
Chap. Ill how the Bohr-Sommerfeld orbits, in field-free space, are
quantized, and how the size and shape of each orbit are given by the
total quantum number n, the azimuthal quantum number l (or k), and the
radial quantum number r. In a magnetic field we have seen how
Fig. 20.2.—Classical vector model for the hydrogen atom in a weak electric field.
when in a magnetic field. The state m,• = § in Fig. 20.3, for example,
has the same energy as the state m} — — f. Similarly the states m, = +J
and my = have the same energy. Instead of a level j = $ being
split up into four components as in the Zeeman effect, there are but two
levels. The reason for this can best be understood by referring to the
classical orbital model, or to the quantum-mechanical model of electron
clouds (see Figs. 9.8 and 9.10). The nature of the forces acting on the
electron are purely electrostatic so that the energy of the electron in an
orbit of given n and l depends only on the inclination of the orbit plane
with respect to the electric field, or to the distribution of charge on the
quantum-mechanical model, and not on the direction of rotation or
motion of the electron in its orbit. The states with +my and — my
correspond to the same inclination of the orbital plane, or to the same
charge distributions, and therefore have the same distortion, or energy
change, due to the applied field. In a magnetic field, on the other hand,
406 INTRODUCTION TO ATOMIC SPECTRA [Chap. XX
the energy depends on the direction of rotation, and the energies change
sign when m,- changes sign. It shouid be noted in Fig. 9.8 that for given
n, l9 and j each pair of states ±m7- have the same charge distributions.
Again each pair of states with the same n, j7 and tn, but different l have the
same angular distribution but different radial distribution.
The energy levels of the hydrogen atom for the two states n — 2
and n = 3 are shown in Fig. 20.3. The field-free levels and theoretical
pattern for H« are given at the left and the weak-field levels and pattern
are given at the right. Theoretical treatments of the Stark effect show
that wherever two or more levels with the same n and j but different
Fig. 20.3.—Fine structure and weak-field Stark effect for hydrogen H«. (After Schlapp.)
I lie close together the first-order Stark effect predominates over the
second-order Stark effect. At the left in the figure it is to be noted that
states for which n and j are equal fall together (see Fig. 9.3), and that
these same levels at the right show a large first-order splitting. The
only unpaired state for each n is in each case the one at the top, 2p,
2P§, and 3d,I 22>i, and these show only second-order splittings too small
to show in the figure. The reason why a first-order effect occurs in
the double levels and not in the single may be explained as follows:
Owing to the overlapping of angular wave functions, and the mutual
energy shared by the two states, the atom in either one of the shared
states will possess a net electric moment. When a state is unpaired.
Sec. 20.5] THE STARK EFFECT 407
the electric field direction but for a different reason. We may, there¬
fore, ascribe quantum numbers mi and m, to these two motions where
mi takes values from +1 to — Z, and m. takes the values +§ and — £.
[This is the mi of Eq. (20.3).]
In such a strong field Epstein and Schwarzschild, and others, have
shown that the first-order Stark effect is many times larger than the
so-called second-order Stark effect (see Sec. 20.2), and that the inter¬
action energy of Z* with F is given by the first term of Eq. (20.1).
where a is the l*s* coupling coefficient (see Sec. 8.6) determined from the
fine structure, and where F is measured in volts per centimeter.
Let us now apply Eq. (20.9) to the specific example of the first three
states of hydrogen and then the selection rules to the first member
of the Balmer series Ha. The energies for each quantum state are
easily calculated by tabulating as in Table 20.1.
The allowed values of n, ni, and n2 are first written down in columns
to insure obtaining all possible combinations. The quantities n(n2 — n0
are next computed. From Eq. (20.3) the values of mi are obtained.
For every possible state mi there are the two possibilities m, = +£
and m, = —The a coefficients in the last column are not all equal
as each is a measure of the spin-orbital interactions and there are several
different Z’s. We see from this table that there is exactly the same
number of possible states of the atom as in the Zeeman effect. In the
Stark effect, however, certain levels fall together. Plotting the energies
as given by Cols. 5 and 8, each weak-field level ra;* is connected to a
strong-field level with the same m,-, such that no two levels with the same
m, cross.
Sbc. 20.5] THE STARK EFFECT 409
Tablb 20.1.—Energy Factors for the Stark Effect in Hydrogen
1 0 0 0 0 0 ±i 0
+1 +1 0 0 -1 To/2
o +1 0 +1 +2 0 0
0 +1 -1 -2 0 ±i 0
0 0 0 +1 ±* ±a/2
+2 +2 0 -2 ±*
Ta
+2 +1 +1 +3 -1 ±i To/2
+2 0 +2 . +6 0 ±* 0
+1 +2 -1 —3 -1 ±* To/2
3 +1 +1 0 0 ±* 0
+1 0 +1 +3 +1 ±i ±o/2
0 +2 -2 -6 0 ±\ 0
0 +1 -1 -3 +1 ±\ ±o/2
0 0 +2 ±h ±o
Fxg. 20.5.—Stark effect of hydrogen, showing weak- and strong-field energy le\ els for » ■■ 2
and n « 3-
410 INTRODUCTION TO ATOMIC SPECTRA [Chap. XX
Intensity rules have been worked out from the quantum mechanics
by a number of investigators. Those shown in Fig. 20.6 have been
calculated by Schlapp using as a model the Dirac electron as applied
to hydrogen. Although the weak-field patterns have not been observed,
the strong-field pattern for each line has been observed and found to
be in good agreement with theory, in both the intensities and the relative
separations. It should be pointed out that under different electrical
conditions in different types of sources the relative intensities of the
components are different. The microphotometer curve shown for
Ha in Fig. 20.6, taken from the work of Mark and Wierl,1 is observed
Fig. 20.6.—Stark effect of hydrogen, showing calculated and observed structure for H«.
(Observed field-free -pattern after Spedding, Shane, and Grace; observed strong-held pattern
after Mark and Wierl.)
Coilc.
Obs.
Fig. 20.7.—Observed Stark pattern for hydrogen Ha. (After Mark and Wierl.)
AX = - Xo, (20.11)
Fig. 20.9.—Quadratic Stark effect for the potassium doublet — 5p,sP^|f X4047
and X4044. (After Grotrian and Ramsauer.)
1Ladenburq, R., Phya. Zeits., 22, 549, 1921; Zeits. f. Phys., 28, 51, 1924; Ann.
d. Phya., 78, 675, 1925.
414 INTRODUCTION TO ATOMIC SPECTRA [Chap. XX
Ramsauer1 have observed similar shifts, with slight splitting into com¬
ponents of the doublet 4— 5p,aP*,§. A reproduction of their measure¬
ments on this doublet is shown in Fig. 20.9.
If a given orbit is not far from hydrogen-like, the term will show
only a quadratic Stark effect in a weak held. As the field increases,
and the levels from the hydrogen-like orbits of the same n begin to overlap
the one in question, the term will begin to show a first-order Stark
effect. An electric field which, for a given level, gives only a quadratic
effect is sometimes called a weak field. A field strong enough to give a
first-order effect is also called a strong field. We therefore see that what
may be called a weak field for one energy level may be a strong field
for other levels of the same atom. (It should be noted that this reference
to weak and strong fields cannot be applied to hydrogen, for there the
first-order effect is observed first in weak fields and then the second-order
effect in strong fields.) A pure quadratic Stark effect arises when the
conditions for weak field are satisfied for both the initial and final states.
These are the conditions under which Fig. 20.9 was obtained. As the
field increases, the initial state usually goes over to the strong-field case
long before the final state does. When the field is strong for both the
initial and final states the Stark separations are given by AF + BF2
[see Eqs. (20.1) and (20.2)], and the resultant Stark patterns will resemble
those of hydrogen. We shall now apply these principles to the Stark
effect of helium.
20.8. The Stark Effect in Helium.—The appearance of forbidden
lines in helium when the spectrum is produced in an electric field was
first observed by Koch.1 Following Koch’s discovery, Stark3 and his
coworkers were able to produce long series of forbidden lines like ls2s,*S-
1 snst*S, l«2p,*P — 1 snp,zP, etc., where Al = 0 and ls2s,3& — 1 snd,*D,
l«2p,*P — 1 snf,zF, etc., where Al = 2. Later experiments by Foster4 and
others extended the observations and, under improved experimental
conditions, brought out the significance of these new lines. The correla¬
tion between the Stark patterns of helium with its two electrons and the
Stark patterns of hydrogen with its one electron is one of the most
interesting and remarkable achievements of our modern theory of
atomic spectra.
As a specific example of the Stark effect in helium let us consider
the spectrum line X3705, arising from the transition ls2p,3P — ls7d,*D,
and examine the various energy levels, transitions, and observed Stark
components associated with it. From the energy level diagram of
*1 a«
n an
o
13
*1 an
*4
*2 an
0
*5
*3 an
*1
±6
*4
*2
an
0
±5
*3
11
14
12
0
13
U an
12 an
0
11 an
0 an
Fig. 20.10.—Theoretical Stark splitting for the 1«7Z states of helium, where l — 0, 1, 2, 3.
4, 5, and 6, for a, p, d, f, g, h, and i electrons respectively.
Turning first to the final state l«2p,8P, we find that the only other
near-by states are x8 and *S, arising from ls2s several thousand wave
numbers lower. From the discussion in the previous section one should
therefore expect only a quadratic Stark effect for both the lP and *P
terms of ls2p. Even the very highest fields yet attained would not be
strong enough to bring the ls2s and ls2p sets of levels near enough
together to give rise to a first-order Stark splitting. We may therefore
assume little or no splitting and a very small shift in the final state *P
for all ordinary fields.
416 INTRODUCTION TO ATOMIC SPECTRA [Chap. XX
field strengths.
when the field and splitting are large enough, the 3S term will show a
first-order splitting. In these strong fields the spin-orbit interaction
SL is broken down and the quantum number L no longer has meaning.
Only the projections of the orbital and spin momenta on the field direction
are quantized, and we have as quantum numbers ML and Ms in place
of mi and m, as in hydrogen. For each quantum value of Mh there
will be three possible values of MS) viz.} 1, 0, and — 1. These are indi¬
cated at the right in Fig. 20.10. Although there is no coupling between
the spin resultant S and the field P, S will interact with the orbital
magnetic moment projected on F (see Sec. 20.5) and give rise to splittings
of the same order of magnitude as the fine structure of each triplet.
This splitting is too small to show in the figure. The small separations
appearing at the right are shown for the purpose of correlation with field-
free states.
Mention should be made at this point of the singlet terms arising
from each of the configurations shown in Fig. 20.10. These singlets in
field-free space lie just above each corresponding triplet, and these go
over in an electric field to an array of levels almost identical with the one
Sec. 20.8] THE STARK EFFECT 417
shown for the triplets. Due to the spin-spin interaction this set of levels
will be just above but almost superposed upon the triplet array.
In considering the allowed transitions between two sets of levels the
selection rules,
A Ms - 0,
AMl = 0 for p components, (20.12)
AM l = 1 for s components
must be applied. With the set of levels shown at the right in Fig. 20.10
as initial states, and the states Ml — 1, 0, and —1 (from ls2ptzP)
HELIUM
4120 4026 3867 3819 J733 3705
as final states, transitions from the states ML = 0, ±1, and ±2 are the
only ones allowed. Since the final levels show only a quadratic effect,
the three lower levels will be superposed and the observed spectrum
lines should follow the dotted lines in Fig. 20.11. The observed patterns,
as photographed by Foster, are reproduced in Fig. 20.12. These observed
lines are drawn as heavy lines in Fig. 20.11. In very weak fields only
certain of the lines from levels with higher L values appear. Most of
these in zero field are forbidden transitions. As the field increases, the
2ZP — VP transition appears. Quantum-mechanical calculations for
the tracing of various levels from zero field to strong field have been carried
out by Foster1 and shown to be in good agreement with experiment.
1 For a more detailed account of, and references to the subject of, the breadth
,
of spectrum lines the reader is referred to the article published by V. Weisskopf,
Phys. Zeits., 84 1,1983, and also to Max Bom, “Optik,” pp. 421-455, 1933.
418
Sac. 21.1] THE BREADTH OF SPECTRUM LINES 419
t '
Fig. 21.1.—Solar spectra of the east and west limbs, and the north and south poles,
of the sun. Arrows indicate solar lines showing the Doppler shift. Bracketed lines,
Bhowing no shift, are part of the oxygen A band produced by absorption in the earth’s
atmosphere.
ties the probability that the velocity will lie between u and u + du is
given by
dw = ^e-Ko'du, (21.3)
where
* = 25* • (21-4)
This general equation is plotted in Fig. 21.2 for the purpose of showing
the general shape of a spectrum line which has been broadened by the
so-called Doppler effect. To find the two frequencies at which the
intensity drops to half its maximum value1 the exponential term in
Eq. (21.5) is set equal to one-half. Solving this for v — and multiply¬
ing by two we get for the half-intensity breadth, in absolute frequency
units,
(21.6)
This equation shows that the Doppler broadening is (1) proportional to
the square root of the temperature, (2) proportional to the frequency vQf
and (3) inversely proportional to the square root of the molecular weight.
1The half-intensity breadth of a spectrum line is here defined as the interval
between two points where the intensity drops to half its maximum value. (Some
authors take half of this interval and call it the half breadth.)
Sec. 21.2] THE BREADTH OF SPECTRUM LINES 421
(21.7)
Fig. 21.3.—Amplitude-time curve of a single electric charge oscillating about a fixed point.
A damped oscillator.
it decreases. The latter would show that the x-ray region of the spectrum
might be suitable for the measurement of natural breadths (see Sec. 21.4).
21.2. Natural Breadths from Classical Theory.—According to classical
electromagnetic theory a vibrating electric charge is continually damped
by the radiation of energy. The energy E of such an oscillator decreases
exponentially by
E - E^r*1 (21.8)
and the amplitude (see Fig. 21.3) by
y T
422 INTRODUCTION TO ATOMIC SPECTRA [Chap. XXI
Here a>o is equal to 2r times the frequency of the oscillation p0. The
displacement z of the oscillator at any time t is given by
I{v) (21.12)
2r 4rs0o - >)2 + (|)'
Vo ~ ”= i ■ \6'«= (2L13)
which gives for the natural half-intensity breadth
y 2. 4re2v20
(21.14)
n 2r 3mc*'
Since Av/v = AX/X, and c = v\, the half-intensity breadth, in terms
of wave-length, is
c Awe2
AX 1.16 X 10-12 cm (21.15)
2TV** Zme2
which is constant and equal to 0.000116 A for all wave-lengths, a value
many times too small to be measured by ordinary spectroscopic methods.
21.8 Natural Breadths and the Quantum Mechanics.—According
to quantum-mechanical principles an energy level diagram of an atom
is not tb be thought of as a set of discrete levels but as a sort of continuous
Sec. 21.3] THE BREADTH OF SPECTRUM LINES 423
term spectrum in which the probability distribution is concentrated
in regions where the terms are observed (see Fig. 21.5). With such
probability distributions for the different levels of a given atom the
transitions between levels will not give rise to infinitely sharp lines.
From Heisenberg’s uncertainty principle, expressed in terms of energy
and time,
AEAt ~ h, (21.16)
it is not difficult to understand how it is that the energy levels are
not discrete. Writing for AE the uncertainty in the energy hAT, and
for the time At the mean life t, the uncertainty in
the breadth of an atomic state in absolute
frequency units is of the order
AT ~ ~ (21.17)
I(T) = (21.18)
' ' 2x 4r*(Tn - T)2 + (7n/2)2
where Tn is the term value of the center of gravity of the term distri¬
bution, and
.
7»
_ XV
-
&***2(Tn
3mc, - Tmy 1
(21.19)
T
Here Tm is the term value of any lower level to which a transition may
take place. Tn — T represents small Av intervals in the initial term
and Tn — Tm the mean radiated frequency vn,m corresponding to transitions
between the two levels. With the above value of yn the term half¬
intensity breadth is given [see Eq. (21.14)] by
W(7\> - Tm)2
n’m 3mc*
(21.20)
424 INTRODUCTION TO ATOMIC SPECTRA [Chap. XXI
Now the coefficient Cn,m varies from level to level and depends for its
value on all other states of the atom lower than the one in question:
Cn,m = (21.21)
Here gn and gm are the quantum weights (2J + 1) of the levels for the
transition n —> m, and fn,m is the number of so-called dispersion electrons*
for the same transition. The sum is to be taken over all levels m which
lie deeper than the level n.
Although not easily determined the / factors in the above given
equation are a measure of the intensity of a spectrum line and in a few
simple cases have been calculated1 and measured.2
London has shown from Schrodinger’s theory that the number of
dispersion electrons to be associated with any state of the atom is equal
to unity. This means that, if we sum up the f’s for all transitions into
and out of a given state,
2/ = 1. 1.22)
The fa for transitions into a state are taken with a plus sign and those
out of a state with a minus sign. The table of / values given in Table
21.1 will serve as a simple example of this rule. Since these are all
transitions into the normal state, 3s,2S, and there are no transitions
out, the sum, including the rest of the series and the small sum con¬
i'able 21.1.—Dispersion Electrons, / Factors. Observed Values to Be Asso¬
ciated with the Principal Series of Sodium
(After Filippov and Prokofjew)1
Jf N _ 71
7i +
+ 72
72 1 Fio. 21.6.—Showing th
I\v) — 7y * T~ jT ~ \ 2* relation between the natur
T 4 tt2(vi 9 —
4tt2(vi(2 vv)2
)2 +-
4 ( '1 1 breadth of a spectrum lit
V 2
y 2
)
J
/ni Ai\
and the natural breadths
the corresponding energ
(21.24) levels.
From this relation we observe that the line drops to half its maximur
intensity when the first and second terms in the second denominator ar*
equal, i.ewhen
From this we see that the half-intensity breadth of the radiated lin»
7i + 72
K = vi,2 — v = « *x + h
Doppler breadth is not too large it is easy to see how one might measure
the intensity contours far out on the wings and from these calculate the
half-intensity breadth. This has been done for the sodium D lines by
Minkowski.1 In this particular instance almost perfect agreement
between theory and experiment was found. Many measurements of
other lines have been made in this indirect way with varying degrees
of agreement.
In hydrogen the Doppler effect has been the chief limiting factor
m all attempts to resolve the fine structure of the Balmer series lines
(see Fig. 9.5). For Ha, at temperatures between 250° and 300° Abs., the
Doppler half-intensity breadth is about 0.2 cm-1, or 0.09A, a value of the
same order of magnitude as the largest fine-structure interval. By oper¬
ating a hydrogen-discharge tube in a liquid air bath the Hi and Hi lines
have been considerably sharpened. Attempts to excite the lines with the
tube cooled in liquid hydrogen have been unsuccessful. The effect of
atomic mass on the Doppler breadth is very beautifully brought out by
the comparison of the H« lines reproduced in Fig. 9.5. The Hi lines
Fig. 21.8.—Observed natural breadth contour of one of the anomalous copper arc lines.
See Fi«. 19.14. (After Allen.)
This is but one of the effects produced by the collision of two atoms, one
of which is in the process of emitting or absorbing radiation. Following
the early suggestion of Michelson1 in 1895, and Lorentz2 in 1905, the
phenomenon is based upon the following assumptions: //, during the time
an atom is emitting or absorbing radiation of frequency v0 it collides
elastically with another atom, the phase and amplitude of the radiation
have a chance of undergoing a considerable change.
If one assumes that the time two atoms are in collision is large com¬
pared with the mean time between collisions, the emitting or absorbing
atoms will most of the time be under the influence of strong atomic
fields. These conditions, which exist in a gas at relatively high pressures,
give rise in general to a red shift of the spectrum lines. This phenomenon
will be discussed in the following
section under the heading of Asym¬
metry and Pressure Shift.
In this section we are concerned
with the damping effect, or broaden¬
ing of a spectrum line, produced by
the sudden change in the atomic
radiation by collision. The assump¬
tions usually made are: (1) the
mean time between collisions is large
Fig. 21.8a.—Undamped oscillation in- compared with the collision time,
stantly stopped after a Bhort interval of and (2) with every impact the oscil-
time. Ideal model for collision damping. , ... . . . , ~
lations are either completely cut off,
or they are momentarily interrupted during impact only to resume the
same frequency again with possibly a phase and amplitude change
(see Fig. 21.10). That the phase of the oscillations may be changed
in the short collision time may be seen from the following rough calcula¬
tion. For normal temperatures and normal-effective-collision cross
sections the collision time is of the order of 10'13 sec. For ordinary
visible light with a period of about 10~15 sec. there will be some hundred
modified oscillations during impact.
In giving a classical treatment of the phenomenon of collision damp¬
ing, it is customary to assume the radiation to be an undamped oscillation
of the form shown in Fig. 21.8a, and given by
x — Ao cos (wot + <p) (21.27)
(where «o = 2tv0), which is instantly stopped after an interval of time r.
Now r is the interval between the time the atom begins radiating and
the time a collision occurs. A Fourier analysis of a single finite wave
1 Michelson, A. A., Astrophys. Jour., 2, 251, 1895.
* Lorentz, H. A., Koninklijke. Akad. Wetenschappen Amsterdam, Versl., 14*
518, 577, 1905-1906.
Sac. 21.5] THE BREADTH OF SPECTRUM LINES 420
t ’ i
where r0 is the mean flight time, i.e., the average time between collisions.
When a Fourier analysis of the set of wave trains has been made, the
relative intensity 7, as a function of the frequency v, is found to be given
by the relation
= 7TT
~ o
(21.30)
(21.31)
where R is the universal gas constant = 8.3 X 107 ergs mole-1, T the
absolute temperature, and m the molecular weight. The mean free
path Z0 is just
Zo * t>oTo, (21.32)
430 INTRODUCTION TO ATOMIC SPECTRA [Chap. XXI
1 l8ET (21.33)
7rlo\ Tfi
Source h2 He No A N.
» SchCtz, ,
W., Zeits. /. Phy45 30, 1927.
Since the observed values povt are from two to three times those cal¬
culated from kinetic theory the term optical cross section has been applied
in the one case and kinetic cross section in the other.
In some instances the effect of a foreign gas on the spectrum lines
of a given element is to give very broad lines and large optical cross
sections. In favorable cases popt may be from 10 to 100 times pkll). Th}s
is interpreted to mean that even at great distances another gas atom
may act on the radiating atom strongly enough to change the phase of
the oscillation. Such large actions often arise when no foreign gases
are present.
A comparison of Doppler half-intensity breadths with collision half¬
intensity breadths [Eqs. (21.6) and (21.35)] shows that for visible light,
Sac. 21.6] THE BREADTH OF SPECTRUM LINES 431
and room temperature and pressure, the two effects are of the same order
of magnitude. In mercury, Fabry and Perot reduced the pressure to
1 mm in order to measure the Doppler broadening without large distor¬
tions from collision damping. With increased pressures, on the other
hand, collision damping becomes the predominant cause for the broaden¬
ing of lines. The effect of pressure or density [see Eq. (21.35)] on the
breadth of a line is shown in the specific case of the mercury resonance
line X2537 in Fig. 21.9. Here the half-intensity breadth, as it is observed
in absorption, is plotted against the relative density, the latter being
defined as the density of the gas relative to its density at normal tem-
Fig. 21.11.—Potential curves for two states of an atom at close approach of another atom.
Illustrating the cause for asymmetry and red shift. (After Jabldnski and Weisskopf.)
Fig. 21.12.—Asymmetry and broadening of the mercury resonance line X2536 with increased
pressure. (After Fdchtbauer, Joos, and Dinkelacker.)
observed is spread out more on the long wave-length side than it is on the
short (see Fig. 21.12).
At normal temperatures and pressures the collision times are small,
and the broadening and shift are small. With increasing pressure
the mean collision time increases and the time between collisions decreases
1 Jabl6nski, A., Zeil8.f. Phpe., 70, 723, 1931; Bee also Margenau, H., Phys. Rev.,
,
43 129, 1933.
Sec. 21.6] THE BREADTH OF SPECTRUM LINES 433
with the result that, as the line is shifted to the red, it is broadened
asymmetrically. The mercury resonance line X2537 is shown in Fig.
21.12, as observed in absorption by Fiichtbauer, Joos, and Dinkelacker,1
at a pressure of 10 and 50 atm., respectively. The foreign gas used in
these observations was nitrogen. To show that the pressure shift
depends upon the foreign gas used,
curves have been reproduced in Fig.
21.13. In general it is observed that
the shift of a line is very closely pro¬
portional to the relative density. The
relative density is here defined as the
density of the gas relative to the
density at normal temperature and
pressure.
That the lowering of an energy
level due to pressure effects is greater SO 40
Relative Density
for high levels than it is for the low Fig. 21.13.—Pressure shift of the
is confirmed by Babcock’s2 interfer- mercury resonance line X2537, in absorp-
ometer measurements on the iron F^bauer, Joo,. and
Fig. 21.14.—Relative shifts of the energy levels of iron as a function of their term values.
(After Babcock.)
term values, relative shifts are all that can be determined. The results
from over 100 lines are best shown by the graph reproduced in Fig. 21.14,
where each term is plotted relative to the ground state as zero. . Weight¬
ing each plotted point, Babcock expresses the belief that the quintet and
septet terms are depressed slightly more than the triplets. Whether
this is a real effect or not, the major effect is in good agreement with what
one would expect from polarization effects explained above.
21.7. Stark Broadening.—In an ordinary arc of high current density
many ions are produced which upon collision with other atoms give
rise to strong electric fields. The effect of these intermolecular fields
is to produce a Stark-effect broadening of the observed spectrum lines.
In addition to this effect by ion fields, dipole or quadrupole moments
of gas atoms or molecules may also produce relatively strong inter¬
molecular electric fields. The hydrogen atom and molecule are good
examples of a dipole and quadrupole, respectively. Being of the same
order of magnitude as collision damping and Doppler broadening (see
Secs. 21.5 and 21.1), the Stark effect plays a relatively important part
in the general subject of the breadth of spectrum lines. In attempting
to calculate Stark-effect broadening one is immediately confronted
with the problem of continually changing, inhomogeneous, electric
fields. At the approach of two atoms in collision the electric field at
each of the respective atoms increases, reaches a maximum at closest
approach, and then diminishes as the two atoms recede.
Since the Stark effect for an inhomogeneous electric field has never
been worked out, Holtsmark1 and Debye2 made the simplifying assump¬
tion that there will be an average intermolecular field F to which, one
can assume, the ordinary Stark-effect formulas apply. The problem
of calculating average fields has been divided into three classes, (1)
the field due to charged atoms or molecules, i.e.} ions, (2) the field due
to dipole moments, and (3) the field due to quadrupole moments. Although
the theoretical treatment of each of these three possibilities is out of
place here, we shall write down the derived formulas for the three average
field strengths:
(For ions) F = aicn*. (21.36)
(For dipoles) F = a*tin. (21.37)
(For quadrupoles) F = ci^qn*. (21.38)
Here the a’s are constants, e is the ionic charge, m the dipole moment,
q the quadrupole moment, and n the number of corresponding particles
per cubic centimeter.
Holtsmark made the next simplifying assumption that a spectrum
line is spread out symmetrically into a continuous frequency band, the
1 Holtsmark, J., Phys. Zeits., 20, 162, 1919; 26, 73, 1924.
* Debye. P.. Phvs. Zeits.. 20. 160. 1919.
Sec. 21.7] THE BREADTH OF SPECTRU 1ft LINES 435
The essential features of the above given equations are the expressions
for the half-intensity breadths of the observed spectrum lines. For the
latter two types of field these are:
For ion 8t = 3.25A^enK (21.42)
For quadrupole 8, = 5.52A^qn*, (21.43)
As one might well expect, all half-intensity breadths are functions of the
density n.
Of the many attempts made to check these formulas with observa¬
tion those made by Holtsmark are perhaps the best. Of these his
comparison of Eq. (21.43) with Michelson’s observations on the hydrogen
1 For a derivation of the equations given here the reader is referred to M. Bom,
,
“Opfifc,” 1933; also V. Weisskopf, Phys. Zeits84 1, 1933.
436 INTRODUCTION TO ATOMIC SPECTRA [Chap. XXI
line H« has been pointed out by several investigators as giving the best
general agreements. The quadrupole moment of the hydrogen molecule
is known from theory (by Debye) to be 3.2 X 10-26 gm* cm* sec.-1.
This value substituted in Eq. (21.43) gives the half-intensity breadth as
a function of n. Am is calculated from Eq. (20.2). In Fig. 21.15 both
the calculated and observed half-intensity breadths have been plotted
for comparison. Subtracting the Doppler breadth the agreement, even
here, is not very good.
In general the first-order Stark effect will account for the broadening
of many spectrum lines. The second-order, or quadratic, Stark effect
which only becomes appreciably large in strong fields must also produce
an observable effect.
One of the best confirmations of St ark-effect broadening may be
found in the chief series of the alkali atoms where historically the terms
sharp and diffuse series originated. Due to the penetration of the s and
p orbits involved in the sharp-series lines, both in the initial S states
and in the final P states show no first-order Stark effect. As a result
the lines, despite strong intermolecular electric fields, are sharp. For the
diffuse series, however, the d and / orbits are nearly hydrogen-like
and, being not far removed from the hydrogen-like /, g, h, . . . orbits,
show a first-order Stark effect. Under suitable excitation conditions
even the diffuse and fundamental series have been observed as sharp
lines.
Finkelnburg1 has studied the Balmer series under pressures of from
1 to 30 atms. of hydrogen. Using a high-voltage condensed-spark
discharge, between metal electrodes placed very close together, each
member of the series is observed as a very broad line. At a pressure
of about 2 atms. the H«, H^, and HT lines are symmetrically broadened,
without an appreciable shift of the maximum, to the extent that the half¬
intensity breadths are about 25, 100, and 200 A, respectively. At from
10 to 30 atms. the lines are so broad that they overlap one another and
form a continuous spectrum. These enormous breadths are attributed
by Finkelnburg to a Stark effect, the enormous fields being produced
by the very high ion density in the spark discharge. That the Stark
effect is chiefly responsible for the broadening is confirmed by (1) the
general contour of the lines, (2) the symmetrical broadening without
appreciable shift, and (3) the increased broadening with higher members
of the series.
It should be pointed out that asymmetry and pressure shifty discussed
in Sec. 21.6, may well be classified as a second-order Stark effect. At
close approach of a foreign atom the associated electric field causes a
polarization and a displacement of the energy levels of the atom in
question (see Fig. 21.11).
1 Finkelnburg, W., Zeite. f. Phys , 70, 375, 1931.
APPENDIX
Table I.—Values of the Physical Constants
(After Birge)1
Velocity of light. c - (2.99796 ± 0.00004) X 1010 cm sec.”
Mass of the electron (spectro-
scopic) . m© (9.036 ± 0.010) X 10”*g
Mass of the proton. Mp (1.6608 ± 0.0017) X 10”Mg
Charge on the electron. e (4.770 ± 0.006) X 10”10 abs-e.s.u.
Planck’s constant. h (6.647 ± 0.008) X IQ'*7 erg. sec.
Rydberg constant (hydrogen1). Rhx (109677.769 ± 0.06) cm”1
Rydberg constant (hydrogen2). Rh* (109707.56 ± 0.05) cm'1
Rydberg constant (helium) Rue (109722.403 ± 0.05) cm*1
Rydberg constant (infinite.
mass). R (109737.42 ± 0.06) cm*1
Ratio Mp/mo (spectroscopic).... 1838 ± 1
Wave number per absolute volt. (8106 ± 3) cm"1 abs-volt”1
Wave-length per absolute volt. (12336 ± 5) X 10”* cm”1 abs-volt“
Energy of one abs-volt electron (1.5910 ± 0.0016) X 10”1* erg
Speed of one abs-volt electron. . (5.9346 ± 0.0017) X 107 cm sec.^
Fine-structure constant.. . . a = 2ire*/hc (7.283 ± 0.006) X 10"*
a* (5.305 ± 0.008) X 10”»
1/a 137.29 ±0.11
Unit angular momentum. h/2ir (1.0419 ± 0.0013) X 10”27 erg. sec.
Magnetic moment (one Bohr
magneton). mi = (0.9174 ±0.0013) X 10”*° erg gauss”
Magnetic moment (one nuclear
magneton). mi/1838 = 4.991 X 10~84 erg gauss”1
Zet>man displacement per gauss. - (4.674 ± 0.003) X 10"# cm-1 gains”
Bohr magneton Mi
Ratio = (0.8805 ± 0.0005) X 107 gauss”1 sec”
Bohr mechanical moment h/2r
1 Biroe, R. T.t Phya. Rev., Supplement, X, 1, 1929.
7 - 0 1 2 3 4 7 - 0 1 2 3 4 5 6 7
71 - i, it - I 11 9 m 8
ji - 1, it - 2 4 14
sd 7*i - I* 7*i - 2 (8) 1 (S)
ii - 1. 71 - 1 18 I
7i - i, h - 3 2 I 12 14
dd
7i - |. 71 - 2 (2) (1) (13) (34)
8f ji - 4. it - 1 1 2!
7*1 - I. it - K jH (1) 1 (X) X (1)
>i - *. y* - i 21
E ■
- 4. it - 4 BE (2) 7i - 2. 7i - 1 A HI III IIS
it - 4.7* - 1 1 I df 71 - 2, 71 - I §3 9 491 if
PP
71 - 3.7* - 4 (2) <l> 7*i - t, 7*1 - I 0/0 it 93 IS 18 II
7l - I. 7* - 1 0 0 (D X (4) 7*i - I. 71 - I IS 278 II 948 111! I
71 - 4. it - 2 1 44 7*1 " *. 7. - 1 0/0
(9) 9 (9) 9 (9)
71 - It it - | S3 V 2 » - «. 71 - I i i 4! 14 14 SI
Pd
7*1 * 2t it - 2 IX IS 11 7i - h it - 1 (!) (1) (II) (14) (24) (SI)
71 - 3. 7. - « 14 w 111 S h - i. 71 - I 0/0 (f)| # (») 1 (9) f
INDEX
A Aufbauprinzip, Bohr's, 78, 79
Auroral line, 266
Abnormal series, 18, 386, 388 Autoionization, 386, 394
Absorption, Kirchhoff’s law of, 1 in barium, 397
Absorption edges, x-rays, 306, 307 in calcium, 397
Absorption spectra, x-rays, 306, 329 in copper, 395
Actinium, 83 in strontium, 397
Allen, C. W., 396, 427 Azimuthal quantum number, 43, 73
Allison, S., 313 defined, 43
Alternation law of multiplicities, 249, 251
Aluminum, 81, 82 B
anomalous series in, 386, 388
energy level diagram for, 115 Babcock, H. D., 291, 433
nuclear moment of, 372 Bacher, R. F., 364, 378, 382
Anderson, O. E., 370 Back, E., 73, 163, 169, 184, 223, 227, 354.
Angstrom, A. J., 2 357, 376, 380, 382
unit of wave-length, defined, 2 Back-Goudsmit effect, of hyperfine
Angular factor, probability density, 60, structure, 376
64 Biicklin, E., 300
Anomalous series, 18, 386, 388 Ballard, S. S., 37
Anomalous Zeeman effect, 152 Balmer, 3
Antimony, 83, 259, 307 law, 4
nuclear moment of, 372 series in hydrogen, 3, 14, 30, 31, 33
Aphelion, of electron orbit, 44 fine structure of, 132, 135, 137
Arc spectra, photographs, 174, 250, 349 Paschen-Back effect of, 168
Argon, 81, 82, 274 photographs of, 5, 31
fine structure of, 277 fine structure, 132, 138
Arsenic, 83, 259 Stark effect, 410
nuclear moment of, 372 Barium, 83, 176
Asymmetry of spectrum lines, 418, 431, ionized, energy level diagram of, 98
432, 433 nuclear moment of, 372
Atom model, for complex spectra, 251 photograph of triplet in, 174
for doublets, 118 Barkla, C. G., 301
for hyperfine structure, 354, 365 Bergmann, 6, 9
Back-Goudsmit effect, 376, 380 series, discovery of, 6
Faschen-Back effect, 376, 380 Beryllium, 80, 81, 82, 176
for//-coupling, 191, 224 energy level diagram of, 179
for //-coupling, 254 ionized, energy level diagram of, 98
for L5-coupling, 191, 224, 254 Bethe, H., 313
for Paschen-Back effect, 163 Bevan, P. V., 17
complete, 231 Bichowsky, F. R., 121
for Stark effect, 405, 407 Birge, R. T., 17
for Zeeman effect, complex atoms, 286 series formulas, 17
hyperfine structure, 373 Bismuth, 84, 259
one-electron systems, 154 Back-Goudsmit effect, 381
443
444 INTRODUCTION TO ATOMIC SPECTRA
Complex spectra, Paschen-Back effect, Dirac, P. A. M., 124, 139, 206, 312
290 electron, 139
Stark effect in, 412 angular distribution for, 140
Zeeman effect, 286 eigenfunctions for, 139
intensity rules, 289 for hydrogen, 139
photographs of, 288, 291 photographs representing, 146
selection rules, 289 probability-density distribution for,
Compton, A. H., 301 140, 143, 146
Compton, K. T., 93 radial distribution for, 144
Condon, E. U., 72, 211, 217, 269, 280, 390 Dispersion electrons, table of, 424
Continuous spectrum, in hydrogen, 33 Displacement law, complex spectra, 249
Continuum, 33, 392 Doan, R. L., 301
Copper, 82, 250 Dolejsek, V., 306
anomalous series in, 386 Doppler effect, 417, 418, 424
energy level diagram of, 395 formula for, 418, 419
nuclear moment of, 372 photographs of, 418
perturbations in, 391 Dorgelo, H. B., 119, 120, 160
Core, atom, 184 Double-electron excitation, 179
Correspondence theorem, Bohr, 38, 41 Double-electron transitions, optical
Coster, D., 311, 312, 324 spectra, 269
Coupling schemes, for two electrons, 189 x-ray spectra, 327
Crawford, M. F., 363, 368 Doublets, fine structure, 11, 114
Critical potentials, 92, 96 formulas for, 128
of the alkali metals, 95 intensity rules for, 118
selection rules for, 117
D separations for, 128
vector model for, 118
Damping, collision, 427 Druyvesteyn, M. J., 324, 325, 326
Darwin, C. G., 124, 139, 166 Dysprosium, 83
Darwin, K, 227, 241
Dauvillier, A., 331 E
Davis, Bergen, 301
De Broglie, L., 306, 331 Eckart, C., 383
corpuscular wave equation, 54 Edlen, B., 32
Debye, P., 434 Effective nuclear charge, 180
Derivation of spectral terms, from equiv¬ Effective quantum number, 89
alent electrons, 187, 292, 293, 295 Eigenfunctions, 59, 139
^/'-coupling, 196 defined, 57
LS-coupling, 187 Einstein, A., 23, 312
shorthand method, for equivalent Electron orbits, 25
electrons, 296 circular, 25, 29, 30, 34
two arbitrary electrons, 187 elliptic, 42, 47
two equivalent electrons, .^-coupling, in hydrogen, 25, 29, 30, 34, 42, 47, 72
238 in lithium, 103, 111
LiS-coupling, 236 in potassium, 101
Designation of terms, ^-coupling, 196 precession of, 104, 106, 108
L&-coupling, 187 quantum-mechanical picture of, 63,
Dewar, J., 4 68, 71, 100, 101, 103
Diagonal lines of multiplets, 206 rubidium, 100
Diffuse series, discovery of, 6 series of, 113
fine structure in, 12, 20 sodium, 101, 103
photographs of, 174 Electron spin-orbit interaction, 124, 128,
Dinkelacher, O., 433 129
446 INTRODUCTION TO ATOMIC SPECTRA
Lithium, effective quantum number for, Magnetic energy, for Zeeman effect, 157,
89 217
energy level diagram for, 77, 87 Magnetic moment, of atoms, with one
-like atoms, 339 valence electron, 155
nuclear moments of, 372 with two valence electrons, 215
probability-density distributions for, ^'-coupling, 217
101, 110, 111 L*ST-coupling, 216
singly ionized, energy level diagram of nucleus, 359
for, 209 of orbital electron, 52, 155
term values for, 88 of proton, 361
Liveing, G., 4 of spinning electron, 155
^-coupling, 185 Magnetic quantum number, 158, 218
12-coupling, 252 complex spectra, 292
London, F., 424 Magnusson, T., 301
Lorentz, H. A., 152, 428 Manganese, 82, 250, 270
classical treatment of Zeeman effect, energy level diagram for, 271
152 fine structure in, 271, 273
unit, for Zeeman effect, 158 multiplets in, 272
Loring, R. A., 223 nuclear moment of, 372
Lo Surdo, A., 401 Margenau, IL, 432
JLS-coupling, 186, 190 Mark, H., 410
g factors for, 157, 216, 240, 241, 287 Masaki, O., 393
g permanence rule, 239, 240 Masurium, 83
g sum rule for, 241 Mechanical moment, defined, 52
T factors, 190, 191, 246 Meggers, W. F., 265, 353, 364
T permanence rule for, 243 Meissner, K. W., 393
r sum rule for, 244, 245 Mercury, 83, 176
energy level diagram, 179
intensities for, 206
hyperfine structure, 384
interaction energy for, 191
term perturbations in, 399
Pasehen-Back effect for, 224
isotope structure in, 384
selection rules for, 203
nuclear moments of, 372
sum rules, for intensities, 206 Metastable state, 96
to .//-coupling, 279, 281, 283, 284 Meyer, H. T., 313
vector model for, 191 Michelson, A., 132, 428
Zeeman effect for, 286 Minkowski, R., 401, 426
Lutecium, 83 Modified Moseley diagram, 335
nuclear moment of, 372 for isoelectronic sequence, 341, 344, 348
Lyman, T., series, in helium, He II, 31, 32 Moller, F. L., 93
in hydrogen, 14, 30, 33 Molybdenum, 83
More, K., 353, 382
M Morse, P. M., 72, 211, 217
Moseley, H. G. J., 302
flf-series x-ray lines, photographs of, 305 diagram, isoelectronic sequence, 333,
Mack, J. E., 350 341, 344, 348
modified, 335
McLay, A. B., 363, 368
for a single atom, 348
McLennan, J. C., 363, 368
x-rays, 311, 321
MacMillan, 353
A-series, 302
Magnesium, 81, 82, 176
L-scries, 305
energy level diagram for, 179 satellites, 325
ionized, energy level diagram for, 98 Multiplets, in calcium, 181
photographs of triplets in, 174 defined, 179
INDEX 451
Multiplets, in manganese, 272, 273 Orthohelium, 209
photographs of, 260, 340, 346, 346 Osgood, T. H., 301
in titanium, 269 Osmium, 83
x-ray structure, 323 Oxygen, 81, 82, 266
Murphy, G. M., 37
P
N
Palladium, 83
Natural breadths, 420 Parhelium, 209
of energy levels, 423 Paschen, F., 11, 13, 18, 31, 132, 139, 153,
formulas for, 422 163, 169, 227, 267, 331
observed in copper, 427 Paschen-Back effect, complete, 231-235
of spectrum lines, 425, 426, 427 complex spectra, 290
Nebular lines, 266 hydrogen, 168
Neodymium, 83 hyperfine structure, 376
Neon, 81, 82, 274 interaction energy for, 225
fine structure of, 277 ^-coupling, 223, 228
series in, 18 LS-coupling, 223, 224
Nickel, 82, 250 magnetic energy for, 164
Nitrogen, 81, 82, 259 normal triplet, 168
energy level diagram for, 260 one electron, 162
fine structure in, 260 pattern for jj-coupling, 227
multiplets in, 262 for LS-coupling, 227
nuclear moment of, 372 photograph of, 169
Nonpenetrating orbits, 100, 103, 105 principal series doublet, 165
Normal doublets, 123 selection rules for, 168, 227
Normal states, of atoms, 277 vector model for, 163
Normal Zeeman triplet, 151, 152, 219 Pauli, W., 124, 222, 231, 239, 353
Notation, electron orbit, 48 exclusion principle, 188, 236, 293
for hyperfine structure terms, g permanence rule, 239
in jj-coupling, 356 g sum rule, 222, 241
in LN’-coupling, 356 Pauling, L., 102, 106, 109, 231, 318, 322,
jy-coupling, 196 330, 358
JW?-coupling, 185 Penetrating electron orbits, 100, 103, 104,
odd and even terms, 204 178, 339
series, 10, 13 Pentad, of multiplets, 189
two-electron systems, 185 Perihelion, of electron orbit, 44
x-rays, 308 Periodic table, 79, 81, 82, 83, 86
Nuclear g factors, 361, 363 Perturbations, conditions for, 390
calculation of, 371 in copper, 391
table of, 372 of energy levels, 387
Nuclear magneton defined, 359 nature of, 390
Nuclear moments, table of, 372 of terms, 386
Nuclear spin, 353, 361, 363, 365 Pfund series, in hydrogen, 14, 30, 33
Phase integral, 24, 27, 43, 49, 50
O Phase space diagram, 24, 27, 45
Phosphorus, 81, 82, 259
Odd terms, defined, 187, 204 nuclear moment of, 372
Optical cross-section of atoms, 430 Photographs, of absorption spectra, 17
Optical doublets, 332 of anomalous Zeeman effect, 152, 291
irregular doublet law, 337 of Back-Goudsmit effect, 381
regular doublet law, 338 of Balmer series of hydrogen, 531
Omstein, L. &, 119, 160, 220 of complex spectra, 250
452 INTRODUCTION TO ATOMIC SPECTRA
Vector model, for Zeeman effect, 154, X-rays, non-diagram lines, 324
286, 373 notation, 308
complex spectra, 286 photoelectrons, 307
hyperfine structure, 373 photographs, 303, 305
one-electron systems, 154 reduced terms, 316, 321
Von Traubenberg, H. Rausch, 411 regular doublet law, 316, 318
relativity doublets, 315, 316
W satellites, 324
energy level diagram for, 328
Waller, I., 69 explanation of, 326
Walter, B., 300 screening constants, 317, 318
Wave numbers, defined, 6 screening doublets, 319
Weak field, Stark effect, 404 selection rules, 301, 324
Zeeman effect, 149, 217 shells, electron, 307, 309
Weisskopf, V., 418, 435 soft radiation, defined, 301
Wentzel, G., 106, 318, 322, 325, 394 Sommerfeld’s fine-structure formula
Weyl, A., 72 for, 315
White, H. E., 36, 70, 142, 267, 278, 296, spectra, photographs, 303, 305
332, 335, 339, 343, 349, 357, 364, spin-relativity doublets, 315, 316
subshells, 310
Whitelaw, N. G., 106 wave-lengths, 304
Wierl, R., 410 Xenon, 83, 274
Wilber, D. T., 296 energy level diagram for, 275
Williams, J-H., 313 fine structure for, 277
Wills
WilBorif E. b7Tio44 . Y
Wilson, W„ %
Ytterbium, 83
Wind, C. H., % I
Wolfe, H. C., 327 *’
Z
Wollaston, W. H., 1
Wood, R. W., 404 Zacharias, J. R., 361, 373
Wulfe, J., 380, 382 Zeeman, P., 149, 380
Zeeman effect, anomalous, 152
X classical explanation, 149
complex spectra, 286
X-rays, absorption spectra, 306 hyperfine structure, 373
explanation of, 329 energy level diagram for, 373
discovery of, 299 selection rules for, 375
electron shells, 308, 309 vector model for, 373
subshells, 310 intensity rules, 160, 220
energy level diagram, 308, 310 complex spectra, 289
fine structure, 314 ^/-coupling, 217, 218
formula for, 315 .LS-coupling, 217, 218
irregular doublet law for, 319 normal Zeeman triplet, 151, 219
K radiation defined, 301 defined, 151
L-doublets, 319 one-electron systems, 149
L radiation defined, 301 p and s components, defined, 151
Moseley diagram for, 311 photographs of, 220
Moseley law, 321 polarization of lines, 151, 160
X-series, 302 Preston’s law, 151
L-series, 305 selection rules, 159, 219
Af-series photographs, 305 complex spectra, 289
multiplet structure, 323 two electrons, 215
INDEX 457
Zeeman patterns, 161, 222 Zeeman patterns, triplets, 222
calculation of, 221 vanadium, 291
* chromium, 291 Zinc, 82
complex spectra, 291 energy level diagram, 179
doublets, 161 nuclear moment of, 372
jy-coupling, 223, 224 photographs of triplets, 174
LS-coupling, 223 Zeeman effect, 152, 176, 220
rapid calculation of, 221 Zirconium, 83