Yavorsky, Seleznev - Physics - A Refresher Course - Mir - 1979

Download as pdf or txt
Download as pdf or txt
You are on page 1of 664

B.M. Yavorsky, Ya.A.

Seleznev

PHYSICS
A Refresher Course

M ir P ubl i sh e r s M o s co w
B. M.YAVORSKY
Y a A. SELEZNEV

physics
A Refresher
Course
B. M. Hbopckhh, K). A. Cejie3HeB

OIPABOHHOE PyKOBO^CTBO
n o (DH3HKE
AJiH nocTynaiom ax b By3U
h caMooSpaaoBaHHH

H3ffaTeJibCTBo «HayKa»
MocKBa
m sits
A Refresher
Course
by
B. M. YAVORSKY
and
Yu. A. SELEZNEV

Translated
from the Russian by G. Leib

MIR PUBLISHERS
MOSCOW
First published 1979

Ha dHBAUUCKOM X8UK€

© PisA&TejibCTBO «HayKa», MocKBa, 1975


© English translation, Mir Publishers, 1979
PREFACE
The fundamental role played by physics in modern
scientific and technical progress has resulted in an important
revision of the content of physics courses in secondary educa-
tional institutions. The new syllabi and teaching methods
used in secondary physics courses are much broader in scope
than those used a few years ago. The interest in physics has
also grown considerably among people studying on their
own. This interest has been stimulated by popular science
literature, series of television and radio broadcasts, and
public lectures. Because of this, there is an increasing need
for various refresher courses in the subject.
The present book gives definitions of the basic physical
concepts and quantities studied in an elementary physics
course, formulates physical laws, and briefly explains the
essence of the phenomena they describe. Some chapters
contain problems and their solutions.
>The book is intended for a very broad circle of readers:
secondary school pupils, students at vocational schools and
special secondary educational institutions, and students
Inking preparatory courses at institutes. It will be a great
help to students preparing for entrance examinations to
higher educational establishments. Having in mind people
studying physics independently, the authors also included
s o m e information on classical and modern physics that is
beyond the scope of secondary school syllabi. These chapters
ii nd sections are marked with an asterisk. Attention is con­
st nntly given to the interpretation of the physical meaning
of laws and the phenomena they describe. The mathematics
needed to use the book is within the scope of a secondary
mathematics course.
6 Preface

The units of physical quantities and the systems of units


are given in a supplement (Part VII).
A detailed index and a system giving references to rele­
vant paragraphs should facilitate the finding of required
information. The references indicate the numbers of the part,
chapter, section, and paragraph containing information rela­
ted to the matter under consideration. For example, the refer­
ence (1.4.3.2°) signifies that the reader will find the infor­
mation in which he is interested in Part I, Chap. 4, Sec. 3,
Par. 2. In references to information contained in the same
section, only the paragraph number is indicated, for instance
(Par. 5°). References to the supplement (Part VII) indicate
only the part and section, or the part, section, and paragraph,
for example (VII.3) or (VII.2.10).
Parts I and VII were written by Yu. A. Seleznev, and
Parts II-VI by B. M. Yavorsky. The problems with solutions
in Parts II-VI were selected by N. A. Boguslavskaya.

B. M. Yavorsky
Yu . A . Seleznev
CONTENTS
Preface 5

MECHANICS

Chapter 1. Kinematics 17

1. Mechanical Motion 17
2. Displacement Vector. Distance 21
3. Velocity 23
4. Acceleration 26
5. Uniform Rectilinear Motion 28
6. Uniformly Changing RectilinearMotion 31
7. Free Fall of Bodies 34
8. Motion of a Body Thrown Vertically Upward 35
9. Uniform Circular Motion of ajParticle 38
10. Motion of a Body Thrown at an Angle with the Ho­
rizontal 41
11. Rotation of a Perfectly Rigid Body About a Fixed
Axis 45

Chapter 2. Dynamics of Motion of a Point Par­


ticle 48
1. Newton’s First Law 48
2. Force 50
3. Mass and Momentum. Density 52
4. Newton’s Second Law 54
5. Newton’s Third Law 57
6. Law of Conservation ofMomentum 58
7. Galilean Principle of Relativity 61
8. Forces of Gravity 63
9. Elastic Forces 68
10. Forces of Friction 70
11. Ways of Measuring Mass and Force 72
12*. Non-Inertial Reference Frames 77

Chapter 3*. Elements of Dynamics of Rotation


of a Perfectly Rigid Body About a Fixed Axis 79
1. Moment of Force and Moment of Inertia 79
2. Fundamental Law of Rotational Dynamics 8§
8 Contents

Chapter 4. Statics 84

1. Addition and Resolution of Forces Applied to a


Point Particle and a Perfectly Rigid Body 84
2. Conditions of Equilibrium of a Point Particle and
Perfectly Rigid Body in an Inertial Reference
Frame 88
3. Kinds of Equilibrium 91

Chapter 5. Work and Mechanical Energy 95

1. Work of a Force in the Motion of a Particle and


the Translational Motion of a Perfectly Rigid
Body 95
2*. Potential and Non-Potential Forces. Conservative
and Non-Conservative Systems ofBodies 99
3. Mechanical Energy 101
4. Law of Conservation ofMechanical Energy 105
5. Power 109

Chapter 6. Elements ofFluidMechanics no

1. Mechanical Properties of Fluids 110


2. Fluid Statics 111
3. Motion of Fluids 116
4*. Motion of Solid Bodies in Fluids 121

MOLECULAR PHYSICS AND FUNDAMENTALS OF


11 THERMODYNAMICS

Chapter 1. Fundamentals of Molecular-Kine­


tic Theory 125

1. Basic Concepts and Definitions 125


2. Brownian Motion 127
3. Diffusion 128
4. Forces of Interaction Between Molecules 129
5. Potential Energy of Interaction of Two Molecules 131
6. Structure of Gases? Solids, and Liquids 13?
Contents 9

Chapter 2. Molecular-Kinetic Theory of Ideal


Gases # 136
1. Ideal Gas 136
2. Velocities of GasMolecules 137
3*. Mean FreePath of aMolecule 139
4. Fundamental Equation of the Kinetic Theory of
Gases 141

Chapter 3. Ideal Gas Laws 144

1. Equation of State 144


2*. Thermodynamic Processes 148
3. Laws of Isoprocesses in Ideal Gases. Equation of
State of an Ideal Gas 150

Chapter 4. Fundamentals of Thermodynamics 156

1. Total and Internal Energy of a Body (a System of


Bodies) 156
2. Work 158
3. Heat 160
4. Heat Capacity 162
5. The First Law ofThermodynamics 163
6*. Reversible and IrreversibleProcesses 168
7*. Cyclic Processes (Cycles) 169
8*. Carnot Cycle 170
9*. The Second and Third Laws ofThermodynamics 173
10*. Heat Engine 174
11*. Refrigerator 176

Chapter 5. Mutual Transitions of Liquids and


Gases 178
1. Evaporation of Liquids 178
2. Saturated Vapour 178
3. Boiling 179
4. Vapour Isotherm 181
5. Critical State of aSubstance. Liquefaction of Gases 183
6. Humidity of Air 184

Chapter 6. Properties of Liquids 186

1. Energy of Surface Layer and Surface Tension 186


Wetting. Capillary Phenomena J88
10 Contents

Chapter 7. Solids and Their Transformation into


Liquids 193

1. Kinds of Crystalline Solids 193


2. Elastic Properties of Solids 194
3. Thermal Expansion of Solids and Liquids 197
4. Melting, Crystallization, and Sublimation of Solids 200

. FUNDAMENTALS OF ELECTRODYNAMICS

Chapter 1. Electrostatics 203


1. Basic Concepts. The Law of Conservation of Electric
Charge 203
2. Coulomb’s Law 204
3. Electric Field. Field Intensity 208
4. Examples of Electrostatic Fields 213
5. Conductors in an Electrostatic Field 219
6. Dielectrics in an Electrostatic Field 221
7. Work of the Forces of an Electrostatic Field 226
8. Potential of an Electrostatic Field 229
9. Relationship Between Intensity and Potential
Difference of an Electrostatic Field 232
10. Capacitance 235
11. Capacitors 237
12. Energy of an Electric Field 240

Chapter 2. Steady Electric Current 243


1. Basic Concepts and Definitions 243
2. Conditions Needed for a Steady Current to Appear
and Be Maintained 246
3. Electromotive Force. Voltage 247
4. Ohm’s Law 248
5. Temperature Dependence of Resistance 252
6. Branching of Currents. Connections of Conductors 253
7. Work and Power of a Current. The Joule-LenzLaw 262

Chapter 3. Electric Current in Non-Metallic Media 263

1. Current in Electrolytes 263


g. Laws of Electrolysis. Discreteness of Electric Charges 264
Contents 11

3. Current in Gases # 265


4. Semi-Self-Maintained Gas Discharge 267
5. Self-Maintained Gas Discharge 268
6. Plasma 270
7. Current in a Vacuum. Emission Phenomena 271
8. Two-Electrode V alve-D iode 273
9. Three-Electrode Valve—Triode 275
10. Electron Beams. Cathode-Ray Tube 277
11. Electrical Conduction of Pure Semiconductors 279
12. Impurity Electrical Conduction of Semiconductors 281
13. Electrical Properties of p- and n-Type Semiconductor
Junctions 283

Chapter 4. Magnetic Field of a Steady Current 285

1. Magnetic Field. Induction Vector of a Magnetic


Field. Magnetic Flux 285
2. Ampere’s Law 291
3. Magnetic Field of an ElectricCurrent 292
4. Interaction ofParallel Currents 296
5. Action of a Magnetic Field on a Moving Charge.
Lorentz Force 297
6. Specific Charge of Particles 301

Chapter 5. Electromagnetic Induction 302

1. Phenomenon and Law of Electromagnetic Induction 302


2. Induced E.M.F.’s in Moving Conductors 304
3. Induced Electric Field 307
4. Induced Currents in Solid Conductors 308
5. Self-Induction 308
6. Mutual Induction. Transformer 310
7. Energy of a Magnetic Field 312

Chapter 6. Magnetic Properties of a Substance 314

1. Magnetic Moments of Electrons and Atoms. Electron


Spin 314
2. Classification of Magnetic Substances 316
3. Diamagnetism 317
4. Paramagnetism 318
5. Ferromagnetism 3g0
12 Contents

IV. OSCILLATIONS AND WAVES

Chapter 1. Mechanical Oscillations 325

1. Basic Concepts and Definitions of Oscillatory Proc­


esses 325
2. Velocity and Acceleration of Harmonic Oscillation 328
3. Harmonic Oscillations of a Spring Pendulum 331
4. Harmonic Oscillations of a Mathematical Pendulum 333
5. Energy of Harmonic Oscillatory Motion 335
6. Addition of Harmonic Identically Directed Oscil­
lations 337
7. Damped Oscillations 338
8. Forced Oscillations 340
9. Auto-Oscillations 343

Chapter 2. Electromagnetic Oscillations 346


1. Free Electromagnetic Oscillations in an Oscillator
Circuit 346
2. Forced Electromagnetic Oscillations. Alternating
Current 350
3. Alternating-Current Circuit. Resistance 351
4. Inductive Reactance 352
5. Capacitive Reactance 353
6. Ohm’s Law for an Alternating-Current Circuit 354
7. Power of Alternating Current. Effective Values of
Current and Voltage 355
8. Resonance in an Alternating-Current Circuit 356
9. Valve Generator 358

Chapter 3. Elastic Waves. Sound 359


1. Preliminary Concepts 359
2. Transverse and Longitudinal Waves 361
3*. Wave Velocity 363
4. Wavelength 364
5*. Equation of a Plane Wave 365
6*. Energy and Intensity of a Wave. Equation of a
Spherical Wave 366
7. Some Characteristics ofSound Waves 367
8. Ultrasounds 369
9*. Interference 370
10*. Standing Wave§ 373
dontentt 13

Chapter 4. Electromagnetic Waves 376


1. Relationship Between Varying Electric and Magnetic


Fields 376
2. Velocity and Some Basic Properties of Electromag­
netic Waves 377
3. Energy and Intensity of Electromagnetic Waves 380
4*. Emission of Electromagnetic Waves 381
5. Concept of Radio Communication, Television, Ra­
dar and Radio Astronomy 385

V. OPTICS

Chapter 1. Geometrical (Ray) Optics 390

1. The Rectilinear Propagation of Light 390


2. Laws of Reflection and Refraction of Light. Total
Reflection 391
3. Plane Mirror. Plane-Parallel Plate. Prism 395
4. Spherical Mirrors 397
5. Lenses 400
6. Concept of Photometry 404
7. Optical Instruments 407

Chapter 2. Wave Optics (Light Waves) 414

1. The Speed of Light 414


2. Interference 416
3. Diffraction 420
4* Diffraction by aSlit. Diffraction Grating 422
5. Polarization 425
6. Dispersion 427

Chapter 3, Radiation and Spectra 429

1. Thermal Radiation. Blackbody 429


2. Energy Distribution in the Spectrum of a Blackbody 431
3. Luminescence 434
4. Kinds of Spectra 435
5. Infrared and Ultraviolet Radiation 436
6. X -R ays 437
7. The Electromagnetic Spectrum 441
u dontenis

Chapter 4*. Fundamentals of the Special Theory


of Relativity 442

1. Laws of Electrodynamics and the Classical Princi­


ple of Relativity 442
2. Postulates of the Special Theory ofRelativity 444
3. Concept of the Length of a Body 446
4. Simultaneity of Events. Synchronization of Time­
pieces 447
5. Relative Nature of Simultaneity ofEvents 449
6. Lorentz Transformations 450
7. Relativity of Lengths 451
8. Relativity of Time Intervals 453
9. The Relativistic Law of Velocity Addition 456
10. Relativistic Dynamics. Dependence of Mass on Ve­
locity 457
11. The Mass-Energy Relation 458

Chapter 5. Quantum Optics 462


1. Basic Concepts 462
2. The Photoelectric Effect 464
3. Laws of the Photoemissive Effect. Einstein’s Pho­
toelectric Equation 466
4. Applications of the Photoelectric Effect 468
5. Light Pressure 470
6. Chemical Action of Light. ThePhotographic Process 472

VI. ATOMIC AND NUCLEAR PHYSICS

Chapter 1*. Elements of Quantum Mechanics 474


1. De Broglie’s Ideas on the Wave Properties of Particles 474
2. Wave Properties of Electrons, Neutrons, Atoms, and
Molecules 476
3. The Physical Meaning of de Broglie Waves 479
4. Linear Harmonic Oscillator. Motion of an Electron
in a Limited Region of Space 480
5. Uncertainty Relation 484
6. Part Played by Uncertainty Relations in Studying
the Motion of Microparticles 488
7. Zero-Point Energy of a LinearHarmonic Oscillator 490
8. The Degeneracy of Gases 491
Contents 1&

Chapter 2. The Structure of Atoms 493



1. Rutherford’s Nuclear Model of an Atom 493
2. Predicaments in the Classical Explanation of the
Nuclear Model of an Atom 496
3. Line Spectrum of a Hydrogen Atom 498
4. Bohr’s Postulates 499
5. Bohr’s Model of the Hydrogen Atom 501
6*. Substantiation of Bohr’s Postulates and the Phys­
ical Meaning of the Electron Orbit in Quantum
Mechanics 504
7*. Quantization of the Angular Momentum of an Elec­
tron and of Its Projection 505
8*. Electron Spin. The Pauli Exclusion Principle 508
9. Mendeleev’s Periodic System of the Elements 510
10*. Lasers and Masers 513

Chapter 3. Structure and Spectra of Molecules 517

1. General Characteristicof ChemicalBonds 517


2. Ionic Molecules 519
3. Molecules with a CovalentChemical Bond 521
4*. Molecular Spectra 522

Chapter 4. Structure and Basic Properties of


Atomic Nuclei 526

1. General Characteristic 526


2. Binding Energy. Mass Deficiency 528
3. Nuclear Forces. Drop NuclearModel 531
4. Natural Radioactivity 534
5. Displacement Law and Fundamental Law of Radio­
active Decay 535
6. Experimental Methods of Studying Particles and
Radioactive Radiation 539
7. Origin of Alpha, Beta* and Gamma Rays 542
8. Nuclear Reactions 545,
9. Interaction of Neutrons with a Substance 548
10. Artificial Radioactivity 550
11. Fission of Heavy Nuclei 551
12. Nuclear Fission Chain Reactions. ANuclear Reactor 554
13. Applications of Nuclear Energy and Radioactive
Isotopes 557
14*. Biological Action of RadioactiveRadiations 959
15*. Fusion Reactions 561
16. Accelerators 565
i6 Contents

Chapter 5. Elementary Particles 568

1. General 568
2. Classification of Elementary Particles and Their
Interactions 570
3*. Cosmic Rays 575
4*. Information on Selected Elementary Particles 576
5. Antiparticles 580
6*, Concept of the Structure of a Nucleon 583

•SUPPLEMENTS

1. Units and Dimensions of Physical Quantities.


Systems of Units 586
2. Basic and Derived Units of the SI System 588
3. Units of Physical Quantities in Mechanics 588
4. Units of Physical Quantities in Molecular Physics
and Thermodynamics 590
5. Units of Quantities in Electrodynamics 599
6. Units of Selected Quantities in Wave Processes and
Optics 603
7. Selected Units in Atomic and Nuclear Physics 603
8. Selected Physical Constants 607
9. Ways of Measuring Physical Quantities 607
10. Errors in Measuring Physical Quantities 614
11. Processing the Results of Direct Measurements 617
12. Processing the Results of Indirect Measurements 619
13. Approximate Calculations Without Accurate Account
Taken of Errors 623
Name Index 626
Subject Index 627
PART MECHANICS

Chapter 1

Kinematics
1. Mechanical Motion
1°. Mechanics studies the simplest kind of motion—me­
chanical motion. By mechanical motion is meant the change
in the position of a given body (or its parts) relative to
other bodies. It is sometimes a simple matter to observe
mechanical motion: an electric locomotive moves relative to
the station platform or the railway track, motor vessels
move and change their position relative to the banks of
rivers or the shores of seas and oceans. Some mechanical
motions do not lend themselves to direct observation. For
example, the atoms and molecules of gases move relative to
the walls of the vessels containing them. We are convinced
that such invisible mechanical motions occur by the physical
phenomena associated with them (11.1.2.1°; 11.1.3.1°).
Newtonian, or classical, mechanics deals with the mechan­
ical motions of bodies occurring at velocities that are much
smaller than the speed of light in a vacuum (IV.4.2.1°).
2°. Kinematics is the branch of mechanics concerned with
the mechanical motions of bodies in time without regard to
the action of other bodies or fields on these bodies.
2-0211
1$ M echanics t.1.1

3°. To describe mechanical motion, it is necessary to indi­


cate’ the body relative to which the motion is being consid­
ered. For example, a passenger sitting in a plane and the
fuselage of the plane travel relative to the Earth, but are
stationary relative to each other. The body relative to which
a given mechanical motion is being considered is called a
reference body.
A coordinate system is related to a reference body. The
rectangular, or Cartesian, coordinate system X Y Z shown in
Fig. 1.1.1 is a very simple coor­
dinate system. The combination
of a reference body and a coordi­
nate system is called a reference
frame. For different reference fra­
mes, see 1.2.1.3°, 6°.
4°. When solving some prob­
lems in mechanics, we may dis­
play no interest in the shape and
dimensions of a body. A body
whose dimensions may be disre­
garded in a given problem is
called a point mass, or point particle, or a particle. For in­
stance, when considering the annual motion of the Earth
about the Sun, the Earth may be considered as a point particle.
In other cases (for example when analysing the diurnal
motion of the Earth about its axis), we may not disregard
the dimensions of a body. A body whose shape and dimen­
sions may be considered constant regardless of all kinds of
external action on it is defined as a perfectly rigid body,
or a rigid body. The latter can be considered as a system of
rigidly connected point particles at constant distances from
one another.
5°. The position of the particle M in the Cartesian system
of coordinates is determined by three coordinates (x, y,
and z) (see Fig. 1.1.1). It is also possible for the position of
the particle to be given by the radius vector, or position
vector, r drawn from the origin of coordinates O to the
point M (see Fig. 1.1.1).
6°. Mechanical motion takes place in time. To determine
the moments of time which different positions of a moving
body (or particle) correspond to, the reference frame must
t.1.1 Kinematics 19

he provided with a clock showing the time intervals from an


arbitrarily chosen initial moment of time. Time always
changes from the past to the future. For the synchronization
of timepieces, see V.4.4.30.
7°. When the particle M (see Fig. 1.1.1) moves, the end
of the position vector r travels along a certain line in space.
When a body of finite dimensions moves, different points
of it travel along different lines in the general case. The line
along which a particle travels is called its path, or trajectory.
The equation showing how the position vector of a moving
particle depends on time
r = r (t)
or the system of equations
x = / (t), y = <P (0. and z = (t)
equivalent to it are called the equations of motion of the
particle.
8°. With respect to the shape of the trajectory, mechan­
ical motions are defined as rectilinear and curvilinear. In
the former case, the trajectory of motion in the given refer­
ence frame is a straight line, in the latter, a curve. For
example, a body dropped from the hands at a small height
above the Earth’s surface moves rectilinearly, while a body
thrown horizontally moves curvilinearly (1.1.10.1°).
If all the points of a trajectory are in one plane, the
motion is called plane, or two-dimensional.
The trajectories of a given mechanical motion can have
different shapes in different reference frames. For instance,
a body dropped from the hands in a coach of a uniformly
moving train (1.1.5.1°) travels rectilinearly along a vertical
line in a reference frame associated with the coach, and
along a parabola in a reference frame associated with the
railway track.
Example 1. Assume that a point particle moves in the
plane XO Y and the equations of its motion have the form
x — at and y = bt
where a and b are constants differing from zero.
Find the shape of the trajectory along which the particle
moves.
io________________________ Mechanics_______________________t J J

Excluding the time t from the given equations, we have

i.e. the trajectory of the particle is a straight line passing


through the origin of coordinates.
Example 2. The motion of a point particle in the plane
XOY is described by the following equations:
x = a sin co£ and y — a cos cot
where a and co are constants differing from zero.
What is the shape of the trajectory of the particle?
We find from the above equations that

a
= since/ and —
a
= cos(ot
Squaring these equations and adding the expressions ob-

Fig. 1.1.2

tained, we get an equation of the trajectory not containing


t (because sin2 cot + cos2 ayt = 1), i.e.
x2 + y2 = a 2
The particle travels along a circle of radius a. The centre
of the circle coincides with the origin of coordinates.
9°. The simplest kinds of mechanical motion of a perfectly
rigid body (1.1.1.4°) are translational and rotational motions.
The motion of a body is translational if all its particles
have congruent trajectories. Any straight line connecting
two arbitrary points (A and B) of a body when travelling
remains parallel to its former positions (Fig. 1.1.2.) The
piston in a cylinder of an internal-combustion engine or
the drawer of a desk being pulled out performs translational
motion.
1.1.2. Kinematics 21

The translational motion of a perfectly rigid body can


be characterized by the motion of any point of it, for instance
by the motion of its centre of mass (1.2.3.4°).
In the rotational motion, or rotation, of a perfectly rigid
body, its points travel along circles in parallel planes. The
centres of all the circles are on a single straight line per­
pendicular to the planes of the circles and called the axis of

rotation. The axis of rotation 0 0 can be inside a body (Fig.


1.1.3) or outside it (Fig. 1.1.4). The axis of rotation may be
stationary or moving depending on the reference frame.
For instance, in a reference frame associated with the Earth,
the axis of rotation of a generator rotor at a power plant is
stationary, while the axes of the wheels of a travelling auto­
mobile move.

2. Displacement Vector. Distance


1°. When a particle moves, the location of its position
vector (1.1.1.5°) in space changes. The difference
Ar = r2 — rx
of the position vectors characterizing the final (2) and ini"
lial (1) positions of a particle moving during the time inter­
val At = t2 — tx is called the displacement vector (displace­
ment) (Fig. 1.1.5). The projections of the displacement vec­
tor onto the coordinate axes OX, OY, and OZ can be ex­
pressed through the differences between the coordinates of its
22 Mechanics 1.1.2

tail and head:


Arx = Ax = x2 — x1
Ar y = Ay = y 2 — yt
Arz = Az = z2 — z1
The coordinate differences Ax, Ay, and Az between two po­
sitions of a moving point particle are often defined as the
displacements of the particle along
the relevant coordinate axes.
Plots of the relationships
Ar* = Ar* (t), Ary = Ar y (t),
and Arz = Ar2 (t) are called the
displacement graphs along the
corresponding coordinate axes.
Displacement vectors are add­
ed geometrically, according
to the parallelogram or polygon of vectors law (the law of
vector addition).
2°. The distance, or path length (S or AS), is a scalar
quantity equal to the length of the portion of a trajectory
travelled by a moving particle dur­
ing a given time interval. The !dis-
tances travelled by a particle dur­
ing consecutive time intervals are
added arithmetically.
The magnitude (numerical or
absolute value) Ar* of a displace­
ment vector in the general case does
not equal the distance AS travelled
by a particle during a given time
interval.
A plot of the relationship S =
= S (t) is called a distance graph.
Example 1. A particle moves vertically upward from
the Earth’s surface and when it reaches the maximum height
H drops back to the Earth. The displacement vector of the
particle equals zero, and the distance travelled by the par­
ticle equals the double height 2H.
* The magnitude of a vector quantity A is represented by placing
vertical bars on both sides of the symbol for the vector: | A |, or by
the use of italics: A.
1.1.3 Kinematics 23

Example 2. A cyclist travels along a circular path of


radius R. During a certain time interval, he travels half the
length of the circumference. The#magnitude (absolute value)
of the displacement vector of the cyclist equals the diameter
of the circle (2/?), and the distance equals half the length
of the circumference (rci?).
Example 3. A particle sequentially moves from the
position O to A, then to B, C, and so on (Fig. 1.1.6). The
distance travelled by the particle equals the sum of the
portions of its path: S = OA + AB + BC + CD + DE.
The’Misplacement vector Ar = OE joins the initial position
0 of the particle to its final position E. The magnitude of
the displacement vector Ar = | OE | does not equal the
distance S travelled by the particle.

3. Velocity
1°. The average velocity (vav) during the time interval
At = t2 — tx is the physical quantity equal to the ratio of
the displacement vector Ar = r2 — rx of a particle to the
time interval At:

The direction of the average velocity vector coincides with


that of the displacement vector Ar.
The average velocity characterizes motion during the
entire time interval At for which it has been determined.
Example. A small heavy ball drops vertically. The
coordinates z t of the ball along the axis OZ directed verti­
cally downward at different moments tt measured from the
beginning of motion are indicated below:
th s 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 1.00
zh m 0.049 0.196 0.441 0.784 1.22 1.76 2.40 3.14 3.97 4.90

Find the magnitudes of the average velocities of the ball


^av. i - 2» ^av. 5- 6» ^av,5- 10. and yav, 1_10 during the relevant
time intervals from tx = 0.10 s to t2 = 0.20 s, from t5 =
24 Mechanics 1.1.3

= 0.50 s to t6 — 0.60 s, from t5 = 0.50 s to Z10 1.00 s,


and from tx = 0.10 s to t10 = 1.00 s.
0.196—0.049 - K .
y av, 1-2 0 . 2 0 — 0 .10 m ^s
1.76 — 1.22
^av, 5-6 — 0.60 — 0.50 5.4 m/s
4.90 — 1.22 ,
Uav, 5-10 j >00 — 0#50 ^ 7.4 m/s
4.90—0.049
yav, 1-10“ 1>00_ o .io ; 5.4 m/s

2°. By the velocity (instantaneous velocity, velocity at


a given moment of time) is meant the physical quantity equal
to the limiting value reached by the average velocity (Par. 1°)
as the time interval At tends to zero:
Ar
v = lim vav lim
\t-+ o M-+0 AT
The velocity equals the limit of the ratio of the elemen­
tary displacement Ar to the elementary time interval At

during which this displacement occurs. The velocity vector


is directed tangentially to the path. The direction of the
velocity is called the direction of motion of a particle
(Fig. 1.1.7).
3°. The motion of a point particle is defined as uniform
if the magnitude of its instantaneous velocity does not
change with time (v = const). If the instantaneous velocity
of a particle does change with time, the motion is called
non-uniform (or variable).
4°. The average speed (vSt av) is the physical quantity de­
termined by the ratio of the distance AS travelled by a
particle during the time interval At to the length of this
1.1.3 Kinematics 25

interval:
45
US, a v — -7
At7 -
As the time interval At tends to zero, the instantaneous
value of the speed
,. AS
Vs = Ahf -m0 1 f
coincides with the magnitude v of the instantaneous velocity
of a particle because
lim = lim I Ar[ v At
A f-0 A t-0
In the general case, the average speed v St av does not
equal the magnitude of the average velocity of a particle
yav. The equality v St av = vRV is observed only in rectili­
near motion of a point particle without any change in the
direction of motion.
The average speed is convenient for describing motion
along a closed path or along a path whose different portions
intersect.
Example. A point particle during the time interval At
performs a complete revolution along a circle of radius
/?. The average velocity of the particle is zero (vav = 0
because Ar = 0), whereas the average speed during the
same time interval differs from zero (vp%av = 2nR/At).
5°. A velocity graphTis a plot of the time dependence of
the projection of the velocity vector onto a coordinate axis
\vx = vx (t), or v v'= v T/ (t), or v2 = vz (J)l. The time measured
from a conventionally chosen initial moment (t0 = 0) is
laid off along the axis of abscissas at a definite scale p*,
and the values of the projections of the velocity vector onto
the given coordinate axis are laid off along the axis of ordi­
nates at the scale
A plot of the relationship v = v (£)~when the magnitudes
of the velocity of a particle at different moments are laid
off along the axis of ordinates at a definite scale is also
called a velocity graph.
* Tt will not be mentioned in the following, as a rule, that when
plotting graphs the values of physical quantities are laid off at defi­
nite scales along the coordinate axes.
26 Mechanics 1.1.4

Example. A point particle moves uniformly along a circle


in the plane XOY. The magnitude of the velocity of the
particle is v. The initial position
A of the particle at the moment
t 0 = 0 and the orientation of the
coordinate axes are shown in
Fig. 1.1.8a. Graphs of the rela­
tionships v = v (t), Vx= v x (t),
and Vy—Vy (t) at the scales fip and
Ht during the time T of one com­
plete revolution of the particle
along the circle are shown in
Fig. 1.1.86, c, and d.

4. Acceleration
1°. The average acceleration
(aaav) is the physical quantity equ-
al to the ratio of the change in
the velocity Av Vo — v, of a
point particle to the length of the time interval At = t2
during which the change occurred:
Av
v —"At
The directions of the vectors aav and Av coincide.

Fig. 1.1.9

Example. Figure 1.1.9a shows portion 1-2 of the trajec­


tory of a point particle. At the moment t±J the particle has
the velocity v1? and at the moment t 2, the velocity v 2.
1.1.4 Kinematics 21

The average acceleration vector aav is directed in the same


way as the vector of the change in velocity Av = v2 — vx
(Fig. 1.1.96). In the general case, the direction of the vector
aav coincides neither with that of the vector vlt nor with
that of the vector v2, nor with that of a tangent to any
point of the trajectory on its given portion.
2°. The acceleration (instantaneous acceleration) of a
point particle at the moment t is defined to be the physical
quantity a equal to the
limiting value of the ave- _?t
rage acceleration (Par. 1°)
during the time interval
from t to t + At as At
tends to zero: Av>0
(a)
,. Av
a = lim aa >= lim —
Af-0 A/-+-0 Fig. 1.1.10

The acceleration (at a given point of a trajectory or at a


given moment of time) equals the limit of the ratio of the
elementary change in the velocity Av to the elementary time
interval At.
In a given reference frame, the acceleration vector can be
set by its projections onto the relevant coordinate axes (by
the projections a x, a yy and az).
3°. The component aT of the acceleration vector directed
along a tangent to a trajectory at a given point is called the
tangential acceleration . The latter characterizes a change in
the magnitude of the velocity vector. The vector a x runs in
the direction of motion of a particle when its velocity grows
(Fig. 1.1.10a) and in the opposite direction when the velo­
city diminishes (Fig. 1.1.106).
The component an of the acceleration vector directed
along a normal to a trajectory at a given point is called the
normal acceleration . The latter characterizes a change in the
direction of the velocity vector in curvilinear motion.
Inspection of Fig. 1.1.10 shows that the quantities
a, a x, and an are related by the expression
a = Y a\ + al,
4°. When mechanical motions of a point particle are
classified according to two features simultaneously—accord-
28 Mechanics 1.1.5

ing to the shape of the trajectory and according to the


nature of the change in the velocity—four kinds of motion
are distinguished:
(a) uniform rectilinear,
(b) non-uniform rectilinear,
(c) uniform curvilinear, and
(d) non-uniform curvilinear.
Table 1.1.1 characterizes the accelerations and velocities
of these kinds of motion.

Table 1.1.1

Uniform motion Non-uniform motion

= = f-- 0,
ax — flu —0
Rectilinear a = 0, a = 0 a = at , a =£0
motion v = const v=£ const
v = const const

ax — 0, 0 an =fc0

Curvilinear a = a ni a =£0 a = V a t + an ’ a=£ (>


motion i^ c o n s t
v = const
v=£ const \=j= const

5. Uniform Rectilinear Motion


1°. In the uniform rectilinear motion of a point particle,
its instantaneous velocity does not depend on the time
(v = const and v = const) and is directed along the trajec­
tory at each of its points. The average velocity during any
time interval equals the instantaneous velocity of the parti­
cle: v av — v. Thus,

2°. The plot v = v (t) for uniform motion is a straight


line parallel to the time axis Ot (Fig. 1.1.11).
1.1.5 kinematics 29

The form of the plots vx= v x (/), vy^ v (J (t), and vz=^vz (t)
depends on the direction of the vector v and on the choice
of the positive sense of the relevant coordinate axis.

----- ^
0 t
Fig. 1.1.11

Example 1. A particle M moves uniformly and recti-


linearly with the velocity v, and the positive sense of the

M v M v
-O- - »»------X *
(a) (a)

>0 0
}ux<0
h
( b ) ( 6)
Fig. 1.1.12 Fig. 1.1.13

coordinate axis OX coincides with the direction of the vec­


tor v (Fig. 1.1.12a). The plot vx = vx (t) for this case is shown
in Fig. 1.1.126.
Example 2. If the positive sense of the coordinate axis
OX is opposite to the direction of the velocity vector v of
the particle M (Fig. 1.1.13a), then the plot vx = vx (t) has
the form shown in Fig. 1.1.136. The ordinates of all the
points of this graph are equal and negative.
3°. In uniform rectilinear motion with the velocity v,
the displacement vector Ar of a point particle during the
time interval At = t — t0 equals Ar = v At.
4°. The form of the displacement graph (1.1.2.1°) of a
point particle along any of the axes of a Cartesian coordinate
system for uniform rectilinear motion depends on the sign
of the projection of the velocity vector of the particle onto
the given coordinate axis. For example, if the projection
30 M echanics i.i.5

vx of the velocity of a particle onto the coordinate axis OX


is positive (Fig. 1.1.14a), then the displacement graph
Arx along the axis OX has the form shown in Fig. 1.1.146.

Fig. 1.1.14

If vx < 0 (Fig. 1.1.15a), then the graph Arx = Arx (t) has
the form shown in Fig. 1.1.156. In both cases, the slope
(tan a) of the graph with respect to the axis Ot (the angle

Fig. 1.1.15

a is measured from the positive sense of the coordinate axis


counterclockwise) corresponds to the projection vx of the
velocity of the particle onto the coordinate axis OX (see
Figs. 1.1.14 and 1.1.15):

ItHt = vx
where pr and ut are the scales along the corresponding axes of
the displacement graph (1.1.3.5°).
5°. The distance S travelled by a point particle in uni­
form rectilinear motion during the time interval At =
= t — t0 equals the magnitude Ar of the displacement vec­
tor of the particle during the same time interval. Con-
1.1 .6 kinematics Si

sequently,
S = v At = v (t — t0)
or, if t0 = 0,
S = vt
Example. A particle moves uniformly and rectilinearly
from a position with the coordinate x1 = 3 m to a position

D
o----------------- —
I----O----------- _►
x2 0 Xj X
Fig. 1.1.16

with the coordinate x2 = —5 m (Fig. 1.1.16). The displace­


ment of the particle along the coordinate axis OX is
Arx = x2 — xx = —5 m — 3 m = —8 m
(the minus sign signifies that the particle moves in the nega­
tive sense of the axis OX).
The distance travelled by the point is
S = I x2 — xx | = | —5 m — 3 m | = 8 m

6. Uniformly Changing Rectilinear Motion


1°. Uniformly changing rectilinear motion is a particu­
lar case of changing motion when the magnitude and direc­
tion of the acceleration remain constant (a = const). The
average acceleration aav equals the instantaneous accelera­
tion a (aav = a). The acceleration a is directed along the
trajectory of a particle. The normal acceleration (1.1.4.3°)
is absent (an = 0).
If the directions of the acceleration a and the velocity
v of a particle coincide, the motion is defined as uniformly
accelerated. The magnitude of the velocity of uniformly
accelerated motion of a particle grows with time.
If the directions of the vector a and v are opposite,
the motion is defined as uniformly decelerated (motion with
a negative acceleration). The magnitude of the velocity of
uniformly decelerated motion diminishes with time.
32 Mechanics 1.1.6

2°. The change in the velocity Av = v — v0 during the


time interval At = t — t0 in uniformly changing rectilinear
motion is
Av = a A£
or
V — v0 = a (t — t0)
If at the initial moment of time (t0 = 0) the velocity of
a particle is v 0 (the initial velocity) and the acceleration
a is known, then the velocity v at the arbitrary time t is
v = v0 + a*
The projection of the velocity vector onto the axis OX
of a Cartesian coordinate system is related to the correspond­
ing projections of the vectors of the initial velocity and the
acceleration by the equation
Vx = Vox 4”
The equations for the projections of the velocity vector onto
the other coordinate axes are written in a similar way.
The above expressions for v and vx hold not only for
uniformly accelerated or decelerated motion, but also for
uniformly changing motion when the direction of the velo­
city is reversed.
3°. The displacement vector Ar of a particle during the
time interval At = t — t0 for uniformly changing rectili­
near motion with the initial velocity v0 and the acceleration
a is
A a , , sl (At)2
Ar = v0 AH---- y 2-
while its projection onto the axis OX of a Cartesian coordi­
nate system (or the displacement of a particle along the
relevant coordinate axis) when t0 = 0 is
Arx = v0xt + -^f-
The equations for the projections of the displacement vector
onto the axes OY and OZ are written similarly.
The above expressions for Ar and Ar* hold for any recti­
linear uniformly changing motion, including motions in
which the direction of travelling is reversed.
1.1.6 Kinematics 33

4°. The distance S travelled by a particle during the


time interval At = t — t0 in uniformly accelerated recti­
linear motion with the initial velocity v0 and the accelera­
tion a, when t0 = 0, is
S = vQt + —
When v0 = 0, the distance
equals
at2
S=
A plot of *S = S (t) for this
uniformly accelerated motion is
shown in Fig. 1.1.17.
5°. For uniformly decelerated rectilinear motion, the
formula for the distance

So = v0t*----at2
y

holds only up to the moment when motion stops (or up to


the moment when the direction of
motion changes) tCl which can be
found from the condition
vc = v0 — atc = 0
whence
tc ~ a
The distance travelled by a par­
ticle during the time interval
At = tc — t0 is
o _ v\
2^r
A plot of S — S (t) for uniformly decelerated motion
is shown in Fig. 1.1.18.
6°. In uniformly changing rectilinear motion with a
reversal of the direction of motion at the moment tc, the
distance travelled by a particle by the moment t is

s V* , g (At)*
2a 2
34 Mechanics 1.1.7

where the first addend is the distance travelled by the parti­


cle in uniformly decelerated motion from the initial moment
t0 = 0 to the moment tc, and the second addend is the dis­
tance travelled by the particle in uniformly accelerated
motion during the interval At = t — tc.

A plot of S = S (t) for such motion is shown in Fig. 1.1.19.


7°. The distance travelled by a particle in uniformly
changing rectilinear motion during the arbitrary interval
At = i2 — tx can be found from the given plot v = v (t)
(Fig. 1.1.20).
The hatched area in the figure corresponds to the distance
required.

7. Free Fall of Bodies


1°. Free fall is defined as the motion which a body would
perform only under the action of the force of gravity
(1.2.8.3°) without account taken of the air resistance. When
a body freely falls from a small height h from the surface
of the Earth (h < i?E, where i?E is the Earth’s radius),
it moves with a constant acceleration g directed vertically
downward.
The acceleration g is called the acceleration of free fall.
It is the same for all bodies and depends only on the altitude
and the geographical latitude.
2°. If at the initial moment (i0 = 0) a body had the velo­
city v 0, then after the arbitrary interval At = t — tu the
U .8 Kinematics 35

velocity of the body in free fall is


v = v0 + g t
At an initial velocity of falling equal to zero (v0 = 0),
the velocity of the body at the arbitrary time t is
v = gt
3°. The distance h travelled by a body in free fall by
the time t is
h=v*t + ^ ~

If the initial velocity of the body is zero (v0 = 0), then

4°. The magnitude of the velocity of a body after freely


falling over the distance h is
v = V v20+ 2gh
or, when v0 = 0,
v =
If the coordinate axis OY is directed vertically downward,
then the magnitude of the velocity of a body at an arbitrary
point of its trajectory with the coordinate y is

v = V v l + 2s ( y —yo)
5°. The duration At of free fall without any initial velo­
city (v o = 0) from the height h is

8. Motion of a Body Thrown Vertically Upward


1°. A body is moving vertically upward with the initial
velocity v 0. If the air resistance is disregarded, then the
acceleration of the body equals that of free fall g (1.2.8.4°).
On the section up to the maximum height, the motion of
36 Mechanics 1J.8

the body is uniformly decelerated, and after this point is


reached it is free fall without any initial velocity (1.1.7.1°).
2°. The velocity of a body at the arbitrary moment t
from the beginning of motion, regardless of whether we are
considering only the ascent of the body or its descent after
the maximum height has been reached, is
v = v0 + gt
Problem. A body is thrown vertically upward with an
initial velocity of 20 m/s. Determine the velocities of the
body \ 1 and v 2 at the moments tx = 1.5 s and t2 = 3.2 s
from the beginning of motion. Consider that the coordinate
axis OY is directed vertically upward.
Given: v0 = 20 m/s, tx = 1.5 s, t2 = 3.2 s.
Required: v2, v 2.
Solution: The expressions for the projeclions vly and
v2y of the required velocities are:
V\y == Vq and v%y = Vq g ^2
whence
yly = 20 —9.8 x 1.5 « 5 m/s
y2y = 2 0 - 9 .8 x 3.2 11 m/s
The signs of the projections vly and v2y indicate that
the velocity vector v 1 will be directed vertically upward
(the body has not yet reached its maximum height), and the
vector v 2 vertically downward (at the moment t2 the body
moves downward).
3°. The displacement vector Ar of a body during the
arbitrary time interval At = t — t0 provided that = 0

Ar = Vo* +

This equation holds for any moment t when a body is rising


or falling after reaching its maximum height.
4°. At the moment tc corresponding to the maximum
height of a body over the point from which it was thrown
(when y = y mSLX or the height h = ftmax = yraax— y0)> we have
vc = v0 — gtc = 0
1.1.8 Kinematics 37

whence

At this moment, the direction of motion of the body is rever­


sed.
The maximum height of the body over the point from
which it was thrown is

^max = ^max Vo ~ “9“


Problem. A body is thrown vertically upward from the
distance = 1.5 m above the ground at the edge of a pit
having a depth of h = 3.5 m. The
initial velocity of the body is 2.3
m/s. Determine the moment of time
tf from the beginning of motion (/0=
= 0) at which the body will reach
the bottom of the pit, and find the
distance S travelled by the body dur­
ing this time.
Given: h0 = 1.5 m, h = 3.5 m,
v0 — 2.3 m/s.
Required: tt, S.
Solution: Let us make the origin
of coordinates 0 coincide with the
point from which the body was
thrown and direct the coordinate axis Fig* 1-1*21
OY vertically upward (Fig. 1.1.21).
The time dependence of the coordinate y of the body will
have the form
y = y o + v ^ —-y -
where y 0 is the initial coordinate.
Assuming in this equation that t = Jf, y 0 = 0, and
y = — (h0 + h), we have

— (ho 4" h) = Vfjft---- 2~


whence
— 2vqtf — 2 (h,Q-f- ft) = 0
38 Mechanics 1.1.9

and
, _ V , ± V v l + 2 g ( h 0+ h )
tf' 1,2 ~ ------------ i ------------
The root
. _ Up — Y ~l~ (hp + h)
*’2 g
in the given conditions has no physical meaning because
we see that ttf 2 < 0. Thus,
_ V p + V v l + 2 g ( h 0+ h ) _
i t - f t . i --------------- ----------------
_ 2.3+1^(2.3)* + 2 X 9.8 (1.5+3.5) _ 4 0 „
--------------------- o ---------------------- S
The distance S travelled by the body is (Fig. 1.1.21)
S = 2hmax + h0-f- h
where femax is the maximum height of the body over the
point from which it was thrown. We know that hmax =
= yJ/2g. Hence,
S= 2 Jr h 0-\-h = -\~h0~{~h =

= J | | ! + 1.5 + 3 ,5 « 5 .5 m
5°. The magnitude of the velocity vr of a body at the
moment when it returns to its initial point after moving
vertically upward and downward equals the magnitude of
the initial velocity of the body: vT = v0. The vectors vr
and v 0 have opposite directions.
The time it takes a body to travel from the initial point
to the highest one (Ath) is the same as that which it spends
from the highest point to the initial one (A£r)*
Afh = AJr = -y-

9. Uniform Circular Motion of a Particle


1°. Circular motion is the simplest example of curvi­
linear motion.
The velocity v of circular motion is defined as the linear
{circular) velocity. In uniform circular motion, the magnitude
Kinematics 39

v of the instantaneous velocity of a point particle (1.1.3.2°)


does not change with time: u = const (vA = vb = vCj in
Fig. 1.1.22). A moving particle traverses arcs of the circle
that are equal in length in equal intervals.
Tangential acceleration (1.1.4.3°) is absent in uniform
circular motion of a particle (ax = 0).

Fig. 1.1.22 Fig. 1.1.23

The change in the direction of the velocity vector v is


characterized by the normal acceleration an (1.1.4.3°), also
called the centripetal acceleration. At each point of the trajec­
tory, the vector an is directed along a radius to the centre
of the circle (Fig. 1.1.22), and its magnitude is

where R is the radius of the circle.


2°. In describing mechanical motion, particularly circu­
lar motion, a polar coordinate system is used apart from the
Cartesian (rectangular) coordinate system (1.1.1.3°). The
position of the particle M on a plane (for instance XOY)
is determined by two polar coordinates (Fig. 1.1.23): the
magnitude r of the position (radius) vector r of the particle
and the angle cp, the angular coordinate, or the polar angle.
The angle (p is measured from the axis OX counterclockwise
up to the position vector r. The point O here is defined as
the pole of the coordinate system.
3°. Let us make the pole of the coordinate system coincide
with the centre of the circle along which a point particle is
moving. As a result, r = R (Fig. 1.1.24), and the change in
40 Mechanics 1.1.9

the position of the particle on the circle can be characterized


by the change A<p in the angular coordinate of the particle
(within the limits of a change in the angle from 0 to 2jt):
A(p = qp2 — <Pi

The angle A(p is defined as the angle of rotation of the posi­


tion vector of the particle.

4°. The average angular velocity of a particle moving in


a circular orbit around a given centre (or axis) is the quanti­
ty <°av equal to the ratio of the angle of rotation Acp of the
position vector of the particle during the interval At to
the length of this interval:

5°. The angular velocity (instantaneous angular velocity)


o) is the limit which the average angular velocity tends to
(Par. 4°) as the time interval At tends to zero:

0) = lim (oav = lim


A*-0 Af-*0 M
6°. In uniform circular motion of a particle, the angles
of rotation of its position vector are the same during any
equal time intervals. Hence, in such motion, the instanta­
neous angular velocity equals the average angular velocity:
0) = CDgy.
1.1.10 Kinematics 41

The angle of rotation A(p of the position vector of a


particle performing uniform circular motion is
Aqp= coAf
7°. The time interval T during which a particle completes
one full revolution around a circle is defined as the period
of revolution, and the reciprocal of the period

is defined as the frequency of revolution.


The angle of rotation of the position vector of a particle
during one period is 2n radians, therefore 2n = coT7, whence
T = 2ji/cd, or
2n 0
co = — = 2jtv

8°. The distance S travelled by a particle in uniform


circular motion during the interval At = t — t0l when
t0 = 0, is
S = vt
9°. The distance travelled by a particle during one period
in a circular orbit of radius R is 2nR, and the angle of rota­
tion of the position vector of the particle during the same
interval is 2n rad, i.e. 2jii? = vT and 2n = coT. Hence,
we find the relationship between the linear and angular
velocities:
v = coR

10. Motion of a Body Thrown


at an Angle with the Horizontal
1°. If the initial velocity v0 is imparted to a body at
the angle a with the horizontal (—Jt/2 < a < jt/2 rad),
then its motion will be curvilinear. Provided that h < REj
where h is the distance from the body to the Earth’s surface,
and R e is the radius of the Earth at the given location, and
disregarding the air resistance, we can consider that the
trajectory of the body is a parabola, for instance, in the
plane XOY. The motion will be uniformly changing: the
42 Mechanics L1.10

acceleration of the body is constant and at any moment


of time equals the acceleration of free fall g (Fig. 1.1.25).
2°. It is convenient to analyse this motion in two iner­
tial reference frames (1.2.1.3°): in the fixed one XOY asso­
ciated with the Earth, and in the moving one X'O'Y'
(Fig. 1.1.26). At the initial moment t0 = 0, the origins of

coordinates 0 and O' coincide. Further, the moving refer­


ence frame uniformly travels along the axis OX of the fixed
frame with the velocity vh, and it is simple to see that
vh = v0 cos a.
The body in the moving reference frame has uniformly
changing rectilinear motion with the initial velocity v ov
(i>ov = u0 sin a) and the acceleration g.
3°. The velocity of the body v in the fixed reference frame
at the arbitrary moment t from the beginning of motion
is (1.2.7.2°):
v = vv + vh, or v = vov + g* + vh
If the positive senses of the coordinate axes OX and OY
are chosen as shown in Fig. 1.1.26, then
ux = vo cos a, vy = p0’sin OL— gt
v = V vl + vl = V (vo COS a)2+ (y0 s i n a — gt)2
The angle p of the velocity vector v with the horizontal
(Fig. 1.1.27) is
v0 sin a — gt
P —arc tan
v0 cos a
4°. With the same directions of the coordinate axes
OX and OY as in Paragraph 3°, for a body thrown horizon­
1.1.10 Kinematics 43

tally with the initial velocity v 0 (a = 0), we have


V = V0 + g * , Vx = Vo, V y = — g t ,

v = Y v20+ (gt)2 and p = arc tan

5°. The displacements of a body thrown at the angle a


with the horizontal along the axes OX and OY of a fixed

Fig. 1.1.27

inertial reference frame (see Fig. 1.2.26) during the interval


At = t — t0 when t0 = 0 are
Arx = v0t cos a and Ary = u0t sin a —

The coordinates x and y of the body in a fixed inertial


reference frame in the same conditions are
0^2
x = v0t cos a and y = v0t sin a —
If #0 =t^ 0 and y0 0, then
0^2
x = -x 0~\-v0t cos a and y = y0+ v0t s i n a - y
6°. The moment £c corresponding to the maximum' height
of a body over the point from which it was thrown when
0 < a < n/2 rad is determined from the condition v y = 0,
i.e. (Par. 3°)
v0 sin a — gtc = 0
whence
* __ vosiRa
te~~r~
The maximum height hmax of a body over the point from
which it was thrown is
v% sin 2 a
^TOax = y max £/o
44 Mechanics 1.110

Problem. A body is thrown at the angle a = j i /6 rad


with the horizontal from a position with the coordinate
y0 = 5.0 m above the ground (Fig. 1.1.28). The initial
velocity of the body is 10 m/s. Determine the coordinate
z/max of the maximum height of the body, the coordinate
xs of the point where the body strikes the ground, and the
velocity vs at this point.
Given: a = jt/6 rad, y0=
= 5.0 m, and v0 = 10 m/s.
Required: ymax, xs, v8.
Solution: The coordinate
of the highest point of the
body’s trajectory in the sys­
tem of coordinates XOY
shown in Fig. 1.1.28 is
v% sina a
£/max —yo 2g
(10)2 X (0.50)2 ^
= 5.0- 6.3 m
2X9.8 "
To determine xs from the condition y8 = 0, we can
find the time ts during which the body travels until it strikes
the ground:
gt*
0 = y0+ vots s m a ---£-
whence
t>0 3 i n g + V ^ i ) j s i n a g + 2 g y 0 _
S
_ 10 X 0 .5 0 + / (10)2 X (0.50)a+ 2 X 9.8 X 5.0 _ „ a
-------------------------- O -D s
The second root of the equation has no physical meaning
in the given conditions. Consequently,
x s = v oh cos o = 10 X 1.6 X 0 .8 7 « 1 4 m
The final velocity of the body vs is

vs = y (v0 cos a)2+ (v0 sin a — gts)2-


= |/(1 0 x 0.87)2 + (10 x 0 .5 0 -9 .8 X 1.6)* « 14 m/s
L l.ll Kinematics 45

and the angle p between the axis OX and the vector v8


is
*10X0.50—9.8X1.6
P = arc tan - v0 cos a
=arc tan 10 X 0.87
-5.4 rad

11, Rotation of a Perfectly Rigid Body


About a Fixed Axis
1°. To describe kinematically the rotation of a perfectly
rigid body about a fixed axis, the same quantities (and the
equations relating them) are used as to describe the circular
motion of a particle: the angular
coordinate of a point of the body (q>),
the angle of rotation of the position
vector r of a point of the body (Acp),
the average and instantaneous angular
velocities (coay and oo), and the linear
velocities of various points of the
body (v). The interval T during which
the body completes one revolution
about its axis is called the period of
revolution, and the quantity v that is
the reciprocal of the period, the fre­
quency of revolution.
2°. When a perfectly rigid body
rotates about a fixed axis, the angles
of rotation of the position vectors of
various points of the body during the interval A* are the same
(in Fig. 1.1.29 we have A<px = Acp2). The angle of rotation
A(p, the average coav and instantaneous co angular veloci­
ties characterize the rotation of the entire rigid body as
a whole.
3°. The linear velocity of a point of a rigid body is pro­
portional to the distance R from the point to the axis of
rotation:
v = ©i? = 2nvi? = -Tp R
4°. In uniform rotation of a rigid body, the angles of
rotation of the body during any equal intervals of time are
46 Mechanics 1.1.11

the same (Acp = const), and the instantaneous angular


velocity of the body equals the average angular velocity
(co = coav). The various points of a rigid body have no
tangential accelerations aT (1.1.4.3°) (at = 0 ), while the
normal (centripetal) acceleration an (1.1.4.3° and 1.1.9.1°)
of a point of the body depends on its distance R from the
axis of rotation:

aa = 4 = = R
The vector an at each moment is directed along the radius
of the trajectory of the point toward the axis of rotation.
5°. In non-uniform, or variable, rotation of a rigid body,
the angles of rotation of the body during equal time inter­
vals are not equal. The angular velocity co of the body
changes with time.
6°. The average angular acceleration a av in the interval
At = t2 — tx is defined as the physical quantity equal to
the ratio of the change in the angular velocity Aco = co2 —
— coj of a rotating body during the interval At to the length
of this interval:
Ao

If the angular velocity changes identically during arbi­


trarily equal intervals of time (Aco12 = A(o34, etc.), then
a av — const (uniformly changing rotation).
7°. The angular acceleration (instantaneous angular accel­
eration) of a rotating body at the moment t is defined as the
quantity a equal to the limit which the average angular
acceleration tends to during the interval from t to t + At
as At tends to zero:

a = lim a aT = lim
A f-0 A*-*0 m

The angular acceleration equals the limit of the ratio


of the elementary change in the angular velocity Aco to the
elementary time interval At.
8°. When the angular velocity of a body grows, the rota­
tion is called accelerated, and when the angular velocity
diminishes, it is called decelerated.
Kinematics 47

In uniformly changing rotation (Par. 6°), the instanta­


neous acceleration of a body remains constant and coincides
with the average angular acceleration: a = a av = const.
9°. The change in the angular velocity Aco of a perfectly
rigid body during the interval At = t — t0 in uniformly
changing rotation with the angular acceleration a is
Aco = a A* = a (t —10)
If the initial angular velocity of a body equals co0 at t0 = 0,
then at the arbitrary moment t the angular velocity of the
body is
(0 = C00“f- Q*t
When o)0 = 0, the angular velocity of a body at an arbitra­
ry moment is
co = a t
10°. The angle of rotation Acp of a body about its axis
during the interval At = t — t0 in uniformly changing
motion is
a
Acp= , a (AO2
cd0 a * h— y -

When t0 = 0, we have
at2
Acp = (o0; 2
When co0 = 0 at t0 = 0, then
a at2
Chapter 2

Dynamics of Motion
of a Point Particle

1. Newton’s First Law


1°. Dynamics deals with the influence of interactions
between bodies on their mechanical motion (1.1.1.1°).
The fundamental task of dynamics consists in determining
the position of a body at an arbitrary moment according
to the known initial position of the body, its initial veloci­
ty, and the forces (1.2.2.1°) acting on the body.
2°. A body on which no other bodies or fields act is de­
fined as a free {isolated) body.
When solving some problems, a body may be considered
free if external forces act on it, but they balance one another
(1.4.1.2°).
Similarly, a point particle is considered to be free (iso­
lated) if no forces act on it or such forces balance one another.
When studying the translational motion of a rigid body,
the motion of its centre of inertia (1.2.3.4°) is considered.
3°. Newton's first law: every point particle continues in
its state of rest or of uniform motion in a straight line unless
it is compelled by forces to change that state.
Newton’s first law establishes the fact of the existence
of inertial reference frames and describes the nature of the
motion of a free point particle in an inertial reference frame.
Reference frames in which a free point particle is in its state
of rest or of uniform motion in a straight line are defined
as inertial reference frames (or Newtonian reference frames).
The uniform motion of a free point particle in a straight
line in an inertial reference frame is called inertial motion
{motion by inertia). In inertial motion, the velocity vector of
a point particle changes in time neither in direction nor in
magnitude (v = const). The state of rest of a particle is a
particular case of inertial motion (v = const = 0).
1.2.1 Dynamics of Motion of a Point Particle 49

4°. An inertial reference frame must be associated with


an inertially moving reference body. When choosing such a
body in the conditions of a concrete problem, one must assess
the external actions on it if present, the fact of compensa­
tion of these actions or the possibility of disregarding them,
and also the nature of the
body’s motion in the given
conditions.
5°. To describe many ( a )
mechanical motions near V/MW.
the Earth, an inertial re­
ference frame is associated
with the Earth (a geocen­ Id -
^___________v a
tric reference frame). Here
we disregard the rotation
of the Earth about its own
axis and about the Sun.
Newton’s first law is
observed more strictly in
a heliocentric reference frame. (c)
The origin of coordinates
of this frame is made to Fig. 1.2.1
coincide with the Sun’s
centre, and the coordinate
axes are directed toward definite stars, which can be
assumed to be fixed.
6°. Reference frames in which a free point particle or
free body does not retain a constant velocity (non-inertial
motion) are defined as non-inertial reference frames.
A reference frame travelling with acceleration relative
to an inertial reference frame is a non-inertial one. In a
non-inertial frame, even a free body can perform non-iner­
tial motion, i.e. travel with acceleration.
Example. A stationary cart carries a vessel with water
in which a wooden bar floats (Fig. 1.2.la). Describe the
behaviour of the bar in accelerated rectilinear motion of the
cart to the right using two reference frames: (1) a stationary
inertial frame associated with the surface over which the
cart travels (the coordinate axes OX and OY of this frame are
shown in Fig. 1.2.lft), and (2) a non-inertial reference frame
50 Mechanics 1.2.2

associated with the accelerating cart (axes O'X' and


O'Y' in Fig. 1.2.1c).
The bar can be considered as a free body because the
force of gravity (1.2.8.3°) of the bar is balanced by the
buoyant force (1.6.2.3°), while all the other actions on the
bar may be ignored.
It is known from experiments that when the cart moves
in this way the bar will approach the left wall of the vessel.
In the first case, the behaviour of the bar is interpreted
on the basis of Newton’s first law: the free bar continues in
its state of rest (its unchanged position in the coordinate
system XO Y), whereas the cart together with the vessel
travels to the right (the left side of the vessel approaches
the bar with acceleration).
In the second case, the bar moves with acceleration
(non-inertially) to the left without any actions whatsoever
on it in this direction, while the cart with the vessel is at
rest in the coordinate system X'O'Y'. Here Newton’s first
law is not observed for the bar (the bar performs non-inertial
motion although it may be considered as a free body).

2. Force
1°. Force is a vector physical quantity that is a measure
of the mechanical action exerted on a point particle or a
body by other bodies or fields. A force is defined completely
if its magnitude, direction, and point of application are given.
The straight line along which a force is directed is called
the line of action of the force.
The action of a force results in a given body changing
the velocity of its motion (it acquires acceleration) or deform­
ing (11.7.2.1°). Forces are measured (1.2.11.2°,6°) on
the basis of ti ese experimental facts.
2°. The various interactions known in modern physics can
be classified under four headings:
(a) gravitational interaction appearing between all bodies
in accordance with the law of universal gravitation (1.2.8.1°);
(b) electromagnetic interaction—between bodies or par­
ticles having electric charges (111.1.3.1°, VI.5.2.6°);
1.2.2________ Dynamics of Motion of a Point Particle__________ 51

(c) strong interaction existing, for example, between the


particles which atomic nuclei consist of, and also between
mesons and hyperons (VI.4.3.5°)* and
(d) weak interaction characterizing, for example, the
processes of transformation of some elementary particles
(VI.5.2.7°).
Force as a quantitative characteristic allows us to assess
only gravitational and electromagnetic interactions. In the
exceedingly small regions of space and in the processes in
which strong and weak interactions manifest themselves,
such concepts as the point of application, the line of action,
and together with them the concept of force itself lose their
meaning.
3°. In problems of mechanics, gravitational forces (forces
of gravity) (1.2.8.1°) and two varieties of electromagnetic
forces—elastic forces (1.2.9.1°) and friction forces (1.2.10.1°)—
are taken into consideration.
4°. The forces of interaction between portions of a system
of bodies being considered are called internal forces.
The forces exerted on bodies of a given system by bodies
not included in this system are called external forces.
5°. A system of bodies on each of which no external
forces act is called a closed (isolated) system.
6°. If several forces act simultaneously on a point par­
ticle (Fj, F2, . . ., Fn), they can be replaced by one force
Fs called the resultant force and equal to their sum:

The projections of the resultant force onto the axes of a


Cartesian coordinate system equal the algebraic sums of the
corresponding projections of all the forces:

i= i
Fix* = 1=1
2 ^*2/* i= l
Fiz
Example. The forces Fx, F2, and F3 act on the particle
M in the plane XO Y (Fig. 1.2.2a). Their resultant Fx
can be found as the closing side of a polygon constructed
on the forces Fx, F2, and F3 as on its sides (the polygon of
vectors) (Fig. 1.2.26). The projections Fx* and Fx„ of the
52 Mechanics 1.2.3

Fig. 1.2.2

resultant force onto the coordinate axes OX and OY are


Fz,x = Fix ~b F2x + ^3X and Fzy = Fiy + F2y + F3y
The magnitude F^ of the resultant force is
F* = V f ix + K y

3. Mass and Momentum. Density


1°. The property of a body to retain its velocity in the
absence of interaction with other bodies is defined as inertia.
The physical quantity that is a measure of the inertia of a
point particle or a measure of the inertia of a body in trans­
lational motion is called the inert mass (mln).
2°. The mass characterizes still another property of bod­
ies—their ability to interact with other bodies in accor­
dance with the law of universal gravitation (1.2.8.1°). In
these cases, the mass is a measure of gravitation, and is
called the gravitational mass (mgT).
Modern physics has established the identity of the values
of the inert and gravitational masses of a given body with
a high degree of accuracy (mln = mgT). For this reason, they
are not distinguished from each other, and we speak simply
of the mass of a body (m).
For the measurement of the mass of a body see 1.2.11.2°,
3°, 7°.
3°. In Newtonian mechanics, it is considered that
1.2.3 D y n a m i c s of M o tio n of a P o in t P a rticle 53

(a) the mass of a body does not depend on its velocity;


(b) the mass of a body equals the sum of the masses of all
the particles which it consists of; and
(c) the law of conservation of mass is obeyed for a given
combination of masses, namely, the mass of a system of
bodies remains constant no matter what processes occur in it.
4°. The centre of mass {centre of inertia) of a system of
point particles is defined as the point whose position vector
re is determined by the expression
n
2 miri
2=1
y1 mi
2= 1

where mt = mass of the i-th particle of the system


= its position vector
n — number of particles.
The centre of mass is the point at which we can consider
that the mass of a body is concentrated upon its transla­
tional motion.
5°. The momentum p* of a point particle is the vector
quantity equal to the product of the mass mt of the particle
and the velocity of its motion:
Pi =
The momentum p of a system consisting of n particles
equals the sum of the momenta of all the particles of the
system:
n n
P= 2=1
2 Pi= 2=1
2 m*vi
or the product of the total mass of the particles of the sys-
n
lem m = mi and the velocity Vc of translational motion
2=1
of its centre of mass:
54 Mechanics 1.2.4

6°. The coordinates x, y, and z (or the position vector


r) and the velocity v of a point particle in the given refer­
ence frame determine the mechanical state of the particle.
When even one of these quantities changes, the particle
acquires a new mechanical state.
A quite definite momentum corresponds to each mecha­
nical state of a given particle in the reference frame being
used. This momentum depends neither on the processes that
caused the particle to acquire the given mechanical state
nor on its preceding or subsequent mechanical states. There­
fore, the momentum is a function of the mechanical state
of a particle.
Inertial motion (1.2.1.3°) of a point particle and inertial
translational motion of a body of finite dimensions are not
attended by changes in their momenta. Upon the non-iner-
tial motion (1.2.1.6°) of a body or particle, their momenta
change.
7°. By the average density of a body is meant the quantity
P av equal to the ratio between the mass m of a body and
its volume:
_ m
P ay — ~

The density p of a body at a given point equals the limit


of the ratio between the mass Am of an element of the body
selected in the vicinity of the given point and its volume
AV as AF tends to zero:
Am
p = lim aF
AV-0
If a body is homogeneous, then
m
P — Pav — y

4. Newton’s Second Law


1°. Newton's second law: the acceleration acquired by a
point particle in an inertial reference frame is directly pro­
portional to the force acting on the body, inversely pro­
portional to the mass of the particle, and coincides in direc-
1.2.4 Dynamics of Motion of a Point Particle 55

tion with the force:

a -i-
m
In this form, the law holds for a particle whose mass does
not change while the force acts. It also holds for the trans­
lational motion of a body of finite dimensions whose mass
is constant. In the latter case, m stands for the mass of the
body, and a for the acceleration of any point of the body,
for instance the acceleration of its centre of mass (1.2.3.4°).
The projections of the acceleration of a point particle or
of a translationally moving body of finite dimensions onto
the axes of a Cartesian coordinate system are expressed by
the relationships
1 X 1y and az ———
a* = m m * m
2°. The equation of Newton’s second law in a more gene­
ral form is
A (mv)
F = lim - tt- = lim M
A*-*0 A<-0
If the force F is constant, then
F = _Ap_ = A (m\)
At to

where At = time interval during which the force F acts


on a point particle or a body performing transla­
tional motion
Ap = A (m\) = change in the momentum of the
particle or body during this interval (the
quantity Ap is sometimes defined as the
impulse).
The force acting on a point particle (body) equals the
change in its momentum (or the impulse) in a unit time.
The force is a measure of the change in the momentum of
a particle (body) in a unit time.
Newton’s second law can be written in a different way
as follows:
F A* = Ap = A (m\)
56 Mechanics 1.2.4

The product of the force F and the time interval At during


which this force acts on a particle or body is defined as the
impulse and equals the change in the momentum of the
particle or body.
The impulse is a measure of the action of a force in time
and a measure of the change in the momentum of a particle
(body).
3°. Newton’s second law in a more general form also
holds when the mass m of a point particle (or of a transla-
tionally moving body of finite dimensions) changes not
only with time (as, for example, during the flight of a rocket),
but also as the velocity of the particle or body changes. This
occurs at high velocities approaching the velocity of light
in a vacuum (IV.4.2.1°).
We can conclude on the basis of Newton’s second law that
changes in the velocities of particles or bodies occur not
instantaneously, but during finite time intervals. For exam­
ple, it follows from the equation of Newton’s second law for
a body having a constant mass (F A* = m Av) that we can­
not have A* = 0 when Av =^= 0.
4°. If several forces act simultaneously on a point par­
ticle, then each of them imparts the same acceleration to
the particle (determined by Newton’s second law) as it
would in the absence of other forces. This forms the prin­
ciple of the independence of action of forces.
The resultant acceleration as acquired by a point of the
mass m when several forces act on it is determined accord­
ing to Newton’s second law:

where Fs is the resultant force (1.2.2.6°).


5°. The force or the resultant of several forces causing
a point particle (or body) to move along a circle is called
the centripetal force. It is directed toward the centre of the
circle.
6°. If several external forces whose vector sum is F^, ext
act on a system of point particles, then the acceleration
ac of the centre of mass of the system is determined accord-
1. 2.5 D yn a m ics of M o tio n of a P oin t P a rticle 57

ing to Newton’s second law:

ac

where m is the total mass of all the particles of the system.

5. Newton’s Third Law


1°. Newton's third law: the forces of interaction of two
point particles in an inertial reference frame are equal in
magnitude and opposite in direction:

Fto = — F hi

where Fih = force exerted on the i-th particle by the Zc-th one
Fhi = force exerted on the Zc-th particle by the i-th
one.
The minus sign indicates that the force vectors are direct­
ed oppositely.
Newton’s third law reflects the circumstance that inter­
acting point particles have equal rights. The forces Fik
and Fki are applied to different particles and mutually bala­
nce each other only when both particles belong to the same
perfectly rigid body.
2°. In connection with the circumstance that the velocity
of propagation of gravitational and electromagnetic inter­
actions is finite and equals the speed of light in a vacuum
(V.4.4.40), the application of Newton’s third law to point
particles or bodies that are not in direct contact is restrict­
ed. For example, if the Moon for some reason or other
were to suddenly pass over to a new orbit, the forces of inter­
action between the Moon and the Earth would change not
instantaneously, but after a certain time elapses. Within
the limits of this interval, Newton’s third law is not appli­
cable.
When assessing the “contact” interactions of particles or
bodies, such restrictions vanish.
58 Mechanics 1.2.6

6. Law of Conservation of Momentum


1°. If we consider a system consisting of n point particles
or n translationally moving perfectly rigid bodies in an
inertial reference frame, then on the basis of Newton’s
second law we have
A (MiVi) Fl2 + F l3+ • • • + Fln + Flt cxt
At

A (m2y2)
At
—F2i + F23 + • • • + F2n + F2t ext

A (y n) = Fnl + F„2 + . . . + FB(B_n + Fn>ext

where symbols such as Fih stand for the internal forces of


interaction of the bodies of the system, and F(, ext is the
resultant of the external forces applied to the i-th body of
the system. Addition of the left-hand and right-hand sides
of these equations yields
n
^ A (m(Vi) -p ,
2j ------ = ^2, int “1 * 2, ext
i= l
or
-^ - = F2f int + Fj:, ext

where Ap = change in the total momentum of the system


during the interval At
Fs, mt = sum of all the internal forces of interaction
of the parts of the system
Fz.ext = sum of all the external forces acting on the
bodies of the system.
Since on the basis of Newton’s third law we have Fik =
= —Fki, then F2tlnt = 0, and
Ap pi
"AT = * 2’ ext
i.e. the change in the total momentum of a system is deter­
mined by the sum of only the external forces.
1.2.6 D y n a m ic s of M o tio n of a P oin t P a rtic le 59

The changes in the projections of the momentum of a


system of bodies onto the axes of a Cartesian coordinate
system are determined by the equations
&Px
At = ext*
APy
At
Fz, ext

Ap z
At
= ^ 2, ext 2

2°. If a system is closed, then F^ext — 0 (because exter­


nal forces act on no body of the system), and
= 0, or Ap = 0, or p = const

This is the essence of the law of conservation of momentum


for a closed system of bodies: in an inertial reference frame,
the total momentum of a closed system of bodies does not
change with time.
The interaction between the bodies of a closed system
may result in a change in the momenta of separate bodies,
in the transfer of momentum from one body to another, but
this never affects the change in the total momentum of the
entire system.
Since the momentum of a system equals the product of
the mass m of the system and the velocity vq of its centre of
mass, i.e. p = m \c (1.2.3.5°), then for a closed system we
have
p = m \ c = const
whence
Vc = const
i.e. the centre of mass of a closed system of bodies in an
inertial reference frame moves rectilinearly and uniformly.
3°. The law of conservation of momentum may be used
in describing the behaviour of open systems of bodies in the
following particular cases:
(a) the external forces acting on any body of the system
are balanced (in this case Fs, Cxt = 0), and
(b) the projection of the sum of all the external forces
onto a coordinate axis equals zero. Here we speak about
6) M ech a n ics 1. 2.6

the law of conservation of the projection of the momentum of an


open system onto a given coordinate axis. For example, when
Fs, ext^ = 0, we have p x = const although the projections
of the momentum onto the other coordinate axes may change
with time.
Example 1. A system of two carts between which a com­
pressed spring is placed (Fig. 1.2.3a) is not closed because
external forces act on the carts: the forces of gravity Px
and P2 (1.2.8.3d) and the reaction forces Rxand R2. But if the

Vk Vi v

Fig. 1.2.3 Fig. 1.2.4

forces Px and R1? and also P2 and R2 are balanced, then the
law of conservation of momentum can be applied to the
system. In accordance with this law, for instance, the
position of the centre of mass of the system of carts will
remain unchanged in an inertial reference frame (XOY)
after the spring causes the carts to move (Fig. 1.2.36).
Example 2. After a shot from a gun on a carriage, the
latter recoils (Fig. 1.2.4). The velocity vx of recoil of the
carriage can be found from the law of conservation of the
projection of the momentum onto the coordinate axis OX.
The projection of the force of g r a v ity of the projectile P2
(which is an external force) that is unbalanced after the
shot onto the axis OX equals zero (P2* = 0).

i
1 .2 .7 Dynamics of Motion of a Point Particle 61

7. Galilean Principle of Relativity


1°. The positions of a point* particle in two arbitrary
inertial reference frames are related by the Galilean trans­
formation:
r ' - r —(r0+ ut)
where r and r' = position vectors of the particle in the
first and second reference frames
u = constant velocity of uniform and rectili­
near motion of the second reference
frame relative to the first one
r0 = position vector conducted from the
origin of coordinates of the first frame
to the origin of coordinates of the sec­
ond one at the moment t0 = 0.
Figure 1.2.5a shows the coordinate axes X YZ and
X'Y'Z' of two reference frames and a point particle M at

Fig. 1.2.5

the initial moment £o = 0, an(l Fig* 1-2.56 shows the axes


and the particle after the interval At = t — t0.
The relationship between the coordinates of the particle
in the two reference frames is described by the system of
equations
x’ = x — (x0 + uxt), y‘ = y — (y0+ uvt), and
z' = z — (z0+ u1t)
(the Galilean coordinate transformation).
62 Mechanics 1.2.7

It is considered that the time in all the inertial reference


frames flows identically:
At \ = At
(the absolute nature of time in Newtonian mechanics).
2°. The velocities of a point particle in the two reference
frames are related by the expression
v' = v — u
or in projections onto the coordinate axes
v'x = vx — uxi Vy = Vy — uyJ and v'z = vz — uz
(the velocity addition law in Newtonian mechanics).
3°. Owing to the constancy of the velocity of the second
frame relative to the first one (u = const), the accelerations
of a particle in both reference frames are the same:
a' = a
or
ax — ay —6^^, and az — flj
The acceleration of a particle is determined by the forces
acting on it in accordance with Newton’s second law and
does not depend on the velocity of the inertial reference
frame (the absolute nature of the acceleration in all inertial
reference frames).
4°. The forces of interaction between point particles
or bodies depend only on their relative arrangement or on
their relative velocities and do not depend on the velocity
of the inertial reference frame. For example, in any inertial
reference frame, the forces of gravitational interaction
(1.2.8.1°) of two particles are inversely proportional to the
square of the distance between them, and this distance will
be the same in all inertial reference frames regardless of
their velocity. Similarly, the viscous force (1.6.3.3°) depends
on the relative velocity of the contacting layers of fluids
that does not depend on the velocity of inertial motion of
the reference frame itself. The elastic force (1.2.9.1°) depends
on the extension or compression of a spring, but does not
depend on the velocity of the inertial reference frame in
which the experiment with the given spring is being conduct-
1.2.8 Dynamics of Motion of a Point Particle 63

ed. Consequently, F = F', where F is the force acting on


a given body in one inertial reference frame, and F' is
the force acting on the body in a different inertial reference
frame.
The mass of a body in Newtonian mechanics does not
depend on the reference frame it is being considered in
(m = m'), therefore in both inertial reference frames the
form of the laws of mechanics remains unchanged.
5°. The uniform motion of the reference frame in a straight
line does not affect the course of the mechanical phenomena
proceeding in this frame. In an inertial reference frame,
it is impossible to distinguish the state of rest from uniform
motion in a straight line. All inertial reference frames are
equivalent for any mechanical phenomena. These state­
ments express the mechanical (classical) principle of rela­
tivity (Galilean principle of relativity).
The principle of relativity is one of the most general
laws of nature because in the special theory of relativity
it also extends to non-mechanical phenomena (V.4.2.10).
6°. The acceleration of a given point particle, and also
the forces exerted on it by other particles or bodies do not
depend on the velocity of the inertial reference frame.
The position vector r of a particle (or the Cartesian coordi­
nates x , y, and z) and its velocity v are relative quantities.
For example, even at the initial moment (£0 = t'Q = 0),
their values for a given particle may be different in different
inertial reference frames. This is why the shape of the trajec­
tory of a particle is also relative (1.1.1.8°).

8. Forces of Gravity
1°. Law of universal gravitation: forces of mutual attrac­
tion (forces of gravity, or gravitational forces) act between
two point particles that are directly proportional to the
masses of these particles and inversely proportional to the
square of the distance between them. The magnitude of
the force of gravity is determined by the expression
64 Mechanics 1.2.8

where m1 and m.z = masses of the interacting particles


r — distance between them.
The proportionality constant G is defined as the gravita­
tional constant. It is determined experimentally and equals
the force of interaction of two point particles of unit masses
and at a unit distance from each other. In the SI system
(VI1.8), the gravitational con­
stant has the value
G = 6.6732 x lO '11 N .m 2/kg2
Forces of gravity are directed
along the line connecting the
interacting points and are there­
fore called central forces. .They
depend only on the coordinates
of the interacting particles. In
a gravitational field (Par. [7°)
that does not change with time
(a stationary gravitational field),
the work of the force of gravity
(1.5.2.2°) acting on a given mov­
ing particle depends only on the
coordinatesjof the initial and final positions of the particle
and is independent of the shape of its trajectory. Gravity
forces are therefore potential forces (1.5.2.1°).
2°. The law of universal gravitation in the above form
holds not only for two particles, but also for
(a) bodies of an arbitrary shape whose dimensions are
only a small fraction of the distances between the centres
of mass (1.2.3.4°) of the bodies, and
(b) bodies having a spherically symmetrical distribution
of their mass.
In these cases, r is the distance between the centres of
mass of the interacting bodies.
3°. A body at the point B on the Earth’s surface charac­
terized by the latitude (p (Fig. 1.2.6) is acted upon by two
forces: the force of gravity Fgr and the force of the reaction
of the Earth’s surface (or the force of the reaction of the
support) N whose direction is determined not only by the
force of gravity, but also by the rotation of the Earth. The
resultant F of these two forces ensures the rotation of the
1.2.8 Dynamics of Motion of a Point Particle G5

body along a circle with its centre at O' upon the diurnal
rotation of the Earth about its axis. The force P acting on
the body owing to its attraction #to the Earth and equal in
magnitude to the reaction force N but directed oppositely
to it is called the force of gravity. The latter can be measured,
for example, with the aid of a dynamometer provided that
the body and the dynamometer are at rest relative to the
Earth.
4°. In a reference frame associated with the Earth, any
unsupported body when moving from a small height h above
the Earth’s surface (h < i?E, where i?E is the Earth’s
radius) acquires the acceleration of free fall g (1.1.7.1°) under
the action of the gravity force P. This acceleration does not
depend on the mass m of the body and in accordance with
Newton’s second law (1.2.4.1°) can be expressed through
the gravity force P:

5°. The centre of gravity of a body of finite dimensions


is defined as the point relative to which the sum of the
moments of the forces (1.3.1.4°) of gravity of all the par­
ticles of the body equals zero. The force of gravity of the
body is applied at this point.
The centre of gravity of a body (or system of bodies)
generally coincides with the centre of mass of the body (or
system of bodies) (1.2.3.4°).
6°. By the weight of a body W is meant the force due to
the attraction of the rotating Earth which a bodly exerts on
a support or a suspension that prevents its free fall. The
weight is applied not to the body being considered itself,
but to its support or suspension. If the support (suspension)
is fixed relative to the Earth, then the weight equals the
force of gravity of the body. The weight and the force of
gravity are also equal in uniform and rectilinear motion of
the support (suspension) in a reference frame associated
with the Earth.
Example 1. In a reference frame associated with the
Earth, the cabin of a lift (elevator) moves with the accelera­
tion a directed vertically downward (Fig. 1.2.7). A body of
mass m is on the floor of the cabin. In accordance with
66________________________ Mechanics_______________________1.2.8

Newton’s second law, the body moves together with the


cabin of the lift with the acceleration a under the action of
two forces: the force of gravity P and the force of the reaction
of the cabin floor N. Hence,
ma = P + N
whence with account taken of the positive direction of the
axis O Y and the direction of the vectors a, P, and N, we
have
—ma = — P + N "
The magnitude of the force of reac­
tion of the cabin floor is
N = P — ma
or, since P = mg,
N = m (g — a)
The magnitude of the body’s
weight W equals the force of reac­
tion of the floor (W = N), but the direction of these forces
is opposite.
It follows from the last formula that when the lift cabin
moves with the acceleration a directed vertically downward:
(a) the weight of the body equals the force of gravity
if a = 0, i.e. when the cabin is at rest or is uniformly moving
in a straight line;
(b) the weight of the body is less than the force of grav­
ity when 0 < a < g; and
(c) the weight of the body vanishes when a = g. In this
case, we have a state of weightlessness {zero gravity, or impon­
derability).
Example 2. If the body together with the cabin of the
lift is moving with the acceleration a directed vertically
upward (Fig. 1.2.8), then from Newton’s second law we
have
N = m {g + a)
whence it follows that at any value of the acceleration
(a =^= 0) the weight of the body will be greater than its force
1.2.8 Dynamics of Motion of a Point Particle 67

of gravity (W > P). This is the state of overload (or simply


overload).
7°. Bodies interact gravitatioifally by means of a gravita­
tional field. This held together with other fields and substance
is one of the forms ofj! matter. A gravitational held is
inseparably linked with every body. It manifests itself
in that a particle placed in the held is acted upon by a gra­
vitational force proportional to the mass of the particle. The
body whose gravitational held is being
studied is called the source of this
held.
A force characteristic of a gravita­
tional held is its intensity I equal to the
ratio of the force of gravity Fgr acting
on a point particle placed in the held
to the mass m of this particle:
igr
1=

The intensity I of the gravitational


held at a given point of space equals
the acceleration of free fall g of a particle that is at this
point of space:
i = g
The magnitude of the intensity of the gravitational
field of a particle of mass M at a distance r from it is

/ = G—

This formula also holds when the source of the gravitational


held is a spherical body homogeneous in density. The dis­
tance r is measured from the centre of mass of the body, and
the radius of the body’s surface must be smaller than the
distance r.
8°. The orbital velocity is defined as the velocity vj that
must be imparted to a body for it to become a satellite of
the Earth (or another planet, and also of the Moon or another
massive celestial body) and move along a circle whose plane
passes through the centre of the Earth (Moon, etc.) and whose
centre coincides with that of the Earth (Moon, etc.) (for the
fi8 M e c h a n ic s 1.2.0

escape velocity, see 1.5.4.4°):

» i= j/ = 'Y Jak « 7.9 x 103 m/s

9. Elastic Forces
1°. The forces appearing in the elastic deformation of
bodies (11.7.2.3°) are defined as elastic forces. They act bet­
ween the contacting layers of the body being deformed, and
also at the place of contact of a
body being deformed with the
body causing its deformation.
For example, the elastic force
Fel acts from the side of the elas­
tically deformed board D on the
bar C lying on it (Fig. 1.2.9).
Elastic forces are of an elec­
Fig. 1.2.9 tromagnetic nature (111.1.3.1°).
2°. The majority of the prob­
lems in the elementary course
of physics deal with one-dimensional (linear) deformations
or strains of tension or compression. Here the elastic forces

Fig. 1.2.10

are directed along the line of action of the external (deform­


ing) force, i.e. along the axes of longitudinally deformed
filaments, helical springs, bars and the like, or at right angles
to the surfaces of the bodies in contact (see Fig. 1.2.9).
1.2 .9 Dynamics of Motion of a Point Particle 69

3°. The elastic force acting on a body considered in a


given problem from the side of the support or the suspension
is defined as the force of reaction of the support (suspension)
or the force of tension of the suspension. Figure 1.2.10 shows
examples of the application to bodies of the forces of reac­
tion of supports (the forces Nlt N2, N3, N4, and N3) and the
forces of tension of suspensions (the forces Tx, T2, T3, and
T*).
4°. Hooke's law for one-dimensional tension (or compres­
sion) characterized by the elongation (compression)\ vector
Al: the elastic force is propor­
tional to the elongation (comp­ W rm w i
ression) vector and is opposite
to it in direction (Fig. 1.2.11):
Fei = —/c Al I
where k is the quasi-elastic
force coefficient of a body (or Iw ra V 1 —
the spring constant for springs) Fixt
—a quantity determined by Fig. 1.2.11
the elastic force appearing
upon a unit deformation of the given body (or by an exter­
nal force resulting in a unit deformation of the body):

5°. The elongation (compression) vector of an elastic


body is related to the external force Fext by the expression
Al = I s s l
k
6°. The elastic force depends only on the change in the
distances between the interacting parts (particles, layers
or elements) of the given elastic body. The work of the
elastic force (1.5.2.3°) does not depend on the shape of the
trajectory and equals zero for movement along a closed
trajectory. Consequently, elastic forces are potential forces
([. 3.2.1°).
70 Mechanics 1.2.10

10. Forces of Friction


1°. External, or contact, friction is the interaction be­
tween different contacting bodies preventing their relative
displacement. For example, external friction exists be­
tween a block and the inclined plane on which the block
is lying or down which it is sliding. If friction manifests
itself between the parts of the same body, it is defined as
internal, or viscous, friction (1.6.3.3°).
2°. The friction between the surfaces of two contacting
solid bodies in the absence of a fluid layer between them
is called dry friction. The friction between the surface of a
solid body and the fluid medium surrounding it in which
the body moves is called liquid friction (see also 1.6.3.3°).
The force of friction Ffr directed along the surfaces of the
bodies in contact opposite to the velocity of their relative
displacement appears in all kinds of friction.
Dry friction is subdivided into:
(a) static friction—friction in the absence of relative
displacement of bodies in contact with each other; and
(b) sliding friction—friction with relative motion of
the bodies in contact.
3°. The force of friction Ffr, 0 preventing the appearance
of motion of; one body over the surface of another one is
called the force of static friction.
If a growing external force Fext is applied to a body in
contact with another body parallel to the plane of contact,
no motion of the body will be initiated when Fext changes
from zero to a certain value. This points to the ambiguity of
the force of static friction: when we try to bring the body
out of its state of rest, the force of static friction changes
from zero to its maximum value F™**.
The relative motion of a body begins when Fext > *.
The force is called the maximum force of static friction.
When we speak of the force of static friction, we usually
have in mind the maximum force of static friction.
4°. The force of static friction is due to the engagement
of irregularities on the surfaces of bodies, elastic deforma­
tions of these irregularities, and cohesion of the bodies
where the distances between their particles are small and
1.2.10 Dynamics of Motion of a Point Particle 71

sufficient for the appearance of intermolecular attraction


(11.1.4.1°). In this connection, the force of static friction
can be considered as a variety of the manifestation of elas­
tic forces (1.2.9.1°).
In approximate calculations, it is assumed that

Fftr* = foN, or
The force N (Fig. 1.2.12) acting on a given body from the
side of its support perpendicular to its surface is called the
force of normal reaction, and the
force Pn exerted by the body on
its support is called the force of
normal pressure. The dimension­
less coefficient of proportiona­
lity / 0 is called the coefficient of
static friction. It depends on the
material of the bodies in contact,
on the quality of machining of the Fig. 1.2.12
contacting surfaces, the presence
between them of foreign substances, and many other factors.
The values of the coefficient of static friction are obtained
experimentally.
5°. The force of sliding friction Ffrt sl between the sur­
faces of contacting bodies upon their relative motion depends
on the force of normal reaction N or on the force of normal
pressure Pn:
Ftr,s\ = fN , or F u , si = f P n
where / is the coefficient of sliding friction depending on the
same factors as the coefficient of static friction / 0, and also
on the velocity of relative motion of the contacting bodies.
The coefficient of sliding friction is determined experi­
mentally and in the majority of cases at low velocities of
relative motion of contacting bodies is less than the coeffi­
cient of static friction (/ < / 0).
6°. Forces of friction, unlike gravitational forces (1.2.8.1°)
and elastic forces (1.2.9.1°), do not depend on the coordi­
nates of the relative arrangement of the bodies. Forces of fric­
tion may depend on the relative velocities of the contacting
bodies. The work of the forces of sliding friction (1.5.2.5°)
72 Mechanics 1.2.11

depends on the shape of the trajectory of relative displace­


ment of the contacting bodies and does not equal zero when
the trajectory is closed. The forces of friction are non-po­
tential forces (1.5.2.4°).

11. Ways of Measuring Mass and Force


1°. Newton’s second law contains two independent quan­
tities—the mass m of a body and the force F—and can be used
for measuring only one of them. The other quantity, in
addition to the acceleration
a of the body, has to be
measured independently of
this law. To measure
masses and forces, we must
take advantage of some
II l l l e III file III Il l s
other physical laws in addi­
tion to Newton’s second
(a) (&) (c) law.
Fig. 1.2.13 There are different ways
of comparing forces and the
masses of bodies, but as soon as one of them is used as the
basis of measuring forces or masses, we have to arbitrarily
choose the mass of a certain body as the unit of mass (the
standard of mass) (VII. 1.1°) or a certain force as the unit of
force (the “standard force").
2°. One of the ways of comparing the masses of bodies
is based on the simultaneous use of Newton’s second and
third laws. If we ensure an elastic collision (1.5.4.1°) be­
tween two bodies suspended on long thin unstretchable and
weightless strings (Fig. 1.2.13a) (the sequential stages of
conducting such an experiment are shown in Fig. 1.2.136
and c), then on the basis of Newton’s second and third laws
we have
ma = m^a3, whence ws
= —
a

We find the ratio of the masses of the interacting bodies


according to the experimentally measured values of as and a.
If the mass of one of the bodies (for instance ms) is taken
1.2.11 D yn a m ics of M o tio n of a P o in t P a rticle 73

as a standard, then as a result of conducting a similar expe­


riment we can measure the mass m of any body:
m = ms 0'S
a

After this, the force F acting on a body of known mass


m is determined on the basis of Newton’s second law accord­
ing to the experimentally meassured acceleration a acquired
by the body under the action of the force F:
F = ma
A variety of this procedure is the compa­
rison and measurement of the masses of bodies
or particles on the basis of the law of con­
servation of the total momentum of a closed
system of interacting bodies or particles P
(1.2.6.2°). We use the experimentally meas­
ured changes in the velocities of the inter­ Fig. 1.2.14
acting bodies to determine the ratio of their
masses. If the mass of one of the bodies is known, we can
determine the mass of the other one.
3°. The simultaneous use of the law of universal gravita­
tion, Hooke’s law, and Newton’s second law is a more con­
venient way of comparing and measuring masses.
A body of mass m is suspended on a vertical stationary
helical spring at a point on the Earth’s surface (Fig. 1.2.14).
The condition of equilibrium of the body (1.4.2.2°):
P + Fel = 0
where P = force of gravity of the body
Fel = elastic force exerted by the spring on the body.
If we disregard the Earth’s rotation, then on the basis
of the law of universal gravitation (1.2.8.1°)
mM-g,
P = G -R * -

According to Hooke’s law (1.2.9.4°)


Fei = k AZ
where k = spring constant
AI = linear stretching of the spring.
74 Mechanics 1.2.11

Thus,

whence

where B = kR%/GMv is a constant quantity (when using


a given spring at a given point on the Earth) regardless
of what body is suspended on the spring.
The value of B changes when account is taken of the
rotation of the Earth, the altitude of the place of conducting
the experiment, and when a spring having a different spring
constant k is used, but it will remain the same when differ­
ent bodies are suspended on the spring.
The masses m and ms of two bodies suspended in turn
on the given spring are compared according to the experi­
mentally measured elongations of the spring AZ and AZs:
m = AIB and ms= AlsB
whence
m

The spring constant k does not have to be measured here.


If the mass of one of the bodies, for example ms, is
chosen as a standard, then the mass m of any body is

If the spring is provided with a pointer and a scale, then


the latter can be marked with graduations corresponding
to the elongations of the spring when bodies are suspended
on it whose masses are multiples or different fractions of
ms (graduation of the spring).
4°. A spring may be used for measuring the masses of
bodies only at thejpoint on the Earth where it was graduated.
The elongations of a given spring under the action of the
same body having the mass ms at different points will be
different:
rcis = AZSli?i and m5= AZS22?2
and AZs> AZS| because Bx =^= B%.
1.2.11 D yn a m ics of M o tio n o f a P o in t P a rtic le 75

5°. In this way of measuring masses, the force is mea­


sured on the basis of Newton’s second law
F = ma
Particularly, the force of gravity P of a body of known
mass m can be determined in this way:
P = rng
at the point on the Earth where the spring was graduated.
6°. An elastic spring is also used for comparing and
measuring forces. This is done on the basis of the law of
universal gravitation, Hooke’s law, and Newton’s second
law. To measure forces, a certain body is chosen whose force
of gravity is taken as the “standard force” (the unit of force).
When a body is in equilibrium on a stationary spring
(see Fig. 1.2.14), we have
P = k A1
where A1 is the spring elongation vector.
If bodies whose forces of gravity are P and Ps are at­
tached in turn to the end of the given spring, then
P = k AZ and Ps = k Als
whence
P = Al
P s A ls

Two forces can be compared according to the experimentally


measured elongations AI and AZs of the given spring.
If the force Ps is taken as the standard one, then the
force of gravity P of any body suspended on the given
spring is

A graduated spring intended for measuring forces is


called a spring dynamometer. The latter can be used to
measure not only the forces of gravity of bodies or vertically
directed forces, but also arbitrarily directed forces. Without
any additional recalculations, these results will hold only
for the point on the Earth at which the dynamometer was
graduatedr
76 Mechanics 1.2.11

When forces are measured with the aid of a spring dyna­


mometer, the mass is measured on the basis of Newton’s
second law according to the known force F and the experi­
mentally measured acceleration a acquired by the given
body under the action of this force:

7°. The measurement of the mass of a body with the aid


of a beam balance is based on the use of the law of universal
gravitation, Newton’s second and third laws, the condition
of equilibrium of a body hav­

n r 3
a
*3
ing an axis of rotation
(1.4.2.3°), and on the choice of
a standard for measuring mass.
The equilibrium of an
equal-arm beam balance (Fig.
W w* 1.2.15) is achieved provided
Fig. 1.2.15 that
w = w.
where W and Ws are the weights of the bodies on the pans.
The forces of gravity of the bodies P and Ps equal to their
weights are also the same: P = Ps, whence
mg = msg and m = ms
because the accelerations of free fall are the same at the place
where both bodies are.
By choosing a body of mass ms as a standard of mass
(weights are used for this purpose), we can measure the mass
m of any body.
The masses of the weights are indicated on them, and a
beam balance can be used to measure the mass m of a body
regardless of the place on the Earth where weighing is being
performed. To determine the force of gravity (or weight)
of a body with the aid of a beam balance and weights, we
must know the acceleration of free fall at the place of weigh­
ing.
1 .2 .1 2 D yn a m ics of M o tio n of a P o in t P a rticle 77

12*. Non-Inertial Reference Frames


1°. An elementary course of physics treats very simple
examples of noil-inertial reference frames moving in a
straight line. Newton’s laws are not obeyed in non-inertial
reference frames (1.2.1.6°). The state of rest or the motion of
point particles and bodies in non-inertial reference frames is
described by equations whose form is similar to the equa­
tion of Newton’s second law ___________ _
(1.2.4.1°), but forces of inertia
(inertial forces) are introduced ( a ) Qw.
into the equations.
The forces of inertia are not
caused by the action of other
bodies or fields on a given par­
ticle or body. The appearance
of inertial forces reflects the °nrf
non-inertial nature of motion
of the reference frame itself.
Otherwise, the forces of inertia
are characterized by the same
Fig. 1.2.16
features as conventional forces
in mechanics (their mag­
nitude, direction, and point of application). Particularly,
they are always proportional to the masses of bodies.
The forces of inertia applied to a system of particles or
bodies are always external ones. This violates the isolated
nature of the given system and results in the fact that the
law of conservation of momentum (1.2.6.2°) and the law of
conservation of mechanical energy (1.5.4.1°) are not obeyed
for it.
2°. When describing the translational motion of a body
in a non-inertial reference frame moving in a straight line,
the force of inertia F*n is introduced equal to
Fin — ^lanrf
where m = mass of the body
a nrf = acceleration of the given non-inertial reference
frame relative to an inertial one.
Example 1. A ball of mass m is on the smooth horizontal
floor of a wagon. The ball is fastened to the front wall of the
wagon with a spring (Fig. 1.2.16a). In Fig. 1.2.166, the system
78 Mechanics 1.2.13

XOY is a fixed inertial reference frame and X'O'Y' is a


non-inertial reference frame associated with the wagon. When
the latter moves with the acceleration a nrf relative to the
inertial reference frame, the spring will stretch by AZ, the
position of the ball will change, but it will be stationary
relative to the wagon. In the absence of the force of friction,
the state of rest of the ball relative to the wagon is described
by the equation
P + N + Fel + Fln = 0
where P = force of gravity of the ball
N = force of reaction of the floor
Fel = elastic force exerted on the ball by the stretched
spring
Fm = force of inertia.
We know that
P = N
Fm —Fei (a)
Since Fel = k AZ, the force of inertia can be measured
t according to the stretching
of the spring.
(a) The motion of the ball
together with the wagon rel­
ative to the inertial refer­
ence frame obeys Newton’s
Y% N second law
a nrf
P + N + Fei = manrf
while
F ei = manTt (b)
From Eqs. (a) and (b), we
Fig. 1.2.17 get
Fln = manTt
and with account taken of the directions of the vectors
Fln and anrf, we get
F,n= - m a nr,
Example 2. If in an experiment similar to that described
in Example 1 there is no spring (Fig. 1.2.17a), then when
the wagon moves with the acceleration anrf the ball
will move with the acceleration a' relative to the wagon
(.3 .1 Dynam ics of M o tio n of a P oin t P article 79

(Fig. 1.2.176). To describe this motion of the ball in a non-


inertial reference frame associated with the wagon, the
force of inertia Fln is introduced jso that in the absence of
friction the condition
P + N + Fln = ma'
is observed. Since P = — N, we have
Fln = ma'
The ball remains stationary relative to the inertial
reference frame (XOY) upon the accelerated motion of the
wagon because the balancing forces P and N do not impart
any acceleration to it. The behaviour of the sphere in the
inertial and non-inertial reference frames can be brought
into agreement provided that
a' + anr{ = 0, or a' = — anrf

whence it follows that


F,n= — manrf

Chapter 3*
Elem ents of Dynamics of Rotation
of a Perfectly Rigid Body
About a Fixed Axis

1. Moment of Force and Moment of Inertia


1°. The main task of dynamics of rotation is the deter­
mination of the angular coordinates of the points of a rotat­
ing body at any moment of time according to the known
initial angular coordinates, the angular velocities, and the
given moments of the external forces acting on the body.
80 Mechanics 1.3.1

2°. A perfectly rigid body having a fixed axis of rota­


tion does not change the angular velocity of its rotation
without the action of moments of external forces. In an
inertial reference frame, the body is either in a state of rest
(g) = 0) or rotates with a constant angular velocity (co =
= const and a = 0).
The rotation of a perfectly rigid body in an inertial ref­
erence frame changes when moments of external forces
(Par. 4°) act on it. If, for example, the external force Fext
is applied to a body having the
axis of rotation O'O' at the point
O at the distance R from the axis
of rotation (Fig. 1.3.1), then its com­
ponent Fj, in a plane perpendicular
to the axis of rotation and perpen­
dicular to the radius R of the point
of application of the external force
causes a change in the rotation of
the body.
The components F|| and F0 in
the absence of forces of friction do
not affect the rotation of a perfectly
rigid body. The first of them, F|f,
which is parallel to the axis of rota-
tion, may induce only accelerated translational motion of
the body along the axis OZ. The second component, F0,
whose line of action intersects the axis of rotation, may
induce accelerated translational motion of the body together
with the axis of rotation along the coordinate axis OX.
3°. The arm of a force relative to an axis is defined as the
shortest distance d from the axis of rotation to the line of
action of the force. Figure 1.3.2 shows the arms dv d2, and
d3 of the forces Fx, F2, and F3 applied to a body at the
points 7, 2, and 3.
4°. By the moment of a force or torque is meant the
quantity M equal to
M = Fd
where F = magnitude of the force applied to a body
d = arm of this force relative to the given axis.
The total moment of several forces acting on a body
1.3.1 Dynam ics of R o t a t i o n of a R i g i d R o d y 81

equals the algebraic sum of the moments of all the forces


relative to the given axis:
n

The moments of forces rotating a body about a given axis


clockwise and counterclockwise are taken with different
signs. For example, the moments
of the forces Fx and F3^(see Fig.
1.3.2) are considered positive,
and the moment of the force F2
negative.
5°. The moment of inertia of
a point particle relative to a
given axis is defined as the scalar
quantity I t equal to the product
of the mass mt of the particle
and the square of its distance R\
from the axis:
Ii = miR\
The moment of inertia of a body relative to an axis is
defined as the quantity I equal to the sum of the moments of
inertia of all the n particles of the body:

/ = S miR\
i= 1
The calculation of this sum for bodies having an arbi­
trary shape is very complicated, and their moments of iner­
tia are determined experimentally.
The moment of inertia is a measure^of the inertness of a
body in rotation. It plays the same part as the mass in de­
scribing the motion of a body in a straight line. But if the
mass of a given body in problems in Newtonian mechanics
is considered to be constant, then the moment of inertia of
a given body depends on the position of the axis of rota­
tion.
6°. The moments *of inertia of selected homogeneous
bodies having a^ very simple shape arej as follows (m is the
mass of the body):
62 Mechanics } .3 i

(1) A solid sphere of radius R whose axis of rotation


passes through the centre of mass of the sphere:
I = 4D-m R 2
(2) A solid cylinder (disk) of radius R whose axis of
rotation coincides with the longitudinal axis of the cylinder
and passes through its centre of mass:

(3) A cylindrical shell (hoop) of radius R whose axis of


rotation coincides with the longitudinal axis of the cylin­
der and passes through its centre of mass:
I = mR2
(4) A thin straight rod of length I whose axis of rotation
is perpendicular to the longitudinal axis of the rod and
passes through its centre of mass:

(5) A thin straight rod of length I whose axis of rotation


is perpendicular to the longitudinal axis of the rod and
passes through the end of the rod:

t — T mp

2., F un d am en tal L aw of R otational D ynam ics


1°. The fundamental law of rotational dynamics: in an
inertial reference frame, the angular acceleration a acquired
by a body rotating relative to a fixed axis is proportional
to the total moment M ext of all the external forces acting
on the body and is inversely proportional to the moment of
inertia I of the body relative to the given axis:
a = Mext

2°. The angular momentum of a particle (the moment of


momentum of a particle) relative to a fixed axis is defined
1.3.2 D yn a m ics of R o ta tio n of a R ig id Body 83

as the quantity L t equal to the product of the moment of


inertia I t of the particle and the angular velocity co of its
motion about this axis: •
‘ Li = Iiio
For a particle, I t = mtRi and co = vJR i, therefore

L i = miR \ ^ = PiR i

where p t = ntiVi = magnitude of the momentum of the


particle (1.2.3.5°)
R t = distance- from it to the axis.
The angular momentum of a body (the moment of momentum
of a body) relative to a fixed axis is defined as the quantity L
equal to the sum of the angular momenta of all the n parti­
cles of the body relative to this axis:

L = S L i = S/*co
ir= 1 i= l
or
L = I co
where I is the moment of inertia of the body relative to the
given fixed axis.
3°. When using the angular momentum, the equation of
the fundamental law of rotational dynamics acquires the
form
ALt jk/r A (/co) -nr
- J f = M exU or ^ = ^ext

where AL = A (Ico) = change in the angular momentum of


the body during the interval At
Mtxt = total moment of all the external forces acting
on the body in the given inertial reference
frame.
In this form, the fundamental law of rotational dynamics
may be applied to a body whose moment of inertia changes
during motion, or to a system of bodies performing rotation
about a given fixed axis.
It follows from the fundamental law of rotational dyna­
mics that a change in the angular momentum (or in the
84 * Mechanics 1.4.1

angular velocity with a constant moment of inertia) cannot


occur instantaneously.
4°. The total moment 7l/ini of all the internal forces of
interaction between the parts of a body relative to its axis
of rotation always equals zero. The moments of the forces
of interaction of the parts of the body balance one another
in pairs and do not cause the angular momentum of the
body to change.

Chapter 4
Statics

1. A ddition and R esolution of Forces


A pplied to a P oint Particle
and a P erfectly R igid B ody
1°. Statics studies the equilibrium of point particles,
bodies, or systems of bodies.
2°. A system of several forces simultaneously acting on a
particle may be replaced by the resultant force Fs that can
be found according to the rule of the polygon (1.2.2.6°).
When finding the resultant of two forces, we also use the
rule of the parallelogram: the resultant force Fs equals the
diagonal of a parallelogram whose sides are the two forces
Fx and F2 being added (Fig. 1.4.1).
A system of forces is said to be balanced (in equilibrium)
if the resultant of this system equals zero (Fs = 0).
If the resultant of a system of forces acting on a particle
does not equal zero, then this system can be balanced by
applying to the particle a balancing force Fbai equal to
F b a l= — F s
3°. The point of application of a force to a perfectly rigid
body may be transferred along the ‘line of action of this
force (transfer of the point of application of a force in a per-
1.4.1 Statics 85

fectly rigid body). For example, the force acting on a rigid


body at'the point A (Fig. 1.4.2) may be applied at the point
B or at the point C, etc. Such a transfer is possible because
of the circumstance that the deformations of a rigid body
are disregarded when considering it.

Fig. 1.4.1 Fig. 1.4.2

When treating elastic deformations of a body, for in­


stance the stretching of an elastic spring, the point of appli­
cation of a force may not be transferred because the same
force applied at different points may cause different defor­
mations of the given body.
4°. Bodies restricting the motion of a given body being
considered are called constraints, and the forces exerted by
the constraints on a given body
are called the forces of reaction
of the constraints. Constraints
include, for instance, various
supports or suspensions (1.2.9.3°).
5°. The resultant Fj of sev­
eral forces Ff applied to a rigid
body when the lines of action of
these forces intersect at different
points (points 0 2, and 0 3 in
Fig. 1.4.3) is found either by
addition of the forces in pairs
(for instance, first the resultant
of the forces Fx and F2 is found, then they are added with
the force F3, etc.), or by addition of the projections of all
the n forces onto the coordinate axes of the chosen ref­
erence frame:

F 2x= 2 Fixi P zy— 2 Fiyi an(^ — 2 & iz


i= l i= l
86 Mechanics 1.4.1

whence
Fi. = V F\x + F\y + F \z
6°. The resultant of two parallel forces Fj and F2
(Fig. 1.4.4) equals their sum, while the line of its action

divides the distance between the points 1 and 2 of applica­


tion of the forces in a ratio inversely proportional to the
magnitudes of the forces:
Fs = Fi + F2 and -
The resultant Fs of two antiparallel forces Fx and F2
(Fig. 1.4.5) equals the vector sum of the forces Fj and F2:
Fs = F1 + F2
It acts in the same direction as the great­
er of them, and the line of its action
intersects the continuation of the straight
line connecting the points 1 and 2 of
application of Fthe forces at the point O
for which the following equation is ob­
served:
Jl_=
h Pi
7°. A system of two antiparallel .
forces Fj and F2 equal in magnitude
(Fig. 1.4.6) is defined as a couple of forces (or simply a
couple). The resultant of a couple equals zero. A couple is
characterized by the moment of the couple M , and
M = Fid = F2d
1.4.1 Statics 87

where d is the shortest distance between the lines of action


of the forces forming the couple and is called the arm of
the couple. •
A couple imparts an angular acceleration (1.1.11.7°),
but cannot change the translational motionfof a body.

8°. The components of the force vector F are defined as


the vectors F* whose sum equals the given vector F:

§ Fj = F
i= 1

where n is the number of components.

The problem of the resolution of the vector F into two


components lying in one plane has unambiguous solutions
in two cases:
(a) one of the components is known. Figure 1.4.7 shows
at a definite scale the vector F and one of its components
I’V Making the tails of the vectors coincide at the point 0
(Fig. 1.4.8) and drawing through their tips the straight
line NNj we find the second component F2 as the length of
88 Mechanics 1.4.2

the segment NN confined between the tips A and B of the


given vectors and directed from the tip of the component Fx
to that of the vector F; and
(b) the directions of both component vectors of F are
known (lines'TVTV and M M in Fig. 1.4.9a). Conducting straight
lines N'N' and M 'M r parallel to the lines N N and MM
through the tail and the tip of the vector F, in the adopted
scale we find the components Fx and F2 as the sides^AB and
BC of the triangle ABC (Fig. 1.4.96).

2. Conditions of Equilibrium
of a Point Particle and Perfectly Rigid Body
in an Inertial Reference Frame
1°. The condition for the equilibrium of a point particle
is the equality to zero of the sum of all the forces F* acting
on the particle:

2F, = 0
«=1
or in projections onto the axes of a Cartesian coordinate
system:

= Yi Fiy = 0, and 3 ^ = 0
{=1 2=1 2= 1
where n is the number of forces acting on the given particle.
2°. If the constraints permit only translational motion
of a perfectly rigid body, it will be in equilibrium provided
that

S F , = 0,
i= l
or

2 ^ = 0, 2 / ^ = 0, and 2 ^ = 0
i= l i= l 1=1

where n is the number of forces acting on the body.


Here, all the forces Ff are considered to be applied at
the centre ofJ mass of the body regardless of the actual ar­
rangement of the points of application of these forces.
1.4.2 Statics 89

3°. A perfectly rigid body with a fixed (stationary) axis


of rotation is in equilibrium if the sum of all n moments M t
of the external forces relative tfl this axis equals zero (the
rule of moments):
2 a#i = o
t= l
4°. If a perfectly rigid body can move translationally
and also rotate about a certain axis, equilibrium of the body
is achieved when two conditions are observed simultaneou ly:

S F , = 0 and 2 Mt= 0
1=1 1=1
where F* = external force acting on the body
Mi = moment of this force
n = number of external forces.
Example 1. A block B can be in equilibrium on the
inclined plane D (Fig. 1.4.10) under the action of the force
of gravity P, [the force of frstatic
friction Ffr>0, and the force of reac­
tion N if the condition
P -T N -f- FfFto = 0
is observed and if, in addition, the
line of action of the force of reac­
tion N passes through the point K
Fig. 1.4.10
of intersection of the lines of action
of the forces P and Ffr.o- Only when
Ihe last condition is observed will the sum of the moments of
(lie forces P, N, and Ffr>0 relative to any axis (for instance
relative to the horizontal axis perpendicular to the plane
of the drawing and passing through the centre of mass C
or the point K) vanish. If the line of action of the force of
reaction N passed through the centre of mass C of the block,
then the moment of the force Ffr,9 relative to the axis pass­
ing, for instance, through the centre of mass C would not
equal zero, and as a result the block would have nothing
left to do but turn over clockwise about its edge A.
The general conditions of equilibrium of a body can be
used not only to find the lines of action of forces, but also
lo determine unknown forces.
90 Mechanics 1.4.2

Example 2. One end of a beam at the point B is supported


on a stationary cylindrical roller, and its other end is an­
chored in a wall (Fig. 1.4.11). The external force Q applied
vertically downward at the point E acts on the beam in
addition to the force of
gravity P. Find the force NB
of the reaction of the roller
and the force ND of the
reaction of the wall.
The line of action of the
force Nb can be only verti­
cal because otherwise the
P force exerted on the roller
Fig. 1.4.11
by the beam would have a
horizontal component that
would move the roller to
the right or the left. This would contradict the fact
that the roller and the left end of the beam are stationary.
If the forces P, Q, and NB are vertical, then the force of
reaction ND also cannot have a horizontal component, i.e.
this force should be directed along a vertical line.
From the condition of the equality to zero of the sum of
the four forces acting on the beam:
P + Q + Nb + Nd = 0
or
P y Jr Q y ~ \ - N B y - \ - N D y = Q
or
- P _ Q + n b + Nd = 0
it is impossible to find the two unknown quantities N b
and Nd .
The second equation can be obtained from the condition
that the sum of the moments of the forces acting on the
beam relative to any fixed axis of rotation equals zero. If
this axis is chosen so that it passes through the point of
application of one of the unknown forces, we can obtain an
equation in one unknown. Thus, if we write the equation for
the moments of the forces relative to the horizontal axis
passing through the point B and directed backward out of
the plane of the drawing, we get
— Ph — Q (^i + h) + (h + h + h) = 0
1.4.3 Statics 91

where lly l2 and l3 are the distances between the lines of


action of the forces indicated in Fig. 1.4.11. It is simple to
find N d from this equation, after which we find N b from
the preceding equation.

3. Kinds of Equilibrium
1°. The equilibrium of a body in a certain position is
ealled stable if forces or moments of forces appear that tend
to return the body to its initial position upon any small
deviations of the body from this position
permitted by the constraints (1.4.1.4°).
Examples. Figures 1.4.12 to 1.4.14
show the positions of stable equilibrium
of selected bodies and small deviations
of the bodies from these positions.
1. The ball A suspended on a string
(Fig. 1.4.12a) is in stable equilibrium
under the action of the force of gravity
P and the force of tension of the string
T so that when the ball is slightly devi­ Fig. 1.4.12
ated, for example to the right (Fig. 1.4.126)
an unbalanced force F appears that returns the ball to its
initial position.
2. The sleeve B (Fig. 1.4.13a) that can slide without
friction along a smooth horizontal rod and that is fastened
to an undeformed spring is in stable equilibrium under the
action of the force of gravity P and the force of reaction N
of the rod. When the sleeve is displaced slightly along the
rod (Fig. 1.4.136), the elastic force Fei appears that is direct­
ed toward the initial position of equilibrium of the sleeve.
3. The non-homogeneous sphere D (Fig. 1.4.14a) floats
in a liquid under the action of the force of gravity P applied
at the centre of mass C of the sphere and the buoyant force
Fb (1.6.2.3°) applied at the point 0 1 that is somewhat lower
than the geometrical centre of the sphere O. If the sphere is
submerged deeper into the liquid (Fig. 1.4.146), the buoy­
ant force grows, and the resultant of the forces P and Fm
makes the sphere return to its initial position. If the sphere
is turned about its centre O (Fig. 1.4.14c) a moment of the
02 Mechanics 1 .4 .3

couple P and Fb appears that returns the sphere to its ini­


tial position of stable equilibrium.
2°. The equilibrium of a body in a certain position is
called unstable if forces or moments of forces appear that
N

Fig. 1.4.13 Fig. 1.4.14

tend to deviate the body still more from its initial position
upon even small deviations of the body from this position
allowed by the constraints (1.4.1.4°).

Fig. 1.4.15 Fig. 1.4.16

Examples. Figures 1.4.15 to 1.4.17 show positions of


unstable equilibrium of selected bodies and slight deviations
of the bodies from these positions.
1. The small ball A is in equilibrium at the top point of a
spherical supporting surface (Fig. 1.4. 15a). Upon a slight
deviation of the ball (Fig. 1.4.156), the unbalanced force F
appears that moves the ball away from its initial position.
2. The non-homogeneous sphere D floats in a liquid
(Fig. 1.4.16a). The density (1.2.3.7°) of the hatched part of
1.4.3 Statics 93

the sphere is greater than that of its remaining part. The


force of gravity of the sphere P is applied at the centre of
mass C, while the buoyant force»Fb (1.6.2.3°) is applied at
the point Ox below the geometrical centre of the sphere 0 .
The equilibrium of the sphere is unstable because when it is
rotated slightly about its centre O a moment of the couple P
and Fb appears that results in the further deviation of the
sphere from its initial position (Fig. 1.4.166).
3. The ball B is on a horizontal surface and is pressed
by a stretched spring^against the projection D having a

( c)

height equal to the radius of the ball (Fig. 1.4.17a). The


forces Fel and N2, P and Nx are balanced in pairs. If the ball
is displaced horizontally to the right, the elastic force of
the spring will return it to its initial position. But this does
not yet point to stable equilibrium of the ball because a
body must return to its position of stable equilibrium after
any small deviations from this position. If the ball B is
displaced slightly upward (Fig. 1.4.176), then the moment of
the force Fe'i relative to the point of contact of the ball may
exceed the moment of the force of gravity bP, as a result of
which the ball will be in the position shown in Fig. 1.4.17c.
3°. The equilibrium of a body in a certain position is
called neutral, or indifferent, if no forces or moments of
forces appear that tend to return the body to its initial posi­
tion or move it still farther from its initial position upon
any small deviations of the body from this position allowed
by its constraints (1.4.1.4°).
Examples. Figure 1.4.18 shows^the positions of indiffer­
ent equilibrium of selected bodies.
1. The ball A on a horizontal surface (Fig. 1.4.18a) is in
indifferent equilibrium. Displacements of the ball over the
94 Mechanics 1.4.3

surface do not lead to the appearance of any other forces


apart from the balanced ones P and N.
2. The equilibrium of the block B in the position shown
in Fig. 1.4.186 will be indifferent if upon slight displacements
of the block to the right or the left from this position the
elastic force of the spring does not reach the value of the
maximum force of static friction (1.2.10.3°) (Fel < Ff?*o)-
3. The sphere D of a substance having the same density
as the liquid in which it is floating (Fig. 1.4.18c) is in in­
different equilibrium under the action of the two balanced

Fig. 1.4.18

forces P and Fb applied at the centre of mass C of the sphere.


Any displacements or rotations of the sphere will not be
followed by a return to its initial position or by a further
displacement from it.
4°. Assume that the equilibrium of a point particle, body,
or system of bodies is due to the action of only potential
forces (1.5.2.1°)—forces of gravity, elastic, or electrostatic
forces (111.1.2.2°). Consequently, the minimum value of
the potential energy (1.5.3.5°) corresponds to the position
of stable equilibrium in comparison with the values of this
energy at the closest adjacent positions permitted by the
constraints (the principle of the minimum potential energy).
The potential energy grows upon any slight deviations of a
particle, body, or system of bodies from the position of stable
equilibrium.
For example, when the ball A (see Fig. 1.4.12) deviates
from its position of stable equilibrium, its potential energy
in the Earth’s gravitational field grows. The minimum value
i.5 .1 . Work and Mechanical Energy 95

of the potential energy of elastic interactions in the given


system corresponds to the position of stable equilibrium of
the block B (see Fig. 1.4.13). When#the sphere D is rotated (see
Fig. 1.4.14c) its potential energy in the Earth’s gravitational
field grows, while when this sphere is submerged to a greater
depth (see Fig. 1.4.146) its potential energy in the Earth’s
gravitational field diminishes somewhat, but the potential
energy of its elastic interaction with the liquid grows.

Chapter 5
Work and Mechanical Energy

1. Work of a Force in the Motion


of a Particle and the Translational Motion
of a Perfectly Rigid Body
1°. By the elementary work AA of the force F over the
elementary displacement Ar of a point particle is meant the
scalar physical quantity equal to
AA = F Ar cos a
where a is the angle between the vectors F and Ar.
In the Cartesian (rectangular) system of coordinates, the
elementary work is
A^4 = Fx Ax + Fy Ay + Fz Az

where Fx, Fyi and Fz = projections of the force F


Ax, Ar/, and Az = projections of the elementary dis­
placement vector Ar of the par­
ticle onto the coordinate axes.
The value of the elementary work of a force depends on the
choice of the reference frame. In addition, depending on the
mutual orientation of the vectors F and Ar, the elementary
work may be positive, negative, or equal to zero.
Mechanics 1.5.1

2°. If a system of n forces acts on a point particle, then


the elementray work A^4E of all these forces when the par­
ticle is displaced by Ar is
n
A^42 = 2 ^iAr cos a*
i= l

where a t is the angle between the force F* and the displace­


ment Ar of the particle, or
A^4s = F%Ar cos a
where Fs = magnitude of the resultant F^ of all the forces
acting on the particle
a = angle between the vectors Fs and Ar.
3°. In the translational motion of a perfectly rigid body,
the elementary work of a force (or of the resultant of a sys­
tem of forces) is
AA = F Ar cos a
where a is the angle between the force F (or the resultant of a
system of several forces) and the elementary displacement
Ar of the centre of mass (1.2.3.4°) of the perfectly rigid body
or a different point of it.
4°. The work A of a force (or of the resultant of a system
of forces) F over a finite displacement Ar is
n
A = 2 F Art cos a*
<=i
where n = number of elementary displacements Ar f into which
the total displacement Ar has been divided
a t = angle between the force F and the elementary
displacement Arf.
In the general case, the calculation of such a sum is very
complicated. In the elementary course of physics, the cal­
culation of the work is restricted to finite displacements of a
particle or a perfectly rigid body for some particular cases:
(a) The force F is constant, and the trajectory of a par­
ticle (or body) is a straight line (F = const and a t = a =
= const) (Fig. 1.5.1). Here
n
A = F cos a Ar,-
i=i
1.5.1 Work and Mechanical Energy 9?

or
A = F\cos a Ar i2 == Fjzos a S l2
where Ar12 is the magnitude of the displacement vector ArJ2
of the particle (or body) from its initial position 1 to its
final position 2 equal in this case to the distance S12.
(b) The force F is constant, the trajectory of a particle
(or body) is curved (F — const, a* const) (Fig. 1.5.2).

Fig. 1.5.1 Fig. 1.5.2

The work of the force F upon the displacement of the particle


(or body) from the position 1 to the position 2 is
A = FAr Ar 12
where FAr is the projection of the force vector F onto the
direction of the total displacement Ar12. The quantity

FAr = F cos a; here a is the angle between the vectors F


and Ar12.
If a particle (or body) moves along a closed trajectory
(Fig. 1.5.3), then the total work A of the force F equals
zero because Ar12 = 0.
98 Mechanics 1.5.1

(c) The force is constant in magnitude and forms identi­


cal angles a with elementary displacement vectors Arf at
any point of the trajectory (F = const and a t = a = const).
In Fig. 1.5.4, we have Ft = Fh = Fm = F = const. Here
the work of a force is
n
A = F cos a Art = F cos a S i2

where S12 is the path of a particle or body from its initial


position 1 to its final position 2.

Fig. 1.5.5 Fig. 1.5.6

If in such cases the trajectory of a particle is an arbitrary


closed curve (Fig. 1.5.5), then the work of a force differs
from zero notwithstanding the fact that the total displace­
ment vector of the point of application of the force is zero:
Ari 2 = 0, but A =^= 0 because S12 ^ 0.
Some examples of calculating the work of a variable
force will be given below (1.5.2.2°, 3°).
5°. If we have a plot of the relationship F&r = f (Ar)
(Fig. 1.5.6), then the elementary work AA of a force is depict­
ed by the area of a curvilinear trapezoid with the base
Art. This area is hatched in Fig. 1.5.6. The total work A
when a point particle or perfectly rigid body travels from
the position 1 to the position 2 is depicted by the area of a
curvilinear trapezoid with the base Ar12.
1.5.2 Work and Mechanical Energy ftfl

2*. Potential and Non-Potential Forces.


Conservative and Non-Conservative
Systems of Bodies
1°. Forces are defined as potential when their work depends
only on the initial and final positions of a moving point par­
ticle or body and does not depend on the shape of their tra­
jectory. With a closed trajectory, the work of a potential
force always equals zero.
Potential forces include gravitational, elastic, and
electrostatic forces.
2°. The work A g of the gravitational force (1.2.8.1°) when
a point particle of mass m moves relative to another particle
of mass M and placed at the origin of coordinates is

As = G m M (-L — L )

where rx and r2 are the magnitudes of the position vectors


characterizing the initial and final positions of the moving
particle (Fig. 1.5.7).

When the distance between interacting particles grows,


the work of the gravitational force is negative (A g< 0 when
r2 > rj). When the particles approach each other, the work is
positive (^4g > 0 when r2 < rj). When one of the inter­
acting particles travels along a closed trajectory, the work of
the gravitational force equals zero (A g = 0 when r2 = r^.
If we disregard the rotation of the Earth, then the work
A gr of the force of gravity (1.2.8.3°) when a body of mass m
100 Mechanics t.5.i

is lifted to the altitude h is


A _ GrtiM^h
*r"“ ~ Re (*E+fc)
where M e and R e are the mass and the radius of the Earth,
respectively.
When h < the work of the force of gravity in
lifting a body of mass m is approximately equal to
Agr ~ — mgh
where g is the value of the acceleration of free fall near the
Earth’s surface (1.2.8.4°).
3°. The work ^4el of an elastic force (1.2.9.4°) in one­
dimensional tension (or compression) characterized by the
elongation (compression) vector A1 is
Aei= - F elM =
Should one of the coordinate axes (for instance OX) of
the selected reference frame coincide in direction with the
vector Al, then
>i _ l kx\ kx\ \
- \ ^ 2 ------- r )
where xx and x2 are the coordinates of the tail and tip of
the vector Al.
When a point of an elastically deformed body travels
along a closed trajectory, the work of the elastic force equ­
als zero (^4ei = 0 when Al = 0, or when x2 = xx).
4°. Forces whose work depends on the shape of the tra­
jectory are defined as non-potential. When a point particle or
body travels along a closed trajectory, the work of a non­
potential force does not equal zero.
Non-potential forces include friction forces and some
others (see for example 111.4.2.1°).
5°. In the majority of cases, when the angle between
the force of friction Ffr and the elementary displacement Ar of
a body is n radians, the work of the force of friction is nega­
tive and is
Afr = — FtrSl2
where S12 is a path of a body between points 1 and 2.
Sometimes, the angle between the force of friction Ffr
L5.3 W ork and M ech a n ica l E nergy 101

and the elementary displacement Ar equals zero, and the


work of the force of friction is positive:
Afr = Ffr5i2
Let us assume, for instance, that the external force F acts
on the block B that can slide along the cart D (Fig. 1.5.8).
If the cart moves to the right,
then the work of the force of
Ffr,1 B Ffr,2 F
sliding friction Ffrt2 exerted on
the cart by the block is positive. .___ =----- D

c a lle d L Z I Z i f t h S e r n a l
and external forces (1.2.2.4°) Fig. 1.5.8
acting on the bodies of the sys­
tem are potential ones (Par. 1°).
In a closed conservative system, only internal potential
forces act between the bodies.
If even one of conditions needed for a system to be
conservative is violated, then it is called non-conservative.
Internal non-potential forces act between the bodies of a
closed non-conservative system in addition to the internal
potential forces.
Sometimes, one body instead of a system of bodies is
considered as a conservative or non-conservative system.

3. M echanical E nergy
1°. By energy is meant a scalar physical quantity that is
a single measure of various forms of motion of matter and a
measure of the transition of motion of matter from some
forms to others. To characterize different forms of motion
of matter, the relevant kinds of energy are introduced, for
example mechanical energy (Par. 2°), internal energy
(1.5.4.2°, 11.4.1.2°), the energy of electrostatic (111.1.7.5°),
intranuclear (VI.4.2.2°) interactions, etc.
Energy obeys the law of conservation, which is one of
Uie most important laws of nature.
2°. The mechanical energy E characterizes the motion and
interaction of bodies and is a function of the velocities and
mutual arrangement of the bodies. It equals the sum of ttyp
kinetic and potential Ep energies.
102 Mechanics 1.5.3

3°. The kinetic energy of a point particle or body is a


measure of its mechanical motion depending on its velocity
in the given inertial reference frame.
The kinetic energy Z?k, t of a particle of mass mt and
moving in the given inertial reference frame with the velo­
city Vi, or of momentum p t = is
~ mivi Pi

The kinetic energy Ek of a system is the sum of the kine­


tic energies Ekt t of all the n particles which the system
consists of:
n n n
m t i>?
jL
£ k= 2 £ k,i = 2 2 = 2 2mt
i= 1 i= 1

In the translational motion of a body, its kinetic energy


equals half the product of its mass m and the square of the
velocity v of any of its points (for example the centre of
mass), or the square of the momentum p of the body divided
by its double mass:
j-j m v2 p2
^ = - = -2^
The values of the kinetic energy of a particle or body
depend on the choice of the reference frame, but cannot be
negative (Ek ^ 0).
4°. Theorem on the kinetic energy: the change AE k in
the kinetic energy of a body when it passes over from one
position to another equals the work A of all the forces acting
on the body:
A = A2?k = Ekt 2- £ k)1
where i?kt x and Eki 2 are the kinetic energies of the body in
its initial and final positions, respectively.
The work of any forces is a measure of the change in the
kinetic energy of a body or a point particle. The action of
forces whose work on a given portion of a trajectory is posi­
tive leads to an increase in the kinetic energy of the body
(AZ?k > 0, or E kt2 > /?k,i). The action of forces whose
work is negative leads to a reduction in the kinetic energy
pf the body (AEk < 0, or Eky2 < E ktl).
1 .5 .3 W ork and M e ch a n ic a l E nergy 103

5°. By the potential energy Ep is meant the part of the


mechanical energy that depends on the configuration of a
system, i.e. on the mutual arrangement of its parts and their
position in an external force field. The potential energy
depends on the relative arrangement of the interacting point
particles, bodies (or their parts) and relates to the totality
(system) of interacting objects. This is why it is called the
mutual potential energy, or the energy of potential interac­
tions (for example the potential energy of the gravitational
interaction of given bodies, the potential energy of elastic
interaction, etc.). When we speak of the potential energy
of a particle or a body, we always have in mind the other
particles or bodies interacting with the given ones.
Since in all practical problems we are interested in the
difference between the values of the potential energy, the
zero or reference point of the potential energy is chosen
arbitrarily from considerations of simplifying the solution
of the problem in hand. Hence, potential energy can be posi­
tive, negative or equal to zero (Ev ^ 0).
6°. A measure of the change in the potential energy of
a system when it passes from one state to another is the
work of the potential forces interacting between the ele­
ments of the system. The work Ap of the potential forces
equals the change AEV in the potential energy of the system
when it passes from the initial state to the final one taken
with the reverse sign:
Ap = A2?p — (-®p, 2 ^p» i)

where EVt x and EVt 2 are the potential energies of the body
in its initial and final states, respectively.
With a view to the formulas for calculating the work of
some potential forces given in 1.5.2.2°, 3°, we can obtain
expressions for the potential energy of interaction of very
simple mechanical systems.
Example 1. The potential energy of the gravitational in­
teraction of a system of two particles of masses m and M.
and at a distance r from each other is

„ GmM
E \>= r
104 Mechanics 1.5.3

where G is the gravitational constant (1.2.8.1°), and the


zero point of the potential energy (Ev = 0) is taken at r =
= oo.
The potential energy of the gravitational interaction of
a body of mass m with the Earth is

p _ GmMjsJi
p“ B E ( R B + h)

where h = altitude of the body


M e — Earth’s mass
R e = Earth’s radius.
The zero point of the potential energy is taken at h = 0.
With the same choice of the zero point of the potential
energy, for small altitudes h (h R e) the potential energy
of the gravitational interaction of a body of mass m with
the Earth is
Ev = mgh

where g is the acceleration of free fall near the Earth’s sur­


face (1.2.8.4°).
Example 2. The potential energy of elastic interactions is

k (Air

where k = quasi-elastic force coefficient, or spring con­


stant (1.2.9.4°)
AI = magnitude of the elongation or compression
vector (1.2.9.4°).
If the coordinate axis OX coincides with the line of
action of the elastic force, then

flere Ev = 0 when A = 0, or when x = 0. The zero point


of the potential energy corresponds to the state of the sys­
tem when the forces of elastic interaction between the bo­
dies (or parts) of the system equal zero (see also 11.4.1.4°).
1.5.4 Work and Mechanical Energy 105

4. L a w o f C o n se r v a tio n o f M e c h a n ic a l E n e r g y
1°. The law of conservation *of mechanical energy: the
mechanical energy of a conservative system remains con­
stant during motion of the system:
E = i?k + Ev = const
This law holds for both closed and for open conservative
systems (1.5.2.6°).
Example 1. A central perfectly elastic collision of two
balls. A collision is defined as the phenomenon of the change
in the velocities of bodies during
the very small interval of time (a)
of their contact. A collision is
perfectly elastic if the forces of
interaction of the colliding bod­
ies are potential ones and the
mechanical energy of the system JTip
does not change as a result of (b)
Q S
the interaction. A collision is
central if the velocities of the "L
0 s
bodies prior to it are directed «/ X
along the line connecting the Fig. 1.5.9]
centres of mass of the bodies.
Two balls of masses mx and
m2 travel along a horizontal straight line with the veloci­
ties vx and v 2 (Fig. 1.5.9a). Determine the velocities of
the balls Uj and u2 after a perfectly elastic collision.
Although the system of balls is not closed (each ball is
acted upon by the force of gravity and another force balanc­
ing the force of gravity), we may apply to the system the
law of conservation of the projection of the momentum onto
the coordinate axis OX (1.2.6.3°):
m iV ix " b ^ 2 y 2a: = m iu ix “b m 2U 2x

Let the velocities of the balls ux and u2 after the collision


he oriented as shown in Fig. 1.5.9ft. With the selected orien­
tation of the given vectors v,, v 2, and the axis OX, we have
niyVy + m2v2 = + m2u%
106 Mechanics 1.5.4

The system of balls is conservative, and therefore the


law of conservation of mechanical energy can be applied
to it in the form
m1v\ , m2v\ _ , m2u\
~TT~ 2 2 2
because the potential energy of the balls in the Earth’s field
m of gravitation does not change in hori-
/^ - — r zontal motion.
From the last two equations, we have
Pm (raj — ra2) vx+ 2m2u2
u{ =
mi -j- ra2
2&~T and
(ra2—raj) y2+ 2raji;j
u2'■
noi ~r ',t,2
If, particularly, colliding balls have
identical masses (m1 = m2), then
F ig .1.5.10
ux = v2 and u2 = v1
In this case, the balls “exchange” velocities as a result of
the collision.
If the second ball was at rest before the collision (v2 = 0),
then
mj —ra2 2raj
ui = ui raj + ra2 and u2 = vl raj + ra2

When m1 > m2, the first ball continues to move to the right
after the collision, but with a lower velocity. When m1 <
<C m2 ,the first ball moves to the left after the collision. The
second ball in both cases moves to the right after the colli­
sion.
If the condition ml m2 is observed when v 2 = 0, then
« —Vj and
The direction of the velocity vector of the light body will
be reversed after a perfectly elastic collision with a massive
stationary body.
Example 2. A perfectly rigid body of mass m moves from
the point 1 to the point 2 near the Earth’s surface under the
action of the gravity force P (Fig. 1.5.10). The point 1 is at
1.5.4 Work and Mechanical Energy 107

the elevation hx, and the point 2 at the elevation h2. Find
the velocity v 2 of the body if its initial velocity is zero
(vx = 0).
In this case, we can consider that the system consists of
only one body of mass m since the state of the Earth does
not virtually change during its motion. The system is open
because there is an unbalanced external force P, but it is
conservative because the force of gravity P is a potential
one. Let us apply the law of conservation of mechanical
energy. The work A gT of the force of gravity equals the
change in the kinetic energy of the body:
Agv = A 2 E^f *
In addition, this work equals the change in the potential
energy of the body taken with the reverse sign:
Agr = A/?p = — (^p, 2 — E Pf i)
Hence,
Ek, 2 — ^k, 1 = —(^p,2 — Ep, i), or E^t 2-j-EPl 2 = E^t i~l-Ept i
In an inertial reference frame associated with the Earth
and when the zero potential energy is selected at the Earth’s
surface, we have
J U f + mgh2 = ^ - + mghl, or 4 + +

Since v1 = 0, then
v2 = Y 2 g (K — ht)
(the second root of the quadratic equation has no physical
meaning here). The vector v2 is directed vertically downward
(cf. 1.1.7.4°).
2°. If a system of interacting bodies is closed, but non­
conservative, its mechanical energy is not retained. The
change in the mechanical energy of such a system equals the
work of the internal non-potential forces:
^E = A\ntt n0n
A system in which forces of friction act in addition to
potential forces is an example of such a system.The forces
qf friction reduce the kinetic energy of the system when i\
108 Mechanics 1.5.4

moves. As a result, the mechanical energy of a closed non­


conservative system always diminishes, passing over into
the energy of non-mechanical forms of motion.
The decrease in the mechanical energy of a non-conser-
vative system leads to a growth in the internal energy U of
the system (11.4.1.2°). For a closed non-conservative sys­
tem, the law of conservation of the total energy % (11.4.1.1°)
is observed. The total energy equals the sum of the mecha­
nical E and internal U energies of a system:
% = E + U = const
3°. The law of conservation of the mechanical energy is
not observed in an open non-conservative system. The change
AE in the mechanical energy of such a system equals the
total work of the internal and external non-potential forces:
AE = Aint, non -f- Aexit non
and is attended by a change in the internal energy of the
system.
Example. A central completely inelastic collision of two
balls. In a completely inelastic collision between bodies, non­
potential forces act, and after
such a collision the bodies
move like a single one with
the common velocity.
Assume that the veloci­
ties of translational horizontal
motion of balls of masses m^
and m2 were vx and v2, respec­
tively, before colliding (Fig.
1.5.11a), while after the com­
pletely inelastic collision their
common velocity is u (Fig.
1.5.116).
The system of balls is non­
conservative and open because
each ball is acted upon by external forces—the force of grav­
ity and another force balancing it. Using the law of con­
servation of the projection of the momentum (1.2.6.3°) onto
the horizontal coordinate axis <9X,rwe get
i n f i x + m f z x = (m i + ™zf\ux
7.5.5 W ork and M ech a n ica l E nergy 109

or with account taken of the direction of the velocity vectors


and the axis OX,
m&i + m2v2 = {m1 + m2) u
whence

m1-f- m2
The change AE in the mechanical energy of a system of
two balls is determined in the given case by the change in
the kinetic energy of the balls as the result of the inelastic
collision:
AE — Ekt 2— ^k, i
where

Ek, 2 = (mi+p ^ -- and Ek,i = ^ - + ^ .

Using the expression obtained above for u, we have


A I? _ m l + m 2 I m l v l - h m 2v 2 \ 2 m l Vl m 2Vl __
a 2 \ m1+ m2 ) 2 2

The reduction in the mechanical energy of a system of two


balls is attended by a growth in the internal energy of the
system.
4°, The escape (second cosmic) velocity v esc is defined as
the smallest velocity that must be imparted to a body for
it to overcome the gravitational attraction of the Earth (or
the Moon, or some other massive celestial body) and move
away from it to an infinitely great distance t

= Y 2gR ^ « 11.2 km/s

5. P ow er
1°. The average power P^wis defined as the physical quan­
tity determined by the ratio of the work A^4 done by a force
or a system of forces during the finite time interval At to
no Mechanics i AJl

this interval:
p __AA
^ av~"Ar
2°. The power (instantaneous power) P is defined as the
physical quantity equal to the limit which the average power
tends to when the time interval At tends to zero:

P = lim Pay = lini -^7 -


A*->0 A*-0 At
If a particle or a body moves with the velocity v, then
P = Fv cos a
where a is the angle between the vectors F and v.

Chapter 6
Elem ents of Fluid Mechanics

1. Mechanical Properties of Fluids


1°. Fluid mechanics is the branch of physics that studies
the laws of equilibrium and motion of fluids (liquids and
gases), and also the interaction of fluids with solids.
Fluid statics considers the conditions and laws of
equilibrium of fluids under the action of the forces applied
to them and, in addition, the conditions of equilibrium of
solid bodies in fluids.
Fluid dynamics studies the laws of motion of fluids, and
also the interaction of fluids with solids upon their relative
motion.
2°. The concrete structure of a fluid is not taken into
account in fluid mechanics, and they are considered as
continuous media continuously distributed in space. The
model of a continuous medium cannot be applied to greatly
rarefied gases (11. 1 . 2 . 2 °).
t.6.2 E le m en ts of F lu id M ec h a n ics ill

3°. Fluids are distinguished by their fluidity that is


connected with the small forces of friction upon the relative
motion of the contacting layere. When the velocity of
relative motion of the layers tends to zero, the forces of
friction between them tend to vanish. This explains the
absence of the forces of friction of rest (1.2.10.2°) in fluids.
Unlike solids, fluids do not retain their shape, but acquire
the shape of the vessel they are confined in. Liquids differ
from gases in the presence of a surface layer (free surface)
(11.6.1.2°), a greater density in the same conditions (except
for the critical state, 11.5.5.1°), and the nature of the depen­
dence of the density on the pressure (the virtual incompres­
sibility of liquids and the noticeable compressibility of
gases).
Elastic forces hinder the change in the volume of a con­
tinuous medium. Since interactions between the layers of a
fluid and also interactions of fluids with solidft occur not at
separate points, but over an area, and the elastic forces are
always perpendicular to the areas being considered, these
interactions are characterized by the pressure (11.3.1.5°) in
fluid mechanics.
4°. The use of the term “fluid” shows that many relation­
ships and laws in fluid mechanics hold both for liquids and
for gases. When it is necessary to underline the difference
between a liquid (characterized by a surface layer) and a
gas (having no surface layer), the former is sometimes called
a dropping liquid.
An incompressible fluid is one for which the pressure de­
pendence of its density may be ignored in the given prob­
lem.
If we cannot ignore the pressure dependence of the densi­
ty of a gas, the latter is called a compressible fluid.

2. Fluid Statics
1°. In the absence or upon the compensation of external
actions on a fluid in an inertial reference frame, a particle of
a continuous medium is in equilibrium if the resultant
(1.2.2.6°) of all the forces exerted on it by the adjacent par­
ticles (1.4.2.1°) equals zero. The same condition must be
112 Mechanics 1.6.2

observed upon the equilibrium of a small element of volume


of any shape separated in the fluid. This leads us to Pas­
cal's law: at a given point in a fluid, the pressure is the same
in all directions.
Example. A small element of volume having the shape of
a right triangular prism is taken inside a fluid (in Fig. 1.6.la

w (b)

Fig. 1.6.1

the elastic forces exerted on the faces of the prism by the


surrounding fluid are shown conditionally).
The condition of equilibrium of the volume taken is
Fx + F2 + F3 + F4 + F5 = 0
It can be written in the form of two equations:
f4 + f5 = 0

(the condition of equilibrium of the forces acting on the


base of the prism), and

Fx + F2 + F8 = 0
(the condition of equilibrium of the forces acting on the
faces of the prism).
A closed triangle of forces corresponds to the last condi­
tion that is similar to the triangular base of the prism
(Fig. 1.6.16). We can therefore write that
Ft = F2 = F8
Ax Ay AI
1.6.2 Elements of Fluid Mechanics 113

or, multiplying the denominators by the altitude AL of


the prism,
Fi = F* F,
Ax AL Ay AL AI AL
Since the denominators equal the areas of the corresponding
faces of the prism, we conclude that
Pi = P2 = P 3
i.e. the pressures on all the side faces of the prism are the
same.
Consideration of the equilibrium of the forces acting on
the base of the prism leads to a
similar conclusion.
The last equation expresses
Pascal’s law.
2°. If a liquid is in a gravita­
tional field, then when determining
the conditions of equilibrium of a
particle of the liquid we must take
into account not only the forces
of its elastic interaction with the
Fig. 1.6.2
adjacent particles, but also the
force of gravity of the given parti­
cle. It will be seen that the pressures inside the liquid at
different levels will not be the same. Regardless of the
shape of the element of volume of the liquid being considered
(Fig. 1.6.2), it will be in equilibrium provided that

P i — Pi = p g K —
where px and p 2 = pressures in the liquid at the depths hx
and h2 from its surface, respectively
p = density of the liquid
g = acceleration of free fall.
If we know the pressure p 0 at the level of the liquid’s
surface (h± = 0) (for instance it equals the pressure of the
surrounding air, or the pressure on the liquid of a piston
contacting its surface), then the pressure p at an arbitrary
depth h will be
P = Po + Pgh
114 Mechanics t.6.2

and at the given point, according to Pascal’s law, it will


be the same in all directions.
The pressure due to the force of gravity of a liquid and
depending on the depth under the liquid’s surface is called
the hydrostatic pressure.
The hydrostatic pressure is taken into consideration when
determining the forces of action of a liquid on the bottom
and walls of a vessel, on solids
in the liquid, when deriving the
conditions of equilibrium of liq­
uid columns in communicating
vessels, etc.
Example. Two different liq­
uids whose densities are px
and p2 are poured into commu­
nicating vessels (Fig. 1.6.3).The
Fig. 1.6.3 —"^“ ■freely moving piston B separates
the liquids to prevent their
mixing.
The following condition corresponds to equilibrium of
Ihe liquid columns and the piston:
POI + Pi£^i = P02 + P2#^2
If the vessels are open, then p Q1 = p 02 (the pressures at the
surface layers of the two liquids are the same and equal
atmospheric pressure), and
P i ^ i = p2g^2, or p ihi = pA
The law of communicating vessels: the heights of columns
of different liquids in communicating vessels are inversely
proportional to the densities of these liquids:
Jh_ _ P2
h2 Pi
If communicating vessels contain a homogeneous liquid
(Pi = P2)» then its free surface in all the vessels will be at
the same level (h± = h2).
3°. Archimedes' principles a body submerged in a liquid
is acted upon by a buoyant (.Archimedean) force FA equal to
the force of gravity (weight) of the liquid displaced by the
body.
1.6.2 Elements of Fluid Mechanics 115

If a body is completely submerged in a liquid, then

where piq = density of the liquid


Vh = volume of the body.
If a body is only partly submerged in a liquid and a por­
tion of the volume 6f the body remains over the free surface
of the liquid, then
F A = P lq £ ^ sub

where VSub is the volume of the portion of the body sub­


merged under the free surface of the liquid.
The point of application of the buoyant force is called
the centre of pressure (the centre of mass of the submerged

r-A—
i —

t UE:
1
portion of the body). The centre of pressure D coincides with
the centre of mass C (1.2.3.4°) of a body only when the latter
is homogeneous and completely submerged in a homogeneous
liquid (Fig. 1.6.4a). For a homogeneous body only partly
submerged in a liquid (Fig. 1.6.46) and for non-uniform bodies
(in Fig. 1.6.4c the density of the hatched portion is greater
than that of the remaining portion of the body), the centre
of pressure and the centre of mass of the body do not coincide.
The buoyant force is due to the fact that the hydrostatic
pressures at different depths are not the same.
4°. A body submerged in a liquid is in equilibrium if
the force of gravity (or weight) of the body P is balanced by
the buoyant force FA:
P = Fa
116 Mechanics 1 .6 J

If with a given submersion no other forces act on a body


and Fa > P , then the body rises until the condition
Plq^^sub ~ P

is observed. When P > FA, the body sinks.

3. M otion of F luids
1°. The motion (flow) of a fluid is called steady if the veloc­
ity of the fluid at the given points of space is independent
of time. The velocity of a fluid may differ at different points
of space.
If the velocity of a fluid changes with time at fixed points
of space, the flow is called unsteady.
The flow of a fluid when its contacting layers move with­
out mixing is called laminar. When the layers of the fluid
mix, the flow is defined as turbulent. Laminar flow maybe
either steady or unsteady. Turbulent flow is always un­
steady.
2°. A line a tangent to which at a given point coincides
in direction with the velocity of the fluid at this point at a
given moment of time is called a streamline. In the steady
flow of a fluid, the streamlines coincide with the paths of
the fluid particles.
A flow tube is defined as the surface formed by the stream­
lines passing through all the points of a small closed contour
(an area element) in a fluid. The fluid flowing through an
entire complex of flow tubes is called a stream.
In the steady flow of a fluid, the flow tubes do not change
in shape with time, and the fluid particles do not leave the
confines of definite flow tubes during their motion.
If the velocity of a fluid changes only slightly from point
to point of the cross section of a stream, then the pipe or
bed through which the fluid is flowing is assumed to be a
single large flow tube. The average velocity over a cross
section is considered to be the velocity of the fluid at this
cross section of such a flow tube.
3°. By internal friction (viscosity) is meant the phenom­
enon of the appearance of forces preventing the relative dis­
placement of the layers of a fluid. The forces of internal fric-
1.6.3 Elements of Fluid Mechanics 117

tion are directed along the layers in contact (and not at


right angles to their surfaces like#the elastic forces) and depend
on their relative velocities.
The cause of internal friction in gases is the transfer by
gas particles of the momenta between contacting layers.
Figure 1.6.5 conditionally shows the contacting layers 1

SI 1
1B
2 ,}A \f
--- ^ ►

Fig. 1.6.5

and 2 of a gas moving with different velocities vx and v 2.


One layer slides over the other, and v1 > v2. The transfer of
a particle (4) from the bottom layer to the top one leads to
retarding of the latter. The transfer of a particle (B) from
the top layer to the bottom one is associated with the trans­
fer of a momentum to the bottom layer which has a compo­
nent coinciding with the velocity v 2l i.e. the transfer of the
particle B will be attended by a growth in the velocity of
the bottom layer. Hence, a friction force directed to the
left will act on the top layer of the gas, and one coinciding
in direction with the velocity v 2 will act on the bottom
layer.
The appearance of forces of internal friction in liquids
at high temperatures close to the critical ones (11.5.5.1°) is
explained in a similar way. At a temperature close to the
freezing point (11.7.4.6°), the mechanism of the appearance
of internal friction forces in a liquid has a more intricate
nature.
4°. An ideal (non-uiscous) fluid is a continuous medium in
which viscosity is absent or it may be disregarded. Other­
wise a fluid is called viscous.
To maintain the flow of a viscous fluid, the pressures in
different cross sections of a flow tube must be different—the
work of the pressure forces must compensate or exceed the
work of the forces of internal friction.
5°. In steady flow, the mass of a fluid passing through any
cross section of a flow tube in a unit time remains constant.
118 Mechanics 1.6.3

The fluid does not accumulate in separate parts of the flow


tube, does not form voids, and does not pass over into adja­
cent flow tubes. This makes it possible to write a continuity
equation for the steady flow of a fluid:

pvS = const

where p = density of the fluid


v = magnitude of the fluid’s velocity in an arbitrary
cross section of a flow tube of area S.
If a fluid is incompressible, then the density p in all the
cross sections of a flow tube is the same (p = const), and
the continuity equation acquires the form

vS = const

6°. Bernoulli's equation is a corollary of the law of con­


servation of mechanical energy for the steady flow of an
incompressible non-viscous fluid through a flow tube. This
equation is

P + pgfr + = const

where p = density of the fluid


v = magnitude of the velocity of the fluid in a cross
section of the flow tube at the height h from a
conventionally chosen level
p = pressure in the same cross section of the flow tube
due to the elastic forces of the fluid.
Bernoulli’s equation becomes simpler for a horizontal
flow tube (h = const):

p - f = const

The quantity p in this equation is called the static pressure,


py2/2 the dynamic pressure, or the kinetic head, and the sum
p + pv2/2 the total pressure {head) (p0).
It follows from this equation that in the cross sections of
a flow tube where the velocity of the fluid is greater, the
1.6.3 Elements of Fluid Mechanics 119

static pressure is lower, the latter increasing in the cross


sections where the velocity of the fluid diminishes.
Bernoulli’s equation may be applied to real gases flowing
without eddies if their velocities
are not great (approximately up
to 100 m/s) because only in these
conditions may we disregard the
compressibility of gases without
a large error.
Example 1. Using Bernoulli’s
equation, find the velocity of a
non-viscous incompressible liquid
as it emerges from a small hole
in the side of an r open vessel
(Fig. 1.6.6).
Bernoulli’s equation for the sections 1 (the surface of the
liquid in the vessel) and 2 (the cross section through the
hole) is
P'1
Pi + pghi - f — - = />2 + PSh2

If we disregard the change in the atmospheric pressure


within the limits of the height of the liquid column in
the vessel, then p x — p 2 and, consequently,

**t+ 4 = * * 2 + 4
Since the cross-sectional area of the hole is much smaller
than the area of the free surface of the liquid, then on the
basis of the continuity equation vx <C vV% <€ vV^i and
the second addend in the left-hand side of the equation may
be ignored. Hence,
ghi = gh2~ \-^
whence
v2 = Y 2g (hr — h2) (Torricelli1s formula)

Example 2. The static pressure in a liquid flow can be


measured with the aid of a thin tube arranged along the
flow and having a hole in its side wall (Fig. 1.6.7a). If the
120 Mechanics 1.6.8

tube communicates with the atmosphere, then the static


pressure is
P — Patm + Pgh

where p atm = atmospheric pressure


p = density of the liquid
h = height to which the liquid rises in the ver­
tical portion of the tube.

3-SS

(a) (b) (c)


Fig. 1.6.7

The total pressure p 0 in a liquid flow is measured using a


tube with an open end facing upstream—a Pitot tube
(Fig. 1.6.76):
Po = Patm + PgH

The dynamic pressure (kinetic head) is measured with


the aid of a Pitot-Prandtl tube (Fig. 1,6.7c) according to the
difference AH between the liquid levels in the tubes
showing the total and static pressures:

^ - = 9^H

This method is used to measure the velocity of a fluid, and


also to measure the velocity of various bodies in fluids.
When used in a gas flow, the tubes showing the total and
static pressures are connected to a manometer that registers
1.6A Elements of Fluid Mechanics 121

the difference between the pressures Ap in these tubes. The


pressure difference is used to find the velocity of the gas
flow (or the velocity of the body):

4*. Motion of Solid Bodies in Fluids


1°. Instead of considering the motion of a solid body
in a stationary continuous medium, on the basis of the me­
chanical principle of relativity (1.2.7.5°) we can study the
processes occurring when a fluid flows around a stationary
solid body. The effects of interaction of the body with the
flow are identical here (the forces of interaction, the distri­
bution of the pressures over the body’s surface, etc.), but the
second way is simpler in practice. For example, before send­
ing an aircraft of a new design on a flight, its model (full
size or at a reduced scale) is tested in a wind tunnel by blow­
ing air around it.
2°. When a viscous fluid flows around a solid body, the
flow is deformed. The layers of the fluid in direct contact
with the body adhere to-j its surface. A boundary layer is
formed on*the surface of the body, i.e. a region within whose
confines the velocity of the fluid changes from zero to the
velocity yUf of the undisturbed flow. Figure 1.6.8a schemati­
cally shows the distribution of the velocities in the fluid
flowing around a flat plate; the heavy dash line shows the
boundary layer.
At a certain point of the body’s surface (point 0 in
Fig. 1.6.8), the boundary layer may break away. The fluid
from the boundary layer is thrown out into the main flow,
and after the point 0 a vortex flow is formed (Fig. 1.6.86
shows the approximate path of the fluid streamlines before
and after the point 0 ).
The resistance to the motion of a body in a fluid depends
mainly on the characteristics of the boundary layer—its
thickness, whether it is laminar or turbulent (1.6.3.1°),
etc.—and on the eddies or vortices formed after breaking
away of the boundary layer.
.122 Mechanics 1.6.4

3°. The resistance upon the motion of a body in a vis­


cous fluid is considered to consist of two components: the
friction resistance and the pressure resistance.
The friction resistance is due to the forces of internal fric­
tion (1.6.3.3°) appearing upon considerable velocity changes
in the boundary layer. These forces depend on the shape and

(n) I

(b)
Fig. 1.6.8.

dimensions of the body, the viscous properties of the fluid,


and are proportional to the relative velocity of the body
and the fluid.
The pressure resistance is determined by the difference
between the pressures at the front and rear of the body
around which the fluid is flowing. The force of the pressure
resistance depends on the shape and dimensions of the body,
and is proportional to the density of the fluid and the square
of the relative velocity of the body and the fluid. In other
words, the pressure resistance is proportional to the dynamic
pressure (1.6.3.6°).
Since the friction resistance and the pressure resistance
depend differently on the velocity of a body, the former pre­
dominates at very low velocities and the latter at very high
ones. Within the broad range of intermediate velocity va­
lues, the calculation of the total resistance of a body to mo-
1.6.4 Elements of Fluid Mechanics 123

tion is an exceedingly complicated task. In the long run, this


resistance is determined experimentally for different bodies
in different conditions of their motion in a fluid.
4°. A difference between the static pressures at different
points on the surface of a solid body moving in a fluid may
result not only in a resistance
a
force, but also in the so-called lift­
ing force, or lift.
The reason why a lift appears
is the vortex motion of the fluid
around the body in a flow. Such
motion is defined as circulation
(icirculation of the velocity). In vor­
tex motion, the fluid streamlines
are closed, and it is always pos­
sible to find such a mentally
closed contour beyond whose limits
the streamlines do not pass.
Example 1. Lift when a viscous fluid flows around a
rotating circular cylinder (the Magnus effect).
If a viscous fluid flows from the left with the velocity v
onto a cylinder rotating clockwise about its horizontal axis
(Fig. 1.6.9), then the lift Fj ap.
pears that is directed upward
The cylinder carries along the
layers of the viscous fluid ad­
joining its surface and causes
them to rotate-circulation ap­
pears. If the streamlines a and
b are the boundaries of the un­
disturbed flow, then in the sec­
tion ac of the flow tube the re­
sultant velocity of the fluid will
be greater than in the section
db. The static pressure at the bottom points of the cylin-
der will be greater than at the top ones. This is exactly
what leads to the appearance of the lift F*.
Example 2. Lift of an aerofoil.
The sections of aerofoils intended for flights with a velo­
city lower than that of sound (Zhukovsky sections) have the
form conditionally shown in Fig. 1.6.10.
124 Mechanics 1.6.4

When air flows around such an aerofoil, breakaway of the


boundary layer occurs only on its rear edge. If the stream of
air comes in from the left, then vortex motion at the rear
edge of the aerofoil will occur counterclockwise. Clockwise
vortex motion of the air around the aerofoil section must
appear here in the system “aerofoil-main stream”. This leads
to the circumstance that the resultant velocity of the air
stream over the aerofoil will be greater than under it. The
static pressure at the bottom surface of the aerofoil will be
greater than at its top surface. It is exactly this pressure
difference that ensures the force F acting on the aerofoil. Its
vertical component F* is the lift, and its horizontal compo­
nent Fd ,the drag force, or drag. The latter is balanced or
overcome by the tractive effort of the aircraft’s motor.
MOLECULAR
PHYSICS
PART AND
FUNDAMENTALS
OF THERMODY­
NAMICS

Chapter 1
Fundamentals
of Molecular-Kinetic Theory

1. Basic Concepts and Definitions


1°. The molecular-kinetic theory is the name given to the
science that explains the structure and properties of bodies
by the motion and interaction of the atoms, molecules, and
ions they consist of. The theory is based on three very im­
portant principles that are confirmed completely both exper­
imentally and theoretically:
(a) all bodies consist of particles—molecules, atoms, and
ions whose composition includes smaller elementary parti­
cles (VI. 5.1.1°);
(b) atoms, molecules, and ions are in continuous chaotic
motion; and
(c) forces of interaction—attraction and repulsion
(11.1.4.3°, 4°)—exist between the particles of any body.
These initial principles are confirmed by the phenomena of
diffusion (11.1.3.1°), Brownian motion, the features of the
structure and the properties of liquids and solids, and also
126 Molecular Physics and Thermodynamics

by investigations in the field of the physics of elementary


particles.
2°. An atom is defined as the smallest particle of a given
chemical element. Quite definite atoms correspond to every
chemical element; they retain the chemical properties of
the relevant element.
An atom consists of a positively charged nucleus and
negatively charged electrons moving in the electric field
of the nucleus. The electric charge of a nucleus (VI.2.1.1°)
equals the absolute value of the total charge of all the atom’s
electrons, hence an atom is electrically neutral (see VI.2.1.1°-
5° for greater detail on the structure of an atom).
3°. A molecule is defined as the smallest stable particle of
a given substance possessing its basic chemical properties.
A molecule may consist of one or several atoms of the same
or different chemical elements.
Atoms combine into a molecule at the expense of chemi­
cal bonds based on different interactions of the outer {valen­
ce) electrons (VI.2.9.2°). The number of atoms in a molecule
varies within very broad lim its—from two (CO, 0 2, NO,
H2), three (C02, S 0 2), and four (NH3) to hundreds and even
thousands. Like an atom, a molecule is electrically neutral,
i.e. contains an equal quantity of electrically charged par­
ticles of both signs.
4°. By the amount of substance is meant the physical
quantity determined by the number of specific structural
elements—molecules, atoms, or ions (111.3.1.1°)—which the
substance consists of (see also VII.4.3°). Since the masses
of the separate structural elements (for instance molecules)
differ from one another, then identical quantities of differ­
ent substances have different masses. For example, 1025 mol­
ecules of hydrogen and 1025 molecules of oxygen are con­
sidered to be identical amounts of substance although they
have different masses equal to 33.45 X 10“3 and 531.45 X
X 10"3 kg, respectively. Mass is not a measure of the amount
of substance. The unit of the latter is the mole (VII, Table
VII.2), and also multiple and submultiple units of it.
(For the information given in Paragraphs 4°-7°, see also
VII.4.3°, 4°.)
5°. The number of atoms (molecules or other structural
units) contained in one mole (kilomole) of a substance is
II.1.2 Fundamentals of Molecular-Kinetic Theory_________427

called the Avogadro constant (number) N a •

A a = 6.022 x 1023 mol"1 =J6.022 x 1026 kmol"1

6°. The volume of a mole (kilomole) is called the molar


volume Vm:
F m= i ^ = - -

where v = 1/p = specific volume (11.3.1.6°)


p = density of the substance (1.2.3.7°)
(x = mass of one mole (kilomole) (the molar mass).
In standard conditions (t = 0 °C, p = 101 325 N/m2 =
= 1 atm = 760 mm Hg), the molar volumes of all ideal
gases (11.2.1.1°) are the same: Fm = 22.4 m3/kmol =
= 22.4 dm3/mol (see also Avogadro’s law—11.3.3.6°).
7°. The molar mass p = mNa, where m is the mass of
one structural element (atom, molecule, or ion).
The amount of substance n (number of moles or kilo-
moles) is
n = Ml [i

where M is the mass of the substance.


8°. The dimensions of an atom are determined by the
distances from the centre of its nucleus at which the outer
valence electrons or the outer filled electron shells of the
atom are (VI.2.8.6°).

2., Brownian Motion

1°. Brownian( motion is the continuous chaotic motion of


minute particles that are in such conditions in a fluid in
which the force of gravity does not affect their motion (sus­
pended particles).
2°. The cause of Brownian motion is the different action
of the molecules of a fluid on the Brownian particles from
different^ sides. Owing to the chaotic nature of the motion of
the fluid molecules, a Brownian particle at any moment of
128 Molecular Physics and Thermodynamics II.1.3

time is subjected to an unbalanced action that continuously


changes in magnitude and direction. As a result, the particle
moves chaotically along the
sections of an intricate broken
line (Fig. II.1.1).
" '
Brownian motion is a di­
K - rect experimental proof of the
existence of molecules of a
fluid and of the chaotic na­
#
4f —
== '
ture of their thermal motion.
3°. Brownian motion is
" i characterized by the mean val­
7 ue of the square of the displace­
ment of a particle in an
--- V arbitrary direction x2. As a
v r result of the chaotic nature of
Brownian motion, the mean
displacement of a particle
x = 0 because the probabi­
lity of a particle moving in
Fig. II.1.1 any direction is the same.
Here x2 0.
4°. The laws of Brownian motion:
(a) it continues incessantly without any visible changes;
(b) x2 does not depend on the substance of the particles,
but depends only on their linear dimensions and geometrical
shape; and
(c) x2 grows with increasing temperature and decreasing
viscosity of the fluid.

3. Diffusion
1°. Diffusion in the simplest case is defined as the process
of levelling out of the densities (or concentrations) of two
substances when they are mixed with each other. The mutual
penetration of the substances is a result of the chaotic mo­
tion of their particles and the change in density along a cer­
tain direction. It is exactly in this direction that diffusion
occurs. Diffusion is observed in fluids and in solids.
I I . 1.4 Fundamentals of Molecular-Kinetic Theory 129

Examples of diffusion:
(a) If we carefully pour water into a vessel containing a
solution of blue vitriol (copper sulphate), then the sharply
defined interface between the solution and the water will
vanish with time. The water will gradually become blue,
and after a rather long time a liquid homogeneous in colour
is obtained.
(b) The stratosphere is a homogeneous mixture of oxy­
gen, nitrogen, carbon dioxide, water vapour, and inert
gases. If there were no diffusion, all the constituent parts of
the stratosphere would be stratified owing to the different
force of gravity of the molecules of these gases (1.2.8.3°).
(c) Plates of gold and lead ground in and pressed against
each other “intergrow” during five years as a result of the
mutual penetration of the particles into the plates to a depth
of about one millimetre.
2°. A measure of diffusion is the mass AM of a substance
that has diffused in a unit time through a unit area of con­
tact of the substances. The quantity AM grows with an
increasing change in the density (or concentration) per unit
length in the direction of diffusion.
3°. Diffusion accelerates with elevation of the tempera­
ture. For example, in hot water, sugar and salt dissolve
more rapidly than in cold water, mainly owing to diffusion.
The acceleration of diffusion is associated with the fact
that elevation of the temperature is attended by a growth
in the velocities of the molecules (11.2.2.4°).

4. Forces of Interaction Between Molecules

1°. Forces of mutual attraction and repulsion simulta­


neously act between the molecules of any substance (inter-
molecular interaction). The existence of stable liquids and
solids is associated with the forces of intermolecular interac­
tion. An experiment with two pieces of lead which when
closely fitted against each other with their freshly made cuts
withstand a considerable tensile force proves the existence
of forces of attraction. The relatively low compressibility of
liquids and the ability of solids to withstand compressive
130 Molecular Physics and Thermodynamics I I .1.4

deformations (11.7.2.1°) confirm the presence of forces of


repulsion between molecules.
2°. The nuclei and electrons of atoms in molecules expe­
rience electric forces of interaction, although a molecule as a
whole is electrically neutral. A molecule, as regards its
electrical properties, can be approximately considered as an
electric dipole (111.1.4.6°). Electric interaction takes place
between adjacent electric dipoles (111.1.4.7°). Therefore,
the forces of intermolecular interaction have an electrical
origin*. The division of the forces of interaction between
molecules into “forces of attraction” and “forces of repulsion”
is conditional. It is customary practice to consider forces
of attraction negative and forces of repulsion positive.
3°. At distances r between the centres of molecules of the
order of magnitude of 10“9 m, forces of attraction act that
grow as the molecules approach one another. The intermole­
cular forces of attraction / att (r) are short-range ones—they
rapidly diminish with an increasing distance r between the
molecules:

/att ( r ) = ----jr

The coefficient a depends on the structure of the interacting


molecules and the kind of forces of attraction.
4°. At distances r between the centres of molecules that
are compatible with the linear dimensions of small inor­
ganic molecules (10“10 m), forces of repulsion manifest them­
selves mainly owing to the mutual repulsion of the positively
charged nuclei of the atoms in a molecule. The forces of
repulsion / rep (r) diminish with an increase in r more rapidly
than the forces of attraction:

/r e p (r) = ~p&

The coefficient b depends on the same factors as the coefficient


a (Par. 3°).

* Forces of a quantum nature also exist that are not considered


in an elementary course of physics.
ii.i.5 Fundamentals of Molecular-Kinetic Theory 13i

5°. The simultaneous existence of forces of attraction


and repulsion signifies that a molecule is acted upon by the
resultant force of intermolecular interaction / (r) (Fig. II.1.2):

/(* ■ )= ---+ -£

When r ^ r0, the attraction (repulsion) of the molecules


exceeds their repulsion (attraction). When r = r0, the forces
of attraction balance the forces
of repulsion, and / (r) = 0 . The f(r)
most stable arrangement of the
interacting molecules corresponds
to this case. B y /(r ) here, strictly
speaking, is meant fr (r)—the pro­
jection of the force / (r) onto the
direction of the vector r determin­
ing the distance between the cen­
tres of the molecules.
6°. The curve / (r) allows us
to qualitatively explain the mo­
lecular mechanism of the appear­
ance of elastic forces in solids.
When an external stretching force acts on a solid, its
particles move away from one another over distances
exceeding r0. The action of the forces of attraction between
the particles prevents stretching of the body and helps the
particles return to their initial positions. Upon great com­
pression of a solid body, its particles are brought together
to distances between them less than r0, and the forces of
repulsion between the particles prevent further compression
and also help the particles return to their initial positions.

5. Potential Energy of Interaction


of Two Molecules
1°. Two molecules free of any actions except for the force
/ (r) of intermolecular interaction must approach or move
away from each other depending on what force predomi­
nates—the force of attraction or that of repulsion. Only at the
132 M o lecu la r P h ysics and th e rm o d y n a m ic s 11.1.5

distance r = r0 between the molecules at which / (r0) = 0


are the molecules in equilibrium.
The potential energy Ev (r) of interaction of two mole­
cules is defined as the part of the energy of a system of two
molecules that depends on the distance r between their
centres. The quantity Ev (r) is measured by the work done
by the force / (r) when the distance between the molecules
changes from r to infinity, where Ev (r) is considered to equal
zero [Ev (r) = 0 when r oo]. This choice of the zero value
of the potential energy (1.5.3.5°) is due to the circumstance
that two molecules do not interact when the distance r
between them is very great.
2°. If two molecules approach each other over the small
distance Ar (Ar < 0 ) under the action of forces of attraction
[/att (r) < 0 ] , then the system of two molecules does the
positive work AA = /att (r) Ar > 0 (1.5.1.1°). It follows
from the law of conservation of energy (1.5.4.1°) that here
the kinetic energy of the molecules grows, while their poten­
tial energy diminishes. But when the molecules move infinite­
ly far apart from each other (r->- oo) they no longer inter­
act, Ev (oo) 0 (Par. 1°). Hence, when two molecules
approach each other in the region where forces of attrac­
tion act, potential energy of their interaction is negative.
3°. If two molecules approach each other over the dis­
tance Ar (Ar < 0) in the region where forces of repulsion act
t/rep (r ) > 0],* the system does negative work for over­
coming the forces of repulsion. This process is associated with
a reduction in the kinetic energy of the molecules and a
growth in the potential energy of their interaction. It follows
from the choice of the reference point of the potential energy
that Ep (r) must be a positive quantity in the region where
forces of repulsion act. The closest approach of the molecules
is achieved at the distance r = d when the entire kinetic
energy of the molecules is completely used up for doing work
against the forces of repulsion. The potential energy 2?p (r)
equals the total energy E of the system of two molecules
and has the maximum value 2?p, max (d) = E. The value of
E determines the distance d that equals the abscissa of the
point of intersection of the curve Ev (r) with the horizontal
* Such approach is impossible without any external action.
11.1.6 F u n dam en tals of M o le cu la r-K in etic T heory 133

straight line Ev (d) = E (Fig. 11.1.3). The lack of symmetry


of the curve Ev (r) near its minimum 2?p, 0 is explained by
the joint action of the forces ©f attraction and repulsion
between the molecules that, depend differently on the dis­
tance r. The distance d is the
smaller, the greater is the
kinetic energy of the mole­
cules (i.e. the higher is the tem­
perature), and depends on
the structure of the mole­
cules. The rapid growth of the
potential energy in the re­
gion of distances close to d
(see Fig. 11.1.3) indicates that
the forces of repulsion be­
tween the molecules increase
very rapidly with a de­
crease in the distance between them. The distance d is the
diameter (collision diameter) of a molecule determining the
linear dimensions of the region into which another molecule
cannot penetrate. Thus, the forces of repulsion between mole­
cules determine the size of a molecule.
4°. When r = r0, a system of two molecules is in the
state of stable equilibrium. The smallest value of the poten­
tial energy corresponds to this—at this point the curve
Ev (r) has a minimum Ev% 0 (see Fig. II.1.3). If e is the
mean kinetic energy of a molecule (11.2.4.3°), then a com­
parison of e with EVt 0 makes it possible to distinguish three
states of aggregation of a substance: gaseous, solid, and
liquid:
e > £ p, 0- g a s , e < E Pt 0— solid, e « E Vt 0— liquid (*)

6. Structure of G ases, Solids,


and Liquids
1°. Thermal motion is defined as the chaotic motion of
molecules, atoms, and ions in gases, solids, and liquids.
The velocities of the thermal motion of the particles of a
substance grow with elevation of the temperature (11.2.4.4°).
134 M o le c u la r P h y s ic s and T h erm o d yn a m ics I I .1.6

The nature of the thermal motion of molecules, atoms, and


ions depends on the state of aggregation of the substance and
is determined by the forces of intermolecular interaction.
2°. In gases, the forces of attraction between their mole­
cules cannot keep them next to one another, and the mole­
cules fly apart in all directions, occupying the entire volume
of the vessel containing the gas. Gases have no definite
volume and shape and are readily compressed under the
action of an external pressure.
3°. Solids are defined as bodies distinguished by the
constancy of their shape and volume. The explanation is that

(a) (b)
Fig. II.1.4

the forces of mutual attraction of the particles of a solid


body are very great in comparison with the relevant forces
in gases. A particle of a solid cannot move away from its
neighbours over a considerable distance. The thermal mo­
tion of the particles in solids consists in chaotic oscillations
of them relative to their equilibrium positions—the points of
a crystal lattice (Par. 5°). The oscillations of the particles of
a solid are not strictly harmonic (IV. 1.1.4°). The reason is
the same as that of the lack of symmetry of the curve Ev (r)
(see Fig. II.1.3) near the minimum—the different dependence
of the forces of attraction and repulsion on r.
4°. Solid bodies having an ordered periodic arrangement
of their particles in space are called crystals. The latter are
confined by flat faces arranged orderly relative to one another.
The faces converge at the edges and points, or apices. Fig-
II.1.6 F u n d a m en ta ls of M o le cu la r-K in etic T heory 135

ure II.1.4 shows the crystal lattices of table salt (a) and
graphite (b). •
Single crystals having the shape of regular polyhedrons
are called monocrystals. Most solids have a fine crystalline
structure {polycrystals). Such bodies consist of a great num­
ber of intergrown fine, chaotically arranged crystals (crystal­
line grains, crystallites). Examples of polycrystalline solids
are metals, stones, and sand. For the types of crystals see
11.7.1.2°.
5°. The particles which a crystal consists of form a regu­
lar crystal lattice in space (a space lattice). The basis of a
crystal lattice is an elementary cell of a definite geometrical
shape at whose apices—the lattice points, or sites—particles
(atoms, molecules, or ions) are arranged. The elementary
cell repeating at spacings equal to integral lengths of its
edges forms a complete crystal. The length of an edge of an
elementary cell is called the period of the crystal lattice.
The periods may be different in different directions in a
crystal.
6°. In crystalline solids, there is long-range order in
the arrangement of the particles which a crystal cell is con­
structed of: the ordered arrangement of the particles repeats
within the limits of hundreds, thousands, and tens of thou­
sands of cells. In this sense, an entire crystal can be con­
sidered asVsingle "giant particle having a structure of a single
type.
7°. Liquids are defined as fluids that have a definite vol­
ume but do not have their own shape, taking on that of the
vessel containing them. If gases are characterized by com­
plete disorder in the arrangement of their molecules, and
solids by the presence of long-range order, then liquids occu­
py an intermediate position as regards their structure and
the nature of thermal motion in them. This is connected
with the circumstance that the condition (*) (11.1.5.4°) is
most complicated for liquids: the] potential energy of in­
teraction of the liquid particles is commensurable with their
kinetic energy. The strong intermolecular interaction of the
molecules leads to the fact that the particles in liquids are
very close to one another. This arrangement, however, is not
strictly ordered throughout the entire volume as in solids.
Short-range order is observed in liquids—an ordered relative
136 M o le c u la r P h ysics and T h erm o d yn a m ics 11.2.1

arrangement (or mutual orientation) of adjacent particles of


the liquid.
8°. The molecules of a liquid perform chaotic oscilla­
tions about definite equilibrium positions. These oscillations
occur inside the “free volume” placed at the molecule’s
disposal by its neighbours. After a certain average time t elap­
ses, called the mean free lifetime (the mean relaxation time),
the equilibrium positions of the molecules are displaced
over the distances equal in their order of magnitude to the
average distance between adjacent molecules (10~10 m).
These displacements have the form of jumps and are asso­
ciated with the expenditure of energy for overcoming the
bonds of a molecule with its neighbours. The time t rapidly
diminishes with elevation of the temperature.
9°. Liquids do not possess the anisotropy characteristic
of solid crystalline bodies (11.7.1.1°)—the physical proper­
ties of liquids are the same in all directions inside a liquid
(the isotropy of liquids).
10°. A very important feature of liquids is their fluidi­
ty—liquids do not resist a change in their shape (Par. 7°).
At the same time, they are characterized by a low compres­
sibility (11.1.4.1°).

Chapter 2
Molecular-Kinetic Theory
of Ideal Gases

1. I d e a l G a s

1°. The molecular-kinetic theory considers an idealized


model of real gases—an ideal gas.
An ideal gas is defined as a gas between whose molecules
forces of mutual attraction are absent (11.1.4.3°). It is as­
sumed that when the molecules of such a gas collide with one
n .2.2 M o le c u la r -K in e tic T heory of Ideal G ases 137

another and with the walls of the vessel they behave like
absolutely elastic balls of finite but very small dimensions
(the model of elastic balls). The£e collisions occur according
to the laws holding for a perfectly elastic collision (1.5.4.1°).
The existence of molecules of finite, though small, dimen­
sions is associated with the action of forces of repulsion
between the particles (11.1.5.3°). The sharp growth of the
repulsion forces when the molecules approach one another
to small distances does not permit the particles to penetrate
into one another and determines the finite dimensions of
the molecules.
The elementary course of physics treats ideal gases whose
molecules consist of a single atom.
2°. Actually existing gases at not too low temperatures
and sufficiently low pressures—rarefied gases—are close in
their properties to an ideal gas. For example, helium at room
temperature and atmospheric pressure obeys the laws of
ideal gases with a good approximation.

2. V elocities of Gas M olecules


1°. The chaotic nature of the thermal motion of gas
molecules signifies that none of the directions of their pos­
sible motion predominates—all the directions of their mo­
tion are equivalent and are encountered with the same fre­
quency. Among the total sufficiently great number of gas
molecules N contained in a cubic vessel, 7V/3 molecules on
an average will travel along each of the axes OX, OY, and
OZ coinciding with three edges of the cube (Fig. II.2.1).
The number of molecules travelling along each of these
axes in one direction is N/6 and equals their number trav­
elling in the opposite direction.
2°. Collisions between the molecules of a gas result in the
velocities of its molecules constantly changing in magnitude
and direction.
Assume, for example, that the molecule A has the velo­
city v directed along the axis OX. Assume further that as a
result of colliding with another molecule it received such a
momentum (1.2.3.5°) along the axis OZ that a velocity
equal to v was transmitted to it. Hence, the velocity of the
138 M o le c u la r P h y sics and T h erm o d yn a m ics 11.2,2

molecule A will change and become equal in magnitude to


V 2 v (Fig. II.2.2).
3°. The motion of every gas molecule can be described
by the laws of Newtonian mechanics. But the number of
molecules in any gas is exceedingly great, while the forces
acting between the molecules are suchj that it is impos­
sible to describe the properties of this tremendous complex

Fig. II.2.1 Fig. II.2.2

of all the molecules using the laws of mechanics. In such


cases, the statistical method is employed for investigation.
Using a special mathematical apparatus—the theory of
probability—we calculate by this method the average values
of the physical quantities characterizing the motion of all
the molecules (the average velocities of the molecules, the
average value of their energy, etc.). The statistical method
is used to study not only gases, but also liquids and solids.
4°. It is convenient to introduce average (mean) veloci­
ties characterizing the complex of all the molecules of a gas
at the given temperature T.
(a) The magnitude of the arithmetical mean velocity of the
molecules v is
v = V\ + V2 + • ■• + vnN

where N is the total number of gas molecules.


The value of the arithmetical mean velocity of the mole­
cules is
1 1 .2 ,3 M o lecu la r-K in etic Theory of Ideal G ases 139

where R = molar gas constant (11.3.3.7°)


(li = molar mass (11.1.1.7°).
(b) The mean square velocity it of the molecules is
u = V~v2 = ”i + "i+ • •• + vn
Jv
/ 3RT A 7Q1/ ^
— » 1 -7 3 k -jr
Here i;2 is the weara square of the velocity of the molecules. It
must never be confused with the square of the mean velocity,
i.e. v2 =^= (v)2.
Problem. Determine the arithmetical mean and mean
square velocities of air particles at 17 °C. Assume the molar
mass of the air to equal 29 X 10~3 kg/mol.
Given: T = 273 + 17 = 290 K, p. = 29 X 10-3 k g/m ol.
Required: v, u.
Solution: The arithmetical meai elocity of gas mole­
cules is v = Y&RT!n\k. The mean square velocity is u =
= ]/"3i?77p, where R is the molar gas constant equal to
8.31 J/mol* K.
Hence,
8 X 8 .3 1 X 290
/ 3.14 X 29 X lO"3
3 X 8 .3 1 X 2 9 0
463 m/s

/ 29 X 10"3
500 m/s

3*. M e a n F r e e P a th o f a M o le c u le
1°. The mean free path I is the average distance which
a molecule travels without a collision, i.e. between two con­
secutive collisions. A molecule travels uniformly and recti-
linearly over its free path.
2°. The mean free path of a molecule is directly propor­
tional to the arithmetical mean velocity of the molecule
v and inversely proportional to the average number of col­
lisions of a molecule z in a unit time:
140 M o lec u la r P h ysics and T h erm o d yn a m ics I I . 2.3

3°. The mean free time of a molecule x is the time during


which the molecule travels without collisions, i.e. it is the
average time between two consecutive collisions:

T - —
V

4°. The mean free path is

V I nnd2

where n = number of molecules in a unit volume


d = collision diameter of a molecule (11.1.5.3°)
5°. For a given gas at a constant temperature (T = const),
the mean free path is inversely proportional to the pressure
of the gas p (or its density):

P\li —P2 I2

The subscripts 1 and 2 relate to two states of the gas, respec­


tively.
Problem 1. By how much will the mean free path of
a molecule in a vessel at a constant temperature change if
the pressure is reduced by 10 %?
Given: T = const, Ap = p 2 — Pi = —O.lpi.J
Required: AZ.
Solution: According to the relationship between the pres­
sure of a gas and the mean free path of its molecules, i.e.
PiU = Pihi we have Z2 = PtV /V Hence AZ = Z2 — Zx =
— PihKPi — O.lpj) — \ = \ [(1/0.9)— 1] « 0 .1 Zj, whence
AT/Tx = 0.1, or 0.1 X 100 = 10 %.
Th'oblem 2. The mean free path of oxygen molecules
at 27 °C is 4.17 X 10“3 cm. Determine the mean free time
of the molecules in these conditions.
Given: 1 = 4.17 X 10~5 m? T = 273 + 27 = 300 K,
(i = 32 X 10"3 kg/mol.
Required:
11.2.4 M o lecu la r-K in etic Theory of td ea t (rases 14l

Solution: the mean free time of a molecule is x — llu,


where v is the arithmetical mean velocity. Consequently,

I 7 l / * JX |X
x
8R T YM t
/ JX JI

4.17 xlQ~6 1/3.14 X 32 x 10~3


9.3 x 10 8 s
1 /8 x 8 .3 1 x 3 0 0

4. F u n d a m e n ta l E q u a tio n
o f th e K in e tic T h e o r y o f G a se s
1°. The fundamental equation of the kinetic theory of gases
establishes the relationship between the pressure of a gas p,
its volume V, and the kinetic energy of the translational
motion of its molecules Ek:
pV = ^ -E k
N 2
Here Ek = 2 -tt = total kinetic energy of the translational
i=l z
motion of N identical molecules of a gas oc­
cupying the volume V
m = mass of a molecule
Vi = its velocity.
2°. If we introduce the mean square velocity (11.2.2.4°),
then
Ek = -I- Nmv2= -i- M v2

where M = Nm is the mass of the gas. Hence,

= = (*)

Equation (*) allows to express the pressure of a gas:


1 — 1 —
P — Y Pv2= Y nmv2
142 Molecular Physics and thermodynamic$___________ i 1.2.4

where p = nm = density of the gas


n = number of gas molecules in a unit volume (n =
= N/V).
The pressure of a gas is proportional to the product of its
density and the mean square of the velocity of its molecules.
3° For one mole of a gas occupying the volume Vm
(11.1.1.6°), Eq. (*) yields

pv m= i - NAm7*= 4 ArA‘T " = 4 ' jVAti (••)


where N A = Avogadro constant (11.1.1.5°)
e = mv2/2 characterizes the mean kinetic energy
of the chaotic thermal motion of a gas mole­
cule.
4°. A comparison of Eq. (**) with the Mendeleev-Clapey-
ron equation for one mole of a gas pV m = RT (11.3.3.7°)
leads to the molecular-kinetic interpretation of the absolute
temperature (11.3.1.7°):
3 R
T = -— kT
2 Na
In this formula, the quantity k = R /N A is the molar gas
constant (11.3.3.7°) related to one gas molecule, and is
called the Boltzmann constant:
R
na
: 1 .3 8 x l0 ~ 23 J/K

The formula for e permits us to disclose the physical


meaning of the absolute temperature. The latter is a measure
of the mean kinetic energy of the thermal chaotic motion
of ideal gas molecules. The mean kinetic energy of a gas
molecule is proportional to the absolute temperature. In
the region of very low temperatures close to the degeneracy
temperature (VI. 1.8.1°), the preceding statements do not
hold.
Problem 1. What is the pressure of hydrogen in a cylin­
der if the capacity of the latter is 10 litres, and the mean
kinetic energy of translational motion of the hydrogen mole­
cules is 7.5 X 103 J?
Given: Ek = 7.5 X 103 J, V = 101 = lO"2 m
Required: p.
IL2.4 Molecular-Kinetic Theory of Ideal Gases 143

Solution: The fundamental equation of the kinetic theory


of gases is pV = (2/3) Ek. Ilence,

p= -p- = ■§■ = 5 .0 x 105N/m2 = 0 .5 MPa

Problem 2. Under what pressure is a gas if the mean


square velocity of its molecules is 580 m/s, and its density
is 9 X 10 ~4 g/cm3?
Given: V v2 = 580 m/s, p = 9.0 X 10~4 g/cm3 =
= 0.9 kg/m3.
Required: p.
Solution: From the fundamental equation of the molecular-
kinetic theory, we have p = (1/3) pi;2, where p is the density
of the gas, and v2 is the mean square of the velocity of its
molecules. Thus,

P= i - x 0.9 x 5802 = 1.1 xio6N /m * = l.l MPa

Problem 3. Dust particles in the suspended state in a


monatomic gas are in thermodynamic equilibrium (11.3.1.3°)
with the gas. The temperature of the gas is 300 K. Find
the mean kinetic energy of translational motion of one mole­
cule and one dust particle.
Given: T = 300 K.
Required: em, ep.
Solution: It follows from the condition of thermodynamic
equilibrium of the dust particles and the gas molecules that
their mean kinetic energies are the same. From the funda­
mental equation of the molecular-kinetic theory of gases,
we have e = (3/2) kT. Hence,

gm= i p = i - x 1 .3 8 x IQ”2®x 300 = 6.20 x 10-2‘ J

Problem 4. What does the mean kinetic energy of trans­


lational motion of the molecules contained in one mole and
in one kilogram of helium equal at a temperature of 1000 K?
Given: T = 1000K, p = 4 x 10“® kg/mol, M = 1 kg.
Required E kn ,Ek,2.
144 Molecular Physics and thermodynamics 11.3.1

Solution: We determine the mean kinetic energy of


translational motion of one molecule by the formula h =
= (3/2) kT. Since the number of molecules in a mole is
N a = 6.02 X 1023 mol-1, then the mean kinetic energy
of all the molecules contained in one mole will be

E k = iN A = ± k T N A = ± R T
where R is the molar gas constant. Hence,
Ek, , = ± x 8.31 X1000 = 1.25 x 104 J

The mean kinetic energy of the helium molecules in


1 kg is
E2 = — —
2 jll
/tr = -~x-}-X
2 4
10-3 x 8 .3 1 x 1000 » 3 x 10® J
Here M/\x is the amount of the helium (the number of moles).

Chapter 3
Ideal Gas Laws

1. Equation of State
1°. Physical phenomena in gaseous and other bodies are
studied by the thermodynamic method in addition to the
statistical one (11.2.2.3°). Thermodynamics is the name given
to the branch of physics studying the conditions of the con­
version of energy from one kind to another and the quanti­
tative relationships in such conversions (11.4.3.4°). Thermo­
dynamics is based on experimentally established laws of
thermodynamics (11.4.5.1° and 11.4.9.2°). These laws allow
us, without taking into consideration the molecular struc­
ture of substances, to get much information on the properties
of bodies and the laws of the processes occurring with bodies
11.3.1 Ideal Gas Laws 145

in different conditions. Thermodynamics is therefore widely


used in all branches of physics.
2°. A combination of the values of a certain number of
physical quantities characterizing the physical properties
of a body (a system of bodies) determines the state of the
body (the system)—its thermodynamic state. The physical
quantities unambiguously determining the state of a body
(a system of bodies) are called thermodynamic parameters.
The latter include the temperature, density, heat capa­
city, electric resistivity, and many other physical quan­
tities.
For example, the thermodynamic state of a liquid poured
into an open vessel is determined by its density, the atmos­
pheric pressure, and the temperature. The electric resistance
of a metal conductor is characterized by the electrical prop­
erties of the conductor. The resistance will be determined
if we know the resistivity of the conductor (111.2.4.1°)
and its dimensions (length and cross-sectional area).
Two states of a body (a system of bodies) are considered
to be different if the values of even one of the thermodynamic
parameters are different for them.
3°. The state of a system is called stationary, or steady,
if it does not change with time. This signifies that none of the
thermodynamic parameters determining the state changes
with time. The steady state of a system is defined as an
equilibrium one if it is not due to phenomena occurring with
bodies that are external with respect to the given system.
If, for example, one end of a metal rod is placed in melting
ice and the other end in boiling water, the temperatures
of both ends of the rod will not change with time. Such
a steady state will not be in equilibrium, however, because
the constant temperatures are maintained by supplying
energy to the rod from the boiling water and withdrawing
energy from the rod to the melting ice. Heat exchange
(11.4.3.1°) occurs in these conditions. The steady state of
a stone on the bottom of a pit will be an equilibrium one if we
assume that no phenomena violating the state of the stone
occur in the system “Earth-stone”.
4°. The equilibrium state of a system is determined not
by all of the thermodynamic parameters, but only by some
of them called parameters of state (basic parameters of state).
146 Molecular Physics and Thermodynamics it.3 .1

The remaining thermodynamic parameters ol* a system in


the state of equilibrium depend on the parameters of state.
The basic parameters of state of a chemically homogeneous
system are the specific volume v occupied by a unit mass of
the substance, the pressure p, and the temperature T.
5°. The pressure p is defined as the physical quantity equal
to the magnitude of the force acting on a unit area of the
surface of a body at right angles to it:

If the pressure on the surface S is constant over its entire


area, then

6°. The specific volume v is the volume of a unit mass:

where p is the density (1.2.3.7°). The specific volume of a ho­


mogeneous body equals the ratio of its volume to its mass:
V
v= M

7°. The temperature T in thermodynamics is a quantity


characterizing the direction of heat exchange between bod­
ies (11.4.3.1°, see also 11.2.4.4°). When a system is in equilib­
rium, the temperature of all the bodies in it is the same.
The temperature is measured by taking advantage of the
fact that when the temperature of a body changes, almost
all its physical properties change: its length and volume,
density, elastic properties, electric conductivity, etc. The
change in any of these properties of a body (a thermometric
body) can serve as the basis for measuring the temperature
if we know the temperature dependence of the given property.
A temperature scale established with the aid of a thermo-
metric body is called empirical. The Ninth Conference Gene-
rale des Poids et Mesures (The General Conference of Weights
and Measures) in 1948 adopted the international Celsius
II.3.1 Ideal Gas Laws 14 1

temperature scale (formerly called centigrade scale) for prac­


tical use. To construct this scale, i.e. establish the tempe­
rature reference point and the jinit of temperature—the
degree Celsius (°C)— it is assumed that at standard atmo­
spheric pressure, i.e. 1.013 25 X l0 5 N/m2, the temperatures
of melting of ice and boiling of water are 0 °C and 100 °C,
respectively. The Ninth Conference Generate des Poids et
Mesures also established the absolute thermodynamic tem­
perature scale in which the temperature is measured in kelvins
(K) and is denoted by T (the temperature in the Celsius
scale is denoted by t). The relationship between the absolute
temperature T and t is
T = 273.15 + t (see also VII.4.20)

The temperature T = 0 K (or —273.15 °C) is called the


absolute zero of temperature (see also 11.4.9.4°).
8°. In the equilibrium state of a body:
(a) the parameters of its state equal the relevant para­
meters of state of the external medium (11.3.2.2°); and
(b) the parameters of state p, v, and T are identical
in all the parts of the body. For example for a gas in the
equilibrium state, its pressure and temperature are ident­
ical in all the parts of the volume of the gas at the same
heights.
9°. A relationship called the equation of state exists be­
tween the three basic parameters of state of a body. It is
written in the form of a functional relationship between p, v,
and 7\ i.e. / (p, v, T) = 0. This equation describes only
the equilibrium states of a body.
Problem. The pressure in a gas generator (a plant for
the preparation of gas) changed by 1.70 X 10~2 atm. How
did the difference between the levels of the water in a mano­
meter connected to the generator change?
Given: Ap = 1.70 X 10"2 atm = 1.72 X 103 N/m2, p =
= 103 kg/m3 (from tables).
Required: Ah.
Solution: The pressure is calculated by the formula
p = Fn/S , where Fn is the force directed normally to the
surface S. In the given case, Fn is the force of gravity of
(lie water column, and S is the cross-sectional area of the
148_________ Molecular Physics and thermodynamics________II.3J

water column. Hence,

The mass of the water is m = Fp, where F is the volume of


the water in a column of height h, V = hS, and p is the
density of the water. Thus,
P _ QhSg_ _ pgh< A p = p i — p2 = pg(hl - h 2)
whence
7 u _ Ap _ 1.72X 103
n2 — pg — 9.8x103 0.18 m

2*. Thermodynamic Processes


1°. Any change in the state of a body (a system of bodies)
is called a thermodynamic process. In any thermodynamic
process, the parameters determining the state of the body
(a system of bodies) change. For
example, an increase in the pressure
exerted on a gas is attended by a
decrease in its volume; elevation of
the temperature of a metal rod is
attended by its elongation, etc.
2°. An equilibrium (quasi-static)
process is defined as a thermody­
namic process in which a body (a sys­
tem of bodies) passes through a con­
tinuous series of equilibrium states
(11.3.1.3°). In an equilibrium process,
the state of a body must change
infinitely slowly. This signifies that
the body passes through a number of infinitely close states
of equilibrium with the external medium (the surroundings).
Example. An ideal gas is in a cylinder with a movable
piston (Fig. II.3.1). We shall consider that the force of grav­
ity of the piston and the force of its friction against the
walls of the cylinder may be disregarded. As long as the
piston is stationary, the gas is in a state of equilibrium with
11.3.2 Ideal Gas Laws 149

the surroundings. If under the action of an external force


the piston slowly [with a velocity much lower that the velo­
city of sound in a gas (IV.3.3.3°)rmoves downward, a pro
cess of compression of the gas will occur, and it can be con­
sidered as an equilibrium one. The pressure, temperature,
and density of the gas in all the parts of the cylinder volume
will be identical, and we can
disregard the propagation of
sound in the gas.
When the piston travels
with a finite velocity, a region
of increased pressure forms un­
der it. The change in the pres­
sure in the gas will spread
throughout its entire volume
with the velocity of sound.
The equality of the pressures
in all the parts of the volume
of the gas will be violated,
and the more, the higher is
the velocity of the piston. The states of the gas will not
be equilibrium ones because the condition of equilibrium
of the states (11.3.1.8°) will be violated.
3°. All real thermodynamic processes proceed with
a finite velocity and are therefore non-equilibrium ones.
Only in separate cases may the non-equilibrium nature of
real processes be disregarded.
4°. Different thermodynamic processes are studied and
compared by depicting them graphically (process charts,
thermodynamic diagrams). To do this, two changing para­
meters of state, for instance the pressure p and specific
volume v, are laid off along the axes of a two-dimensional
Cartesian coordinate system (a thermodynamic p~v diagram).
The dependence of p on v (Fig. 11.3.2) shows the given ther­
modynamic process. The points (pu v j and A 2 (p2, u2)
characterize the initial and final states of the body. The
third parameter of state, 7\ is determined by the equation
of state for each point of the thermodynamic process. The
solution of the equation of state of the body (11.3.1.9°)
relative to T makes it possible to find the relationship
T = <p (p, v).
150 Molecular Physics and Thermodynamics II.3.3

5°. Only equilibrium states and equilibrium processes


can be depicted graphically. We cannot speak of the para­
meters of state p, v, and T of an entire body for non-equilib­
rium processes because condition (b) for equilibrium of
its state (11.3.1.8°) is violated, and the equation of state
of the body loses its meaning.

3. Laws of Isoprocesses in Ideal Gases.


Equation of State of an Ideal Gas
1°. Isoprocesses are defined as thermodynamic processes
occurring in a system having a constant mass at a constant
value of one of the parameters of state of the system.
An isochoric process is a thermodynamic process occur­
ring at a constant volume of the system (V = const).
An isobaric process is one proceeding at a constant pres­
sure (p = const).
An isothermal process is one proceeding at a constant
temperature (T = const).
2°. An isothermal process in an ideal gas obeys Boyle s
law: for a given mass of a gas at a fixed temperature, the
product of the numerical values of the pressure and volume
is a constant quantity:
pV = const
In a thermodynamic p-v diagram, an isothermal process is
depicted by a curve called an isotherm Fig. (II.3.3).
3°. Gay-Lussac's law holds for an isobaric process in an
ideal gas: at a constant pressure, the volume of a given
mass of a gas is directly proportional to its absolute tem­
perature:
F = F„(1 +<M ), or F = a vF07’ = F0—
10
where V0 = volume of the gas at the temperature T0 =
= 273.15 K
a v = I/T q — 1/273.15 K-1 — cubic expansion coeffi­
cient (at constant pressure), which is considered
to be the same for all ideal gases. .
11,3,3 Ideal Gas Laws 151

It follows from the formula a v = V/V0T that a v cha­


racterizes the relative increase in the volume of a gas when
its temperature changes hy one kelvin. In a V~T diagram,
an isobaric process is depicted by a segment of a straight
line AB (Fig. II.3.4). The line cannot be extended to the

Fig. I I .3.4

region of low absolute temperatures close to the degeneracy


temperature of the gas (VI. 1.8.1°).
4°. An isochoric process in an ideal gas is described by
Charles' law: at a constant volume, the pressure of a given
mass of a gas is directly proportional to its absolute tem­
perature;

P = Po(l +«;><), or p = a pp0T = p0~


10
where p 0 = pressure of the gas at the temperature T0 =
= 273.15 K
a p = p/PoT = thermal pressure coefficient charac­
terizing the relative increase in the pressure of
a gas when it is heated by one kelvin.
In a p~T diagram, an isochoric process is depicted by
a segment of a straight line BC (Fig. II.3.5) that cannot
he extended to the region of low temperatures close to the
degeneracy temperature (VI. 1.8.1°) where the laws of ideal
gases do not apply.
Problem 1. A glass flask 10 cm3 in volume with a nar­
row neck was heated to 114 °C, and then the neck of the
152 M o le c u la r P h ysics and T h e rm o d yn a m ics 11.3.3

flask was immersed in mercury. When the air in the flask


cooled, 36 g of mercury entered it. To what temperature did
the air cool? Consider that the
density of the mercury equals
13.6 X 103 kg/m3.
Given: Vx = 10 cm3= 1 0 x
X 10"6 m3, Tx = 273 +
+ 114 = 391 K, p = 13.6 x
X 103 kg/m3, m — 36.0 X
X 10~3 kg.
Required: T2.
Solution: The process of
cooling the air is isobaric be­
cause the air and the mercury
are under atmospheric pres­
sure. According to Gay-Lussac’s law, V1/T 1 = V2lT 2, whence
T2 = W V v
The volume of the cold air V2 is determined as the differ­
ence between the volumes of the flask and of the mercury
entering it, i.e. V2 = Vx — VmeT. The volume of the mer­
cury T^mer = m!p. Hence,
36.0 X 10"3 \
( 10 X 10-6 13.6X 103 )
X 391
288 K
t 2= 10 x 10-8
Problem 2. The air pressure in a cylinder of a constant
volume at 7 °C was 0.1515 MPa. When the air was heated
to 100 °C, the pressure increased to 0.2020 MPa. Determine
the thermal pressure coefficient.
Given: tx = 7 °C, p x = 0.1515 MPa, t2 = 100 °C, p 2 =
= 0.2020 MPa.
Required: a p.
Solution: The heating process is isochoric, V = const.
According to Charles’ law px = p 0 (1 + ocp^) and p 2 =
= p Q(1 + ocpt2), where p 0 is the air pressure at 0 °C. Joint
solution of these two equations yields
= P2 — P1 0.2020 — 0.1515
= 3 .6 7xl0" 3 A'"1
Pi^2 P2^1 0.1515 X 100 — 0.2020 X 7.00
Problem 3. The outdoor temperature is —13 °C, while
indoors it is 22 °C. By how much will the pressure in a gas
II.3.3 Ideal Gas Laws 153

cylinder change if it is brought indoors. The manometer


on the cylinder showed a pressure of 1.50 MPa indoors.
Given: Tl = 273 - 13 = 260 K, T2 = 273 + 22 =
= 295 K, p 2 = 1.50 MPa.
Required: Ap.
Solution: The process of heating the gas in the closed
cylinder is isochoric, V = const. According to Charles’
law, plT = const. The change in the pressure when the
gas is heated is Ap = p 2 — Pi, or p x = p 2 — Ap.
The equation of Charles’ law for two states is p 2lT 2 =
= (p2 — ApJ/Tj, whence
Ap = Pa {T$-
j 2 T') ^ 1,50 {2^
^yo~ 260) » 0.18 MPa
5°. An adiabatic process is a thermodynamic process
carried out in a system without its heat exchange with
external bodies (11.4.3.1°).
The very rapid expansion of a gas contained in a ther­
mally insulated vessel into the atmosphere when the cock
is opened quickly is an example of an adiabatic process.

Fig. II.3.6 Fig. II.3.7

The adiabatic process should proceed so quickly that when


the cock A is opened (Fig. 11.3.6) there will be no time for
heat exchange between the gas and the atmosphere to occur. v
An adiabatic process of expansion of a gas can proceed only
at the expense of a reduction in the internal energy of the
gas (11.4.5.4°). The pressure of the gas drops sharply, and
its temperature lowers.
154 M o le c u la r P h y s ic s a n d T h e r m o d y n a m ic s II.3.S

An adiabatic process is shown in a p-v diagram by


a curve called an adiabat (Fig. II.3.7).
6°. Avogadro's law: equal volumes of different gases con­
tain the same number of molecules at identical pressures and
temperatures. In other words, moles of different ideal gases
occupy equal volumes at* identical pressures and tempera­
tures (see also 11.1.1.6°). The number of molecules in 1 cm3
of an ideal gas in standard conditions (t = 0°C, p =
= 1.013 25 X 105 N/m2) is called the Loschmidt number
and equals 2.687 X 10lfl cm’3.
7°. The equation of state of an ideal gas for one mole has
the form
PVm = RT
where p, Vm, and T are the pressure, molar volume, and
absolute temperature of the gas, respectively, and R is
the molar gas constant, numerically equal to the work done
by one mole of an ideal gas in the isobaric elevation of its
temperature by one kelvin:
d P (V2m— ^im)

where V2m and F1m are the final and initial volumes of
a mole of a gas. The values of R in different units are:
R = 8.31 X 103 J/kmol*K = 8.31 J/mol-K =
= 0.0821 1-atm/mol-K = 1.99 cal/mol*K
The volume of an arbitrary mass M of a gas of molar
mass (lx is V = (Ml\\) Vm. The Mendeleeu-Clapeyron equation
for an arbitrary mass of a gas has the form
pV = — RT
Since VlM = v is the specific volume of a gas, then

pv = — T = BT

where R = 7?/p is the specific gas constant depending on


the molar mass of a gas.
8°. Another form of the equation of state of an ideal gas is
p = n kj1
II.3.3 Ideal Gas Laws 155

where n = NpL!Vm = concentration of the gas, i.e. the num­


ber of particles in a unit volume of the gas
N a = Avogadro constant #(11.1-1-5°)
fc = Boltzmann constant (11.2.4.4°).
Problem 1. How many particles of air are in a room
with an area of 20 m2 and a height of 3 m at a temperature
of 17 °C and a pressure of 752 mmHg?
Given: S = 20 m2, h = 3m, T = 17 + 273 = 290 K,
p = 752 mmHg = 752 X 133 » 10B N/m2.
Required: N.
Solution: We find the number of particles in a unit volume
by the formula n = plkT , where p is the pressure, T is
the absolute temperature, and k is the Boltzmann constant.
The total number of particles N = nV, where V is the
volume’of the room, V = Sh. Hence, the number of par­
ticles is
106 x 20 x 3
N = nV = — Sh 15 X 1026
1.38 X 10-23 X 290

Problem 2. By how much did the pressure of oxygen in


a 100 -litrecylinder drop if 3 kg of a gas were withdrawn from
it. The gas temperature of 17 °C remained constant.
Given: V = 100 1 = 0.1 m3, AM = M 2 - M x = - 3 kg,
T = 290 K.
Required: Ap = p 2 —
Solution: The Mendeleev-Clapeyron equation for
two states of a gas gives us p 1V1 = ( M j p) R Tiand p 2V2 =
= (M 2/\i ) R T 2.
According to the initial conditions, 7\ = T2 = T and
V1 = V2 = V. Using these values in the above equations
for the two states and subtracting the first equation from
the second one, we get (p2 — pj) V = (M 2 — M J /?77|li,
whence Ap = AMRT!V\i, where p is the molar mass of
oxygen equal to 32 X 10“3 kg/mol. Hence,
a — 3 x 8 . 3 1 X 290 q np i /-kg tvj t o
= 0 . 1 X 3 2 X 10 ^ = - 2 • 26 X 10« N/m-

Problem 3. A 100-litre cylinder contains 8.9 X 10“3 kg


of a gas in standard conditions. Find the molar mass of this
?as*
156 Molecular Physics and Thermodynamics II.4.1

Given: V — 100 1 = 0.1 m8, p = 760 mmHg = 1.01 X


X 10* N/m2, T = 273 K, and M = 8.9 X 10"8 kg.
Required: ji.
Solution: From the Mendeleev-Clapeyron equation pV =
= (M /\i) RT, we get
MRT
^= —
8 . 9 X l 0 - * X 8 . 3 1 X 273
1.01 X l 0 5 X 0.1
= 2 .0 X10“8 kg/mol

Chapter 4
Fundamentals of Thermodynamics

1. Total and Internal Energy of a Body


(a System of Bodies)
1°. The total energy % of a system consists of
(a) the kinetic energy Ek of its macroscopic motion as
a whole,
(b) the potential energy Ev, ext due to the presence of
external force fields, for example an electric or gravitational
field; and
c) the internal energy U.
The total energy % is
^ + ^ p , ext + U

2°. The internal energy of a body (a system of bodies) U is


defined as the energy depending only on the thermodynamic
state (11.3.1.2°) of the body or system of bodies.
The internal energy of the system depends on the nature
of the motion and interaction of the particles in the system.
It consists of the following parts:
(a) the kinetic energy of thermal chaotic motion of the
particles forming the system (molecules, atoms, ions, etc.);
(b) the potential energy of the particles due to the forces
pf their interpaolecular interaction;
tlA.i fundamentals of Thermodynamics 157

(c) the energy of the electrons in the electron shells


(VI.2.8.6°) of the atoms and ions; and
(d) the nuclear energy (VIA .2.1°).
3°. All the parts of the internal energy depend on the
state of a system. The internal energy is determined by the
thermodynamic state of a system and does not depend on
how the system acquired the given state. Consequently, the
internal energy is not associated with the process of a change
in the state of a system. In two or more identical states of
a system (11.3.1.2°), its internal energy is the same.
The internal energy is counted from the state of a system
in which U equals zero. It is generally assumed that the
internal energy of a system equals zero at the absolute zero
of temperature (T = 0). Of practical interest, however, is
not the internal energy itself, but its change A U when a sys­
tem passes from one state to another. Therefore, the choice
of the reference point for internal energy is of no significance
[see the arbitrary nature of the choice of the reference point
for the potential energy in mechanics (1.5.3.5°)].
4°. When we may disregard the change in the parts (c)
and (d) of the internal energy (Par. 2°), the internal energy
of a system consists of the kinetic energy of the thermal
motion of the substance particles,and their potential energy.
Example 1. The internal energy of one mole of a monato­
mic ideal gas depends only on its absolute temperature T:

U = — RT

where R is the molar gas constant (11.3.3.7°).


The change in the internal energy of a mole of such
a gas AZ7 when it passes from one state to another is pro­
portional to the change in the temperature AT7:

Af/=r/2-c/1= -|/?7’2 -T fl7’1=T i?A7’


where AT = T2 — TV
Example 2. The internal energy of a deformed spring
is its potential energy (1.5.3.6°):
158 Molecular Physics and thermodynamics II.4.2

where k — spring constant (1.2.9.4°)


x = linear deformation (tension or compression)
(11.7.2.4°).
The change in the internal energy of a spring when it is
stretched (compressed) from x1 to x2 is
A/T kx\ kxi
a 2 2

2., W ork
1°. Our concepts of work and energy treated in mechanics
(1.5.1.1°, 1.5.3.2°) are further developed in thermodynamics.
An essential condition for the performance of work by a body
(or a system) is displacement of the body when acting forces
are present. We can speak of work only when the state of
a body (or a system) changes.
We distinguish the work A done by a system on external
bodies and the work A' done by external bodies on a system.
They are numerically equal and opposite in sign: A =
= —A '. The work A is assumed to be positive, and A*
negative.
2°. The work of expansion of an ideal gas is the work which
the gas does against the external pressure (see Fig. II.3.1).
The elementary work AA is determined by the formula
AA = pS Ax = p AV
where p = external pressure
S = area of the piston under which the gas is
Ax = elementary displacement of the piston from the
position 1 to the position 2
AV = S Ax — change in the volume of the gas.
The same formula expresses the elementary work done
not only by a gas, but also by any body against the external
pressure.
An expanding gas does positive work against the external
forces (AV > 0). Compression of a gas is attended by nega­
tive work (AV < 0). It is done by the external bodies that
created the external pressure.
II.4.2 Fundamentals of Thermodynamics 159

3°. The work of expansion when the volume of a gas


(or any body) changes from Vx to V2 equals the sum of the
elementary works: •
A = %p AV
For example, the work of expansion in an isobaric process
(11.3.3.1°) in which p = const is
A = P (V2 - V,)
where V1 and V2 are the initial and final volumes of the body,
respectively.
In an isobaric process, the work of expansion is depicted
in a p-V diagram by the area of the hatched rectangle
(Fig. II.4.1).
In a p-V diagram, the work of expansion in any process is
measured by the area confined by the curve of the process,

the axis of abscissas, and the vertical straight lines V = Vx


and V = V2 (in Fig. II.4.2 these areas are hatched).
4°. The work of expansion done by a body (a system of
bodies) depends on the nature of the process of the change in
its state. This can be seen in Fig. II.4.2 where the areas
under the curves of the processes 1-1-2 and 1-II-2 are differ­
ent. This underlies the action of heat engines (11.4.10.1°).
5°. The work done by a system in a process is a measure
of the change in its energy in the process (see also 11.4.9.3°).
When work is done, the energy of the ordered motion of one
body transfers into the energy of the ordered motion of an­
other body (or its parts). For example, a gas expanding in
160 Molecular Physics and Thermodynamics 11.4.3

a cylinder of an internal-combustion engine moves a piston


and transfers energy to it. Electric motors driving various
machines also transfer energy.
Problem 1. Determine the work of expansion of 20
litres of a gas in isobaric heating from 300 to 393 K. The
pressure of the gas is 80 kPa.
Given: = 20 1 = 20 X 10"3 m3, Tx = 300 K, T2 =
= 393 K, p = 80 kPa = 80 X 103 N/m2.
Required: A .
Solution: The work of expansion of a gas in isobaric
heating is A = p (V2 — Fj).
Since in an isobaric process V1IT1 = V JT 2, then V2 =
= V1T2IT1. Hence,

A — pVi — 1) = 80 x 103x 20 x 10~3 — 1) = 496 J


Problem 2. Nitrogen having a mass of 280 g was heated
at a constant pressure by 100 K. Determine the work of expan­
sion.
Given: M = 280 g = 0.28 kg, A77= 100 K, p = 28 x
X 10"3 kg/mol.
Required: The work of expansion.
Solution: The work A = p (V2 — Fi). The volumes Fx
and F2 are found from the Mendeleev-Clapeyron equation:
pVx = (M/\x) R TU whence Fx = (M /\i) (R T Jp); similarly
F2 = ( M/ p ) (R T jp ). Hence,
A = p (iL_2Z!2— 3 h . ) —E . R {T Ti) —— R AT =
^ \ H /» g P I H M-

= 2 8 S ^ X 8 -31 x l0 0 = 8 -31 x l ° 3 J

3. H e a t
1°. Heat is such a form of energy transfer in which the
direct exchange of energy occurs between the chaotically
moving particles of the interacting bodies. As a result of the
energy transferred to a body, the disordered motion of its
particles increases, i.e. the internal energy of the body grows.
For example, when two bodies having different temperatures
11.4.3 Fundamentals of Thermodynamics 161

are brought into contact, the particles of the warmer body


transfer part of their energy to the particles of the cooler
body. The internal energy of the first body diminishes, and
of the second one grows, while their temperatures level out.
The process of transfer of internal energy without work
being done is called heat exchange. A measure of the energy
transferred in the form of heat in a heat exchange process
is the quantity called the amount of heat.
In the conventionally adopted terminology, the terms
“work” and “heat” have a double meaning. On the one hand,
work and heat are two different forms of transferring energy,
and on the other, measures of the transferred energy. It is
more correct in the second case to speak not of “the work done”
or of “the amount of imparted heat”, but of the energy trans­
ferred, respectively, in the form of work or in the form of
heat. Where this will not cause confusion, the short terms
will also be used in the present book.
2°. Heat, like work, is not a kind of energy, but a form
of its transfer. Therefore use of the expression “the store of
heat in a body” is wrong. The error consists in confusing
one of the kinds of energy (internal energy) with a form of
energy transfer—heat.
Different amounts of heat must be imparted to a body to
transfer it from one state to another depending on the inter­
mediate states which it passes through. For instance, the
heating of a given mass of a gas by AT K in an isochoric
process (11.3.3.4°) requires a smaller amount of heat than
the same heating in an isobaric process (11.3.3.3°). Heat,
unlike energy, is not a function of state of a system, but
depends on the process of changing of this state.
3°. Heat and work have the common property that they
exist only in a process of energy transfer, and their numerical
values (Par. 1°) depend on the kind of this process.
Heat and work are not equivalent forms of energy transfer
from a qualitative viewpoint. The energy of ordered motion
is transferred in the form of work. If work is done on a body,
it may lead to an increase in any kind of energy of the given
body or other bodies. Assume, for example, that a moving
ball collides inelastically with a wall. After the inelastic
collision, the body stops, and all the energy of its ordered
motion transforms into a change in the internal energy of the
162__________Molecular Physics and Thermodynamics________II.4.4

body and the wall. The energy of llie ordered motion of the
body is transformed into the energy of disordered motion
ol‘ the particles. Such a transfer of energy is irreversible
(11.4.9.3°).
If energy is transferred to a body in the form of heat,
this increases the energy of chaotic thermal motion of its
particles and directly leads only to a growth in the internal
energy of the body. For instance, when a gas contained in
a vessel of constant volume is heated, the velocity of its
molecules increases, and its internal energy grows.
For an increase in kinds of energy other than the internal
energy to occur when heat is supplied to a body, at least
partial transformation of the chaotic motion of the particles
of the body into ordered motion is needed, or as we often
but not accurately say, the “transformation of heat into
work”. This occurs in heat engines (11.4.10.1°, see also
11.4.9.3°).
4°. In real conditions, both ways of transferring energy
to a system (in the form of work and in the form of heat)
accompany each other. For example when a metal rod is
heated, its internal energy grows, and at the same time ther­
mal expansion of the rod occurs (11.7.3.1°) which means that
work of expansion is done.
In the International system of units (SI), the amount
of heat, like work, is measured in joules (J). A non-system
unit—the calorie—is also used to measure the amount of
heat. One calorie (VII.4.1°) is equivalent to 4.19 joules of
work. The mechanical equivalent of heat is
E = 4.19 J/cal - 4.19 X 107 erg/cal
The reciprocal quantity, 1IE = 0.239 cal/J, is called the
thermal equivalent of mechanical work.

4. H e a t C a p a c ity
1°. Heat capacity is the name given to the physical
quantity numerically equal to the amount of heat AQ
that must be added to a body to heat it by one kelvin:
11.4.5 Fundamentals of Thermodynamics 163

The heat capacity of a body depends on its mass, chemical


composition, the thermodynamic state of the body (11.3.1.2°),
and the kind of process in which aiergy is transferred to the
body in the form of heat. Different amounts of heat are
needed to heat a given mass of a gas by one kelvin if heating is
conducted in different conditions, for example at a constant
volume or at a constant pressure. In the second case, a
greater amount of heat is needed (11.4.3.1°).
It follows from the definition of heat capacity that in an
adiabatic process when AQ = 0 the heat capacity equals
zero.
In an isothermal process (AJ = 0), the concept of heat
capacity has no meaning (C = oo). Thus, the temperature
does not change when liquids boil (11.5.3.1°) or solids melt
(11.7.4.1°), i.e. AT = 0, and there is no sense in using the
concept of heat capacity here.
2°. The specific heat capacity c is the heat capacity of
a unit mass of a homogeneous substance: c = C/Af, where M
is the mass of the substance.
The specific heat capacity of a body is not a constant
quantity, and tables of heat capacities indicate the condi­
tions for which the tabulated data hold.
The molar heat capacity Cm is the heat capacity of one
mole of a substance: Cm = cp, where p is the molar mass
of the substance (11.1.1.7°).
3°. The amount of heat AQ needed to heat a body from
the temperature T to T + AT is
AQ = C AT
For a body of mass A/, we get AQ = Me AT = (M/\x) Cm AT
where A/Yp is the number of moles—the amount of the sub-
stance (11.1.1.7°).

5. The First Law of Thermodynamics


1°. The change in the internal energy AC/ of a body (sys­
tem) when it passes from one state to another equals
the sum of the work A' done on the body and the amount of
beat. AQ it received:
AU = A' + AQ
164_________ Molecular Physics and Thermodynamics_______ 11.4.5

This version of the first law of thermodynamics takes into


consideration the existence of two forms of energy transfer—
work and heat.,
2°. If we replace the work A' done on a body (system) by
external forces with the work A = —A \ numerically equal
to A ' but opposite in sign, done by the body (system) itself on
external bodies, then
AQ = AC/ + A
The amount of heat AQ received by a body (system) is used
to change the internal energy A U and for the work A of the
body (system) against external forces (a different version
of the first law of thermodynamics).
For example, if a certain amount of heat is added to a gas
contained in a vessel under a piston, then it can be spent for
elevating the temperature of the gas, i.e. for increasing its
internal energy and for work against the external pressure
p when the piston moves from position 1 to 2 (see Fig. 11.3.1).
3°. The first law of thermodynamics as applied to an iso­
thermal process (11.3.3.1°) in an ideal gas has the form
AQ = A. The internal energy of the ideal gas will not change
because AT1 = T2 — Tx = 0, and AU = 0 (11.4.1.4°). All
the added amount of heat goes for work against the external
pressure.
In an isochoric process (11.3.3.4°), the amount of heat AQ
goes only to increase the internal energy of the gas A £/,
i.e. AQ = AC/. Consequently, the temperature of the gas
increases by AT. The gas does no work of expansion (A = 0).
In an isobaric process (11.3.3.3°), the amount of heat sup­
plied to a gas also goes to increase the internal energy and to
do the work of expansion of the gas against the external
pressure: AQ = AU + A.
4°. In an adiabatic process (11.3.3.5°), AQ = 0, and
the first law of thermodynamics acquires the form A =
= —AC/. In conditions of the absence of heat exchange with
the surroundings, the work done by a body against external
forces is at the expense of a reduction in its internal energy.
For example, if an ideal gas expands adiabatically, overcom­
ing the external pressure, then the work of expansion of
the gas is attended by diminishing of its internal energy
and cooling.
11 .4 .5 Fundamentals of Thermodynamics 165

If external forces do work on a gas in adiabatic condi­


tions, then it is compressed, its internal energy grows, and
heating of the gas occurs. •
5°. The heat capacity of a gas in an isobaric process
exceeds its heat capacity in an isochoric process by the amount
of the work of expansion done by the gas against the external
pressure. The difference between these heat capacities for
one mole of an ideal gas equals the molar gas constant R
(11.3.3.7°).
6. The first law of thermodynamics asserts that it is
impossible to construct such a periodically functioning
machine that would do more work than the energy supplied
to it from outside: a perpetual motion machine of the first
kind is impossible.
Indeed, if a gas, vapour or some other working substance*
performs a cyclic process (11.4.7.1°) in a periodically func­
tioning machine, then its internal energy does not change
(11.4.1.3°), AC/ = 0, and A = AQ (Par. 2°). The work done
by this substance cannot exceed the amount of heat supplied
to it.
Problem 1. A kilomole of a monatomic gas is heated by
100 K at a constant volume. Find the amount of heat added
1o the gas.
r' 'Given: AT = 100 K, V = const.
^Required: AQ.
Solution: No work is done when the gas is heated in the
conditions of a constant volume. All the heat added to the
gas is spent for increasing its internal energy, i.e. AQ =
-= AC/.
U = (3/2) kTNA, where N A is the Avogadro constant, or
V = (3/2) RT.
If the gas is heated by AT = T2 — 7\, then
AC/ = C/2 — Ui = (3/2) R AT. Hence,
A<?= (3/2) R AT = (3/2) x 8.31 x 103x 100 =
= 1.16 x 10° J= 1.16 MJ

* By a working substance (or medium) in thermodynamics is meant


n gas or some other thermodynamic body that performs a cyclic process
nnd exchanges energy with other bodies or media.
166 M olecular Physics and Therm odynam ics 1 1.4.5

Problem 2. Air in a room having a capacity of 90 m3


is heated by 10 K. What volume of hot water must pass
through the radiators of the water heating system? The
water cools by 20 K. The losses of heat are 50%.
Given: V1 = 90 m3, AT, = 10 K, AT2 = 20 K, r) =
= 50%.
Required: V2.
Solution: The amount of heat received by the air is
= C\P\V\ AT,. The amount of heat given up by the water
is Q2 = c2p2 V2 AT2, where c, is the specific heat capacity of
the air, c, « 103 J/kg-K, p, is the density of air, p, =
= 1.29 kg/m3, p2 is the density of water, p2 = 1 X 103kg/m3,
Co is the specific heat capacity of water, c2 = 4 .1 8 X
X 103 J/kg-K.
With a view to the losses of energy, we write Q, = t)()2,
or ^ p !^ AT, = y\c2p2V2 AT2. Hence,
y __ c\P \V \k T , 103 X 1.29 X 90 X 10
27.7 X l0 ’3 m3
2 "" T]c2p2A712 0 . 5 x 4 .1 8 X 103 X103 X 20
Problem 3. Air in a volume of 0.50 m3 is under a pres­
sure of 0.10 MPa. Determine the amount of heat needed for
isochoric heating of the air from 30 to 130 °C. The specific
heat capacity of the air is c = 0.71 X 103 J/kg-K. The
average molar mass of the air particles is to be taken equal
to 29 X 10“3 kg/mol.
Given: V = 0.50 m3, p = 0.10 MPa = 105 N/m2, T, =
= 273 -f 30 = 303 K, T2 = 273 + 130 = 403 K, p =
= 29 X 10”3 kg/mol, c = 0.71 X 103 J/kg-K.
FRequired: AQ.
Solution: The air is heated at a constant volume. The
amount of heat needed for heating the air is AQ = cM (T2 —
— Tj), where Afusthe mass of the^air. It is found from the
equation pV = R T T Hence,
AQ = c ^ ( T 2- T i) =

= 0.71 x 103 x 29x °— x 100 = 4 -0fi x 104 J


Problem 4. What amount of heat is needed to heat
64 g of oxygen by 50 K at a constant pressure? What amount
of heat is used for increasing the internal energy and for
11.4.5 F undam entals of Therm odynam ics 167

doing the work of expansion? (We take the specific heat


capacities of oxygen from tables.)
Given: M = 64 X 10“3kg, p, 32 X 10“3kg/mol, AT =
= 50 K, cp = 9.20 X 102 J/kg-K, cv = 6.53 X 102 J/kg-K.
Required: (?x, Q2, A.
Solution: The amount of heat absorbed by a gas in isobar-
ic heating is ()x = cpM AT, where cp is the specific heat
capacity of oxygen at a constant pressure. Hence,

Qx- 9.20 x 102 x 64 x 10"3 x 50 = 2.94 x 103 J

The amount of heat spent for increasing the internal


energy is Q2 = cvM AT, where cv is the specific heat
capacity of oxygen at a constant volume. Hence,

~<?2 = 6.53 x 102 x 64 x i 0-3 x 50 = 2.08 x 103 J

The amount of heat spent for the work of expansion is

}A = Q , - Q 2 = 2 M x 103- 2 .08 x ! 0 3= 0 .86 x 103 J

Problem 5. Nitrogen is heated at a constant pressure of


100 kPa. The volume of the nitrogen changes by 1.50 m3.
Find: (1) the work of expansion, (2) the amount of heat
added to the gas, and (3) the change in the internal energy
of the gas.
Given: p = 100 kPa = 105 N/m2, AV = 1.50 m3. From
the relevant tables, we find the molar heat capacity of
nitrogen at a constant volume Cv = 5 cal/mol-K =
= 20.90 J/mol*K and at a constant pressure Cp =
- 7 cal/mol-K = 29.30 J/mol-K.
Required: A , Q, AC/.
Solution: (1) The work of expansion of the gas at p =
= const is
A = p (y z — V\) = p AV = 105 x 1.50 = 1.50 x 10BJ

(2) Q — cpM AT, where cp is the specific heat capacity


at a constant pressure, cp = Cp/p , where p, is the molar
mass of the gas. Consequently, Q = (Cp/\i) M AT,
168 M olecular Physics and Therm odynam ics 1 1 .4 .6

We use the Mendeleev-Clapeyron equation for two states


of the gas: = (Ml\x)R T1 and pV 2 = (M ljx) RT^. Sub­
tracting the first equation from the second one yields
p{y 2- \ V i) = M - ( T z - T i)R
pA V
R

\i
AT,’ or —
p,
AT = 4R -
Hence,
1.50 X 105
Q = Cp^ - = 29.30x 8.31
:5.25 x 105 J
(3) The change in the internal energy equals the amount
of heat needed for heating the gas by A77 at a constant vol­
ume, AU = Qi, and Ql = CV (M!\i) A7\ or Q1 = Cv (AIR);
hence,
AU = 20.90 x 1 8 ° I t 0>'= 3-75 X 105 J

6*. Reversible and Irreversible Processes


1°. A thermodynamic process (11.3.2.1°) is called revers-
ible (a reversible process) if it permits returning of the body
(system) to its initial state without any changes remaining
in the surroundings. A process is reversible if when it is con­
ducted first in the forward and then in the reverse direction
(11.4.7.3°) both the body (or system) itself and all the exter­
nal bodies which it interacted with return to their initial
states. An essential and sufficient condition for the rever­
sibility of a thermodynamic process is its equilibrium nature
(11.3.2.2°).
Example 1. A perfectly elastic ball drops in a vacuum
onto a perfectly elastic plate. After dropping, as follows
from the laws of a perfectly elastic collision (1.5.4.1°),
the ball will return to its initial position, passing in the
reverse direction through all the intermediate states that
it passed through when dropping. Upon completion of the
process, the ball and the plate return to their initial states.
Example 2. The undamped oscillations (IV. 1.1.5°) per­
formed j[in a vacuum by a body suspended on a perfectly
elastic spring. Friction is absent in the system, and the oscil-
11.4.7 Fundam entals of Therm odynam ics 169

lations occur under the action of the force of gravity of the


body and the reaction of the elastically deformed spring.
When the time T equal to tha period of the oscillations
(IV. 1.1.3°) elapses, the mutual arrangement and the velocity
of the body, the spring, and the Earth repeat completely.
During the time T the closed system (1.2.2.5°) “body-Earth”
returns to its initial state. An oscillating process in an isolat­
ed system does not change the state of other bodies. Conse­
quently, it is reversible.
2°. Any thermodynamic process that does not comply
with the conditions of reversibility (Par. 1°) is called an
irreversible thermodynamic process. All real processes occur
not infinitely slowly, but at a finite rate. They are all attend­
ed by friction (1.2.10.1°), diffusion (11.1.3.1°), and heat
exchange (11.4.3.1°) with a finite difference between the
temperatures of the body (system) and the surroundings.
This is why all real processes are non-equilibrium ones
(11.3.2.3°). Consequently, all real processes are irreversible.

7*. Cyclic Processes (Cycles)


1°. By a cyclic process, or cycle, is meant a thermody­
namic process as a result of which the working substance
returns to its initial state. In
phase (equilibrium) diagrams
p-V, p-T, and the like
(11.3.2.4°), cyclic processes are
depicted in the form of closed
curves because in any diagram
two identical states are shown
by the same point on a plane.
Cyclic processes are the ba­
sis of heat engines (11.4.10.1°).
2°. The work against the
external pressure done by the Fig. 11.4.3
working substance in an arbit­
rary cyclic process is measured
by the area confined by the curve c1ac2bc1 of the process
(Fig. II.4.3). Upon expansion along the curve cxac2, the
substance does the positive work A x measured by the area
of the figure F1c1ac2V2. Upon compression along the curve
170 M olecular Physics and Therm odynam ics 1 1 .4 .8

c2bc1? the external forces do the positive work A 2 measured


by the area of the figure V1cJbc2V2. Inspection of Fig. II.4.3
shows that A'2. During the entire cycle, the working
substance does the positive work A = A-^ + A 2 = A l — A 2
measured by the hatched area in Fig. II.4.3 (A2 = —A'2).
3°. A direct, or forward, cycle is defined as a cyclic pro­
cess in which the working substance does positive work
at the expense of the heat added to it. In a p-V diagram,
a direct cycle is depicted by a closed curve circumvented
clockwise (see Fig. II.4.3).
A reverse, or backward, cycle is defined as a cyclic process
in which the working substance does negative work. This
signifies that work is done on the working substance and
an amount of heat equivalent to it is withdrawn from the
substance. In a p-V diagram, a reverse cycle is depicted
by a closed curve circumvented counterclockwise.
In a heat engine (11.4.10.1°), the working substance
(11.4.5.6°) performs a direct cycle, and in a refrigerator
(II.4 11.1°), a reverse cycle.

8*. Carnot Cycle


1°. A Carnot cycle is a direct reversible cyclic process
(Fig. 11.4.4) consisting of two isotherms 1-1' and 2-2' and
two adiabats 1-2 and V-2’. In
isothermal expansion 1-1',
the working substance receives
from the! hot reservoir (heat
source)—a source of jenergy
having the constant tempera­
ture 7^—the amount of heat
Qx. In isothermal compression
2'-2, the working substance
gives up the amount of heat
Q2 to the cold reservoir (heat
sink) having the constant tem­
perature T2 (T2C Ti). In adi­
abatic expansion f'and com­
pression, no energy is given up to or received from the sur­
roundings by the working substance, and these processes
take place at the expense of its internal energy (11.4.5.4°).
1 1 .4 .8 Fundam entals of Therm odynam ics 171

2°. The thermal (thermodynamic) efficiency of an arbit­


rary cycle is the ratio of the work A done by the working
substance in a direct cycle to frho amount of heat trans­
ferred to the working substance by the hot reservoir:
A Q1 (?2
Qi Qi
3°. The thermal efficiency of a reversible Carnot cycle
is independent of the nature of the working substance and
is determined only by the temperatures of the hot reservoir
T1 and the cold reservoir T2:

The efficiency tig is less than unity because it is impossible


in practice to ensure the condition T1 oo and theoretically
impossible to design a cold reservoir in which T« = 0
(11.4.9.4°).
4°. The thermal efficiency r]rev of an arbitrary reversible
cycle cannot exceed the thermal efficiency of a reversible
Carnot cycle conducted between the same temperatures
Tx and T2 of the hot and cold reservoirs:
Tr~T2
'Hrev Tl
5°. The thermal efficiency r)ir of an arbitrary irreversible
cycle is always less than the thermal efficiency of a revers­
ible Carnot cycle conducted between the temperatures Tx and
u
t , - t 2
T1
Paragraphs 3° to 5° form the content of Carnot's theorems
(see also 11.4.9.2°).
6°. In a reversible Carnot cycle, the ratio of the tempera­
tures of the hot reservoir and the cold reservoir equals
the ratio of the amounts of heat transferred from and to them
in the cycle, respectively:

T2 Q2
172 M olecular Physics and Therm odynam ics n .4 .8

This ratio can be used as the basis for comparing the tempe­
ratures of two bodies. If the latter are chosen as the hot
and cold reservoirs in a reversible Carnot cycle, then by
measuring and Q2 we can determine the ratio T J T 2.
The thermodynamic temperature scale is established theoret­
ically in this way. In accordance with Carnot’s theorems
(Paragraphs 3°-5°), this scale is not associated with the prop­
erties of a thermometric body (see also 11.3.1.7°).
Problem 1. An internal-combustion engine has an effi­
ciency of 28 % at a fuel combustion temperature of 927 °C
and an exhaust gas temperature of 447 °C. By how much
is the efficiency of an ideal engine greater than that of this
one?
Given: = 927 + 273 = 1200 K, T2 = 447 + 273 =
- 720 K, T] = 28%.
Required: T)c, t|.
Solution: The efficiency of an ideal engine operating
according to the Carnot cycle is found by the formula
1200—720
*lc X 100 = 40%
1200
TJC—T1= 40% - 28% = 12%

Problem 2. The hot reservoir of a heat engine operating


according to the ideal Carnot cycle receives 2 X 103 calo­
ries and transfers 80% of them to the cold reservoir. Find
the efficiency of the cycle and the work done by the engine.
Given: Qx = 2 kcal, Q2 = 0.8(^.
Required:
Solution: The~efficiency of the cycle is

T1 = 4 - = Q ' n Qt x l O O - f t - 0- ^ X 1 0 0 =
V VI VI

= (1 -0.8) x 100= 20%


The work A = <?, — <?2 = (V i = 0.2 X 2 X 103 = 4 X
X 102 cal. Since 1 cal = 4.18 J, we have

A = 4 X 102 X 4.18 = 1.67 kJ


iiA .9 Fundamentals of Thermodynamics 173

9*. The Second and Third Laws


of Thermodynamics

1°. The first law of thermodynamics (11.4.5.2°) does not


make it possible to determine the direction in which a ther­
modynamic process can proceed. For example, on the basis
of the law of conservation and conversion of energy, we
cannot predict the direction in which the exchange of heat
between two bodies heated to different temperatures will
occur. From the standpoint of the first law of thermodynamics,
the transition of energy in the form of heat from a hotter body
to a cooler one and vice versa is equally probable. The first
law of thermodynamics tolerates the design of a perpetual
motion machine of the second kind. This is the name given
to a machine in which the working substance performing
a cyclic process would get energy in the form of heat from
one external body and completely transfer it in the form of
work to another external body. An example of such a machine
could be a periodically operating device “pumping out” the
internal energy of the oceans and transferring it in the form
of work to other bodies.
2°. The impossibility of creating a perpetual motion
machine of the second kind is a statement following from
the generalization of numerous experiments. It is called
the second law of thermodynamics and has several equivalent
formulations:
(a) Clausius: It is impossible to have a process whose
only effect is to convey heat from a cooler to a hotter body;
and
(b) Kelvin: It is impossible to have a periodic process
whose only effect is the conversion of heat received from a hot
reservoir into an equivalent amount of work.
The second law of thermodynamics can be stated in the
form of Carnot’s theorems (II.4.8.3°-5°).
3°. The second law of thermodynamic points to the irre­
versibility of the process of converting one form of energy
transfer—work—into another form of energy transfer—heat.
For example, in the relative motion of two bodies with
friction, there is an irreversible transfer of the energy of
ordered motion of a body as a whole (no reverse transition
174 Molecular Physics and Thermodynamics II.4.10

is possible) into the energy of disordered thermal motion


of the particles of these two bodies (see also 11.4.3.3°).
In the formulations of the second law, a special signif­
icance is attached to the words “whose only effect”. The
prohibitions imposed by the second law are removed if the
effects mentioned are not the only ones. The transfer of
energy in the form of heat from a colder body to a hotter one
is possible if accompanied by another compensating process
conducted, for instance, in a refrigerator (11.4.11.1°). In
the isothermal expansion of an ideal gas, work is done that
is completely equivalent to the heat added to the gas (11.4.5.3°),
but this is not a violation of the second law. Expansion
of the gas is attended by a growth in its specific volume
(11.3.1.6°), and the state of the gas changes. Therefore, the
conversion of heat into work is not the only result of the
process being considered. In the conventionally used termin­
ology, by “the conversion of heat into work” is meant the
transition of the internal energy of the disordered motion of
the particles into the energy of the ordered motion of bodies
(11.4.3.3°).
4°. The third law of thermodynamics deals with the behav­
iour of a thermodynamic system at T 0. * The third
law leads to the unattainability of the absolute zero of tem­
perature. At absolute zero, the heat capacities and the coeffi­
cients of expansion (11.4.4.1°, 11.7.3.2°, 3°) of all bodies
vanish.

10*. Heat Engine


1°. A heat engine is a machine that converts the internal
energy of conventional or nuclear (VI.4.12.7°) fuel into
mechanical energy. The energy that is liberated upon com­
bustion of the fuel or in nuclear reactions (VI.4.8.1°) is
transferred by heat exchange (11.4.3.1°) to a gas. When
the gas expands, work is done against external forces, and
a mechanism is actuated.

* A strict formulation of the third law of thermodynamics cannot


be given in a reference book in elementary physics.
II.4.10__________ Fundamentals of Thermodynamics_____________ 175

2°. The gas in a heat engine cannot expand infinitely


because the engine has finite dimensions. The engine must
therefore be designed so that after expansion the gas is again
compressed to its initial volume. A heat engine has to op­
erate cyclically. In the course of a cycle (11.4.7.1°), expan­
sion is followed by compression of the gas. Real heat engines
function according to an open
cycle: the gas is discharged
after expansion, and a new por­ H ot r e s e r v o ir
tion of the gas is compressed. I
The thermodynamic processes
(11.3.2.1°) occurring in a heat V2ZZ2*ZZ?5
engine can be considered in
a closed cycle when the same .Working'--'- ^useful
substance 5—
portion of the gas expands
and is compressed.
3°. The work of expansion
of a gas during one cycle %
must exceed the work of com­ T
Cold reservoir
pression done by external forces 72<7,
on the gas. This condition is
needed for the engine to be
Fig. II.4.5
able to do useful work. The
temperature of the gas upon
its compression must be lower than in expansion. As a
result, thejpressure of the gas in all intermediate states in
compression will be lower than in expansion, and the con­
dition needed for the engine to do useful work will be
observed.
4°. Any heat engine regardless of its structural features
consists of three basic parts: a working substance, a hot
reservoir, and a cold reservoir (Fig. 11.4.5). The working
substance—a gas or a vapour (11.5.1.1°)—does work upon
expansion, receiving a certain amount of heat Q± from the
hot reservoir. The temperature Tx of this reservoir remains
constant because of the combustion of fuel. Upon compres­
sion, the working substance transfers a certain amount of
heat Q2 to the cold reservoir—a body having the constant
temperature T2 lower than Tr. The pressure of the gas is
lower in compression than in expansion, and this ensures
the useful work of the engine (Par. 3°). The surrounding
176 Molecular Physics and Thermodynamics 1 1 .4 .1 1

environment may serve as the cold reservoir (internal-com­


bustion engines, jet engines). The efficiency of a heat engine
is calculated by the formula given in 11.4.8.2°.

11*. Refrigerator
1°. A refrigerator is a cyclically functioning machine
that maintains a lower temperature in the refrigerator
chamber than the ambient temperature. This is achieved
by the transfer of a certain
amount of heat from a cold
H ot re s e rv o ir
( atm osphere)
body to one having a higher
temperature. Such a transfer
V r*>v does not contradict the sec­
ond law of thermodynamics
y7ZZZZ2EZZZZZZLm (11.4.9.2°) because this trans­
t .Working z-
—, fer of heat is not the only
*.:substance _1 process. A compensating pro­
cess occurs (11.4.9.3°) involv­
M yyM :
W( ing the conversion of the
mechanical energy of the sur­
~ 7*— rounding bodies into the in­
Refrigerator ternal energy of the hot res­
chamber
ervoir.
2°. A diagram showing how
Fig. II .4.6
energy is converted in a refrig­
erator is contained in Fig.
II.4.6. Upon the isothermal
expansion occurring at the temperature of the refrigerator
chamber T2, the working substance does work and absorbs
the heat Q2 from the chamber. Upon the isothermal com­
pression of the working substance occurring at the higher
temperature T1 of the hot reservoir (the atmosphere), the
heat is transferred to the latter. This occurs as a result
of the work of external forces. The working substance is
transferred from the state with the temperature T1 to that
with the temperature T2 and back by processes of adiabatic
expansion (the temperature drops from T1 to T2) and
adiabatic compression (the temperature rises from T2
to Tx).
t!.4.11 Fundamentals of Thermodynamics ill

3°. Generally, the vapour of a low-boiling liquid such


as ammonia and Freon is used as the working substance in
a refrigerator. Energy is supplied to the machine from elec­
tric mains. This energy is responsible for “pumping” heat
from the refrigerator chamber to a hotter body—the surroun­
dings.
4°. The working substance in a refrigerator is compres­
sed at a higher temperature than that of expansion, and
the work of the external forces in compression i4com is
greater than the work of expansion of the working substance
A exp. During a cycle, the external forces do the positive
work
-^ext = ^4com ^ ex p = Q i Q2

The refrigerating factor of a refrigerator is the ratio of


the amount of heat absorbed during a cycle from the refrig­
erator chamber to the work of the external forces:
ft @2 _ @2
P ^ext Q1 - Q 2

It follows from the second law of thermodynamics that

The equal sign relates to a reversible cyclic process in a re­


frigerator, and the inequality sign to an irreversible one;
the smaller the difference T x — T 2, the less is the amount
of mechanical or electric energy that has to be spent for
“pumping” the heat from the cold body to the hot one.
The refrigerating factor can be greater than 100% whereas
the efficiency of a heat engine is always much lower than
100 % .
5°. A refrigerator can be used as a heat pump for heating.
Flectric energy is used to actuate a refrigerator in which
the hot reservoir is the premises being heated and the refrig­
erator chamber is the outdoor atmosphere. The premises
being heated receive a greater amount of heat than is liberat­
ed when the electric energy is directly converted into the
internal energy of heaters such as electric stoves.
178 Molecular Physics and Thermodynamics I t.5 .2

Chapter 5
Mutual Transitions
of Liquids and Gases

1. Evaporation of Liquids
1°. Vaporization is the process of transition of a substance
from the liquid state to the gaseous one. Vaporization occurr­
ing at any temperature from the free surface of a liquid
is called evaporation. The molecules flying out of a liquid
in evaporation are called the vapour of the given liquid.
Not only liquids, but also solids form a vapour (11.7.4.8°).
Molecules having the highest velocity and kinetic energy
of chaotic thermal motion fly out of the surface layer of
a liquid, hence evaporation results in cooling of a liquid.
A measure of the process of evaporation is the rate of evap­
oration—the amount of liquid transforming into a vapour
in a unit time from a unit surface area of the liquid.
2°. Cooling upon the evaporation of liquids has a great
practical importance. For example, liquid ammonia or liquid
carbon dioxide is evaporated in special devices in the
refrigerator cars used for the transportation of perishable
products. The evaporation of liquid ammonia in the coils
of refrigerators is used for the preparation of ice. The coils
pass through a solution of salt (brine) and cool it to below
0 °C. Moulds of sheet steel filled with water are placed in
the brine, and pieces of ice form in them.

2. Saturated Vapour
1°. If a vaporization process (11.5.1.1°) occurs in a closed
vessel, then after a certain time elapses the amount of the
liquid stops diminishing, although the liquid molecules
capable of leaving its surface continue to pass over into the
vapour. In this case, the vaporization process is attended
by a reverse compensating process of condensation—the trans-
II.5.3 Mutual Transitions of Liquids and Gases 179

formation of a vapour into a liquid. The rate of condensation


is determined by the number of jnolecules passing from the
vapour into the liquid through a unit area of the liquid
surface in a unit time.
2°. After a certain time elapses in a closed vessel contain­
ing a liquid, dynamic (mobile) equilibrium sets in between
the processes of vaporization and condensation—the rates
of these processes become equal. From this moment, the
amounts of liquid and of the va­
pour over it stop changing.
3°. A vapour in a state of
dynamic equilibrium with its
liquid is called a saturated va­
pour. The pressure of a saturat­
ed vapour p s depends only on
its chemical composition and
temperature and does not de­
pend on the free volume of the
vessel containing the vapour.
The explanation why p s
fails to depend on the volume
is that when the volume of a saturated vapour diminishes,
a greater and greater part of the vapour passes over into
the liquid. But this does not violate dynamic equilibrium
between the vapour and the liquid up to the moment when
condensation of the vapour terminates.
4°. The saturated vapour pressure rapidly grows with
elevation of its temperature. Figure 11.5.1 shows plots of
(lie temperature dependence of the pressure for a saturated
vapour and an ideal gas at a constant volume.

3. Boiling
1°. Boiling is defined as the process of intensive vapor­
ization not only from the free surface, but also throughout
Ilie entire volume of a liquid inside the vapour bubbles
formed during this process. The pressure p inside a bubble
is determined by the formula
2 a
P = Po + Pgh-t R
ISO_________ Molecular Physics and thermo&ytiamlck________11.5.8

where p 0 = external pressure


pgh = hydrostatic pressure (1.0.2.2°) of the higher
layers of the liquid
2o/R = additional pressure connected with the curva­
ture of the surface of a bubble (11.6.2.4°)
R = radius of a liquid bubble
h = distance from its centre to the surface of the
liquid
p and a = density and surface tension (11.6.1.3°) of the
liquid.
2°. A liquid begins to boil when

Pn&Po + p g h + j r

where p Q is the saturated vapour pressure (11.5.2.3°) inside


a bubble. At small values of /?, the pressure p s is sufficiently
high, and boiling occurs at comparatively high temperatures.
If a liquid contains centres of vaporization (dust particles,
bubbles of dissolved gases, and the like), then generally
2oIR p0, and boiling begins at lower temperatures.
If pgh p0, then the condition of boiling becomes
P s>P o
3°. The boiling point is the temperature of a liquid
at which its saturated vapour pressure equals or exceeds
the external pressure. The boiling point grows with an
increasing external pressure, and vice versa. This follows
from the condition of boiling. An increase in p 0 is attended
by a growth in the saturated vapour pressure needed for
boiling to begin, and this is possible only at a higher temper­
ature (11.5.2.3°). The difference between the boiling points
of various liquids is due to the fact that their saturated
vapour pressures differ at the same temperature. The higher
the saturated vapour pressure, the higher is the boiling
point.
The gas bubbles inside a liquid are filled with its satu­
rated vapour. When the temperature of the liquid rises,
the pressure of the vapour in a bubble grows, and its volume
increases. The buoyant Archimedean force (1.6.2.3°) acting
on the bubble grows with an increase in its volume. The
bubble rises to the surface under the action of this force,
11.5.4 Mutual Transitions of Liquids and Gases 181

and if the condition of boiling is observed (Par. 2°), it


bursts, discharging the vapour. It is obvious that when ps
grows this entire process becomes easier, and the corre­
sponding liquid boils at a lower temperature.
4°. If a liquid contains no bubbles needed for the boiling
process, then it can be superheated to a temperature that
is higher than the boiling point at the given pressure. The
unstable (metastable) state appearing here is characteristic
of a superheated liquid. The latter can be obtained if we
reduce the external pressure on a liquid by such an amount
that it becomes lower than the saturated vapour pressure
at the given temperature, and the condition of boiling
(Par. 2°) is violated. Violation of the metastable state of
a superheated liquid causes its violent boiling.
5°. In the course of boiling, the temperature of a liquid
remains constant if the external pressure p 0 (Par. 1°) does
not change. The heat supplied from outside to the liquid
is used for vaporization (11.5.1.1°) and the work of external
forces. The amount of heat Lb needed to convert a unit
mass of a liquid heated to its boiling point into a vapour
is called the specific heat of vaporization. It follows from
the law of conservation of energy that in the reverse process—
condensation of a vapour into a liquid—an amount of heat
equal to Lb is liberated.
6°. If a given liquid boils at a higher (or lower) temper­
ature, then the value of Lb diminishes (or grows). This is
associated with the temperature dependence of the saturated
vapour pressure (see Fig. 11.5.1) and the condition of boiling
(Par. 2°). Elevation or lowering of the boiling point as
/i result of an increase or decrease in the external pressure
leads to a reduction or a growth in the amount of heat needed
lo convert a unit mass of the liquid into a vapour in the
conditions of boiling.

4. Vapour Isotherm
1°. A vapour is defined as unsaturated if its pressure is
lower than the saturated vapour pressure ps at the given
temperature. The pressure of an unsaturated vapour depends
on its volume—a reduction of the volume is attended by
a growth in the pressure, and vice versa.
182 Molecular Physics and Thermodynamics 11.5.4

2°. A vapour isotherm is a curve showing how the pres­


sure p of the vapour depends on its specific volume v at
a constant temperature. A general view of a vapour iso­
therm is shown in Fig. II.5.2. The path 0 -> 1 corresponds
to an unsaturated vapour, the path 1 -> 2 to a saturated
vapour, and the path 2 -> 3 to a liquid. By reducing the
volume of the unsaturated vapour, we can bring it into
a state of saturation (point 7). With a further reduction
in the volume (path 7 -> 2), the saturated vapour pressure

Fig. II.5.2 Fig. II.5.3

does not change. Part of it passes over into the liquid, and
at the point 2 the entire vapour condenses. A reduction
in the volume of the liquid on the path 2 -> 3 requires
a considerable pressure increase in connection with the low
compressibility of liquids.
3°. An unsaturated vapour can be brought into the
saturated state not only by its isothermal compression,
but also by lowering the temperature. The unsaturated
vapour when passing over into a state of saturation partly
liquefies. This explains the sweating of cold objects brought
into a warm room, the formation of a fog, dew, etc.
The isotherm can be considered as
a curve of the continuous transition of a liquid into the
state of an unsaturated vapour. A vapour can pass from the
saturated to the unsaturated state not only by isothermal
expansion, but also by elevation of the temperature. The
unsaturated vapour obtained by heating a saturated vapour
is called superheated.
11.5.5 Mutual Transitions of Liquids and Gases 183

Upon the slow isothermal compression of a vapour in


the absence of dust particles, ions, and other condensation
centres, it is possible to obtain a supersaturated vapour
whose pressure exceeds that of a saturated vapour at the
given temperature.
4°. Isotherms constructed for different, constantly grow­
ing temperatures 7\, r 2, T3 (T% > T2 > 7\) (Fig. II.5.3)
show that the state of a saturated vapour is obtained at
smaller and smaller specific volumes of the vapour: i>3' <
< v'2 < v[. The explanation is that elevation of the tem­
perature is attended by a rapid growth in the saturated
vapour pressure, and for the pressure of the unsaturated
vapour to become equal to it the volume of the vapour must
he reduced. Complete condensation of the vapour with
elevation of the temperature occurs at greater volumes:
vl > v'i > v”\. The reason for this is the thermal expansion
of liquids (11.7.3.3°) which at a higher temperature occupy
a greater volume.

5. Critical State of a Substance.


Liquefaction of Gases
1°. A glance at Fig. II.5.3 shows that upon elevation
of the temperature the portion of the curve corresponding
to the saturated vapour (the horizontal portion of the iso­
therm) diminishes; at a certain temperature Tcr no saturated
vapour is formed. The temperature TCr at which the differ­
ence between the specific volumes of a saturated vapour and
a liquid vanishes is called the critical temperature. The hori­
zontal portion of the isotherm transforms into a point
o inflection C (the critical pointV This occurs at a definite
pressure p CT and specific volume z;cr which together with Tcr
are called the critical parameters. The state of a substance
characterized by the critical parameters p CT, vCTl and Tcr
is known as the critical state.
2°. The density of a saturated vapour and its pressure
rapidly grow with increasing temperature (11.5.2.4°). The
density of a liquid in dynamic equilibrium with its vapour
diminishes when the temperature grows as a result of ther­
mal expansion of the liquid (11.7.3.3°). Curves of the tern-
184 Molecular Physics and Thermodynamics 11.5.6

perature dependence of the densities of a liquid and its


saturated vapour intersect at the critical point corresponding
to the temperature TGV (Fig. II.5.4). At the critical point,
the density of the liquid equals that of the saturated vapour
in equilibrium with it.
3°. When T tends to Tct, the difference between the
liquid and gaseous states of a substance vanishes. When
T = Tcr, the specific heat of vaporization (11.5.3.5°) and
the surface tension of the liquid
(11.6.1.3°) vanish.
4°. A gas cannot be convert­
ed into a liquid at tempera­
tures above the critical one Tcr
even if very great pressures are
applied to it. At supercritical
temperatures, a substance can be
only in its vapour state.
5°. The liquefaction of gases—
their conversion into the liquid
state—is possible only at tem­
Fig. II.5.4 peratures below the critical one.
For gases having a critical tem­
perature of the order of magnitude of room temperature
and higher such as NH3 (tCT = 132 °C), C 02 (tCT =
= 31.1 °C), and Cl2 (tCT = 144 °C), the lowering of the
temperature to below T ct and subsequent isothermal com­
pression do not involve any special difficulties.
6°. To liquefy gases having a low critical temperature
such as 0 2 (tCT = - 1 1 8 °C), N 2 (tCT = -1 4 7 .1 °C),
H2 (tCT = —239.9 °C), and especially He (tCT = —267.9 °C),
deep refrigeration is needed. One of the methods carried
out in refrigerating machines—gas expanders—is based
on the cooling of gases when they perform useful work.
The gas expands adiabatically, and this is attended by its
cooling (11.4.5.4°)

6. Humidity of Air
1°. The absolute humidity of air f is defined as the mass
of water vapour contained in 1 m3 of air in the given con­
ditions. The value of / is assessed according to the density
11.5.6 Mutual Transitions of Liquids and Gases 185

of the water vapour in the air. It is customarily expressed


in g/m3 instead of in SI units:

/=Pvap(g/m»)

In meteorology, the absolute humidity is assessed


according to the pressure of the water vapour expressed
in millimetres of mercury: f = p (mmHg). At room tem­
peratures (T « 300 K), p (mmHg) « p (g/m8).
2°. The relative humidity of air (p is the ratio of the absolute
humidity to the amount of water vapour needed to saturate
1 m3 of air at the given temperature. It follows from the
preceding paragraph (Par. 1°) that the relative humidity
can be determined as the ratio of the pressure of the water
vapour contained in the air to the pressure of saturated
water vapour at the given temperature:

Usually cp < 1, and it is expressed in per cent.


3°. The dew point is the temperature at which water
vapour previously not saturating the air becomes saturated.
Knowing the temperature of the air and having determined
the dew point, the humidity of the air is calculated. A table
of the saturated vapour pressure of water at different tem­
peratures is used here.
Problem. The temperature of the air in a room is 20 °C,
and the relative humidity of the air is 60%. At what out­
door temperature will the window panes begin to mist
over?
Given: tx = 20 °C, cp = 60%.
Required: t 2.
Solution: The relative humidity is (p = p/ps, where p
is the pressure of the water vapour in the air, and ps is the
pressure of the water vapour saturating the air at the given
temperature.
At 20 °C, the saturated water vapour pressure is ps =
= 17.5 mmHg (from a table). The pressure p = ps<p =
= 17.5 X 0.6 = 10.5 mmHg,
186 Molecular Physics and Thermodynamics II.6.1

Vapour condensation will begin at the temperature


of the air for which the pressure p will correspond to the
pressure of the vapour saturating the air. We find from
a table that t2 — 12 °C corresponds to a pressure of p =
= 10.5 mmllg.

Chapter 6
Properties of Liquids*

1. Energy of Surface Layer


and Surface Tension
1°. The molecules of a liquid on its surface near the
liquid-vapour interface are subjected to a different inter-
molecular interaction than the
molecules in the body of the
liquid.
The molecule 1 surrounded
on all sides by other molecu­
les of the same liquid expe­
riences on an average iden­
tical forces of attraction
(11.1.4.3°) to all its neigh­
bours. These forces on an
average mutually compen­
sate one another,fand their
resultant equals zero "(Fig.
II.6.1). The molecule 2 expe­
riences a smaller attraction from above by the vapour
molecules and a greater attraction from below by the liquid
molecules. In Fig. II.6.1, the forces of attraction of the
molecule 2 to the vapour molecules are shown by dash arrows.

* For information on the structure and thermal motion in liquids


see 11.1.6.7°, 8°, for the thermal expansion of liquids see 11.7.3.3°.
11.6.1 Properties of Liquids 187

As a result, a molecule in the surface layer will be acted


upon by the resultant R of the forces, which is generally
related to a unit area of the surface layer.
2°. To transfer molecules from the interior of a liquid
to its surface layer, work for overcoming the force R (Par. 1°)
must be done. This work is used to increase the surface
energy. This is the name given to the surplus potential
energy that the molecules in the surface layer have in
comparison with the potential energy that the molecules
in the body of the liquid have.
To isothermally increase the surface layer of a liquid
at the expense of the molecules in its interior, the work
A must be done equal to
A = (Ep, s — EVt v) N
where EV9S = potential energy of one molecule in the
surface layer
Ev%v = potential 'energy of a molecule in the body
of the liquid
N = number of molecules in the surface layer
of the liquid.
3°. The surface tension of a liquid is defined as the work
needed for an isothermal increase of the surface area of the
liquid by one unit:

° = ~Es~ = ^ p’s ^p* v) ~~s~= s n

where n is the number of molecules per unit surface area


of the liquid.
If the surface of a liquid is confined by a wetted perime­
ter (11.6.2.1°), then the surface tension numerically equals
the force acting on a unit length of the wetted perimeter
and directed at right angles to this perimeter:

where I = length of the wetted perimeter


F = force of surface tension acting on the length I
of the wetted perimeter.
188 Molecular Physics and Thermodynamics 11.6.2

The force of surface tension is in a plane tangent to the


surface of the liquid.
4°. A reduction in the surface area of a liquid diminishes
its surface energy. A condition for the stable equilibrium
of a liquid, as of any body, is the minimum potential surfa­
ce energy (1.4.3.4°). This means that in the absence of
external forces a liquid must have the smallest possible
surface area with the given volume and takes on the form
of a sphere.
5°. With an increase in the temperature of a liquid and
its approaching the critical one, i.e. when T T ct (11.5.5.3°),
the surface tension a tends to zero. Far from Tcr, the value
of a linearly diminishes with increasing temperature.
Special admixtures are added to a liquid to reduce its
surface tension. They occupy the surface of the liquid and
reduce the surface energy (surface-active substances, or
surfactants). They include soap and fatty acids.

2. Wetting. Capillary Phenomena *


1°. The phenomena of wetting are observed on the inter­
face between solids and liquids. They consist in the curva­
ture of the free surface of the liquid near the solid walls
of the vessel containing it. The curved surface of the liquid
at its place of contact with the solid is called a meniscus.
The line along which the meniscus intersects the solid is
the wetted perimeter.
2°. Wetting is characterized by the angle of contact 0
made by the surface of the solid and the meniscus at the
points of their intersection, i.e. at points on the wetted
perimeter. A liquid is called a wetting one if the angle of
contact is acute, i.e. 0 ^ 0 < jt/2 (Fig. II.6.2a). For exam­
ple, water wets clean glass, and mercury wets zinc. A liquid
is non-wetting when the angle of contact is obtuse, i.e.
jt/2 < 0 < Jt (Fig. II.6.26). For example, water does not
wet paraffin, and mercury does not wet iron. If 0 = 0,
* The phenomena are considered for the very simple case when the
forces of interaction between the molecules of the liquid and of the
gas over the liquid may be disregarded.
tl.6.2 Properties of Liquids 189

wetting is considered to be perfect, while 0 = ft corresponds


to perfect non-wetting. When 0 = 0 and 0 = ft, the meniscus
has a spherical shape, concave or convex. When 0 = ft/2,
the liquid has a flat free surface. This case is called the
absence of wetting and non-wetting.
3°. The difference between the angles of contact in the
phenomena of wetting and non-wetting is explained by the
ratio of the forces of attraction between the molecules of
solids and liquids and the forces of intermolecular attraction
in liquids (11.1.4.1°, 11.6.1.1°). If the forces of attraction
between the molecules of a solid and a liquid are greater

(<0 Cb)

Fig. II.6.2

than the forces of attraction of the liquid molecules to one


another, then the liquid will be a wetting one. If the mole­
cular attraction in a liquid exceeds the forces of attraction
of the liquid molecules to those of a solid, then the liquid
does not wet the solid.
4°. Curvature of the surface of a liquid creates an addi­
tional (differential, or positive) pressure on the liquid in
comparison with the pressure under a flat surface. For
a spherical liquid surface with the contact angle 0 equal
to 0 or ft, the additional pressure p m is
190____________ Molecular Physics and Thermodynamics_________ 11.62

where o = surface tension


R = radius of the spherical surface.
If the meniscus is convex, then p m > 0, and if it is
concave, then p m < 0 (Fig. II.6.3). With a convex menis­
cus, the pressure in comparison with that under a flat sur­
face of the liquid (for example atmospheric pressure on the
free surface of a liquid) increases by the amount p m. With
a concave meniscus, the pressure in comparison with that
under a flat surface diminishes by the amount p m. The
additional pressure inside a spherical bubble of radius R
is due to the differential pressure on both surfaces of the
bubble and equals p m = 4o/R.
5°. Narrow cylindrical tubes with a diameter of about
a millimetre and less are called capillaries. The level of
a perfectly wettiug (non-wetting) liquid in a capillary

P m


i
T —
o
A i f T f
o
V
-
^ 4

~7 ~ . ______ _
— ■
i
Z T “. -------------- 7 1 ---------------------
£ T :

Fig. I I . 6.3 Fig. I I . 6.4

of radius r is higher (lower) than in a wide vessel communi­


cating with it by the distance h (Fig. II.6.4) equal to
2a
h=
pgr
where p = density of the liquid
g = acceleration of free fall.
The changes in the levels in capillaries are called capillary
phenomena. They are associated with the fact that for
equilibrium of a wetting liquid in a capillary the differen­
tial pressure p m must equal the hydrostatic pressure pgh
(1.6.2.2°). Lowering of the level of a non-wetting liquid
in a capillary is explained by the fact that the differential
IT .6 .2 P ro p e rtie s of L iq u id s 191

pressure p w orces the liquid out of the capillary until the


same equilibrium as for a wetting liquid sets in, i.e. pgh =
= 2o/r.
6°. Capillary phenomena explain the property of a num­
ber of substances (cotton wool, fabrics, the soil, concrete)
to absorb moisture (the hygroscopic!ty of substances). To
retain the subsoil moisture in soil, the capillaries in it are
destroyed in ploughing and harrowing. Otherw'se the mois­
ture in the soil will rise to the surface along the capillaries
and evaporate. To prevent moisture from penetrating into
the rooms of a building along the capillaries, a dampproof
course consisting of layers of tar paper, pitch, and other
substances preventing capillary phenomena is laid between
the foundation of a building and its walls. The rising of
water along capillary fibres in plants ensures their develop­
ment.
Problem 1. Under what pressure is the air inside a soap
bubble 4 mm in diameter and what does the differential
pressure equal? The atmospheric pressure is 752 mmHg.
Given: p 0 = 752 mmHg, d = 4 mm = 4 X 10~3 m,
a = 40 X lO"3 N/m, 1 mmHg = 133 N/m2.
Required: p, p m.
Solution: The air in the bubble is under a pressure
of p = p 0 + pm, where p m is the pressure exercised on the
air by the two spherical surfaces of the bubble film. The
film has a very small thickness. Therefore, the diameters
of the two surfaces are virtually the same. The differential
pressure is

__ o 2a
Pm — ^

where a is the surface tension, and R is the radius of curva­


ture of the surface, R = dl2. Hence, the pressure of the
air in the bubble is

I q ^O’ *7£0 w 4 QQ I ^ X 40 X 10 ®
P = Po + 2 — = 7 5 2 x 1 3 3 + 4 x iQ-3 —

= 100 016 + 80 = 100 096 N/m2


192 Molecular Physics and Thermodynamics 11.6.2

The differential pressure is pm = 80 N/m2 = 0.6 mmHg.


Problem 2. Soapy water drips from a capillary. When
a drop breaks away, the diameter of its neck is 1 mm. The
mass of a drop is 0.0129 g. Find the surface tension of the
soapy water.
Given: m = 0.0129 g = 1.29 X 10"5 kg, d = 1 mm =
= 10~3 m.
Required: a.
Solution: A drop breaks away from the capillary when
P > F, where P = mg is the force of gravity, and F is
the force of surface tension.
The condition of equilibrium is F = P, ol = mg, where I
is the perimeter of the neck of a drop, I = Jtd. Hence,
mg 1.29 X 10-* X 9.81
a = nd 3.14 X IQ"3
= 4 .0 5 x l0 - 2 N/m

Problem 3. Find the difference between the levels of


the liquid in two capillary tubes immersed in a liquid.
The density of the liquid is 0.8 g/cm3. The internal diame­
ters of the capillaries are 0.04 and 0.1 cm. The surface ten­
sion of the liquid is 22 X 10“3 N/m.
Given: d1 = 0.04 cm = 0.4 X 10"3 m, d2 = 0.1 cm =
= 10"3 m, p = 0.8 g/cm3 = 8 X 102 kg/m3, a = 22 X
X 10-3 N/m.
Required: Ah.
Solution: The height to which a liquid rises in a capil­
lary is
^ _ 2a
~Tgr
where r is the radius of the capillary, p is the density of the
liquid, and r = d!2. Hence,
2 X 2a
h=
P gd
and
A fe = A1-
1 f e 2z = -pg
^ - \( 4di— d2
j - )/ =
4X 22X 10-*/ 1
X
8X 10* 9.8 \ 0.4X10-*
16. 8X10“3 m
11.7.1 S o lid s and T h eir T ra n sfo rm a tio n in to L iq u id s 193

Chapter 7
Solids and
Their Transformation
into Liquids*

1. Kinds of Crystalline Solids


1°. Crystalline solids are anisotropic—their physical
properties depend on the direction inside the crystal. Thus,
the thermal expansion of crystals (11.7.3.1°), their mecha­
nical strength (11.7.2.9°), and their optical properties (the
speed of light, the refractive index (V. 1.2.1°) depend on the
direction inside the crystal.
2°. The following four kinds of crystal lattices are dis-
1inguished:
(a) Ionic crystals—most inorganic compounds, for exam­
ple salts (NaCl, etc.), metal oxides, and so on. The lattice
points (sites) of ionic crystals accommodate regularly alter­
nating positive and negative ions (see Fig. II. 1.4a) between
which forces of electrostatic interaction mainly act provid­
ing an ionic (heteropolar) bond (VI.3.2.1°). In the process
of crystallization (11.7.4.6°), some atoms (for example Na)
lose electrons that attach themselves to other atoms (for
example Cl), and two oppositely charged ions appear.
(b) Atomic (valence) crystals—the crystal lattices of
semiconductors (Te, Ge, etc.) (111.3.11.1°), and many
organic solids. Typical examples of such crystals are the
varieties of carbon—diamond and graphite (see Fig. II.1.46).
The points of crystal lattices of atomic crystals accommodate
electrically neutral atoms, most frequently identical ones,
between which there is a specific covalent (homopolar)
bond (VI.3.3.1°) having a quantum-mechanical origin.

* For information on the structure of crystalline solids and the


thermal motion in crystals see II.1.6.3°-6°.
194 M o lecu la r P h ysics and T h erm o d yn a m ics n.7.2

(c) Molecular crystals—Br2, I2, CH4, naphthalene, pa­


raffin, and many solid organic compounds. The lattice
points of such crystals accommodate molecules retaining
their individuality. Forces of attraction act between them
that are characteristic of molecular interaction (11.1.4.3°).
The relatively low stability of molecular crystals and their
low melting points are explained by the fact that the forces
of attraction between their molecules are lower than in
crystals of the other kinds.
(d) Metallic crystals {metals). When metals crystallize
(11.7.4.6°), their outer (valence) electrons (VI.2.9.2°) are
detached from the atoms, and positive ions are formed.
The latter occupy the lattice points of the crystals. The
outer (valence) electrons, become collective—they belong
to the entire crystal as a whole, forming an electron gas
in metals.
The metallic bond in the crystal lattice of metals is
ensured by attraction between the positively charged ions
at the lattice points and the negatively charged electron
gas. The collective electrons of metals “draw together”
the positive ions, as it were, balancing the repulsion between
them. At distances between the ions equal to the crystal
lattice constant (11.1.6.5°), a stable configuration of the
ions appears that corresponds to the condition of equality
between the forces drawing the ions together and those
of their mutual repulsion. The presence of the electron gas
explains the good electric and thermal conductivities of
metals.

2. Elastic Properties of Solids


1°. By the deformation of a solid body is meant the
change in its dimensions and volume usually attended by
a change in the shape of the body*. Deformations occur
upon the heating (cooling) of solids or under the action
of .external forces. In deformations, the particles at the
lattice sites of a crystal are displaced from their equilib-

* An exception is uniform tension (compression) in which the shape


of the body remains unchanged.
11.7.2 Solids and Their transformation into Liquids 195

rium positions. This displacement is hindered by the forces


of interaction between the particles of the solid. Internal
elastic forces appear in a deformed solid.
2°. Elasticity is the property of bodies to restore their
dimensions, shape, and volume after the action of the
external forces that caused their deformation stops.
An elastic force Fel is defined as the force appearing
upon the deformation of a body and directed oppositely
to the direction of displacement of the particles of the
body in deformation (for the elastic force see also 1.2.9.1°).
3°. Deformations that vanish after the action of external
forces stops are called elastic. The particles of the solid body
that were displaced in the process of deformation (Par. 1°)
return to their initial positions of equilibrium, and the
original dimensions and shape of the body are restored.
Thus, an elastically deformed spring after the load is remo­
ved restores its initial length, volume, and shape. The
elastic deformation stops when FeX = F, where F and Fe\
are the magnitudes of the external force causing the defor­
mation and of the elastic force.
Non-elastic deformations of a solid body do not vanish
after the action of the forces causing them stops and lead
to irreversible changes in the crystal lattice of the body.
Such deformations are called plastic. Plastic deformations
are characterized by the appearance of permanent set (resi­
dual deformations) that remains in a body after the action
of the forces stops.
4°. A measure of the deformation of a body is the relative
deformation, or strain, Ax!x equal to the ratio of the absolute
deformation Ax to the initial value of the quantity x charac­
terizing the dimensions or shape of the body. Thus, the
tensile strain of a body is the ratio of the elongation of the
body AI to its initial length I, i.e. AHI (Par. 7°).
5°. The mechanical stress a is the physical quantity
numerically equal to the elastic force per unit of cross-
sectional area of a body:

where Fel = elastic force


S = cross-sectional area of a body.
13*
196 M o lecu la r P h ysic s and T h erm o d yn a m ics II.7.2

If the stress is constant over the entire cross-sectional


area, then a = Fe\/S.
The stress is called normal if the force AFel is perpendi­
cular to the cross-sectional area AS, and it is a shear stress
if the force AFei is directed along a tangent to the area AS.
6°. Hooke's law (see also 1.2.9.4°): the strain is directly
proportional to the stress:
A#_ a
~~~K

where K is the modulus of elasticity. When Ax = x , i.e. when


the strain equals unity, the modulus of elasticity numeri­
cally equals the stress: K = a. Hooke’s law holds for
elastic deformations (Par. 3°) up to definite values of these
deformations. The stress at which Hooke’s law begins to be
violated is defined as the proportionality limit.
7°. The simplest kind of deformation is linear, or unia­
xial, tension (compression). In uniaxial tension (compres­
sion), the tensile (compressive) force causes the length of
a body to grow (diminish). A measure of the deformation
is the tensile (compressive) strain
AHI. Here the modulus of ela­
sticity K is called Young's mo­
dulus and is designated by E.
Young’s modulus numerically
equals the stress at which the
length of a body is doubled
(halved), E = a when A1 = 1.
8°. A graph of the dependen­
ce a = f (AllI) for uniaxial ten­
sion is called a tensile stress-
Fig. I I .7.1 strain diagram (Fig. II. 7.1). The
point A corresponds to the pro­
portionality limit (Par. 6°), and Hooke’s law holds in the
region OA. The elastic limit is defined as the greatest stress
at which permanent set (Par. 3°) does not yet appear. The
point A' in Fig. II.7.1 corresponds to the elastic limit.
The region A 'B of the diagram corresponds to permanent
set. The horizontal portion BC of the diagram shows
yield—such a state of a deformed body when the elonga-
II.7.3 S o lid s and T h eir T ra n sfo rm a tio n in to L iq u id s 197

tion grows without an increase in the stress. The point


B is the yield point.
9°. The strength of a m aterials its property of withstand­
ing the action of external forces without destruction. The
ultimate strength is defined as the mechanical stress corres­
ponding to the maximum load which a body withstands
before its crystalline structure is destroyed.
By the safety factor is meant a number showing how
many times the ultimate strength is greater than the tole­
rated stress. The ultimate strength depends on the property
of the material. The safety factor depends on the nature
of the load, the conditions of utilization of the material,
and many other factors.
Problem. A wire 3.0 m long and 0.8 mm in diameter
is hanging vertically. A load with a mass of 5 kg is suspended
from the free end of the wire. The length of the wire increased
by 0.6 mm. Determine the stress, the tensile strain, and
Young’s modulus.
Given: I = 3.0 m, d = 0.8 mm, m = 5.0 kg, Al =
= 0.6 mm.
Required: a, AZ/Z, E.
Solution: The stress a = F/S. The absolute value of
the elastic force equals the force of gravity, F = mg; the
cross-sectional area of the wire S = ncPIA. Hence,

° - ^ = 3 . » " y i 8XXl.W - = :i2-0 X l0 ,N /m i° 32MPa


The tensile strain is
M _ 0.6X10-3
I “ 3.0
2.0x 10"4
According to Hooke’s law, Young’s modulus is
E = - a = 3,0 * 63x i()-31(>6= 16 • 0 x 1010 N/m2 = 160GPa

3. Thermal Expansion of Solids


and Liquids
1°. The increase in the linear dimensions and volume
of a body occurring when its temperature is raised is called
thermal expansion. Linear thermal expansion is characters-
198 Molecular Physics and Thermodynamics 1 1 .7 .3

tic of solids. Volume {cubic) expansion occurs both in solids


and in liquids when they are heated.
2°. Linear thermal expansion is characterized by the
coefficient of linear expansion (the average coefficient of
linear expansion) a within the given temperature interval.
If l0 is the initial length of a body at the temperature t0,
and AZ = I — Z0 is the increase in the body’s length when
it is heated by A Z = Z — t0 kelvins, then a characterizes
the elongation per unit length AZ/Z0 of the body that occurs
when it is heated by one kelvin:

The length Z of a body at the temperature t is determined


by the formula Z = Z0 (1 + a At). For most solids, the
coefficient a ranges from 10"6 to 10"6 K"1, and we may
consider that a is virtually independent of the temperature.
3°. The volume expansion of solids and liquids is cha­
racterized by the coefficient of volume expansion (the average
coefficient of volume expansion) y within the given tempera­
ture interval. This is the name given to the dilatation
AVIV0 occurring when a body is heated by one kelvin:
1 AV
y~ At V,,

where V0 is the initial volume of the body at the temperature


Z0 , and AF = V — F0 is the increase in the volume of
the body when it is heated by At = t — t0 kelvins. The
volume of a body V at the temperature t is determined by
the formula V = V0 (1 + y At). The coefficients of linear
and volume expansion are related as follows:
y « 3a
This relationship holds for conditions when we can disregard
the terms 3a2 At and 3a AZ2 in the expression (1 + a AZ)3.
4°. Linear thermal expansion is explained by the asym­
metrical shape of the curve showing how the potential
energy Ep (r) of interaction of two molecules depends on
the distance r between them. It is known (11.1.5.3°) that
such a nature of the curve Ev (r) is associated with the
different dependence of the forces of attraction and repulsion
II.7.3 Solids and Their Transformation into Liquids 199

acting between molecules on the distance r. If at a certain


moderate temperature a molecule has the total energy E
then it oscillates about the equilibrium position r0 between
the points a and 6, and ar0 « br0 (Fig. II.7.2). Elevation
of the temperature is attend­
ed by a growth in the total
energy of a molecule E2, E 3,
etc., and it oscillates between
the points a' and b', a" and
b", etc. Here a'r0 < ;r 06',
a"r0 < r06", etc. The inequ­
ality grows with increasing
temperature. This is associa­
ted with the fact that upon
heating the point a' (an, etc.)
on the curve Ep (r) is dis­
placed to the left relative to
the point a by a smaller amount
than the point bf (&", etc.) is displaced to the right relative
to the point b. The result is that the average distance bet­
ween the positions of equilibrium of the particles of a solid
grows with elevation of the temperature, i.e. thermal ex­
pansion occurs.
Problem. What amount of heat was used to heat a copper
sphere from 0 °C if its volume increased by 10 cm3?
Given: tx = 0 °C, AV = 10 cm3 = 10"5 m3, from the
relevant tables: c = 3.80 X 102 J/kg-K, a = 1.70 X
X 10-5 K -1, p 0 = 8.90 X 103 kg/m3.
Required: Q.
Solution: The amount of heat Q = me (t2 — ^), where
t2 is the temperature to which the sphere was heated.
From the formula for volume expansion V = V0 [1 +
I Y (h — *i)I> where y is the coefficient of volume expan­
sion, and V0 is the volume of the sphere at 0 °C, it follows
that
h
y-yo
FoY
Since V — V0 = AP, and V0 = mlp0, we find
. _ AFp0 , ^ _ /AFp0 , \
200 Molecular Physics and Thermodynamics 11,7.4

And, finally, since y = 3a, where a is the coefficient


of linear expansion, and tx = 0 °C, we have
AFp0_ 3.80 X 102 X 10-BX8.90X103
3a ~~ 3 x 1 .7 0 x 1 0 -5
6 .6 5 x l0 4 J

4. M elting, C rystallization,
and Sublim ation of Solids
1°. The melting of solids is their transition from the
solid state to the liquid one. At the expense of the energy
supplied to a solid in melting, the amplitudes of the displa­
cements of the particles oscillating at the lattice points
grow, and become compatible with the lattice constant
(11.1.6.5°). In the process of melting, the crystal lattice
of a solid is destroyed.
2°. Melting occurs at a definite temperature called the
melting point Tm. For most solids, Tm grows with increasing
external pressure. The latter prevents the increase in the
equilibrium distances between the particles in a crystal
lattice needed for melting of the body to begin, and hampers
the destruction of the lattice.
3°. The melting of solids, as a rule, is attended by an
increase in the specific volume or a reduction in the density
of a body. Exceptions to the rule are ice and bismuth in
which melting is attended by a reduction in the specific
volume (an increase in the density). In these substances,
an increase in the external pressure leads to lowering of the
melting point.
4°. In the course of melting of a solid, it simultaneously
exists in both the solid and liquid states. The temperature
of a body does not change in melting and constantly remains
equal to Tm. All the heat supplied to the body is used for
destroying the crystal lattice and for work against external
forces. As a result of melting, the internal energy of a body
(11.4.1.2°) grows, as does the potential energy of intermole-
cular interaction (11.1.5.1°).
5°. The amount of heat needed for the transition of
a unit mass of a solid into the liquid state at the melting
point is called the specific heat of fusion Lt.
I I .7 .4 Solids and Their Transformation into Liquids 201

6°. The transition of a substance from the liquid to


the solid state is called freezing (<crystallization, solidifica­
tion). In crystallization, the mean free lifetime of the liquid
molecules (11.1.6.8°) grows, their motion becomes ordered
and gradually transforms into thermal oscillations about
certain central positions—the points of a crystal lattice.
For any chemically pure liquid, this process occurs at
a constant temperature—the freezing point (crystallization
point, solidification point) Tiv that coincides with the melt­
ing point Tm (Par. 2°). The freezing of a unit mass of a li­
quid is attended by the liberation of a certain amount of
heat—the specific heat of freezing, equal to the specific heat
of fusion.
7°. A liquid begins to freeze near centres of crystalliza­
tion-im purities, dust particles, local violations of the
homogeneity of the liquid. At these places, order sets in
in the arrangement of the particles, and a crystal lattice
forms. In the absence of centres of crystallization, a liquid
can be cooled to a temperature below its freezing point
Tfr (a supercooled liquid). The supercooled state is not
stable and is easily violated. For example, a supercooled
liquid begins to freeze when it is jolted.
8°. A vaporization process (sublimation) may occur in
solids—the direct breaking away of the molecules from the
surface of a solid and their transition into the gaseous state.
For the particles of a solid to be able to leave its surface,
energy must be spent to overcome the intermolecular attrac-
(ion of the particles and tear them away from the surface
of the solid. The specific heat of vaporization or sublimation
of a solid is the amount of heat that is needed to evaporate
a unit mass of it. It follows from the law of conservation of
energy that the difference between the specific heats of
evaporation of solids and liquids at their melting point
equals the specific heat of fusion.
Problem. How much ammonia has to be evaporated and
Mien heated to 0 °C in a refrigerator to produce 40 kg of ice
from water taken at 10 °C at the expense of the absorbed
amount of heat?
The boiling point of ammonia is —33.4 °C, its specific
heat of vaporization at the boiling point is Lb = 1.37 X
X 106 J/kg, the specific heat capacity of ammonia in the
202 Molecular Physics and Thermodynamics 11.7.4

gaseous state is c2 = 2.1 X 103 J/kg-K. The specific heat


of fusion of ice at 0 °C is L{ = 3.35 X 105 J/kg, and the
specific heat capacity of water is cx = 4.187 X 103 J/kg-K.
Given: Water: M x — 40 kg, cx = 4.19 X 103 J/kg-K,
tx = 10 °C, t\ = 0 °C, Lf = 3.35 x 105 J/kg.
Ammonia: c2 = 2.1 X 103 J/kg-K, t2 = —33.4 °C, t'2 =
= 0°C, Lh = 1.37 X 106 J/kg.
Required: M 2.
Solution: The amount of heat given up by the water is
Q i = M i [(^i — ^i) c i + Lf]
The amount of heat received by the ammonia is
Q2 = M 2 [(^2 — h) c2 + ^bl
It is evident that Qx == Q2, whence
u M i [Qi ti) c1 Lf\_
2 (*2—*2)C2+ A)
_ 40[(10 — 0 )x 4 .1 9 X lw » + 3 .3 5 x lO B] ?|
— [U— ( — 3 3 .4 )]X 2 .1 X l >3 + 1.37 X t» 8 ~ ’ g
FUNDAMENTALS
PART OF
ELECTRODYNAMICS

Chapter 1

Electrostatics

1. B a s ic C o n ce p ts.
T h e L a w o f C o n se r v a tio n o f E le c tr ic C h arge

1°. Electrostatics is the branch of electrodynamics


(111.1.3.1°) that deals with the properties and interactions
of bodies or particles having an electric charge that are at
rest in an inertial reference frame (1.2.1.3°).
2°. An electric charge is defined as the physical quantity
characterizing the property of bodies or particles to enter
into electromagnetic interactions (111.1.3.1°) and deter­
mining the values of the forces and energies in such interac­
tions. Electric charges can be positive and negative. A posi­
tive charge appears, for example, on glass rubbed with
leather, a negative one on amber rubbed with fur.
3°. Elementary particles (VI.5.1.1°) and their antipar­
ticles (VI.5.5.1°) are stable carriers of electric charges.
A proton and a positron carry a positive charge, an electron
end an antiproton carry a negative one.
The smallest stable particles having a negative (positive)
electric charge and entering the composition of a substance
204 F u n dam en tals of E le c tro d y n a m ic s 111.12

are electrons (protons). The absolute value of the electric


charge of a proton and an electron is 1.602 X 10“19 C =
= 4.80 X 10”10 cgse (VII.8). The masses of a proton and
an electron are 1.67 X 10~27 and 9.1 X 10“31 kg, respec­
tively. The electric charge of a proton and an electron is
called an elementary charge.
4°. The electric charge of any charged body equals an
integral number of elementary charges. An electrically
neutral (uncharged) system contains an equal number of
elementary charges of opposite signs. Atoms, molecules,
and their collectives—macroscopic bodies—are electrically
neutral.
5°. If the electrically neutral state of a body is violated,
it is said to be electrified. To electrify a body, a surplus
(shortage) of electric charges of one sign or the other must
be created on it. Bodies are electrified in different ways,
the simplest of which is by contact. Upon the contact of
some bodies consisting of different substances, the valence
electrons of atoms (VI.2.9.2°) of one of the substances pass
over into the other substance.
6°. In all the phenomena associated with the redistribu­
tion of electric charges in an isolated system of interacting
bodies, the algebraic sum of the electric charges remains
constant (the law of conservation of electric charge). The law
of conservation of electric charge, like the laws of conser­
vation of energy (1.5.4.1°) and of momentum (1.2.6.2°),
is one of the fundamental laws of physics.

2. C o u lo m b ’s L a w

1°. Electric charges are called point ones if they are


distributed on bodies whose linear dimensions are consid­
erably smaller than any other distances encountered in
a given problem. For example, the charges on two interact­
ing metal charged spheres each having a radius of 1 mm
and 1 m apart may be considered as point ones if the force
of interaction between the spheres is being calculated.
If the field intensity (111.1.3.3°) of one of these spheres is
being calculated for a point at a distance of 15 mm from
Electrostatics 205

its centre, however, then the charge of the sphere can no


longer be considered as a point one.
2°. The forces of electrostatic •interaction depend on the
shape and dimensions of the electrified bodies and the
nature of the distribution of the charges on them. For fixed
point charges qx and g2, and also f°r spherical charged bodies
if their charges qx and q2 are uniformly distributed through­
out their entire volume or over their surface, these forces
obey Coulomb's law: the magnitude of the force F of electro­
static interaction between the charges qx and q2 in a vacuum
is directly proportional to the product of the charges and
inversely proportional to the square of the distance r between
them:

where k is a proportionality constant depending on the


system of units used in calculations. For interacting uni­
formly charged spheres, their radii R x and R 2 may be com­
mensurable with the distance r between their centres. The
force F is called the Coulomb force.
3°. Coulomb forces, like gravitational ones, obey New­
ton’s third law (1.2.5.1°): the forces of interaction between
charges are equal in magnitude and directed oppositely
to each other along the straight line connecting the point
charges or the centres of the spheres. The Coulomb force
is a central one (1.2.8.1°).
Experiments show that unlike electric charges attract
each other, and like ones repel each other. In this respect,
Coulomb forces differ in principle from forces of gravity,
which are always forces of attraction (1.2.8.1°).
4°. The force of repulsion exercised on the charge q2
by the like charge qx is directed the same as the position
vector r drawn to this charge (Fig. 1 1 1 . 1.1a).
The force of attraction exercised on the charge q2 by the
unlike charge qx is opposite in direction (Fig. III. 1.16).
The electrostatic forces of repulsion are conventionally
considered positive, and the forces of attraction negative
(cf. II.1.4.2°-4° on the forces of intermolecular interaction).
The adopted signs of the forces of attraction and repulsion
correspond to Coulomb’s law: the product of like charges
206 F u n d a m en ta ls of F ie c iro d y n a m ic s tti.1 .2

is a positive number, and the force of repulsion has a positive


sign. The product of unlike charges is a negative number,
and this corresponds to the sign of the force of attraction.

5°. Coulomb’s law for the interaction of point charges


or charged spheres (Par. 2°) in a vacuum is written in the
form
Tp_ *Zl92
— 4jie0r2

where e0 is the electric constant (the permittivity of free


space) in the SI system of units (VII.5.1°).
In the absolute electrostatic system of units (cgse)
(VII.5.2°)
F=
r2
6°. The magnitude of the force of electrostatic interac­
tion of point charges or uniformly charged spheres in a ho­
mogeneous and infinitely extending gaseous or liquid
dielectric (111.1.6.1°) is

F = 4ne„er2 <in the SI system)


F= (in the cgse system)

where e is the dielectric constant (the relative permittivity


of the medium).
111.1.2 Electrostatics 201

The quantity e is always greater than unity and shows


how many times the force of interaction between the charges
qx and q2 in a given medium is Jess than in a vacuum. The
definition of e given above cannot be applied to solid die­
lectrics (see also III.1.6.5°-8°).
Problem 1. Point charges of 10“7 and 10~6 C interacted
in a vacuum with a force of 0.36 N. Next the charges were
placed in kerosene. For a vacuum, ex = 1, and for kerosene
e2 = 2.0. By how much must the distance between the
charges be changed for the force of their interaction to
remain the same?
Given: qx = 10"7 C, q2 = 10~6 C, F = 0.36 N, e1 = 1,
e2 = 2.0.
Required: rl7 r2.
Solution: According to Coulomb’s law, F = #ig2/(4Jieoer2)>
where e0 is the electric constant in the SI system, e0 =
= 8.85 X 10~12 F/m, and e is the relative permittivity
of the medium.
The distance between the charges in a vacuum is rx =
= ]/ ^ 2/(4jie0F), and in kerosene is r2 = ]/ qxq2l(^mQE2F).
Hence,

Problem 2. With what force do two identical small


spheres interact in a vacuum if one carries a charge of 6.0 X
X 10“9 C, and the second a charge of —3.0 X 10"e G.
The distance between the spheres is 5.0 cm. With what
force will these spheres interact if they are brought into
contact and then removed to their previous positions?
Given: qx = 6.0 X 10~9 C, q2 = —3.0 X 10“9 C, r =
= 5.0 cm = 5.0 X 10~2 m.
Required: Fx, F2.
Solution:
JP ___ Mi<Ji2
q ___ 6.0
U X1 10~*
.U A V X
A ^(—
---- O3.0
. W AX 10~9)
)

P i~ 4jie0r2 — 4 x 3 .1 4 x 8 .8 5 x l t i - 12x 2 5 x l 0 - 4
« —6.6xl0"5N
208 Fundamentals of Electrodynamics 1II.1.3

The spheres will be attracted (indicated by the minus


sign) with a force of Fx « —6.6 X 10~5 N.
Upon contact of the spheres, part of the charges will
be compensated. A summary charge equal to qx + q2 =
= 6.0 X IO"9 - 3.0 X 10-9 = 3.0 X lO'9 C will remain
on both spheres. This charge will be distributed between
the spheres equally. The charge of each sphere will be
g3= 3 - ( ) X l (>- 8 = 1 . 5 X 1 0 - 9 G

Thus,
jp _ M s _ 1.52 X iO-18
2 4jce0r2 4 X 3.14 X 8.85 X IO"12 X 25 X IO"4
= 0.81 xlO -5 N
The spheres will repel each other with a force of F2 =
= 0.81 X IO"5 N.

3. Electric Field. Field Intensity


1°. Electromagnetic interaction is defined as the interac­
tion between electrically charged particles or macroscop-
ically charged bodies.
The branch of physics studying electromagnetic interac­
tions is called electrodynamics.
An electromagnetic -field is a form of matter by means
of which electromagnetic interactions of charged particles
or bodies that in the general case are moving in the given
reference frame are carried out.
An electric field is a part of an electromagnetic field
whose feature is that it is created by electric charges or
charged bodies and also acts on them regardless of whether
they are moving or are stationary. An electric field is des­
cribed by definite force and energy characteristics (111.1.8.1°).
If electrically charged particles or bodies are stationary
in a given reference frame, then they interact by means
of an electrostatic field. The latter is an electric field unchang­
ing in time (stationary). In the general case, an electric and
an electromagnetic field can change with time (varying,
non-stationary electric and electromagnetic fields).
111.1.3 Electrostatics 209

2°. Modern physics is based on the theory of short-range


force action. According to it, varying electromagnetic fields
propagate in space with a finite speed equal to that of
light and act on charged particles or bodies that are in
space. The finite nature of the speed of propagation of
electromagnetic interactions underlies the special theory of
relativity (V.4.2.10).
In the previously existing theory of long-range force
action, it was considered that all interactions (for example
electromagnetic and gravitational ones) propagate with an
infinitely great speed, i.e. occur instantaneously, directly
between particles and bodies that are remote from one
another. At present, this theory is only of historical interest.
3°. The force characteristic of an electric field is the
field intensity vector E:
E - i
Q
where F is the force acting on the positive charge q placed
at the given point of the field. The intensity of an electric
field at a given point of it is equal to and coincides in direc­
tion with the force acting on a stationary unit positive
point charge (a test charge). It is considered here that the
test charge does not distort the field being studied with its
aid, and its own electric field is ignored.
4°. The force acting on the charge q placed in an electric
field of intensity E is
F = qE
An electric field is homogeneous if its intensity vector E
is the same at all points of the field. Examples of such
fields are the electrostatic fields of a uniformly charged
infinite plane (111.1.4.4°) and of a plane capacitor far from
the edges of its plates (111.1.11.2°).
5°. The method of lines of force, or field lines, is used
for the graphical depiction of an electrostatic field. The
field lines are imaginary lines, the tangents to which at each
point along them coincide with the direction of the intensity
vector at this point of the field (Fig. III.1.2). Field lines
are open—they begin at positive charges and end at negative
ones. Field lines never intersect because at each point of
14-0211
210 Fundamentals of Electrodynamics 111.1.3

a field its intensity has a single value and a definite direc­


tion. (For examples of field lines of selected electrostatic
fields see III.1.4.1°-5°.)
6°. If a positive electric charge moves in a homogeneous
electric field, and its initial velocity is directed along

Fig. III.1.2

a field line, then the path of the charge will coincide with
the field line.
7°. Every electric charge sets up an electric field in
space regardless of the presence of other charges. The prin­
ciple of superposition of electric fields: the intensity of the
electric field of a system of N charges equals the vector sum
of the intensities of the fields produced by each of them
separately:
N
E —Ei + E2 + . . . + Ejy — 2 Ei

where N is an arbitrary positive integer.


8°. If electric charges set up a field in a medium that
is an isotropic homogeneous dielectric (111.1.6.1°), then
with the given arrangement of electric charges in space the
intensity of the electrostatic field in such a medium will
be 1/e-th of that in a vacuum: E = E0/e, where E0 is the
intensity of the field set up by the given system of charges
in a vacuum, and e is the relative permittivity of the me­
dium (111.1.2.6°).
Problem 1. A droplet of water 0.1 mm in diameter is
suspended in oil with an electric field intensity of 104N/C.
The intensity of the homogeneous field is directed vertically
upward. How many elementary charges are on the droplet?
The density of the oil is 8 X 102 kg/m3.
n i .l .3 Electrostatics 211

Given: d = 0.1 mm = 10“4 m, E = 104 N/C, px =


= 1.0 X 103 kg/m3, p2 = 0.8 X 103 kg/m3, e = 1.6 X
X 10"19 c .
Required: N.
Solution: Three forces act on the droplet:
(1) the force of gravity directed vertically downward,
p = mg = (jt<23/6) pjg;
(2) the buoyant Archimedean force (1.6.2.3°) Fx =
= (jtcP/6) p2g directed vertically upward; and
(3) the electric force directed vertically upward, F2 —
= qE = NeE.
The droplet is in equilibrium provided that P = F1 +
+ F 2, i.e.
Ti/CP . *j rp
-a~9ig=-cc92g + NeE
whence
__n<Pg ( Pi — p2) _
— 6
_ 3 .1 4 x 10-12 9.8 X 103 (1.0 — 0.8)
~ 6 1.6 X 10“19 X iO4
6 x l 0 6 ele-
mentaryjcharges
When the charge of the droplet is known, this problem
makes it possible to determine the elementary charge
(111.1.1.3°) (the setup of the Millikan experiment for mea­
suring the charge of an electron).
Problem 2. A pith ball whose mass is 0.5 g and whose
charge is —9.8 X 10"9 C is hanging on a silk thread. Through
what angle will the thread divert if the ball is introduced
into a homogeneous electric field of intensity 5 X 104 N/C
directed horizontally (Fig. III.1.3)?
Given: q = —9.8 X 10“9 C, m = 0.5 g = 5 X 10~4 kg,
E = 5 X 104 N/C.
Required: a.
Solution: The equilibrium angle of deflection a is deter­
mined from the condition tan a = F/P, where F = \q\ E
is the electric force, and P = mg is the force of gravity.
Ifence,
tan a = Imq g\ E _
— 9 . 8 X 10-9 X 5X104
5x10-4x9.8 0.1, whence a = 6‘
14*
212 Fundamentals of Electrodynamics 111.1.3

Problem 3. An electric field is formed by two identical


point charges of 5 nC opposite in sign. The distance between
the charges is 10 cm. Determine the field intensity (1)
at a point between the charges and equidistant from them;
(2) at a point on the continuation of the
{///////////. line joining the centres of the charges that
is 10 cm from the negative charge; and (3)
at a point that is 10 cm from the positive
and negative charges (Fig. III.1.4).
Given: | | = | | = q = 5 nC =
= 5 X 10“9 C, a = 10 cm = 10”1 m.
Required: EA, EB, Ec .
Solution: In accordance with the prin­
ciple of field superposition, the intensity
of the field set up by the two charges is
determined by the sum of the field inten­
Fig. III.1.3 sities set up by the separate charges:
E = E_ + E+, where E_ and E+ are in­
tensity vectors of the fields set up by the negative and
positive charges, respectively.
At the point A, the field intensities E_ and E+ are
directed identically. Therefore, the magnitude of the inten­
sity vector at the point A is

IEA| = 2 1E+1= 2 1E_ | = ------


4jie0
________2 x 5x 10~9______
4x3.14x8.85x10-1*
= 3 . 6 x l 0 4V/m
At the point B, we have | E^ | = <jr+/[4jie0 (2a)2] because
the point is at the distance 2a from the charge q+, and
| E l | = g_/(4ne0a2). Hence,

_ 5xl0-» V 3 —
— 4 X3.14 X8.85 X 10-12X 10-* X 4
= 3 . 4 x l 0 3 V/m
111.1.4 Electrostatics 213

At the point C, we have | Ec | = | E" | cos a +


+ | El | cos a. It follows from the conditions of the prob­

lem that | El | = | E l |, the angle a in an equilateral


triangle is 60 degrees, and cos 60° = 0.5. Therefore,

q
| Ec | = 2 1E+ | cos a = 2 x 4jie0a2 q
X 0.5 = 4j i e0fl2

= 4 . 5 x l 0 3V/m

The directions of the vectors Eb and Ec are shown in


(lie drawing.

4. Examples of Electrostatic Fields

1°. The intensity of the electrostatic field of a point


charge q in a dielectric (111.1.2.6°) is

E = 44-a^ T On the SI system)

E = -^ -y (in the cgse system)


214 Fundamentals of Electrodynamics III.1.4

where r = position vector conducted from the point charge


to the point of the field being studied
8 = relative permittivity of the medium
80 = electric constant in the SI system (VII.5.1°).
Figure III.1.5 shows the electrostatic fields of a posi­
tive (a) and a negative (b) isolated point charges and the

directions of the field intensity vectors. The magnitude


of the field intensity vector of a point charge is

E= ^ r ^ (in the S1 systen,)


E= (in the cgse system)

2°. The intensity of the electrostatic field of a system


of N point charges q^, q2, . . ., qN, according to the super­
position principle (111.1.3.7°), is

E = 4ji80
r ~ 2 -v rTf -r-t (in the SI system)
i= \
N
V qi Vi (in the cgse system)
er? r t
i= l

Figure 111.1.6 shows the electrostatic fields of two unlike


(a) and like (b) charges and the directions of the relevant
field lines.
111.1.4 Electrostatics 215

3°. The intensity of the electrostatic field of a sphere


of radius R having the charge q uniformly distributed over
its surface is (with r ^ R) #

E= i ^ T <in the SI system)

E= ~ - — (in the cgse system)


where r is a position vector conducted from the cent e of
Ihe sphere to the point of the field being studied (for the

Fig. III.1.6

remaining notation see Par. 1°). The magnitude of the


held intensity vector of a sphere is (with r ^ R)
1 Q
E= Z (*n t*ie SI system)

E = -—g- (in the cgse system)


The electrostatic field outside a charged sphere coincides
with the field of a point charge (equal to the charge of the
sphere) placed at the centre of the sphere (Fig. III. 1.7).
The intensity of the electrostatic field inside a sphere whose
surface is charged equals zero (111.1.5.3°).
4°. A uniformly charged infinite plane sets up a homo­
geneous electrostatic field, the magnitude of whose inten­
sity is
E = -^—
Z808 (in the SI system)
E = —5Z. the cgse system)
216 Fundamentals of Electrodynamics IIL1.4

where a is the surface charge density equal to the electric


charge on a unit area of the surface: a = q/S. The field
lines are at right angles to the plane. Figure III.1.8 shows
the electrostatic fields of a uniformly charged positive (a)
and negative (b) infinite isolated planes.
5°. Two uniformly charged infinite parallel planes having
the same charge density a, but opposite charges, produce
a homogeneous electrostatic field. Its
intensity in the space between the
planes has the magnitude
E = -e0B
— (in the SI system)

(jn the CgSe system)

and in the remaining space E = 0


(Fig. III.1.9).
6 °. An electric dipole is defined as
a combination of two equal and oppo­
site point charges +<7 and —q at the distance I from each
other (Fig. 111.1.10). Since the charges of a dipole are in
different points of space, they Jdo fnot compensate each

Fig. III.1.8

other in intensity, and each of them sets up its own


electric field. According to the superposition principle
(111.1.3.7°), the intensity of the electrostatic field of
a dipole equals the sum of the intensities of the fields
III.1.4 Electrostatics 217

created by each of the dipole charges. The magnitude of the


intensity vector at the point O sufficiently removed from
the dipole (r > Z) (see Fig. IH.1.10) is

E= y 3 cos2 0 + 1 (in the SI system)

£ = -^|-]/^3cos2 0 + l (in the cgse system)

Here p e is the magnitude of the vector p e— the electric


dipole moment:
Pe = ql
The vector 1 is directed along the axis of a dipole from the
negative charge to the positive one.

each other with a force whose magnitude equals


p __ 6Pe,iPe,s
4jie0r4
where r = distance between the dipole centres, r > I
I = length of a dipole.
If the dipoles face each other with their opposite charges,
they attract each other. Otherwise they repel each other.
Problem 1. The surface density of the charge on a uni­
formly charged sphere is 6.4 X 10"8 C/m2. Find the electric
field intensity at a point that is six radiuses from the centre
of the sphere.
Given: a = 6.4 X 10“8 C/m2, r = 6R.
218 F u n d a m en ta ls of E le c tro d y n a m ic s 111.1.4

Required: E.
Solution: The electrostatic field of a sphere with a
charged surface outside the sphere is similar to the field
of a point charge at its centre: E = ^/4jte0er2.
The charge on the sphere is q = aS = a4jii?2, where R
is the radius of the sphere. The relative permittivity e = 1;
hence,
o4nR2 _ a _ 6.4 xlO”8 2.0x 102V/m
4jie0e36i?2 ~ 36a0 ~ 8.85 X 10"12 X 36

Problem 2. What force acts on a charge of 0.1 nC placed


in the field of a uniformly charged plane with a charge
surface density of 10"5 C/m2? The relative permittivity
of the medium e = 5.
Given: <7 = 0.1 nC = 1 X 10“l0 C, a = 10-5C/m2, e = 5.
Required: F.
Solution: The force acting on the charge is F = qE, where
E is the intensity of the field produced by the uniformly
charged plane, E = a/2e0e. Hence,
a
F = q 2e0e
IO x 10~5
-1 0
= 1.13 x 10“5 N
2 X 8.85 X 10“12 X 5

Problem 3. A dust particle having a mass of 2.0 X


X 10"12 g is suspended in air between two horizontal oppo­
sitely and uniformly charged plates. The intensity of the
field of the plates is directed vertically upward. The charge
of the particle equals five elementary charges. Find the
charge on the plates. The area of each plate is 100 cm2.
Given: m = 2.0 X 10"12 g = 2.0 X 10"15 kg, e = 1.0,
q = 5 X 1.6 X 10-19 C = 8.0 X 10-19 C, S = 100 cm2 -
- 10-2 m2.
Required: qvl.
Solution: The dust particle will remain suspended pro­
vided that P = F, where P = mg and F = qE. The field
intensity of two uniformly charged plates is E = a/e0e,
where a =■^1/5'. Hence,
_m g __ e ___ 2.0 X 10"15 X 9.8
^P1— ~q ee°° ~ 8.0X10-™ X
X 1 .0 x 8.85X 10-12 x 10-2 «2.2x 10-9 C
III.1.5 Electrostatics 219

5. Conductors in an Electrostatic Field


1°. Conductors are substance in which the ordered
motion of electric charges may occur, i.e. in which an
electric current (111.2.1.1°) may flow. Conductors include
metals, aqueous solutions of salts, acids, etc., and ionized
gases. When a metal is formed, the valence electrons of the
atoms interacting with one
another detach themselves and
become free (collective) electrons
(conduction electrons of metals).
2°. If a metal conductor
is placed in an electric field,
then under the action of this
field ordered motion of the
electrons will be superposed
on their thermal chaotic mo­
tion, and they will move in
a direction opposite to the
field intensity. For example,
in a conductor placed in an
external homogeneous elec­
tric field with the intensity
Eext, the electrons will move
from the right to the left
(Fig. III.1.11). A surplus ne­
gative charge will appear on
the surface AB of the conductor, and a surplus positive
charge on the surface CD. The charges appearing on the
surfaces of the conductor set up an internal electric field
in it whose intensity vector Eint is directed oppositely
to the vector Eext of the intensity of the external electric
Held. The resultant electric field in the conductor will have
the intensity E = Eext + Eint whose magnitude is smaller
than that of Eext:
IE | = | Eext | — | Eint |

When | Eext | = | Elnt |, the force exercised on the conduc­


tion electrons vanishes, and the ordered motion of the
charges in the conductor terminates.
220 F u n d a m en ta ls of E le c tro d y n a m ic s 111.1.5

3°. The conduction electrons in a metal conductor are


redistributed under the action of an external electrostatic
field so that the intensity of the resultant field at any point
inside the conductor equals zero; the non-compensated
electric charges are arranged stationarily only on its sur­
face.
4°. The phenomenon of the redistribution of the charges
in a conductor in an external electrostatic field is called
electrostatic induction. It consists in the separation of the

\g. III.1.12

positive and negative charges, whose number in a conductor


is identical. The charges separated by the electrostatic
field (the induced charges) mutually compensate each other
if the conductor is removed from the field, and the ordinary
state of a metal solid is restored.
5°. If a conductor is hollow inside, then the intensity
of the electrostatic field in this hollow space equals zero
regardless of the field outside the conductor and of how the
latter is charged. The hollow space in the conductor is
screened (protected) from external electrostatic fields. This
underlies electrostatic protection: if an instrument is sur­
rounded by a closed metal surface, then no external electric
fields will act on it. Usually, an earthed copper netting
playing the part of a screen is used for this purpose. The
potential of the screen remains equal to that of the Earth
(111.1.8.4°).
6°. At all points of the surface of a charged conductor,
the intensity of an electrostatic field is perpendicular to the
surface. If this were not the case and there would be a tan­
gent component of the intensity E T directed along the
111.1.6 Electrostatics 221

surface AB of a charged conductor, it would cause motion


of the electric charges along the surface (Fig. III.1.12).
But this contradicts the essential* equilibrium distribution
of charges on the surface of a charged conductor. Conse­
quently, E x = 0, and E = En =£ 0, where En is the normal
(perpendicular to the surface) component of the electrostatic
held intensity.
7°. At all points inside a charged conductor, its poten­
tial cp is the same (111.1.8.1°). The surface of a charged
conductor is an equipotential surface (111.1.9.2°).

6. Dielectrics in an Electrostatic Field


1°. Dielectrics are defined as substances that do not
conduct an electric current (111.2.1.1°). Free electrons
(111.1.5.1°) are virtually absent in dielectrics, and the
ordered motion of electric charges in conventional condi­

E=0 E*0
(Cl) (b)
Fig. III.1.13

tions is impossible. A dielectric becomes conductive when


n very high breakdown voltage (111.2.3.4°) is applied to it.
This phenomenon is called breakdown of the dielectric.
(A dielectric may also be made conductive by heating
it to a high temperature.)
Dielectrics include some solids (glass, porcelain, etc.),
liquids (chemically pure water, methyl chloride CHSC1,
etc.), and gases (hydrogen H2, nitrogen N 2, carbon tetra­
chloride CC14, ammonia NH3, etc.).
A dielectric is called homogeneous and isotropic if all
its properties are the same at any point and in all directions
222 Fundamentals of Electrodynamics 111.1.6

inside il. The valence electrons in the atoms of dielectrics


are tightly bound to their nuclei and in ordinary conditions
cannot detach themselves from them.
2°. The molecules of a dielectric are electrically neutral—
the sums of the positive charges of their nuclei and of the
negative charges of all the electrons are equal. It does not
follow from this, however, that the molecules of a dielectric
have no electric properties. The molecules of a dielectric
produce an electric held equivalent to that of an electric
dipole with the electric moment p e equal to # 1 (the dipole
moment of a molecule) (111.1.4.6°), where q is the positive
charge of a molecule (or the negative charge equal to it),
and I is the distance between the centres of mass (1.2.3.4°)
of the positive and negative charges.
3°. Polar and non-polar dielectrics are distinguished
depending on the structure of the molecules. If in the absence
of an external electric held the centres of mass (1.2.3.4°)
of the positive and negative charges in a molecule of a di­
electric coincide, it is called non-polar.|In the absence of an
external electric held, the dipole moment of a molecule of
a non-polar dielectric equals zero. For example, in a hydro­
gen atom, the electron travels along its orbit with such
a great speed that on the average its position coincides
with that of the nucleus—a proton. Therefore, 1 = 0
(Par. 2°), and pe = 0 (Fig. III. 1.13a). If a hydrogen atom
is put in an external electric held, the latter will act on the
nucleus and the electron with forces whose magnitude is
F = eE. The directions of the vectors F of the forces exer­
cised on the nucleus and the electron will be opposite.
The result will be deformation of the electron’s orbit, and
the centres of mass of the electron and the nucleus will
no longer coincide (1 0) (Fig. III.1.136). An induced
electric dipole moment appears whose magnitude is p e = el,
where e is the absolute value of the charge of an electron.
The molecules of all non-polar dielectrics behave in an
external electric held like a hydrogen atom. When such
dielectrics are introduced into an external electric held,
deformation occurs in the molecules (atoms), and an induced
electric dipole moment of the molecules appears. The mole­
cules of non-polar dielectrics in an electric held are similar
in their electric properties to induced quasi-elastic dipoles.
iil.i.6 E lec tro sta tics 223

4°. The nuclei and electrons in the molecules of polar


dielectrics are arranged so that the centres of mass of the
positive and negative charges da not coincide. Such mole­
cules, regardless of the external electric field, behave like
rigid dipoles having an electric moment whose magnitude p e
is constant (I = const). For example, in a molecule of H20
schematically shown in Fig. III.1.14, the centres of mass
of the positive charges of the nuclei of the three atoms and
all their electrons are spread apart. A water molecule in
its electric properties is similar to a very extended rigid
dipole.
If a polar dielectric is not placed in an external electric
field, then the thermal chaotic motion of the molecules and

cc cVc ccc ©e
V ecc €c
V €c
c
B

Fig. III.1.15

their continuous collisions with one another result in a


lack of order in the arrangement of the rigid dipoles.
Dipoles will always be present whose electric fields will
have intensities directed oppositely. Therefore, although
every dipole creates an electric field (111.1.4.6°), the total
intensity of the field of all the dipoles of a dielectric will
equal zero according to the principle of superposition of
fields (111.1.3.7°).
5°. When a dielectric is introduced into an external
electric field, it becomes polarized. By polarization of a di­
electric is meant its transition into such a state when within
a small volume of the substance the geometric sum of the
electric dipole moment vectors of the molecules differs
from zero. A dielectric in which polarization occurred is
called polarized. The mechanism of polarization differs for
non-polar and polar dielectrics.
6°. If a homogeneous non-polar dielectric is introduced
into a homogeneous electric field whose intensity vector E
224 Fundamentals of Electrodynamics n i.i.6

is directed as shown in Fig. 111.1.15, then the positive


and negative charges will be displaced in the molecules
of the dielectric. Electric surface-bound charges will appear
on the surfaces AB and CD confining the dielectric. Polar­
ization is characterized by the appearance of surface-
bound charges on the surfaces of a dielectric placed in an
external electric field.
The surface charges are called bound because they appear
as a result of deformation of the dielectric molecules and

Pe
0

E
(b)

cannot be torn away from them [cf. the free charges on the
surface of a conductor (111.1.5.2°)]. Bound charges do not
manifest themselves inside any volume of a dielectric:
the total electric charge of the molecules in this volume
equals zero. On the surfaces AB and CD of a dielectric,
the bound charges are not compensated and create the
internal electric field of the dielectric itself. The vector
Elnt of the intensity of this field is directed into the dielec­
tric opposite to the direction of the intensity of the external
electric field that caused polarization. Therefore, the resul­
tant electric field in a homogeneous isotropic dielectric
has an intensity that is 1/e-th of that in a vacuum (111.1.3.8°).
The polarization of a dielectric with non-polar mole­
cules consisting in the appearance of an induced electric
dipole moment (Par. 3°) in the molecules is called electron
or deformation polarization. Polarization of this kind does
not depend on the temperature of the dielectric. The inten­
sity of the thermal motion of the molecules does not affect
the appearance of induced electric dipole moments of the
molecules.
111.1.6 Electrostatics 225

7°. The forces F = \ q \ E equal in magnitude and


directed oppositely will act on each of the charges of a rigid
dipole (Par. 4°) placed in a homogeneous electric field
of intensity E (Fig. III.1.16a). They will produce a moment
of force (1.3.1.4°) that tends to turn the rigid dipole so that
the vector of its electric dipole moment p e is directed paral-

0--O 0—0 0--0


0--© 0 —0 0- -©
0 --0 0 —0 0 --0 0 .-0
0--O 0 —0 ©--©
©~-e 0 —o 0--O 0+-O ^ o _ (5t °
B
(d) (b)

Fig. III.1.17

lei to the field intensity vector (Fig. 111.1.166). When


a homogeneous polar dielectric is placed in an external
homogeneous electric field, each molecule—a rigid d ip o le -
will experience the orienting influence of the field and will
tend to turn as shown in Fig. III.1.166. The thermal chaotic
motion of the molecules of a polar dielectric prevents rota­
tion of the dipoles along the field intensity E. The orienta­
tion of the rigid dipoles will be greatest with a very strong
external electric field*. (The only effect of the thermal
motion in a very strong field is a “trembling” of the dipoles
oriented along the direction of the field.)
Uncompensated bound electric charges of opposite signs
will appear on the boundary surfaces AB and CD of a pola­
rized dielectric (Fig. III.1.17a). In ordinary, not too strong
external electric fields, the dipoles will be oriented mainly
along the field intensity, and a smaller number of bound
electric charges will appear on the surfaces AB and CD
(Fig. 111.1.176) than in a strong field. The bound charges

* A quantitative assessment of a strong and a weak electric fields


cannot be given in a book on elementary physics.
1 5 -0 2 1 1
226_______________ F u n d a m e n t a l s of E l e c t r o d y n a m i c s ____________ 1 1 1 . 1 . 7

in a polar dielectric, as in a non-polar one, create an internal


electric field (Par. 6°). Polarization of this kind is called
orientation polarization. The latter diminishes with eleva­
tion of the temperature.
8°. Ionic polarization is possible in solid crystalline
dielectrics such as NaCl having an ionic crystal lattice
(11.7.1.2°). It consists in that when such dielectrics are
placed in an external homogeneous electric field, the posi­
tive ions of the lattice are displaced in the direction of the
field intensity vector, and the negative ions in the opposite
direction.
9°. Ferroelectrics form a special group of crystalline
dielectrics. An example of them is Rochelle salt
(KNaC4H40 6-4H20). Ferroelectrics have enormous values
of the relative permittivity e depending on the intensity E
of the electric field in which the substance is.*

7. W ork of the Forces


of an E lectrostatic Field
1°. The force F acting on the charge q' in an electrostatic
field of intensity E is F = q'E.
The elementary work AA (1.5.1.1°) of the force F when
the charge q' is displaced by A1 is
AA==F AI cos a = q'E Al cos a
where AI = magnitude of the vector of the elementary
displacement A1 (1.1.2.1°)
a = angle between the directions of the vectors E
and A1 (Fig. III.1.18).
2°. The work done upon the displacement of the charge
q' between the two points B and C of an electrostatic field
equals the sum of the elementary works (Fig. III.1.19):
A —AA\ -j- AA% -f- . . . = ^ Q AIt cos

* T he co n cep t of th e r e la tiv e p e r m ittiv ity e introd u ced in


111.1.2.6° can n ot be a p p lie d to ferroelectrics. In form ation on the
p r a c tic a lly im p o rta n t p rop erties of ferroelectrics is beyon d the scope
o f th e presen t book.
111.1.7 Electrostatics 227

Example 1. The work A upon the displacement of the


positive charge q in a homogeneous field of intensity E
between the points 1 and 2 is •
A = —qE (x2 — Xj)
where xx and x2 are the coordinates of the points 1 and 2
along the X-axis. The minus sign shows that the directions

of the force F = qE and of the projection of the displace­


ment vector 1 onto the X-axis are opposite (Fig. III.1.20).
Example 2. The work A upon the displacement of the
point charge qf between the points B and C in the electro­
static field created by the point charge q is

a ==^ { t; -7 ;) <in the SI system)


A = ^ - ^ j — y -j (in the cgse system)
where rx and r2 = distances from the points B and C to the
charge q (see Fig. 111.1.19)
e0 = electric constant in the SI system
(VII.5.1°)
e = relative permittivity (111.1.2.6°).
3°. The work needed to move an electric charge from
one point of an electrostatic field to another does not depend
on the form of the path, but depends only on the initial
and final positions of the charge (the property of poten­
tiality of electrostatic forces). Thus, in Example 1, the
work needed to move the charge along the path 1-2 equals
that needed to move it along the path 1-3-2. In Example 2,
the work needed to move the charge q' along any of the
paths BC, BIC, and BIIC is the same. The work done by
228 F u n d a m e n ta ls of E le ctro d y n a m ic s in.i.7

electrostatic forces in moving a charge along a closed path


in an electrostatic field equals zero. When xx = x2 in Exam­
ple 1, and when rx = r2 in Example
E 2, the work equals zero.
4°. The work of electrostatic forces
.^ ?2 of repulsion of like charges is pos­
<// -
J A
itive if the charges move away
from each other, and negative if the
charges move toward each other. The
1 / work of electrostatic forces of attrac­
J lL .
rr^ tion of opposite charges is positive
if the charges approach each other,
! and negative if they move away from
each other.
5°. The work done by electrostat­
Fig. I I I . 1.20 ic forces when an electric charge, q
is moved in an electric field equals
the reduction in the potential energy W of this charge:
A = - A W = —(W2 - Wx) = W1 — W2
where Wx and W 2 are the potential energies (1.5.3.5°) of
the charge at the initial and final points of its path, respec­
tively.
Example 1. The potential energy of a charge g in a homo­
geneous field of intensity E (see Fig. III.1.20) is
W = —qEx
where x is the coordinate of the charge if we assume that
W = 0 when x = 0.
Example 2. The potential energy of the charge q' at
a given point of an electrostatic field at the distance r from
the point charge q that has produced the field (the potential
energy of interacting charges) is

(in the SI system)


PF = — - (in the cgse system)

It is considered that W tends to zero when r tends to infinity


(for the notation see Par. 2°, Example 2).
Ifl.1 .8 E le c tro sta tic s 229

6 . The potential energy of repulsion of like charges


is positive and increases if the charges approach each other.
The potential energy of at­
traction of unlike charges is
negative and grows up to zero
if one of the charges is re­
moved from the other over a
very great distance (r -»• oo)
(Fig. III.1.21).
Problem. Two parallel
plates in air each having the
area 2 X 10"2 m2 are charged
with unlike charges of 100 nC.
What work has to be done to
increase the distance between
the plates by 0.1 mm?
Given: S = 2 X 10”2 m2, q = 100 nC = 10“7 C, Ax =
= 0.1 mm = 10"4 m, e0 = 8.85 X 10“12 F/m, 8 = 1.
Required: AA.
Solution: The work AA = F Ax, where F = qE, q is the
charge of one plate, and E is the intensity of its electric
field:
o
E 2e^ S 2Qe 0

Consequently,
\4 — Q* a _ K ) - i 4 X l()-4
~ 2e 0S A X 2 X 8.85 X IQ*"12 X 2 X 10"2
« 3 x l O " 8J

8. P o t e n tia l o f a n E le c tr o s ta tic F ie ld
1°. The energy of an electrostatic field is characterized
by its potential. The potential of a field at a given point
is a scalar quantity numerically equal to the potential
energy W of a unit positive charge placed at this point:
W
cp = —
n
An electrostatic field, every point of which is characterized
by a certain potential, is an example of a potential field.
230 Fundamentals of Electrodynamics III.1 .8

2°. The work needed to move the charge q from the


point 1 to the point 2 (111.1.7.5°) is
A = Wy— W2 = q (cpi — <p2)
The potential difference at the initial (7) and final (2) points
of the path numerically equals the work done by the forces
of the electrostatic field upon the displacement of a unit
positive charge between these points:
A
<Pl ^2 —~
If the point 2 is at infinity, then W 2 = 0, and accordingly
cp2 = 0. The work A ' needed to move the charge q from the
point 1 to infinity is A' = Wx = qcp±, whence (px = A'lq.
The potential of an electrostatic field numerically equals
the work done by electrostatic forces when they move
a unit positive charge from the given point of the field
to infinity. Thus, the potential numerically equals the work
that is done when a unit positive charge repelled from the
positive charge q moves away to infinity.
3°. The potential also numerically equals the work done
by external forces against the forces of the electrostatic
field when a unit positive charge is moved from infinity
to the given point. For example, the potential numerically
equals the work that will be done if, overcoming its repul­
sion from the positive charge q, another unit positive charge
is moved from infinity to the given point of the field.
4°. In all problems, the potential difference between two
points of an electrostatic field has a physical meaning,
and not the values of the potentials at these points. There­
fore, the choice of the point of zero potential is determined
by considerations of simplicity and convenience in solving
problems. It is sometimes more convenient to assume that
the Earth’s potential equals zero instead of that of an
infinitely remote point.
5°. If an electrically charged particle of charge q and
mass m moves in an electric field of potential difference
cpi — (p2 accelerating it, then the particle acquires the
kinetic energy
m v2 / v
-y- = ? (<Pi —<P*)
III.1.8 Electrostatics 231

and the velocity


2q (eft—cp2)
m
6°. The potential difference between the two points 1
and 2 of an electrostatic field at the distances x1 and x2
from a uniformly charged infinite plane is

9i — (P2 = 2 ^ (^ 2 — x i) (in the SI system)

(p4— q)2 = ? ^ (^ 2— Xi) (in the cgse system)

where o = surface charge density


e0 = electric constant in the SI system (VI1.5.1°)
e = relative permittivity,
7°. The potential difference' between uniformly and’
oppositely charged infinite parallel planes is

9i — 92 = -—
e08 0 n the SI system)

q)j — qp2= (in the cgse system}

where d is the distance between the planes. For the remaining


notation see Par. 6°. __
8°. The potential of the electrostatic field of the point
charge q at a point at the distance r from the charge (provided
that cp 0 when r -*■ oo) is

= —
4Jt808 r
(in the SI system)

<p= — (in the cgse system)

9°. The potential of the electrostatic field of a sphere:


of radius R and charge q uniformly distributed over its
surface coincides outside the sphere with the potential of
the field of the point charge q placed at the centre of the
232 Fundamentals of Electrodynamics 111.13

sphere (provided that (p -^ 0 when r oo). Inside the


sphere, there is a constant field potential equal to

<P= 4We„ei? (in the SI system)


cp= - (in the cgse system)

although the field intensity inside the sphere equals zero


(111.1.4.3°).

9. Relationship Between Intensity


and Potential Difference
of an Electrostatic Field
1°. The two quantities characterizing the force of an
electrostatic field (E) and its energy (cp) are related to each
other.
Near any point of an electrostatic field, the potential
changes most rapidly in the direction of a field line
(111.1.3.5°). The intensity at an arbit­
rary point of an electrostatic field
numerically equals the change in the
potential per unit length of a field
line:

Fig. III.1.22 where A1 = vector with the magni­


tude AZ
Ei = E cos a = projection of the vector E onto the direction
A1 (Fig. III.1.22).
The minus sign shows that the field intensity vector
is always directed toward diminishing potential.
2°. The locus of points of an electrostatic field having
the same potential is called an equipotential surface. The
latter has the following properties:
(a) at every point of an equipotential surface, the field
intensity vector is perpendicular to the surface and is
directed toward diminishing potential; and
111.1.9 Electrostatics 233

(b) the work of moving an electric charge along the


same equipotential surface equajs zero.
An example of an equipotential surface is the surface
of a charged conductor. At all points inside such a conduc­
tor, the intensity of the electrostatic field equals zero
(111.1.5.3°), and all points inside the conductor have the
same potential.
3°. An electrostatic field is depicted graphically with
the aid of equipotential surfaces in addition to field lines.
An infinite multitude of equipotential surfaces can be drawn
around any sources of an electrostatic field. They are usually
portrayed so that the potential difference between any two
adjacent equipotential surfaces is the same.
A known arrangement of the field lines (111.1.3.5°) of an
electrostatic field can be used to construct equipotential
surfaces, and, conversely, a known arrangement of the
equipotential surfaces at each point of a field can be used
to determine the magnitude and direction of the field inten­
sity vector.
Figure III. 1.23 shows plane sections of the electrostatic
fields of a positive point charge (a), a dipole (6), two like
charges (c), and a charged metal conductor of an intricate
configuration (d). The dash lines are field lines, and the
solid lines are sections of equipotential surfaces.
Problem 1. A uniformly charged sphere with a radius
of 2 cm in a vacuum has a surface charge density of 5 X
X 10~7 C/m2. Find the potential of the field at a point
at a distance of 0.5 m from the centre of the sphere, and
also the potential and intensity of the field inside the sphere.
Given: R = 2 cm = 2 X 10“2 m, a = 5 X 10”7 C/m2,
r = 0.5 m, e0 = 8.85 X 10”12 F/m, e = 1.
Required: q^, cp2, E.
Solution: The potential of the electric field due to the
charged sphere coincides outside the sphere with the poten­
tial of the field of a point charge concentrated at the centre
of the sphere, (p2 = <//4jte0pr.
Since q = aS = gAtcR2, then

oR2 5 X 10-7 X 4 X 10-4


<Pi 8.85 X 10“12 X 1 X 0,5
50 V
234 Fundamentals of Electrodynamics 111.1.9

The potential of the field inside the charged sphere is


m __ 9 5 X 10~7 X 2 X 10~2 _a v
4jte0e B e 0e 8 .8 5 x l ( r 12X l
The relationship between the field intensity and the
potential of the sphere has the form E = —Acp/Ar. The

potential of the field inside the sphere at all the points


is the same. Its change per unit length equals zero, hence
E = 0.
Problem 2. What is the potential difference between the
points of an electrostatic field in a vacuum at distances
of 0.4 and 1 m from a point charge of 2 X 10"9 C? What
work is done in moving a positive charge of 4 X 10“10 C
from the first point to the second one?
Given: q = 2 X 10”9 C, r, = 0.4 m, r2 = 1 m, qx =
= 4 X 10-10 C, 8 = 1.
Required: <pj — cp?, A ,
HI.1.10 Electrostatics 235

Solution: The potential difference between two points


of an electrostatic field produced by a point charge is

q_______ i__ __ q
<Pi — <P2 4 ne0er2 4jte0e
2 X 10-® I 1
4 X 3.14 X 8.85 X 10“12 \ 0.4 '

The work A = qx ((px — q>2) = 4 X 10”10X 300 « 10”7 J.


Problem 3. The potential difference between the points
at a distance of 5 and 10 cm from a charged plane is 5 V.
What is the charge of the plane in a vacuum if its area is
400 cm2?
Given: xA = 5 cm = 5 X 10”2 m, x2 = 10 cm = 10"1 m,
S = 400 cm2 = 4 X 10~2 m2, <px — <p2 = 5 V, e = 1.
Required: q.
Solution: The potential difference between points of the
field of a charged plane is

<Pi — <P2 — 2 e 0e (x z — ^ i )
where o = q/S.
The charge of the plane is q = oS , or
(q)i—(p2) 2e0s g
Q x2 Xi
5 X 2 X 8.85 X 10-1*
10"1—5X10"*
X 4 x 10-2 = 7 X 10-“ C

10. Capacitance
1°. When the charge q on a conductor grows, the poten­
tial of the conductor cp grows in direct proportion to the
charge. This holds for conductors of any geometrical shape.
The ratio of the charge of a conductor to its potential does
not depend on the magnitude of the charge on the conductor
and is determined by the properties of the conductor itself
and of the surrounding medium. The electric properties
of a conductor determining the possibility of accumulating
charges on it are characterized by its capacitance,
236 F u n d a m en ta ls of E le ctro d y n a m ic s III.1.10

The physical quantity measured by the ratio of the


charge q of an isolated* conductor to its potential (p is called
the capacitance of the isolated conductor:

In other words, the capacitance of an isolated conductor


is the physical quantity numerically equal to the charge
that changes the potential by unity.
The capacitance of a conductor depends on its linear
dimensions and geometrical shape, but does not depend
on the material of the conductor and its state of aggregation.
Geometrically similar conductors have capacitances that
are directly proportional to their linear dimensions. The
capacitance of a conductor is directly proportional to the
relative permittivity of the medium (111.1.2.6°) surrounding
it.
2°. The capacitance of an isolated sphere is
C = 4jteue/? (in the SI system)
C = eR (in the cgse system)
where R = radius of the sphere
e0 = electric constant in the SI system (VI1.5.1°)
e = relative permittivity of the medium surrounding
the sphere.
3°. The mutual capacitance (usually called simply the
capacitance) of two conductors is defined as the physical
quantity numerically equal to the charge q that must be
transferred from one conductor to the other to change the
potential difference (q)j — (p2) between them by unity:

<Pi—<P2
The mutual capacitance depends on the geometrical
shape, linear dimensions and mutual arrangement of the
conductors, and does not depend on the material of the
conductors and their states of aggregation. The mutual

* The term isolated is applied to a conductor that is far from


charged bodies and other conductors.
111.1.11 E le c tro sta tic s 237

capacitance is directly proportional to the relative permit­


tivity of the medium containing the conductors.
Problem. How many electrons are on the surface of an
isolated metal sphere 4 cm in diameter charged in a vacuum
up to a potential of 100 V?
Given: d = 4 cm = 4 X 10~2 m, cpx — cp2 = 100 V,
e = 1.
Required: N .
Solution: The charge of the metal sphere qsvh = C (<pj —
— cp2), where C is the capacitance of an isolated sphere,
C = 4jte0ei?, and R is the radius of the sphere. Hence
7sPh = 4jie0e (d/2) (cpx — cp2).
The number of electrons N = qsvJe, where e is the
absolute value of the charge of an electron. Consequently,
v_ 4ree0ed(q>i—<pa) _
~ 2e ~~
_ 4 x 3.14 X 8.85 X 1Q~12 X 4 x 10~2 X 100 _ ^ x |q 6

11. Capacitors
1°. A capacitor consists of two conductors charged with
unlike charges equal in magnitude. The conductors must
have such a geometrical shape and must be so arranged
relative to each other that the electric field produced by the
conductors will be concentrated in the space between them.
The conductors forming a capacitor are called its plates.
The capacitance of a capacitor is the mutual capacitance
of its plates (111.1.10.3°). Capacitors are accumulators of
electric energy (111.1.12.2°).
2°. A parallel plate capacitor consists of two parallel
flat plates carrying identical charges of opposite signs.
The plates of a capacitor are separated by the distance d
(Fig. III.1.24). When charging a capacitor, one of the
plates can be charged and the other one earthed. A charge
will remain on the earthed plate that is opposite in sign
and equal in value to the charge on the first plate. A charge
with the sign of the first plate will pass into the Earth.
238 Fundamentals of Electrodynamics I1U.11

3°. The capacitance of a parallel plate capacitor is

C = -E
-^ S (in the SI system)
8iS
c = -££cr c^se system)
where £ = area of each plate or of the smaller of them
d = separation distance of the plates
e0 = electric constant in the SI system (VII.5.1°)
e = relative permittivity of the substance between
the plates.
If a parallel plate capacitor consists of a system of n
plates (a multiplate capacitor)v then the formula for the

capacitance contains the product S (n — 1) instead of S.


A capacitor usually has two plates (n = 2).
4°. A variable capacitor in the simplest case consists
of two sets of metal plates. When the knob is turned
(Fig. III.1.25), the plates of one set fit into the spaces
between the plates of the other set. The capacitance of the
capacitor changes in proportion to the change in the area
of the overlapping part of the plates.
5°. The capacitance is increased by the parallel connection
of capacitors. The plates of the latter having like charges
are connected (Fig. III.1.26). The total capacitance of n
capacitors connected in parallel is
c = cx + c2 + . .. + cn
itt.i.u Electrostatics 239

6°. In the series connection of capacitors, their plates


having opposite charges are connected (Fig. III.1.27).
Here the reciprocals of the capacitances are summated:

and the total capacitance of the system connected in this


way is always less than the smallest capacitance of a ca­
pacitor in the system.
Problem l.Find the potential difference in an air capac­
itor if a porcelain plate is placed between its plates that

Fig. III.1.25 Fig. III.1.26

tightly fits against them. The capacitor was initially charged


to 200 Vt and then the source was switched off.

Fig. III.1.27

Given: cpx — cp2 = Aq^ = 200 V, e2 — 1, e2 = 5.


Required: A(p2.
Solution: The capacitance of an air capacitor is C1 =
= g/Aq?!. The capacitance of a capacitor with a dielectric
is C2 = qlAq)2.
Since the charge of the capacitor does not change, then
Aq)2/Aq)j = CyC2; the capacitance of a parallel plate
240 Fundamentals of Electrodynamics 111.1.12

capacitor is C = e0zS/d. Hence, and A<p2 =


= A tp^/e^ whence A(p2 = 200 X 1/5 = 40 V.
Problem 2. Find the capacitance and surface charge
density on the plates of an air capacitor charged to a poten­
tial difference of 200 V. Th area of each plate is 0.25 m2,
the separation distance is 1.0 mm.
Given: <Pi — <p2 = 200 V, S = 0.25 m2, d = 1.0 mm =
= 10~3 m, e = 1.
Required: C, a.
Solution: The capacitanc. of a parallel plate capacitor is
r _ e0eS _ 8.85 X 10"12 X 1 X 0.25
C~ d ~ 10-3
2.2 x 10~9 F

The potential difference between the two charged plates


is (p2 — ip2 = g d/e0e, whence
(q)i — cp2) e0e _ 200 X 8.85 X 10“12
d W*
1.8 x 10"6 C/m2

12. Energy of an Electric Field


1°. To increase the charge of a conductor, a certain
amount of electricity has to be transferred to it. To do this,
it is necessary to overcome the forces of repulsion between
the newly transferred charge and those already on the
conductor. Work has to be done to increase the charge
on a conductor and its potential. The work that has to be
done to impart the charge q and the potential (p to a con­
ductor can be a measure of the energy of the charged conduc­
tor.
2°. The potential energy of interaction of the charges
on a conductor is called the proper, or internal, energy
of the charged conductor. If the conductor is not in an exter­
nal electrostatic field, then its energy is only proper and
is calculated by the formula
\J7 QW ______ Q2 ffiP2
Prop 2 2C 2
111.1.12 Electrostatics 241

where C = capacitance of the conductor


q and (p = its charge and potential, respectively.
3°. If we have a system of m charged conductors, then
the total electric energy of the system consists of the sum
of the proper energies of the conductors and the energy
of their interaction:

w i
i —1

where qt = charge of the i-th conductor


q)f = potential of the i-th conductor set up by both
the field of all the other conductors and the
proper field of this conductor.
4°. The energy of a charged capacitor is the total energy
of a system of two conductors and is calculated by the
formula
TXT Q (<Pl ^ 2 ) __ C (ffl ----(P2)2
2 2

where q = charge of the capacitor


C = its capacitance
cp, — cp2 ^ potential difference between the capacitor plates.
5°. According to the theory of short-range force action
(111.1.3.2°), the energy of any charged bodies is concentrat­
ed in the electric field of these bodies. We therefore speak
of the energy of an electric field, meaning that the energy
of the sources of the field—charged bodies—is distributed
over the entire space where there is an elec ric field. For
example in a parallel plate capacitor (111.1.11.2°), the
energy is concentrated in the space between its plates.
6°. The energy of a homogeneous electric field concentrat­
ed in the volume V of an isotropic medium is

W = e-°-^ - V (in the SI system)

pP'2
= (in the cgse system)
6 -0 2 1 1
242_______________ F u n d a m en ta ls o f E l e c t r o d y n a m i c s ___________ 1 1 1 . 1 . 1 2

where E = field intensity


e0 = electric constant in the SI system (VII.5.1°)
8 = relative permittivity of the medium.
7°. The volume density we of the energy of an electric
field i6 defined as the energy of the field concentrated in
a unit volume of it:
AW
we =
AV

The quantity we is calculated by the formula

we = - °e~ (in the SI system)

We = (in the cgse system)

The expressions of Par. 7° hold not only for a homoge­


neous field, but also for arbitrary ones, including electric
fields changing with time, in a homogeneous isotropic dielec­
tric (111.1.6.1°).
8°. The definition of an electric field as a special form
of matter (111.1.3.1°) is substantiated in that an electric
field has energy. Energy, like mass, is an essential pro­
perty of a substance and fields—the two forms of matter
studied by physics.
Problem. Capacitors with capacitances of 2.0 and 8.0 pF
are connected in series to a source of voltage at 200 V.
Find the potential difference across each capacitor and
the energy of each capacitor.
Given: Cx = 2.0 pF, C2 = 8.0 pF, cpx — cp2 = V =
= 200 V.
Required: C/2, Wly W2.
Solution: The charges on series-connected capacitors will
be the same. The potential difference across the capacitors
is Ux = q!C1 and U2 = q/C2, where q is the charge; q = CU,
where C is the capacitance of the series-connected capaci­
tors.
With series connection, we have i/C = ilC l + 1/C2,
111.2.1 Steady Electric Current 243

or C = C1C2/(C1 + C2). Hence, q — UCiC2l(Cx + C2), and


v c xct uc. 200 X 8.0
U i = (Ct + CJCj. 160 V
Cx + C2 2 .0 + 8 .0
f/2= UCi - 200 X 2.0 _ y
Ci + c 2 2.0+8.0
The energies of the capacitors are
CXU\ _ 2.0 x 160* /
WV 2 2
« 26 pJ

w 8.0 X 40*
2 = 6 .4 pJ

Chapter 2
Steady Electric Current

1. Basic Concepts and Definitions


1°. With respect to their property of electric conduction,
i.e. their ability to conduct an electric current, all sub­
stances are divided into conductors (metals, electrolytes,
and ionized gases), dielectrics (insulators) (111.1.6.1°), and
semiconductors (111.3.11.1°).
An electric current is defined as the ordered motion of
electric charges. The ordered motion of free electric charges
occurring in a conductor is called a conduction current.
Such currents include the electric current in metals produced
by the ordered motion of free electrons, the current in
electrolytes due to the ordered motion of the ions (111.3.1.1°),
and the current in gases (111.3.3.1°), where ions and elec­
trons perform ordered motion. The ordered motion of electric
charges occurring when a charged body moves in space is
called a convection current. For example, the Earth has
a surplus negative charge, and a convection current is pro­
duced during its motion. Apart from conduction and convec-
Iion currents, there are also displacement currents (IV.4.1.3®).
244 Fundamentals of Electrodynamics

2°. The direction of an electric current is considered


to be the one in which the positive charges move orderly.
In metals, the free electrons move in a direction opposite
to this conventional one. In liquids and gases, the ordered
motion of the electric charges may occur either in the direc­
tion of the current or in the opposite one (the ordered motion
of electrons and negative ions).
3°. The current intensity, or simply current, is defined
as the scalar quantity I equal to the quantity of electricity
Aq carried through a cross-sectional area of a conductor
in a unit time:

Particularly, the current in a conductor is defined as the


quantity of electricity Aq that flows through a cross section
of a conductor in a unit time.
4°. A steady current is one that persists for long periods
of time in one direction. For a steady current

where q is the charge that flows through a cross section


of the conductor during the time t.
5°. The steady current in a metal conductor of cross-
sectional area S is
I = envS
where e = magnitude of the charge of an electron
n = number of charge carriers (conduction electrons)
in a unit volume
v = average velocity of the ordered motion of the
electrons.
6°. The average current density vector j* is defined as the
physical quantity whose magnitude equals the ratio of the
current / to the cross-sectional area of a conductor S perpen-

* In the following, the term current density will signify the av­
erage current density.
111.2.1 S te a d y E lectric C urrent 245

dicular to the vector v (Par. 7°):

The current density determines the current per unit


cross-sectional area of a conductor and characterizes the
distribution of the current over the section of the conductor.
The vector j is directed along the current (Par. 2°).
7°. The density of the conduction current in metals is
j = nex
where n = number of conduction electrons in a unit volume
(the concentration of the current carriers)
e = magnitude of the charge of an electron
v = vector of the average velocity of the ordered
motion of the electrons.
An important feature of metals is the virtually constant
concentration n of the free electrons for a given metal. It
is independent of the temperature. This feature noticeably
distinguishes metals from electrolytes (111.3.1.4°) and
semiconductors (111.3.12.5°).
8°. The magnitude of the vector v has values of the order
of 10“4 m/s at the maximum permissible current densities.
For example in a copper conductor (n « 8.5 X 1028 m~3),
we have v « 8 X 10“4 m/s at the highest current densities
of 7 = 1.1 X 107 A/cm2 that are tolerated1 without dan­
gerous overheating of the conductor.
9°. An electric current appears in the entire conductor
virtually at the same time with closing of the circuit.
The time needed for a current to appear in a circuit is t = Lie,
where L is the length of the circuit, and c is the speed of
light in a vacuum. It coincides with the time needed for
a stationary electric field (111.1.3.1°) to appear along the
entire circuit .This time coincides with that of light propa­
gation along the electric circuit. During this time, ordered
motion of the electrons sets in over the entire length of
a conductor and it begins virtually as soon as the circuit
is closed.
The existence of an electric current is detected according
lo its thermal, chemical, and magnetic actions.
246 Fundamentals of Electrodynamics 1 1 1 .2 .2

2. Conditions Needed for


a Steady Current to Appear
and Be Maintained

1°. For a conduction current (111.2.1.1°) to appear in


conductors and be maintained in them, forces must be exerted
on charged particles that ensure their ordered motion during
a finite period of time.
The Coulomb forces of electrostatic interaction of elec­
tric charges result in their being distributed in a conductor
so that the intensity of the electric field inside it equals
zero, and the potentials of all the points of the conductor
are the same. Therefore, the electrostatic field of Coulomb
forces (a Coulomb electric field) cannot ensure a steady elec­
tric current in a conductor (111.2.1.4°).
2°. The following conditions must be observed for a steady
conduction current to exist in a conductor:
(a) the intensity of the electric field in the conductor
must differ from zero and not change with time;
(b) the circuit of the steady conduction current must be
closed; and
(c) apart from Coulomb forces, forces of a non-electro-
static origin called extraneous forces must be exerted on the
free electric charges. The extraneous forces may be produced
by current sources (galvanic cells, accumulators, electric
generators, etc.).
3°. The extraneous forces cause the electric charges to
move inside the current source in a direction opposite
to the action of the forces of the electrostatic field. As
a result, a potential difference is maintained across the
ends of the external circuit, and a steady electric current
flows through the circuit. The work needed to ensure the
ordered motion of electric charges in a conductor when
a steady electric current flows through it is done at the
expense of the energy of the current source.
4°. An extraneous electric field is an electric field of a non-
electrostatic origin in which free electric charges flow in
a conductor in conditions when extraneous forces act.
The intensity of an extraneous electric field Eex is defined
as the physical quantity numerically equal to the extraneous
IIL 2 .3 Steady Electric Current 247

force acting in an extraneous electric field on a unit positive


charge at a given point inside the conductor:

where Fex is the extraneous force acting on the positive


charge q. The directions of Eex and Fex coincide.

3. Electromotive Force. Voltage


1°. The intensity E of the electric field inside a conductor
through which a steady current is flowing, according to the
principle of superposition of fields (111.1.3.7°), is
E = E Coui + E ex
where ECoul and Eex are the intensities of the Coulomb
field and the field of the extraneous forces, respectively.
2°. The work involved in moving a charge along a con­
ductor when a current flows through it is done by the Cou­
lomb and extraneous forces. The total work A is

A = ^Coul + Aex

where A C0u\ is the work done by the Coulomb forces


(111.1.7.1°), and A ex is that done at the expense of the
extraneous forces.
The total work done when a unit positive charge is
moved over the section 1-2 of an electric circuit through
which a steady current is flowing is

^ 1 -2 _ ^ C o u l, 1-2 . ^ c x , 1-2
q ~ q q

where q is the positive charge carried over the section 1-2


(Fig. III.2.1). According to (111.1.8.2°), A Couit l .2/q =
= (Pi — <p2 is the potential difference between the points
7 and 2.
3°. The electromotive force (e.m.f.) acting on the
section 1-2 of a circuit is defined as the physical quantity
248 F u n d a m e n ta ls of E lectro d yn a m ics H I.2.4

numerically equal to the work (lone by extraneous forces


when they move a unit positive charge over the section 1-2:

_^ex, 1-2

4°. The voltage (voltage drop) t/2-i on the section 1-2


of a circuit is defined as the physical quantity numerically

Fig. III.2.1

equal to the total work done by the Coulomb and extra­


neous forces when a unit positive charge is moved along the
section of the circuit from the point 1 to the point 2 (see
Fig. III.2.1):

f / 2- i = — = ( < P i - c p 2) + £ 2-i

The voltage across the section 1-2 of the circuit equals


the potential difference between the points 1 and 2 only
when no e.m .f.’s are applied on this section: £/2-i = q>i — cp2
when &2-i = 0.

4. Ohm’s Law
1°. The electrical resistance (resistance) of the circuit
section 1-2 is one of the characteristics of the electrical
properties of the given circuit section determining the
ordered motion of the current carriers on this section.
The resistance of a metal conductor on a section of an
unbranched circuit (111.2.6.1°) depends on the material
of the conductor, its geometrical shape and dimensions,
and also on the temperature (111.2.5.1°). For a homogeneous
111.2.4 S tea d y E le ctric C urrent 249

cylindrical conductor of length Z12 and cross-sectional area S y


the resistance R ^ is (Fig. 111.2.2)

#2-1 = P~^f"
where p is the resistivity of the conductor. The latter is defined
as the resistance of a homogeneous cylindrical conductor

I 2
Fig. III.2.2

made from the given material and having a unit length and
unit cross-sectional area.
Conductors having a great resistance # 2-i are usually
called resistors.
The reciprocal of the resistivity is called the conduc­
tivity of a conductor: x = 1/p.
2°. Ohm's law for an arbitrary section of a circuit: the
voltage (voltage drop) across a section of a circuit equals
the product of the resistance of this section and the current:
C/2-1 = R2-1I
A different formulation is: the voltage drop across a section
of a circuit equals the sum of the potential difference across
the ends of the section and the e.m.f. applied to it:
# 2- 1/ = ( 9 1 — 9 2 ) 4 - ^ 2-1
3°. Ohm's law for a section of a circuit containing no e.m.f.
(Fig. III.2.3). Here = 0, U2.x = ^ - cp2 (111.2.3.4°),
and
T ^1 — ^2

*2-1
The current is directly proportional to the potential dif­
ference across the ends of the circuit section and inversely
proportional to the resistance of this section.
250 Fundamentals of Electrodynamics 1112A

4°. Ohm's law for an electric circuit consisting of a current


source of e.m.f. % and internal resistance r, and the exter­
nal resistance Z?ext (Fig. III.2.4).
/
7 2
o- -o
R
Fig. III.2.3

For the entire circuit, we have cpr = cp2, the total resis­
tance of the circuit is R = R = R exi -f- r, and Jf2_j =
= <£. Hence,

^ext + r

The current in the circuit is directly proportional to the


e.m.f. acting in it and inversely proportional to the sum
of the external and internal re-
sistances.
H The voltage (voltage drop) U ac­
ross the external circuit is
U = I R ext = %— Ir
^ext ^ext + r
where Ir is the voltage drop in­
Fig. III.2.4 side the current source.
5°. If a circuit includes sever­
al current sources whose e.m .f.’s
are, for instance, %l9 S 2, and (Fig. III.2.5), then
the e.m.f. % acting in the circuit equals the algebraic sum
of the e.m.f.’s of the separate sources of current:
N
» = i=i
53 *1
The value of the e.m.f. is considered positive if an
arbitrarily chosen direction around the circuit (counter­
clockwise in Fig. III.2.5 as shown by the arrow) coincides
with the transition from the negative pole of the source
11 1.2.4 Steady Electric Current 251

to its positive one. The opposite direction is negative.


For the circuit in Fig. III.2.5, we have

6°. For an open circuit, 1 = 0, and according to Par. 2°,


we have = cp2 — cpt. The e.m.f. of a current source
is measured by the potential
difference across its terminals
with the external circuit open.
7°. Ohm's law for the density
of a steady current in metals
(Ohm's law in the electron theo­
ry)-
j = xE = — E
P
The density of a steady cur­ R
rent (111.2.1.6°) equals the pro­ Fig. III.2.5
duct of the conductivity (Par. 1°)
and the intensity of the electric
field at the given point inside a metal conductor carrying
a current. The average velocity v of the ordered motion of
the electrons in a metal (111.2.1.7°) is proportional to the
intensity of the electric field at the given point of the
conductor:
eT -|7i
V = 2^E
where e = magnitude of the charge of an electron
m = mass of an electron
x = mean free time of an electron (11.2.3.3°).
In the classical electron theory of conduction of metals,
it is assumed that the free electrons in a metal conductor
(11.7.1.2°) move with acceleration under the action of an
electric field from one point of the crystal lattice of the
metal (11.1.6.5°) to another without any collisions. The
current density (111.2.1.7°) is expressed by the formula

j = en\ E = kE
252 F u n d a m e n ta ls of E lectro d yn a m ics II 1.2.5

The conductivity of a metal in the classical electron theory is


1 neH
X = ----
p= - 7i----
Zm
Problem. A battery with an e.m.f. of 16 V is connected
to an instrument. The current in the latter is 2 A. The
efficiency of the battery is 0.75. Find the internal resistance
of the battery.
Given: g = 16 V, / = 2 A, t) = 0.75.
Required: r.
Solution: Ohm’s law for a closed circuit is / — <£I{R + r),
or <£ = U + /r, where U = IR is the voltage drop across
the external resistance, and Ir is the same across the inter­
nal resistance.
The efficiency of a battery is r\ = Ul^, hence, U =
= and % = T]& -f- Ir, whence
r_ % (1 — T|) ^ 1 6 ( 1 - 0 .7 5 ) 2 Q

5. Temperature Dependence of Resistance


1°. The resistivity p of conductors depends on the tem­
perature
P — Po (1 +
where p0 = resistivity at 0 °C
t = temperature, Celsius scale
a = (p — Po)/Po£ == temperature coefficient of resis­
tivity—the relative change in the resistivity
of a conductor when it is heated by one degree.
2°. The value of the temperature coefficient of resistivity
a for metals and alloys within the temperature range from
0 to 100 °C varies from 3.3 X 10“3 to 6.2 X 10"3 K-1. It is
usually assumed for pure metals that a = (1/273) K -1.
For electrolytes, a «< 0.
3°. The temperature dependence of the resistivity of
pure metals cannot be explained satisfactorily within the
scope of the classical electron theory of conductivity.
In the modern quantum theory of the electrical conductivity
ili.2 .6 Steady Electric Current 253

of metals, it is proved that at all temperatures except


absolute zero free electrons experience such interactions
with the points of the crystal tettice of a metal that the
mean free time t of the electrons in the region of moderate
temperatures is inversely proportional to the absolute
temperature T of the metal ( t a 1IT), while p is directly
proportional to the absolute temperature (p oc T).
4°. The phenomenon of superconductivity encountered in
some metals and alloys consists in that below a certain
temperature (the critical temperature Tcr of transition of
a conductor to the superconducting state) the resistivity of
these substances becomes vanishingly small. The tempera­
ture Tcr for pure metals varies from 0.14 K (iridium) to
9.22 K (niobium), and for alloys from 0.155 K (Bi2Pt) to
23.2 K (Nb3Ge).
Superconductivity is used to obtain very strong magnet­
ic fields. If the winding of an electromagnet (111.4.3.7°)
is made of a superconducting wire, then a tremendous
current density appears in this winding and, correspondingly,
the electromagnet has a strong magnetic field.
The memory elements of computers sometimes function
on the basis of superconductivity. The operating principle
of such elements made from superconducting films consists
in that a current once started in a superconducting loop
will persist for a very long time. The design of the switching
units of modern electronic computers is sometimes based
on the principle of destroying the superconducting state
by means of a magnetic field.
The theory of superconductivity has been developed
within the framework of modern quantum mechanics.
Acquaintance with it is beyond the scope of a course in
elementary physics.

6.( Branching of Currents.


Connections of Conductors
1°. An electric circuit is a combination of conductors
and current sources. In the general case, an electric circuit
is branched (multiloop) and has junctions. The junction A
254 F undam entals of Electrodynam ics 111.2.6

in a branched circuit is a point at which at least three


conductors converge (Fig. 111.2.6). The calculation of
a branched circuit or network consists in finding the cur­
rents in each branch according to the given resistances
of the branches and the
e.m .f.’s applied to them. Two
rules are used for this purpose.
2°. Kirchhoff's first rule
(the rule of junctions): the al­
gebraic sum of the currents
converging at a junction
equals zero:
n
2h
h= 1
= 0

where n is the number of


conductors converging at the
junction. The currents are considered positive if they flow
into the junction, and negative if they flow out of it.
3°. Kirchhoff's second rule (the rule of loops): in any
loop arbitrarily selected in a multiloop circuit, the alge­
braic sum of the products of the currents I h and the resis­
tances R k of the relevant portions of this loop equals the
algebraic sum of the e.m .f.’s in the loop:
71 m
3 h R k = 2 *!
i=i

If the currents I k coincide with the selected direction


of traversing the loop, they are considered to be positive.
The e.m .f.’s are considered to be positive if they set
up currents flowing in the direction of traversing the loop.
4°. A multiloop network of steady current is calculated
in the following sequence:
(a) the directions of the currents in all the loops of the
network are chosen arbitrarily;
(b) the rule of junctions is used to compile n — 1 inde­
pendent equations, where n is the number of junctions
in the network; and
(c) arbitrary loops are chosen so that each new loop
contains at least one branch of a circuit that has not been
111.2 .6 __________________________________________Steady Electric Current2 55

_
included in the previously considered loops. For a multi­
loop network containing n junctions and m branches of the
circuit between adjacent junctions, the number of inde­

pendent equations compiled according to the rule of loops


is m — n + 1.
5°. The conductors forming a circuit can be connected
in series or parallel.
In the series connection of conductors (Fig. III.2.7):
(a) the current in all the portions of the circuit is the
same:
I = const
(b) the voltage drop in a circuit equals the sum of the
voltage drops in its separate portions:

U = U1 + U2

(c) the voltage drop on conductors is directly propor­


tional to their resistances:

U2 R2

(d) the total resistance of a circuit consisting of n conduc­


tors connected in series equals the sum of the resistances
of the separate conductors:

i? — i?i -f- R 2 + . . . + Rn
256 Fundam entals of E lectrodynam ics 1II.2.6

6°. In the parallel connection of conductors (Fig. III.2.8):


(a) the current in the unbranched portion of the circuit
equals the sum of the currents flowing in the branches of
the circuit:
I = A + 12
(b) the voltage drops in the branches of the circuit
connected in parallel are the same:
U = const
(c) the currents in the parallel branches of the circuit
are inversely proportional to their resistances:
1 l = 1 *.
J2 Rl
7°. The quantity K, which is the reciprocal of the resis­
tance of a branch of a circuit, is called the conductance,

K = MR. The conductance of a circuit consisting of n


conductors connected in parallel equals the sum of the
conductances of all the conductors:
K = Kx + K t + . . . + Kn
or
1 1 1
R
= + ••• + —
Hn
Problem 1. A measuring instrument—a galvanometer G
with a resistance of 50 Q—can measure an electric current
111.2.6 Steady Electric Current 257

up to 0.1 A. How should it be connected to a circuit for


it to become an ammeter measuring a current up to 10 A
or a voltmeter for measuring voltages up to 100 V?
Given: r = 50 Q, i = 0.1 A, / = 10 A, U = 100 V.
Required: R x, R 2.
Solution: To measure currents greater than those which
an instrument is designed for, a resistor R 1 called a shunt

A
Fig. III.2.9

has to be connected to it (Fig. III.2.9). The ammeter is


connected to the circuit in series. The current in the circuit
is / = i + I ly where I x is the current flowing through the
shunt. The resistance of the shunt is R x = UIIX, where U
is the voltage drop equal here to the identical potential
difference across the terminals of the ammeter and the ends
of the shunt: U — ir.
Hence,

To measure a voltage with a galvanometer, an additional


resistor R 2 has to be connected in series with it (Fig. III.2.10).

Fig. III.2.10

A voltmeter is connected to a circuit parallel to the resistor


across which the voltage drop is measured. The additional
resistance is R 2 = U JI, where U1 is the voltage drop.
The potential difference across the terminals of the galva­
nometer is Ux = ir, Ux = U — U2.
258 Fundamentals of Electrodynamics ni.2.6

The potential difference across the additional resistance


is U2 = iR2. Hence,
U — ir 100— 0 . 1 x 5 0
*2 i 0.1 950 Q

Problem 2. An ammeter and a coil are connected in


series to a current source. A voltmeter is connected parallel
to the coil (Fig. III.2.11). The readings of the voltmeter

Fig. H I .2.12

and ammeter are 200 V and 0.5 A, respectively. The resis­


tance of the ammeter may be disregarded.
Find the resistance of the coil for two cases:
(1) The resistance of the voltmeter is infinitely great.
(2) The resistance of the voltmeter is 2000 Q.
Given: U = 200 V, I = 0.5 A, r = oo, r = 2000 Q.
Required: i?lT R 2.
Solution: (1) When r = oo, the current flowing through
the voltmeter equals zero. The resistance of the coil is
jj U 200 /n n o
jR‘ ==T - = -oX = 400 q
(2) When r = 2000 Q, the current in the circuit equals
the sum of the currents flowing through the voltmeter and
the coil: I = i + i±, where i is the current flowing through
the voltmeter, i = Ulr; the current flowing through the
coil is i± = I — i = I — Ulr. Hence,
u U 200
#2 i U 200
500 £2
I— 0.5
r 200o
tll.2 .6 Steady Electric Current 259

The difference between the results of the measurements


will be AR = 500 - 400 - 100 Q.
Problem 3. A potentiometer with a resistance of 3.0 kQ
is connected to a current source whose e.m.f. is 110 V.
Find the voltage drop across the instrument that is con­
nected to one of the terminals of the potentiometer D and
the sliding contact C at the middle of the potentiometer
(Fig. III.2.12). The resistance of the instrument is 10 kQ.
A potentiometer is a device intended for changing the
potential difference across the ends of a section of a circuit.
For this purpose, a wire rheostat with a sliding contact is
connected to the circuit with the aid of three terminals.
Given: R± = 10 kQ = 104 Q, R 2 = 3.0 kQ = 3.0 X
X 103 Q, % = 110 V.
Required: U.
Solution: The voltageJof a circuit of resistance R is
U = /i?, where I is the current flowing through the poten­
tiometer, and R is the resistance of the loop A BCD.
The resistance R of the two resistors Rx and R J 2 con­
nected in parallel is

or r>__ R2Ri
R #2 + 2*!
The current I = S//?', where R' is the resistance of
the series connected resistors R and R J2, i.e. = * +
+ R 2/2 j whence

r % _ 2%
R + J h " 2R + R 2

The voltage U — 2<£RI(2R + * 2). After introducing


the value of i?, we get

TT _ 2 % Rt _ 2X110X10* _ e , v
4* j+ * 2 4 X 104 + 3 .0 X 103 ^ 01 V

Problem 4. A battery consisting of two cells is connected


to a rheostat. Determine the current in the rheostat in two
cases:
260 Fundamentals of Electrodynamics 111.2.6

(1) The cells are connected to the rheostat with like


poles.
(2) The cells are connected with unlike poles.
Given: = 8 V, rx = 1 Q, £ 2 = 4 V, r2 = 0.5 £2,
R = 5 Q.
Required: /.
Solution: We choose a clockwise direction of traversing
the loops. The currents flowing in the direction of traversing
are considered to be positive (Fig. III.2.13).

----------- *|

^2t \ .D
A h 1■ ^ J .2
1 AA 1----------- 11------------

Fig. III.2.13 Fig. III.2.14

(1) According to Kirchhoff’s first rule, for the junction A

/, + / = h (i)
according to Kirchhoff’s second rule for loops:
for the loop %XBRA%X
h r x + IR = %X (2)
for the loop <$2BRA<$2
- / 2r2 + I R = g 2 (3)
From (2), we get I x = (<£x — IR)lrx. From (3), we have
/ 2 = (— “H I R ) / r 2*
Using the values of I x and / 2 in (1), we get
—8 a+ « i r t i —IR
111.2.6 Steady Electric Current 261

or
— ^2r i -\-IRr 1 + Ir ir2 = %ir2— I R i*2
whence
r_ 8X0. 5 + 4 X 1 * a
i?r1+ r ,r 2+ i?r2 '“ 5 X 1 + 1 X 0.5 + 5 X 0.5
The current / in the rheostat, as in the other portions,
could be chosen as flowing in the direction opposite to the
one shown in the figure. Consequently, using the junction
and loop rules, we have

1 2 == -^1 H " I (1 )

( 2)

%2 = (3)
~ % 2- l R _ « ! + //? , r r (6 1+ IR r - g 2- / i ?
-----:-------= ----- Mi yi = ---- ;------, *2 r»
r2 rl rl

— ^2r i — I f i r i = $ i r2 + 2 + I r ir 2
j — % 2r\ — ^ ir2 _ —4 x 1 —8 x 0 . 5 _
A

The minus sign in the answer shows that the actual direc­
tion of the current is opposite to the chosen one.
(2) For the junction A (Fig. III.2.14)

12 + / = h (1)
For the loop %XBRA%X

h r x + IR = %1 (2)
For the loop %^RRA<S2
- / 2r2 + //? = - g 2 (3)

The joint solution of Eqs. (1), (2), and (3) gives


r _ — % 2r i ~ h % i r 2 — 4 X 1 + 8 X 0.5 _ Q
i?ri + rir2+ # r2 5X 1+ 1 X 0 . 5 + 5 X 0.5
262 Fundamentals of E lectrodynam ics 111.2.7

7. W ork and Pow er of a Current.


The Joule-L enz Law
1°. The Coulomb and extraneous electric forces do the
work A when charges are moved along an electric circuit
(111.2.3.4°). If the current is steady and the conductors
forming the circuit are stationary, then the energy W that
is irreversibly transformed during the time £ in a conductor
equals the work done:
W = A == IU t

where I is the current, and U the voltage drop in the con­


ductor.
2°. The irreversible conversions of energy in a conductor
carrying a current are due to the interaction of the conduc­
tion electrons with the points of the crystal lattice of the
metal (11.1.6.5°). Collisions of the electrons with the posi­
tive ions at the lattice points result in the electrons trans­
ferring energy to the ions. This energy goes to heat the
conductor.
3°. The power of a current equals the work it does in
a unit time:
P = ~ = IU

where I is the current, and U is the voltage drop in the


given portion of the circuit.
4°. The amount of heat liberated in a conductor during
the time t is
Q = W = IU t

When Q is measured in calories (VII.4.1°) and the other


quantities in SI units (VII.5.2°), we have
Q = 0.24 IU t
Th e last two equations express the Joule-Lenz law: the
amoun t of heat liberated by a current in a conductor is
directly proportional to the current, the time during which
it flows through the conductor, and the voltage drop in it.
111.3.1 Current in Non-Metallic Media 263

Chapter 3
Electric Current
in Non-Metallic Media

1. Current in Electrolytes
1°. Electrolytes are substances in which ionic conductivity
is responsible for a current in them. Ionic conductivity
is the ordered motion of the ions under the action of an
electric field. Electrolytes are solutions of acids, alkalies,
and salts, and also molten salts. Ions are atoms or molecules
that have lost or attached one or more electrons. Positive
ions are also called cations, negative ones, anions. The electric
field producing the ordered motion of ions is created in
a liquid by electrodes—conductors connected to a current
source. A positively charged electrode is called an anode,
a negatively charged one, a cathode. Positive ions (cations) —
ions of metals and hydrogen ions—travel toward the cath­
ode, negative ions (anions)—acid residues and hydroxyl
groups OH—travel toward the anode.
2°. The flow of an electric current through liquids is
attended by electrolysis—the liberation of the substances
forming the electrolyte on the electrodes. Electrolytes are
also called conductors of the second kind. The current in them
is associated with the transfer of a substance, unlike metal
conductors—conductors of the first kind, in which the current
is carried by the collective electrons of the metals (11.7.1.2°).
3°. The appearance of ions in electrolytes is explained
by the phenomenon of electrolytic dissociation—the decom­
position of the solute molecules into positive and negative
ions as a result of reaction with the solvent. The molecules
of the solutes consist of mutually bound ions of opposite
signs (for example Na+Cl“, H +C1", K+I “, and Cu++SO"~).
The forces of attraction between these ions ensure intact-
ness of such molecules. The interaction of these molecules
with the polar molecules of the solvent, for instance water
(111.1.6.4°), results in weakening of the mutual attraction
264 Fundamentals of Electrodynamics III.3.2

of oppositely charged ions. The thermal chaotic motion of


the solute and solvent molecules is attended by their colli­
sions, and this results in decomposition of the molecules
into ions.
4°. The degree of dissociation a is defined as the ratio
of the number of molecules dissociated into ions to
the total number n0 of solute molecules: a = njn^.
The thermal chaotic motion of ions in a solution may be
attended by recombination of ions of opposite signs into
neutral molecules.
Dynamic equilibrium sets in between the processes of
electrolytic dissociation and recombination of ions in
unchanging conditions—the number of molecules decompos­
ing into ions in a unit time equals the number of pairs of
ions combining into neutral molecules during this time.
In a state of dynamic equilibrium, an electrolyte solu­
tion is characterized by a definite degree of dissociation a
determining the number of current carriers in the liquid,
i.e. ions of opposite signs. The degree of dissociation a
depends on the temperature, the concentration of the solu­
tion, and the relative permittivity e of the solvent (111.1.2.6°).
The ions in the electrolytes move chaotically until electrodes
(Par. 1°) are submerged in the liquid. Now the ordered
motion of the ions toward the relevant electrodes is super­
posed onto their chaotic motion, and an electric current
(Par. 1°) appears in the liquid.
The density of the electric current in electrolytes obeys
Ohm’s law for the current density (111.2.4.7°) j = xE. The
expression for the conductivity x (111.2.4.7°) is more
complicated than for metals, however, and is not treated
in elementary physics.

2. Laws of Electrolysis.
Discreteness of Electric Charges
1°. The first law of electrolysis (Faraday's first law): the
mass of a substance liberated on an electrode is directly
proportional to the electric charge q passing through the
electrolyte:
m = Zq1 or m = ZIt
JII.3.3 Current in Non-Metallic Media 265

(since q = I t, where I is the current flowing through the


solution during the time t).
The constant of proportionality Z is called the electro­
chemical equivalent of a substance. It numerically equals
the mass of the substance liberated when a unit amount
of electricity (unit charge) flows through an electrolyte.
2°. The second law of electrolysis (Faraday's second law):
the electrochemical equivalents of substances are directly
proportional to the ratios of their atomic (molar) masses A
to their valencies v:

The quantity F is called the Faraday constant.


3°. From Faraday's combined law of electrolysis
1 A
m = -=-----It, or
1A
m— F v ’ q
F v
it follows that the Faraday constant numerically equals
the electric charge that must be passed through an electro­
lyte to liberate on an electrode a mass of any substance
equal in kilograms to the ratio A h . The value of the Faraday
constant in the SI system is
F = 9.648 X 107 C/kmol
4°. The electric charge q of any ion is determined from
the combined law of electrolysis: q = ± vF !N A, where v is
the valency of an ion, F is the Faraday constant, and NA
is the Avogadro constant (11.1.1.5°). The charge of a mono­
valent ion (v = 1) is equal in its absolute value to the
charge of an electron:
q = e = 1.602 x 10-19 C = 4.803 x 10"10 cgse units (VII.5.20)
Anv electric charge is a multiple of the elementary charge
e (111.1.1.3°).

3. Current in Gases
1°. Gases, unlike metals and electrolytes, consist of
electrically neutral atoms and molecules and in normal
conditions contain no free current carriers (electrons and
ions). In normal conditions, gases are dielectrics. Current
266 F undam entals of Electrodj/nam ics II T .3 .3

carriers can appear in gases only upon their ionization—


the detachment of electrons from their atoms or molecules.
Here, the atoms or molecules of gases transform into posi­
tive ions. Negative ions can appear in gases if the atoms
or molecules attach electrons.
A current in gases is called a gas discharge. For a gas
discharge to occur, an electric or magnetic field must be
applied to the tube containing an ionized gas (a gas-discharge
tube).
2°. Gases may become ionized under the influence of
external action (external ionizers): strong heating, ultra­
violet and X-rays (V.3.6.10), radioactive radiation (VI.4.4.1°)
and when the atoms or molecules of gases are bombarded
with fast electrons or ions.
A measure of the ionization process is the ionization
intensity determined by the number of pairs of oppositely
charged particles appearing in a unit volume of a gas in
a unit time.
3°. To ionize an atom or molecule, the work of ionization
A{ must be done against the forces of interaction between
the electron being detached and the remaining part of the
atom or molecule. The value of A\ depends on the chemical
nature of the gas and the energy state of the electron that
is detached from an atom or molecule. The work of ioni­
zation A x has the smallest value for valence electrons
(VT.2.9.20^ and grows with an increase in the number of
electrons detached from an atom or molecule. This is asso­
ciated with the fact that after the removal of one electron
the strength of the bond with the atom or molecule of the
remaining electrons grows. For example, the work of ion­
ization of a nitrogen atom (N^ is 14.5 eV, of ;ts monovalent
ion (N +) is 29.5 eV, and of its divalent ion (N++^ is 47.4 eV.
4°. Collision, or impact, ionization (ionization through
a collision with an electron or ion) is the detachment from
an atom or molecule of a gas of one or more electrons due
to collision of electrons or ions accelerated hv the electric
field in a discharge with the atoms or molecules of the gas.
Collision ionization of a monatomic gas by electrons or
ions of velocity v is possible provided that
Jtl.3.4 Current in Non-Metallic Media 267

where m = mass of an ionizing particle


mv2/2 = its kinetic energy
A x — work of ionization •
M = mass of an atom.
For a monovalent ion to cope with collision ionization,
it must pass a greater potential difference qp in an acceler­
ating electric field than an electron. This can be seen from
the fact that mi?2/2 = ecp and ml0n/M melectTOn/M.
5°. The process of recombination—the joining up of
oppositely charged particles (positive ions and electrons)
into neutral atoms or molecules—is opposite to the process
of ionization. When the action of an external ionizer does
not change with time, dynamic equilibrium sets in between
the processes of ionization and recombination—the number
of newly formed pairs of oppositely charged particles equals
the number of pairs joining up into neutral atoms or mole­
cules.

4. Semi-Self-Maintained Gas Discharge


1°. A semi-self-maintained gas discharge is defined as the
electrical conductivity of gases due to external ionizers
(111.3.3.2°). Such a gas discharge is characterized by the
curve OCA of the"1current 7
against the voltage^ U between
the electrodes shown in
Fig. III.3.1 (the volt-ampere
characteristic of a gas dis­
charge) . An increase1in *C7?is"at-
t^nded by a growth fin the
number of 'charged particles
reaching the electrode, and
the current V grows up to Fig. III.3.1
such a value of 7 = 7S at
which all the 'charged par­
ticles formed in the gas in a unit time reach the elec­
trodes. Here U = C7S.
2°. The maximum current 7S possible at a given intensity
of ionization (111.3.3.2°) is called the saturation current:
/ s = eN0, where e is the absolute value of an elementary
268 Fundamentals of Electrodynamics 111.3.5

charge (111.1.1.3°), and N 0 is the maximum number of


pairs of monovalent ions formed in a gas in one second.
The sharp growth of the current on the section AB of the
curve in Fig. III.3.1 is associated with the appearance of
collision ionization (III.3.3.4°).

5. Self-Maintained Gas Discharge


1°. A gas discharge (111.3.3.1°) that continues after the
action of an external ionizer (111.3.3.2°)’ stops is called
a self-maintained gas discharge. It is maintained and devel­
oped at the expense of
I I I III IV the ions and electrons ap­
pearing mainly as a re­
sult of collision ionization
(111.3.3.4°).
A semi-self-maintained
gas discharge transforms
into a self-maintained one
at a voltage U{g between
the electrodes called the
ignition voltage. The pro­
cess of such a transition is
called electric breakdown of
the gas.
Apart from the ionization of atoms or molecules by
collisions with electrons inside a gas (volume ionization),
electrons are knocked out from a cathode when it is bom­
barded with positive ions (surface ionization). [For other
processes of detaching electrons from a cathode, see therm­
ionic emission (111.3.7.3°) and photoelectric emission
(111.3.7.2°).!
2°. Several kinds of self-maintained discharges in gases
are distinguished depending on the pressure of the gas and
the voltage applied to the electrodes. Glow discharge is
observed at low pressures, usually from hundredths to
several millimetres of mercury (VII.3.2°). Four regions are
distinguished in it: I —cathode dark space, I I —negative
glow. I I I —Faraday dark space, and IV —positive column
(Fig. III.3.2). The regions / - / / / form the cathode part of
111.3.5 Current in Non-Metallic Media 269

a discharge. There is a large concentration of positive ions


near the cathode on the boundary between the regions I
and / / , and as a result the potential in the discharge sharply
drops. In the region I I , the accelerated electrons are sources
of collision ionization. The glow in this region is due to
recombination (111.3.3.5°) of the electrons and positive
ions into neutral atoms or molecules. The region IV has
a constant and large concentration of positive ions and
electrons due to collision ionization of the atoms or mole­
cules of the gas by electrons. The positive column is gas-
discharge plasma (111.3.6.1°). The glow of the positive
column is determined by the emission of the excited atoms
or molecules of the gas (V.3.3.2°) and therefore has char­
acteristic colours. This determines the use of a glow dis­
charge in slow-discharge tubes and gas lasers (VI.2.10.1°).
3°. A corona self-maintained discharge is observed at
a normal pressure in a gas that is in a greatly inhomoge­
neous electric field (near the points of rods, the wires of
high-voltage lines, etc.). The ionization of the gas by elec­
tron collisions and its glow, reminding one of a corona, occur
only in a small region adjoining an electrode (a corona
electrode). The glowing layer is called a corona layer. If the
corona is formed by a cathode, then the electrons causing
ionization inside the corona layer are knocked out by the
positive ions from the cathode. If the corona is formed by
an anode, then the electrons appear near the anode owing
to the ionization of the gas under the action of the emission
of the corona layer. At an increased voltage, a corona dis­
charge on a rod point has the form of a glowing brush—
a system of thin glowing lines emerging from the point
and having bends and breaks that change with time (a brush
discharge).
4°. A spark discharge observed at a normal pressure and
a great field intensity between the electrodes has the form
of intermittent bright zigzag filaments-—channels of an
ionized gas. The filaments penetrate the space between the
electrodes and vanish, being replaced with new ones.
Bright glowing of the gas is observed, and a large amount
of heat is liberated. In the spark channels, where a high
pressure and very high temperatures are produced, electron
and ion avalanches appear. They determine all the proper-
2?0 Fundamentals of Electrodynamics 111.3.6

ties of a spark discharge. Lightning is an example of it.


The main channel of lightning has a diameter from 10 to
25 cm. A flash of lightning is up to several kilometres long,
and a current is developed in it in a pulse up to hundreds
of thousands of amperes.
5°. An arc discharge is a form of discharge at a great
current density and a comparatively small voltage between
the electrodes, of the order of magnitude of several scores
of volts. The main cause of an arc discharge is intensive
thermionic emission (111.3.7.3°) of a hot cathode. The
electrons are accelerated by the electric field and cause
collision ionization of the gas molecules; the electric resis­
tance of the gas gap diminishes, and its conductance greatly
grows. A column of a brightly glowing gas (an electric arc)
appears between the electrodes. At atmospheric pressure,
the temperature of the cathode reaches 3000 °C. The bom­
bardment of the anode by the electrons produces a depres­
sion in it—the crater of the arc with a temperature of about
4000 °G at atmospheric pressure. The temperature of the gas
in the channel of the electric arc is 5000-6000 °G. An arc
discharge is used as a powerful source of light in searchlights,
motion picture projectors, and the like.

6. Plasma
1°. Plasma is defined as a special state of aggregation
of a substance characterized by a high degree of ionization
of its particles. The degree of ionization a of a substance
is the ratio of the concentration of the charged particles
to the total concentration of particles. Depending on the
degree of ionization, we distinguish weakly ionized (a is
a fraction of a per cent), partly ionized (a is several per
cent), and fully ionized (a is close to 100%) plasma. An
example of weakly ionized plasma in natural conditions
is the ionosphere—the upper layers of the atmosphere.
The Sun, hot stars, and some interstellar clouds are exam­
ples of fully ionized plasma that is formed at a very high
temperature (high-temperature plasma) [see also thermo­
nuclear reactions (VI.4.15.1°)].
III.3.7 Current in Non-Metallic Media 271

The plasma in gas discharges and gas-discharge lamps is


an artificially created plasma of different degrees of ioniza­
tion. •
2°* The high electrical conductivity of plasma brings
its properties close to those of conductors. Metal conductors
are an example of fully ionized plasma—there are no neutral
atoms or molecules in metals. The electrical conductivity
and thermal conductivity of fully ionized plasma depend
on the temperature according to laws proportional to T3/2
and respectively.
The control of the motion of plasma in electric and
magnetic fields is the basis for using plasma as the working
substance (11.4.5.6°) in various engines for the direct conver­
sion of internal energy (11.4.1.2°) into electric energy
[plasma sources of electric energy, magnetohydrodynamic
(MHD) generators].
3°. The features of plasma permitting us to consider
it as a special state of aggregation of a substance are its
strong interaction with external electric and magnetic
fields due to the high electrical conductivity of plasma,
the specific collective interaction of plasma particles,*
and the presence of elastic properties leading to the possi­
bility of generating and propagating various oscillations
in plasma.

7. Current in a Vacuum.
Emission Phenomena
1°. A vacuum is defined as such a state of rarefaction
of a gas when we may ignore the collisions between its
molecules and consider that the mean free path I (11.2.3.1°)
exceeds the linear dimensions d of the vessel containing
the gas (I d). The conductivity of the gap between the
electrodes in the state of a vacuum is called an electric
current in a vacuum. There are so few molecules that their
ionization cannot ensure the presence of the number of

* This interaction has an intricate nature whose consideration is


beyond the scope of elementary physics.
272 Fundamentals of Electrodynamics 1 1 1 .3 .7

electrons and positive ions needed for electrical conductiv­


ity. The conductivity of the gap between the electrodes
in a vacuum can be ensured only with the aid of charged
particles appearing as a result of emission phenomena on
the electrodes.
2°. The phenomenon of photoelectric emission consists
in the emission of electrons from the surface of a body
(for example a metal) placed in a vacuum or gas under the
action of light. This phenomenon is observed and used in
specially prepared photocathodes having a high sensitivity.
The sensitivity of a photocathode is determined by the ratio
of the number of electrons emitted by the photocathode
to the number of light quanta (photons) it absorbs [see also
the photoelectric effect (V.5.2.10) and the photon (V.5.1.20)].
3°. Thermionic emission is the emission of electrons by
solids or liquids when they are heated. Thermionic emis­
sion from a heated cathode is of significance for a current
in an evacuated gas-discharge tube. The electrons emitted
by a heated body are called thermoelectrons, and the body
itself—an emitter.
A thermionic current (111.3.8.1°) may appear as a result
of thermionic emission. For an electron to be ejected from
a metal, the kinetic energy of the electron must be suffi­
cient to overcome its bond with the metal, i.e. equal a quan­
tity called the work function A . At room temperature,
only a few electrons have the required kinetic energy, and
the thermionic emission is not great. This emission occurs
intensively when the emitter is heated to a high tempera­
ture corresponding to the visible glow of a heated metal.
4°. When the surface of a metal is bombarded in a va­
cuum with electrons that are accelerated by an electric
field, a counterflow of electrons from the surface is observed.
This phenomenon is called secondary electron emission. The
secondary stream consists of electrons reflected by the
surface, and also of electrons ejected from the metal. The
greatest emission of secondary electrons occurs at energies
of the primary electrons of several hundred eV. For some
pure metal surfaces (mercury, platinum), the number n2
of secondary electrons is from 1.75 to 1.78 times greater
than the number nx of primary electrons. The ratio n2lnx = 6
is called the secondary electron emission coefficient. The value
111.3.8 Current in Non-Metallic Media 273

of 8 is greater for dielectrics and semiconductors than for


metals. The phenomenon of secondary electron emission
is used in electron multipliers tlftit amplify weak currents
many times (Fig. I l l .3.3). The electrons that are ejected

from the cathode Ca under the action of light pass consec­


utively to the emitters CaSx, CaS2, . . ., CaSn. Secondary
electron emission occurs on each of them. If a multiplier
has n emitters or n cascades, then a sufficiently strong
stream of electrons is obtained on the last electrode—the
anode A (collector).

8. Two-Electrode Valve—Diode
1°. Electron valves (tubes) are devices based on the use
of the thermionic emission phenomenon (111.3.7.3°). The
simplest kind of electron valve is a two-electrode one—
a directly heated diode. It is depicted schematically as shown
in Fig. III.3.4. If the anode (plate) is connected to the
positive pole of a source of steady (direct) current, and
the cathode (filament) to the negative pole, then a steady
thermionic current I \ will appear in the circuit of the
valve.
At a constant cathode temperature, the thermionic cur­
rent in a directly heated diode depends on the anode voltage
of the diode—the voltage U& applied between the anode
and the cathode (the volt-ampere characteristic), the size and
mutual arrangement of the electrodes, the work function
of the cathode electrons (111.3.7.3°), and its temperature.
The volt-ampere characteristic of such a diode at a constant
274_____________ Fundamentals of Electrodynamics__________ ill.3.8

cathode temperature is shown in Fig. III.3.5. The non­


linear shape of the characteristic indicates that Ohm’s law
a n .2.4.2°) is not observed in the diode.
2°. At small anode voltages, the current I A grows slowly
with increasing UA. This is associated with the circumstance
that at low values of UA not all the electrons emitted by

the cathode reach the anode. Part of the electrons form


an electron cloud—a space negative charge that prevents
the motion toward the anode of the electrons newly emitted
from the cathode. An increase in the voltage UA is attended
by gradual dissipation of the electron
cloud, and the current I A grows. At UA =
= C7S, the thermionic current I A reaches
its maximum possible value at the given
cathode temperature. This value I A = I 8
is called the saturation current. If N is
the total number of electrons emitted by
the cathode at the given temperature in
a unit time, then / s = Ne, where e is the
Fig. 111.3.6^3 absolute value of the charge of an electron
(cf. 111.3.4.2°).
A directly heated diode has an appreciable shortcoming:
if the cathode is heated by means of an alternating current,
then its temperature periodically changes, and this causes
oscillations of the current in the valve circuit. In the hot-
cathode diode shown schematically in Fig. III.3.6, this
shortcoming is eliminated by placing a cathode heater (for
example a tungsten filament) isolated from the cathode
inside it.
tn .3 .9 Current in Non-Metallic Media 275

3°. Diodes have unipolar (unidirectional) conductance:


a current is possible in a valve only if the anode potential
is higher than the cathode one,* i.e. the voltage U± > 0.
If a potential that is negative relative to the cathode is fed
to the anode, i.e. an electric field is produced that will
repel electrons from the anode, then the valve will be cut
off—there will be no anode current, i.e. no current in the
valve circuit. This property of diodes makes possible their
use for rectifying alternating current (IV.2.2.3°). A vacuum
two-electrode valve used to rectify alternating current is
sometimes called a kenotron.

9. Three-Electrode Valve—Triode
1°. Multielectrode (three and more) valves—triodes,
tetrodes, pentodes, etc.—are used for controlling the therm­
ionic current in a valve, A triode has a third electrode—
a control electrode or grid G—between the anode and the

Fig. III.3.7

cathode through which the electrons flying from the cathode


to the anode pass. A directly heated triode (a) and a hot-
cathode triode (111.3.8.2°) (b) are shown schematically in
tig. III.3.7. The grid is arranged near the cathode so that
even with a small voltage UG between it and the cathode
(the grid voltage, or bias) a strong electric field is set up
near the cathode that noticeably affects the motion of the
electrons in the triode.
276 Fundamentals of Electrodynamics 111.3.9

2°. The dependence of the anode current I A on the


voltage UG with the same heating of the valve and a con­
stant voltage UA between the anode and the cathode (the
anode, or plate, voltage) is called the static grid character­
istic of the valve (Fig. III.3.8). The shape of the grid char­
acteristics is explained by the fact that several electric
fields are superposed in a triode. If the grid potential is
higher than the cathode one, i.e. the grid voltage, or bias,
UG is positive (UG > 0), then the electric field produced
by the grid near the cathode coincides in direction with

that existing between the anode and the cathode. The


electrons flying out of the hot cathode will travel to the
anode with speeds greater than when UG = 0. The number
of electrons reaching the anode in a unit time and, there­
fore, the anode current in the triode grow when UG in­
creases. If, on the contrary, UG < 0, then the electric field
of the grid weakens the one between the anode and the
cathode, and the speed of the electrons, as well as the anode
current diminish.
3°. The negative grid voltage at which the anode current
completely vanishes is called the cut-off voltage. Its absolute
value grows with an increasing anode voltage. Thus, by
changing the grid voltage C/G, we can control the current
in the valve. This is why the grid G is called a control
electrode (Par. 1°). The triode is employed in radio equip­
ment for amplifying weak varying currents in a valve gen­
erator (IV. 2.9.1°).
III.3.10 Current in Non-Metallic Media 277

10. Electron Beams, Cathode-Ray Tube


1°. Electron beams (electron o f cathode rays) are a stream
of rapidly flying electrons. Electron beams are formed in an
olectron valve (111.3.8.1°) and various gas-discharge devices.
2°. Properties of electron beams:
1. They cause some solids and liquids to glow [see lumi­
nescence (V.3.3.1°)I.
2. The deceleration of high-speed electron beams in
a substance leads to the appearance of X-rays (V.3.6.10).
3. Electron beams deviate in electric fields. For example,
when flying between the plates of a charged parallel plate
Vertical

capacitor (111.1.11.2°), electron beams deviate toward the


positively charged plate.
4. Electron beams deviate in a magnetic field under the
action of the Lorentz force (111.4.5.1°).
5. Electron beams impinging on a substance heat it and
exercise a mechanical action on it.
3°. A cathode-ray tube (electron-beam value) is a device
with a hot cathode (111.3.8.2°) based on the phenomenon of
thermionic emission (111.3.7.3°). The electron beam in
a cathode-ray tube is controlled with the aid of electric
and magnetic fields. The design of a cathode-ray tube is
shown in Fig. III.3.9. The electrons emitted by the hot
cathode pass through the control grid (the first control
electrode) and two accelerating anodes. This entire system
278 Fundamentals of Electrodynamics 111.3.10

called an electron gun is intended to obtain an electron beam


having the smallest possible cross section (focussing of the
electron beam) on the tube screen coated with a luminescent
substance (V.3.3.20). A negative potential (from —20 to
—70 V) is applied to the control grid to focus the electron
beam. The field of this electrode contracts the electron
beam emitted by the cathode. A positive potential from
+ 250 to + 500 V is fed to the first accelerating anode,
and from +1000 to +2000 V to the second one. The poten­
tials are still higher in television tubes (IV.4.5.5°). The
focussing and brightness of the fluorescence of the electron
beam on the screen are varied by changing the potentials
of the control grid and the anodes.
4°. After the electron gun, the focussed electron beam
passes a system of vertical and horizontal deflecting electrodes
(deflecting plates) which a variable voltage is applied to.
The oscillations of the potential on the plates cause vertical
and horizontal oscillations of the electron beam on the
tube screen. The simultaneous use of vertical and horizon­
tal deflecting plates makes it possible to move the electron
beam on the screen in an arbitrary direction. The small
mass of the electrons in an electron beam ensures a low
inertia of a cathode-ray tube: the electron beam reacts
virtually instantaneously to voltage changes on the deflect­
ing plates. This underlies the use of a cathode-ray tube
in an electron (cathode-ray) oscillograph—an instrument
employed to study periodically varying voltages.
5°. Two kinds of cathode-ray tubes are distinguished:
(a) electrostatic tubes in which the electron beam is
deflected by changing the electric field between the deflect­
ing plates, and
(b) electromagnetic tubes in which the electron beam is
deflected by changing the value and direction of the magnetic
induction vector B (111.4.1.3°) of the magnetic field in
which the electron beam travels. The Lorentz force deflect­
ing the electron beam (111.4.5.1°) changes in accordance
with the change in the vector
IIL3.11 Current in Non-Metallic Media 279

11. Electrical Conduction


of Pure Semiconductors
1°. Semiconductors are defined as substances whose
resistivity p may change within broad limits between the
resistivity of good insulators such as glass and that of
good conductors such as copper, and diminishes very rapidly
with increasing temperature. The chemical elements having
the properties of semiconductors form the group shown in
Fig. III.3.10 in Mendeleev’s periodic table of the elements

50 @57 @52 @ 53 @
Sn Sb Te I

Fig. III.3.10 Fig. III.3.11

(VI.2.9.1°). Germanium Ge, silicon Si, and tellurium Te


are typical widely used semiconductors. These chemical
elements belong to groups IV and VI of the periodic table,
and the outer electron shell of an isolated atom of one of
these elements carries four valence electrons (VI.2.9.2°).
2°. Crystals of germanium and other semiconductors have
an atomic crystal lattice (11.7.1.2°). A plane diagram of the
structure of a germanium crystal is shown in Fig. III.3.11.
The four valence electrons of each germanium atom are
hound to the similar electrons of neighbouring atoms by
chemical electron-pair bonds (a covalent bond) (VI.3.3.1°).
There are no free electrons in a pure crystal of germanium
and in crystals of other semiconductor elements at low
temperatures, and such crystals are good dielectrics in
these conditions.
3°. A chemically pure semiconductor can conduct elec-
tricity when the covalent bonds in the crystals are broken,
280 Fundamentals of Electrodynamics IIL3.il

For example, heating to relatively moderate temperatures


results in the breaking of the covalent bonds, the appearance
of free electrons, and the appearance of intrinsic electron
conduction (n-type conduction—the charge carriers are nega­
tive) in the pure semiconductor. The energy that has to be
spent on creating electrical conduction in the crystals of
pure semiconductors is called the activation energy of intrin­
sic conduction. Its values in eV for various semiconductors
are indicated in Fig. III.3.10 in circles.
4°. Elevation of the temperature is attended by an
increase in the number of broken covalent bonds and in the
number of free electrons in crys­
tals of pure semiconductors. This
signifies that the conductivity
of pure semiconductors grows
with elevation of the tempera­
ture. The resistivity of pure
semiconductors correspondingly
diminishes when they are heated
(Fig. III.3.12). This feature ap­
preciably distinguishes semicon­
ductors from metals, in which
the resistivity grows upon
heating (111.2.5.2°).
Apart from heating, the break­
ing of a covalent bond and
the appearance of intrinsic conduction of semiconductors
may be caused by illumination (photoconduction of semi-
conductors), and also by the action of strong electric fields.
5°. When an electron in a crystalline pure semiconductor
receives the energy needed to break its covalent bonds and
leaves its site, the electrical neutrality of the crystal is
violated at this site. A surplus positive charge appears at
the site left by the electron—a positive hole is formed.
It behaves like a charge equal in absolute value to the
charge of an electron, but positive in sign. An adjacent
electron may move over to the site freed by the electron,
a hole, and this is equivalent to motion of the positive
hole: it will appear at the new site left by an electron.
The nature of the motion of a hole in a crystal can be
understood from the following analogy. Assume that one
111.3.12 Current in Non-Metallic Media 281

boy steps out of a rank of boys and forms a “vacancy”.


If all the boys who stood to his right move onto the vacant
place one by one, then everything will occur as if the vacant
place moves in the opposite direction,
6°. Electrons in an external electric field move in the
direction opposite to that of the electric field intensity.
The positive holes move in the direction of the electric
field intensity, i.e. in the same direction in which a positive
charge would move under the action of the electric field.
The motion of the electrons and holes in an external field
occurs over the entire semiconductor crystal. The electrical
conduction of a pure semiconductor due to the ordered
motion of the holes is called intrinsic hole conduction (p-type
conduction—similar to the motion of positive charges).
The temperature dependence of the resistivity in hole
conduction is similar to that characteristic of electron
conduction (Par. 4°). The total conductivity of a semi­
conductor (111.2.4.1°) is the sum of its n- and p-type con­
ductivities.

12. Impurity Electrical Conduction


of Semiconductors
1°. By the impurity conduction of semiconductors is
meant their electrical conduction due to the introduction
of impurities (impurity centres) into their crystal lattices
(doping with impurities). The impurity centres may be (a)
atoms or ions of foreign chemical elements introduced into
the semiconductor lattice structure; (b) surplus atoms or
ions of semiconductor elements occurring interstitially,
i.e. between the regular lattice sites; (c) various other
defects and distortions in the crystal lattice: vacant sites,
(•racks, dislocations appearing in the deformations of crys­
tals, etc.
2°. Impurities change the conduction of semiconductors.
A change in the impurity concentration changes the number
of current carriers—electrons and holes. The possibility
o f controlling the number of current carriers (by heating
o r by the action of other factors such as illumination)
underlies the widespread use of semiconductors in science
282 Fundamentals of Electrodynamics III.3.12

and engineering. This possibility is absent in metals


(111.2.1.7°).
3°. Impurities can be an additional source of electrons
in semiconductor crystals. Assume, for instance, that one
germanium atom having four valence electrons is replaced
in a germanium semiconductor lattice with an impurity
atom having five valence electrons (phosphorus, arsenic,
antimony). Four of the impurity atom electrons will be
bound by covalent bonds (VI.3.3.1°) to electrons of adjacent

Fig. III.3.14

germanium atoms, while the fifth electron cannot form


a covalent bond. It is a surplus one, bound more weakly
to its own atomic nucleus, and can easily leave the latter
and become a free electron (Fig. III.3.13). An electric
field causes such electrons to begin ordered motion in the
semiconductor crystal, and electron impurity conduction
appears in it. Semiconductors with such conduction are
called electron, or n-type, semiconductors. The impurity
atoms supplying electrons are called donors.
4°. When an atom with four valence electrons is replaced
in a semiconductor crystal with an impurity atom having
three valence electrons (indium, boron, aluminium), one
electron is lacking for the formation of all the covalent
bonds in the lattice structure (Fig. III.3.14). The impurity
atom can create all the bonds, however, if it borrows an
electron from the nearest host atom in the lattice. A posi-
III.3.13 Current in Non-Metallic Media 283

tive hole (111.3.11.5°) will be formed instead of the electron


borrowed from the host atom. This hole, in turn, can be
filled by an electron from the next adjacent atom of the
lattice, and so on. The consecutive filling of the positive
holes by electrons is equivalent to the motion of the hole
in the semiconductor and to the appearance of current
carriers in it. An electric field causes the hole to move
in the direction of the field intensity vector, and hole impu­
rity conduction appears in the semiconductor. Semiconduc­
tors with such conduction are called impurity hole, or p-type,
semiconductors. The impurity atoms resulting in hole con­
duction are called acceptors.
5°. If donor and acceptor impurities are simultaneously
introduced into a semiconductor, then the nature of its
conduction (n- or p-type) is determined by the impurity
having the higher concentration of the current carriers—
electrons or holes. With any kind of conduction in a semi­
conductor, the concentration of the current carriers in it is
considerably smaller than in metals. But the magnitude
of this concentration, like the energy of the current carriers
in semiconductors, in contrast to metals depends quite
considerably on the temperature. Heating is attended by
a sharp growth in the number of current carriers.

13. Electrical Properties


of p - and rc-Type Semiconductor Junctions
1°. The region of a monocrystalline semiconductor
(11.1.6.4°) in which electron conduction changes to hole
conduction (or vice versa) is called an electron-hole junction
(p-n junction). A p-n junction is usually formed in a semi­
conductor crystal where the introduction of the relevant
impurities creates regions having different (p- and n-)
conduction.
2°. When two semiconductors with different kinds of
conduction come into contact, interdiffusion (11.1.3.1°) of
the current carriers through the line of contact (the junction)
of the semiconductors will occur. Electrons from the rc-type
semiconductor will diffuse into the p-type one. As a result,
electrons will leave the region of the rc-type semiconductor
284 Fundamentals of Electrodynamics III.3.13

adjoining the junction, this region will be depleted of


electrons, and an excess positive charge will be formed
in it near the junction. The diffusion of holes from the
p-type semiconductor for similar reasons will result in the
appearance of a surplus negative charge near the junction
in the p-type semiconductor. As a consequence, a barrier
electric layer of thickness I (Fig. III.3.15) is formed at the
p-n junction. The electric field of the barrier layer prevents
the further transition of electrons and holes across the

OQO j+ + - ©©©
©GO ++ ©©©
gog ::!1 ©©©

Fig. III.3.15 Fig. III.3.16

junction between the two types of semiconductor. The


barrier layer has an increased resistance in comparison
with the remaining regions of the semiconductors.
3°. An external electric field affects the resistance of the
barrier layer. If the rc-type semiconductor is connected
to the negative pole of a source and the positive pole of the
latter is connected to the p-type semiconductor, then the
electric field will cause the electrons in the ra-type semi­
conductor and the holes in the p-type one to approach one
another toward the junction between the semiconductors.
The electrons will pass through the junction and “fill” the
holes. With such a forward (passing) direction of the external
electric field, the thickness of the barrier layer and its
resistance continuously diminish (Fig. III.3.16). In this
direction, the electric current passes through the junction
between the two semiconductors.
4°. If the rc-type semiconductor is connected to the
positive pole of the source, and the p-type one to its nega­
tive pole, then the electrons in the ra-type semiconductor
111.4.1 Magnetic Field of a Steady Current 285

and the holes in the p-type one will move in opposite direc­
tions from the junction under the action of an electric
field (Fig. III. 3 . 17). This results in broadening of the barrier
Jayer and in a growth of its resis­
tance. The direction of the ex- u— :__ I
n ^ * P
lernal electric field broadening —---------- -------- E—
the barrier layer is called the + + +
«*-© !+++ — |!© — ~~**
<±H-
barrier (reverse) one. With such f j+++++++++— !| <±H*
a direction of the external field, |^-Q ;+ + +| -----!;©©~ ->h
virtually no current flows through .
the junction of an ra-type ------------- ----------- “—
and a p-type semiconductors.
5°. A p-n junction has uni- Fig. 111,3,17
Flg* III.3.17
polar conduction similar to the
rectifying action of a two-electrode valve—a diode
(111.3.8.3°). Therefore, a semiconductor with one p-n junc­
tion is called a semiconductor diode. Semiconductor diodes
have many advantages over electron two-electrode valves
(a saving in energy for obtaining current carriers, a small
size, high dependability, and a long service life). A short­
coming of semiconductor diodes is the limited interval
of temperatures within which they function (approximately
from —70 to + 125 °C).

Chapter 4
M agnetic Field
of a Steady Current*
1. Magnetic Field.
Induction Vector of a Magnetic Field.
Magnetic Flux
1°. A magnetic field is one of the parts of an electromagnet­
ic field (111.1.3.1°). A feature of a magnetic field is that
it is produced by conductors carrying currents, moving
* All the formulas in this chapter are given only in SI unit*
(VII.5 3° .
286 fu n d a m e n ta ls of E le ctro d y n a m ic s m.4.1

electrically charged particles and bodies, and also by mag­


netized bodies (111.6.2.1°) and a varying electric field
(IV.4.1.3°).
A magnetic field whose characteristics (111.4.1.3°,
111.4.1.8°) do not change with time is called stationary.
Otherwise the magnetic field is a varying (non-stationary)
one. The appearance of a sta­
tionary magnetic field near
a conductor with a current is
illustrated by Oersted's basic
experiment. If a magnetic
needle that can freely rotate
about its vertical axis is placed
under a straight conductor
through which a steady cur­
rent is flowing, the needle will
turn as shown in Fig. III.4.1, tending to take up a posi­
tion at right angles to the conductor. The needle will coin­
cide the more accurately with this direction, the greater
is the current, the closer is the needle to the conductor,
and the smaller is the influence of the Earth’s magnetic
field.
2°. A magnetic field acts on electrically charged particles
and bodies only if they are moving. It acts on magnetized
bodies (111.6.2.1°) regardless of whether they are moving
or stationary.
3°. The force characteristic of a magnetic field is the
induction vector B of the magnetic field (the magnetic induc­
tion vector). The concept of the magnetic induction vector
is introduced on the basis of one of three experimentally
established facts: (a) the orienting action of a magnetic
field on a current-carrying loop (Par. 6°); (b) the deflection
of a current-carrying conductor in a magnetic field
(111.4.2.4°); and (c) the deflection of a beam of electrically
charged particles moving in a magnetic field (111.4.5.1°).
4°. By the magnetic moment of a loop (a closed plane
contour) carrying the current I is meant the vector pm equal
to

pm = IS n0
11 1 A .1 M a g n e tic F ield of a S tea d y C urrent 28 1

where S = area of the surface confined within the loop


n0 = vector whose magnitude equals unity and
directed at right angles to the plane of the loop
(a unit normal vector).
The vectors n0 and pm are at right angles to the plane of
the loop and are oriented so that the current flows counter­
clockwise when looking from the tips of the vectors n0 and
P m (Fig- III.4.2).
The direction of the unit normal vector n0 and the
magnetic moment vector pm is also determined by the

Uir/fc,

Fig. III.4.2 Fig. III.4.3

right-hand screw rule: if the handle of an auger (or the head


of a screw) with a right-hand thread is turned in the direc­
tion of the current in a loop, then the direction of the vectors
n0 and p m will coincide with the direction of motion of the
point of the auger (or screw) (Fig. III.4.3).
5°. A current-carrying loop suspended on an inelastic
thread in a homogeneous magnetic field is acted upon by
a moment of forces (1.3.1.4°) that turns it. The orienting
action of a field on a current-carrying loop is used to choose
the direction of the magnetic induction vector B of the
field in which the loop is placed. The direction of the vec­
tor B at a given point of a magnetic field is taken as the
direction of the normal vector n0 and the magnetic moment
vector pm of a small current-carrying loop* in the vicinity
of the given point (Fig. III.4.4).

* The current-carrying loop must be so small as not to distort


the magnetic field it is probing. It is also assumed that the self-mag­
netic field induced by the current-carrying loop may be disregarded.
288 F u n d a m en ta ls of E le ctro d y n a m ic s 111.4.1

The direction of the vector B coincides with that of


a straight line drawn through the centre of a magnetic
needle from its south (south-seeking) pole S to its north
(north-seeking) pole N if this small* needle is placed in the
vicinity of the given point of the magnetic field (Fig. III.4.5).

Fig. III.4.4 Fig. III.4.5

A magnetic field is called homogeneous, or uniform, if


the vectors B at all its points are the same. Otherwise, the
field is called non-uniform.
6°. A current-carrying loop placed in a homogeneous
field is acted upon by the moment of forces M whose mag­
nitude is
M — p mB sin a

where p m = magnitude of the magnetic moment vector of


the loop (Par. 4°)
B = magnitude of the magnetic field induction
vector
a = angle between the vectors pmand B (Fig. I l l .4.6).
The following definition of the magnetic induction
ensues from this formula: the magnitude of the magnetic
induction vector at a given point of a homogeneous magnetic
field equals the maximum value of the moment of the forces
Mmax acting on a small current-carrying loop having

* A magnetic needle is considered to be small if it does not dis­


tort the magnetic field by its presence and its self-magnetic field may
be disregarded.
111.4.1 M a g n e tic F ie ld of a S te a d y C urrent 289

a magnetic moment p m whose magnitude is unity and


placed in the vicinity of the given point:
__ Mm ax
d ^m ax
Pm IS

The value M = M max corresponds to the orientation


of the loop at which a = jx/2 radians, i.e. the magnetic
induction lines (Par. 7°) are
in the plane of the loop, and
its magnetic moment is per­
pendicular to the induction
lines. In this position, a cur­
rent-carrying loop will be in
unstable equilibrium (1.4.3.2°).
A current-carrying loop
(or any closed contour carrying
a current) will be in a stable
(equilibrium) position when the plane of the loop is per­
pendicular to the induction lines, and the magnetic moment
vector of the loop is parallel to the induction lines (see
Fig. I l l .4.4).
7°. A magnetic field can be depicted graphically if we
introduce the notion of magnetic induction lines (more
often called B-field, or magnetic field, lines). These lines
are imaginary ones tangents to which at every point coincide
with the direction of the vector B at these points of the
Held. The magnetic field lines are closed. This signifies
that there are no free magnetic charges in nature (magnetic
masses) [cf. electrostatic field intensity lines (I II .l.3.5°)!.
Magnetic field lines are drawn with an arbitrary density.
It is generally considered that the magnitude of the magnet­
ic field induction vector is proportional to the number
of magnetic field lines drawn through a unit surface area
perpendicular to these lines. Unlike an electrostatic field,
a magnetic field is a non-potential one (111.1.8.1°). A non­
potential field is called a vortex field (see also 111.5.3.2°).
For examples of some magnetic fields see III.4.3.4°-6°.
8°. The magnetic induction flux (magnetic flux) AO
through a portion of a surface of small area AS is defined as
the scalar quantity equal to
AO = B AS cos a = Bn AS
19-0211
290 F undam entals of E lectrodynam ics 111.4.1

where Bn = B cos a is the projection of the magnetic


induction vector B onto a normal to the area (Fig. III.4.7).
A positive value of the magnetic flux corresponds to an
acute angle a, or to the condition Bn > 0. A negative
value corresponds to an obtuse angle a, or to the condition
Bn < 0 . The magnetic flux through a surface of area S
is found by algebraic summation
of the fluxes AO through the por­
tions of the surface. If the mag­
netic field is homogeneous, then
the magnetic flux through a plane
surface of area S is
O = BS cos a
Problem 1. A wire coil 20 cm
in diameter is placed in a homo­
geneous magnetic field whose induction is 10~3 T. When
a current of 2 A is passed through the coil, it turns
through 90 degrees. What moment of forces acted on the
coil?
Given: d = 20 cm = 2 X 10"1 m, B = 10~3 T, I = 2 A,
a = 90°.
Bequired: M.
Solution: The magnitude of the moment of the forces
acting on a coil carrying a current in a magnetic field is
M = p mB sin a, where p m = IS is the magnetic moment
of the coil, S is its area, and a is the angle between the
vectors pm and B.
Since a = 90°, then
Ind2 g _ 2 X 3.14 X 4 X 10~2 x jq_3
M 6 x 10~5 N-m

Problem 2. A magnetic field with an induction of 0.1 T


contains a bar 1 m long that rotates at right angles to the
direction of the magnetic field lines. The axis of rotation
passes through one end of the bar. Find the magnetic induc­
tion flux through the surface formed by the bar during
every revolution.
Given: B = 0.1 T, I = 1 m, a = 0.
Bequired: O.
III.4.2 M agnetic Field of a Steady Current 291

Solution: The magnetic induction flux through the


area S is O = BS cos a, where a is the angle between
the induction vector and a normal to the area S formed
by the rotating bar.
Hence,
<I>= Bnl2 = 0.1 x 3.14 x 12« 0.3 Wb

2. A m p e r e ’s L a w
1°. A current-carrying conductor placed in a magnetic
held is acted upon by the Ampere force. Ampere's law:
a small length of a conductor carrying the current I and
of length AI placed in a homogeneous magnetic field of
induction B is acted upon by the force AF whose magnitude
is
AF = I AZ B sin a = / AZ B ±
where a is the angle between the vector B and the current-
carrying conductor (Fig. III.4.8). The magnitude of the

force AF depends on the component of the vector B per­


pendicular to the conductor: B ± = B sin a.
2°. The vector AF is perpendicular to the current-carry­
ing conductor and to the vector B. The direction of the
force AF is determined according to the left-hand rule: if
Ihe palm of the left hand is arranged so that the component
B j of the induction vector perpendicular to the conductor
enters it, and the four stretched-out fingers indicate the
292 F undam entals of Flectrodynam ics 111.4.3

direction of the current, then the thumb bent back to form


an angle of 90 degrees will indicate the direction of the
force exerted by the field on the current-carrying conductor
(Fig. III.4.9).
3°. Unlike the Coulomb forces, which are central ones
(111.1.2.3°), the Ampere force is not a central one. It is
directed at right angles to the magnetic field lines.
4°. Ampere’s law can be used to determine the magnitude
of the magnetic induction vector. The magnitude of the
induction vector at a given point of a homogeneous magnet­
ic field equals the greatest force exerted on a conductor
of a unit length placed in the vicinity of the given point
and carrying a unit current (see also 111.4.1.6°):
D_ AFmax
I Al
The value AF = AFmax is reached provided that the con­
ductor is at right angles to the field lines (111.4.1.7°).

3. M a g n e tic F ie ld
o f a n E le c tr ic C u rr e n t
1°. An electric current flowing in a conductor produces
a magnetic field in the space surrounding it. The magni­
tude and direction of the magnetic induction vector at
any point of the magnetic field depend on the current in
the conductor, its geometrical shape, the arrangement of
the given point relative to the conductor, and also on the
magnetic properties of the medium containing the con­
ductor and the point. It must be noted that the magnetic
field at a given point is induced simultaneously by all
the portions of the conductor carrying the current. Accord­
ing to the principle of the superposition of fields (111.1.3.7°),
the magnetic induction vector of a field at an arbitrary
point of it equals the geometrical sum of the magnetic
induction vectors of the fields induced by all the portions
of the conductor.
2°. If a conductor carrying the current I sets up in
a vacuum a magnetic field whose magnetic induction vector
at a given point is B0, then in a homogeneous isotropic
m .4.3 M agnetic F ield of a Steady Current 293

medium filling the entire space that contains a magnetic


Held, a magnetic field of induction B will be produced
at the same point: •
B = pB0
where p is the relative permeability of the medium. In any
medium p ^ T, and it shows how many times the magnetic
induction at the point being considered of a homogeneous
isotropic medium filling the entire field with preset cur­
rents producing the magnetic field is greater or smaller
than in a vacuum (111.6.2.1°).
3°. The direction of the magnetic induction vector of
a field set up by a current-carrying conductor is determined
by the right-hand screw rule (cf. 111.4.1.4°): if the direc­
tion of motion of an auger (or screw) with a right-hand
thread coincides with that of the current in a conductor,
then the direction of the magnetic induction vector coin­
cides with that of rotation of the handle of the auger (or
the head of the screw) (Fig. 111.4.10).
4°. A very long straight conductor (an infinite conduc­
tor*) carrying the current I sets up a magnetic field with
the induction B in a given medium at a distance R from
the conductor. Its magnitude is

where I = current in the conductor


p0 = magnetic constant in the SI system (VII.5.3°)
p = relative permeability of the medium (Par. 2°).
Figure III.4.11 shows the magnetic field lines of such
a field and the directions of the vectors B at various points
of the field. The field lines are concentric circles in planes
at right angles to the current-carrying conductor.
5°. A conductor in the form of a circular loop of radius
R carrying the current / produces a magnetic field whose
induction at the centre of the loop has the magnitude

B= w - w
* In practice, a straight conductor is assumed to be infinitely long
if it can be considered that the distance from its ends t© the point where
the magnetic induction of the field is being sought is much greater
than R.
294 F undam entals of E lectrodynam ics 111.4.3

where j l l 0 = magnetic constant in the SI system


|i = relative permeability of the medium (Par. 2°).
Figure III.4.12 shows the arrangement of the magnetic
field lines and the direction of the vectors B at various
points of the field in a section at right angles to the loop.

Fig. III.4.10 Fig. III.4.11

Any current-carrying loop produces a magnetic field whose


magnetic induction B is directly proportional to the mag­
netic moment p m of the loop (111.4.1.4°) and the relative
permeability of the medium.
The dependence of B on the geometrical shape of the
loop and the distance from it to the point where the indue-

tion of the magnetic field is being determined has a com­


plicated nature in the general case and is not considered
in elementary physics.
6°. A very simple single-layer solenoid is a long wire
wound in a tight helix (many circular loops, Fig. 111.4.13).
If the length of a solenoid I R, where R is the radius
in .4.3 M a g n e tic F ield of a S tea d y C urrent 295

of a loop, then the solenoid has the magnetic field shown


in Fig. III.4.13. The field inside a solenoid far from its
ends is homogeneous (111.4.1.5°). The magnetic field out­
side a solenoid is similar to that of a permanent bar magnet
(111.6.5.6°) (Fig. III.4.14). The end of a solenoid from
which the field lines emerge is similar to the north pole
N of a magnet. The other end of the solenoid entered by
the field lines is similar to the south pole S of a magnet.

The location of the poles in a solenoid is determined ac­


cording to the known direction of the current in the turns
with the aid of the right-hand screw rule. If a screw with
a right-hand thread is turned in the direction of the current,
then the motion of its point will indicate the direction
of the magnetic field lines (cf. Par. 3°). The north pole
ol* a solenoid will be the end toward which one must look
lor the current in the turns to flow counterclockwise. The
opposite end of the solenoid corresponds to the south pole
(Fig. III.4.15).
7°. For a sufficiently long solenoid (Par. 6°) of N turns
and of length Z, the magnetic field induction at points
on its axis remote enough from the ends of the solenoid
is
B = p,0|Lira/

where / is the current flowing through a turn, and n — N /l


is the number of turns per unit length of the solenoid.
For the remaining symbols see Paragraphs 4° and 5°.
296 F u n d a m e n ta ls of E lectro d yn a m ics III A A

A solenoid containing a ferromagnetic bar (111.6.5.1°)


is called an electromagnet.
Problem 1. What is the induction of the magnetic field
at the centre of a circular loop with a radius of 0.1 m if
the magnetic moment of the loop is 0.2 A-m2?
Given: R = 0.1 m, p m = 0.2 A-m2.
Required: R.
Solution: The induction of a magnetic field at the centre
of a circular current is R = pp0//2/?, where p0 is the mag­
netic constant in the SI system, p0 = 1.26 X 10~6 H/m.
The magnetic moment of a current-carrying loop is
Pm = IS = InR 2, whence I = p m/nR2. Hence,

Pm 0.2
B = pp0 2nR3 1 X 1.26 X 10"6 X 2X3.14X10"3 4 X 10"5T

Problem 2. At a current of 0.5 A, the induction of the


magnetic field on the axis of a sufficiently long solenoid
is 3.15 X 10“3 T. Find the diameter of the wire which
the single-layer solenoid winding is made of. The winding
turns tightly fit against one another. The solenoid has
no core.
Given: B = 3.15 X lO"3 T, / = 0.5 A, p = 1,
p0 = 1-26 X 10"6 H/m.
Required: d.
Solution: The induction of the magnetic field on the
axis of a long solenoid is R = pp0n /, where n is the num­
ber of turns per unit length.
If the turns tightly fit against one another, then d = 1In,
or
, PPq/ _ 1 X 1.26 X 10“6x 0.5
a ~~ D ~ 3 .1 5 X 1 0 -3
= 0.2 x 10"3 m

4. Interaction of P arallel Currents


1°. A force of interaction appears between two parallel
infinitely long conductors (111.4.3.4°) carrying steady cur­
rents. Conductors with identically directed currents are
attracted, and with oppositely directed ones are repelled
/ f 1 . 4 .5 M a g n e tic F ield of a S te a d y C urrent 297

Irom each other. Such interaction of conductors with par­


allel currents is explained by the left-hand rule (111.4.2.2°).
If a conductor carrying #
(lie current I 2 is in a magnet­
ic* field of induction Bx pro­
duced by another conductor
carrying the current / x, then
Ihe Ampere force F is directed
as shown in Fig. III.4.16.
2°. Two unit lengths of par-
aIlei infinite conductors at
a distance of R from each
other and carrying currents of
/, and 12* attract each other
when the directions of the
currents are the same, and
repel each other when the directions are opposite. The mag­
nitude of the force of interaction F is

where p0 = magnetic constant in the SI system (VII.5.3°)


[i — relative permeability of the medium containing
the current-carrying conductors (111.4.3.2°).

5. Action of a Magnetic Field


on a Moving Charge. Lorentz Force
1°. An electric charge moving in a magnetic field is
n c l e d upon by the Lorentz force whose magnitude is
F l = qvB sin a = qv±B
where q — absolute value of the moving charge (q > 0
for a positive charge, and q < 0 for a negative
one)
v — velocity of the charge
B = magnitude of the magnetic field induction
a = angle between the vectors B and v (Fig. III.4.17).
* It should be remembered that the direction of a current is taken
•in l lie direction in which the positive charges move (111.2.1.2°).
298 F u n d a m en ta ls of E lectro d yn a m ics 111.4.5

The magnitude of the Lorentz force is determined by


the component of the speed perpendicular to the vector B,
i.e. u± = v sin a. The direction of the Lorentz force acting
on a positive charge is deter­
mined according to the left-
hand rule (111.4.2.2°). The Lo­
rentz force acts in the opposite
direction on a negative charge
moving with the same veloc­
ity and in a similar mag­
netic field.
The Lorentz force makes
it possible to introduce the
concept of the magnetic induc­
tion vector (see also 111.4.1.6°
and 111.4.2.4°): the magnitude
of the magnetic induction
vector at a given point of a magnetic field equals the max­
imum Lorentz force Fht max acting on a unit positive
charge moving with a unit velocity:
B -- F j j ' max
qv
When a = jx/2 radians, we have FL = FL, max-
2°. In a homogeneous magnetic field whose induction
vector is perpendicular to the direction of the velocity
of a charged particle, the Lorentz force curves the trajec­
tory of motion. The particle moves along a circle of con­
stant radius R in a plane at right angles to the vector B:

where m = mass of the particle


q = absolute value of its charge
v = velocity of the particle
B = magnetic field induction.
The sign of the particle charge determines the direction
of its deviation in a magnetic field whose induction lines
are perpendicular to the plane of the drawing (Fig. III.4.18).
The Lorentz force in this case is a centripetal one (1.2.4.5°)
and does no work when the particle moves along the circle.
If 1.4.5 M a g n e tic F ie ld of a S te a d y C urrent 299

3°. The time T needed for a charged particle to complete


one revolution around the circle (its period) in a homo­
geneous magnetic field is
rp 2n rn
~~~b ~ T

(All the symbols are given in Par. 2°.) The time T is in­
dependent of both the radius of the circle and the veloc­
ity of the particle [at a velocity of v < c, where c is the
speed of light in a vacuum, when the dependence of the
mass on the velocity does not tell (V.4.10.30)]. The opera­
tion of a cyclic accelerator of
charged particles—a cyclotron
(VI.4.16.5°) is based on this fact.
4°. A charge moving simulta­
neously in an electric and a mag­
netic fields experiences a force also
called the Lorentz force (generalized
Lorentz force). Its magnitude is
Fl = qE -f qvB sin a
where E is the magnitude of the
electric field intensity vector
(111.1.3.3°). The direction of the
generalized Lorentz force, according
to the rule of addition of forces (1.2.2.6°), depends on
the directions in which the two forces forming the gen­
eralized Lorentz force act.
Problem 1. An alpha particle having a velocity of
106 m/s flies into a homogeneous magnetic field whose
induction is 0.3 T. The velocity of the particle is perpen­
dicular to the direction of the magnetic field lines. Find
Ihe radius of the circle along which the particle will move
and its period of revolution.
Given: v = 106 m/s, B = 0.3 T, a = 90°, q = 2e =
- 3.2 X 10-19 C, m = 6.64 X 10"27 kg.
Bequired: r, T.
Solution: In a magnetic field, the Lorentz force acts
on a moving charge. Since v _L B, this force will be a centri­
petal one: FL = Fc and, consequently, qvB = mv2lr,
300 Fundam entals of Electrodynam ics I1I.4.5

whence the radius of the circle


m v _ 6.64 XlO"27 w 106
r ~ q B 3 . 2 X lO " 19 X 0.3
7 X 10"2 m

The period of revolution is


2nr _ 2 X 3.14 X 7 X 10"2
v ~ 106
4 X 10”7 S

Problem 2. An electron travels with a velocity of


107 m/s parallel to a long straight conductor carrying
a current at a distance of 2 mm from it. With what force
will the magnetic field of the current act on the electron
if a current of 10 A flows through the conductor?
Given: v = 107 m/s, I = 10 A, e = 1.6 X 10"19 C,
d = 2 X 10"3 m, p = 1, p,0 = 1.26 X 10“6 H/m, a = 90°.
Required: Fl-
Solution: The magnitude of the Lorentz force acting
on an electron moving in a magnetic field is Fl = evB sin a.
The magnetic field induction of a steady current is
B = pp0//2jid.
Hence,
FL = = 1.6 X 10"19 x 107 x 1.26 x 10~6 x
10
X 2 X 3.14 X 2 X IQ"2 = 10“1B N

Problem 3. Find the kinetic energy of an electron


travelling along the arc of a circle with a radius of 8 cm
in a homogeneous magnetic field whose induction is 0.2 T.
The magnetic field induction is directed at right angles
to the plane of the circle.
Given: r = 8 cm = 8 X 10"2 m, B = 0.2 T, m = 9.1 X
X 10-31 kg, e = 1.6 X 10-19 C.
Required: Ek.
Solution: The kinetic energy of an electron is Ek =
= mv2/2. An electron in a magnetic field experiences the
Lorentz force. Its magnitude is FL = evB because v 1 B;
the force F l is a centripetal one: evB = mv2lr. Hence,
the velocity v = eBrlm. Therefore,
e 2B 2r 2 _ (1.6 X 10"19)2 X 1 0 '2 X 22 X 82 X 10”4
4 pJ
2m ~ 2 X 9.1 X IQ"*1
ni.4.6 Magnetic Field of a Steady Current 301

Problem 4. A beam of singly charged neon ions flies


into a homogeneous magnetic field at right angles to the
field lines after passing through an accelerating potential
difference of 10 kV in an electric field. The ions travel
in the magnetic field along two arcs of circles whose radii
are 14.6 and 15.3 cm. Find the mass numbers of the neon
isotopes (VI .4.1.2°).
Given: (px — <p2 — 10 kV, B = 0.14 T, rx = 14.6 cm,
r2 = 15.3 cm, 1 amu (atomic mass unit) = 1.66 X 10~27 kg,
q = 1.6 X 10-19C.
Required: M x, M 2.
Solution: The kinetic energy acquired by an ion in an
accelerating electric field with the potential difference
cpi — (p2 is mv2!2 = g ((pi — <p2), whence
_ (qpx—<p2)
~ i?* (1 )

According to the conditions of the problem, F l = Fc,


i.e. qvB = mv2/r, whence v2 = q2B2r2lm2. Substitution in
(1) yields m = r2B2q/2 ((px — <p2).
The mass numbers of the isotopes expressed in atomic
mass units (VII.7.1°) will be
0.1462 X 0.142 X 1.6 X 10“19
Mi 20 amu
2 X 10s X 1.66 X 10-27
0.1532 X 0.142 x 1.6 X 10~19
M 2= 22 amu
2 X 103 X 1.66 X 10“27

6. Specific Charge of Particles


1°. The specific charge of a particle is the ratio of its
charge q to its mass m, i.e. q!m. The determination of
the specific charge is based on the deviation of charged
particles in a magnetic and an electric fields. To measure
qlm, we must know the velocity of a particle v and the
radius of its trajectory R in a magnetic field (111.4.5.2°).
The velocity v is created by the electric field with a known
potential difference and is determined by the formula
given in 111.1.8.5°. The radius R is determined experimental­
ly. For the specific charges of an electron and a proton,
see VII.8.
302 F u n d a m en ta ls of E le c tro d y n a m ic s H i.5 .1

2°. The joint action of an electric and a magnetic fields


on particles is used to find the specific charges and
masses of positive ions. The instruments used to perform ac­
curate measurements of the relative atomic masses (VI 1.4.3°)
of isotopes of chemical elements (VI.4.1.2°) are called
mass-spectrographs, or mass-spectrometers. The particles are
classified by masses in these instruments in accordance
with the mass spectrum—the combination of the values
of the masses of the given particles. All these instruments
operate on the principle that all particles having a defi­
nite value of the specific charge qlm1 regardless of their
velocities will be focussed and separated from the particles
having other values of qlm2, q/m3l etc. This is achieved
by deviating them in properly selected electric and mag­
netic fields.

Chapter 5
Electromagnetic Induction

1. Phenomenon and Law


of Electromagnetic Induction
1°. If a conducting loop is in a varying magnetic field
(111.4.1.1°), then an induced electric field (111.5.3.1°)
appears in the loop and is characterized by the induced
e.m.f. In these conditions, an electric current called an
induced one appears in a conducting loop, and the entire
phenomenon is called electromagnetic induction.
This phenomenon was discovered by M. Faraday
experimentally. Faraday's experiments were based on the
idea of the close relationship between electrical and mag­
netic phenomena. If a magnetic field appears around con­
ductors carrying a current, then there should also be the
reverse phenomenon—the appearance of an electric current
in a loop under the action of a magnetic field. Faraday
i i i.5 .i Electromagnetic Induction 303

showed by a series of experiments that an induced current


is produced in conducting loops in a varying magnetic
held regardless of how the change of the magnetic flux
(111.4.1.8°) through the area of the surface confined by
the loop is achieved in time. The magnetic flux through
the area of a loop may change with time owing to defor­
mation or motion of the loop in an external magnetic held,
and also because the magnetic held induction may vary.
2. Faraday's law of electromagnetic induction: the e.m.f.
of electromagnetic induction ^ in a loop is numerically
equal and opposite in sign to the rate of change in the
magnetic flux through the area confined by this loop:
ACD

The induced current in a conducting loop of resistance


R (111.2.4.1°) is

The e.m.f. is considered to be positive if the magnetic

moment p m (111.4.1.4°) of the induced current / t in the


loop corresponding to it forms an acute angle with the
lines of the held producing this current. Figure III.5.1
shows a positive (a) and a negative (b) e.m .f.’s
. If a closed circuit consists of N series-connected turns
[for instance in a solenoid (111.4.3.6°)], thenifi is determined
304 Fundamentals of Electrodynamics 111.5.2

by the change of the magnetic flux through the areas con­


fined by all the turns in a unit time.
The quantity of electricity Aq flowing through a cir­
cuit of resistance R in the phenomenon of electromagnetic
induction is

3°. The minus sign in the law of induced e.m.f. expres­


ses Lenz's law: the direction of the induced current in a
loop is always such that the magnetic flux of the field pro­
duced by this current through the area confined by the
loop will reduce the changes in the field that set up the
induced current. Figure III.5.2. shows the directions of
the induced currents in a coil produced by the movement
of a bar magnet in accordance with Lenz’s law.

2., Induced E.M.F.’s in Moving Conductors


1°. When a conductor of length I travels with the veloc
ity i; in a stationary (111.4.1.1°) homogeneous (111.4.1.5°
magnetic field, the induced e.m.f. in the conductor is
jfi = Blv sin a = B lv±
where B = magnitude of the magnetic induction vector
a = angle between the vectors v and B.
A glance at the formula shows that is proportional
to the component of the velocity vector perpendicular to
the field: v± = v sin a.
When a conductor moves in a magnetic field, the Lo-
rentz force acts on its positive and negative charges. This
force causes the charges to separate in the conductor: the
positive and negative charges accumulate on opposite
ends of the conductor (Fig. III.5.3). These charges produce
a Coulomb field (111.1.3.1°) inside the length of the con­
ductor. The charges will move under the action of the
Lorentz force until the force acting on a charge in the Cou­
lomb field balances the Lorentz force FL = qBv± (111.4.5.1°).
III.5.2 Electromagnetic Induction 305

The induced e.m.f. in a length of a conductor is the


work done for moving a unit positive charge along the
conductor. •
2°. The action of the Lorentz forces on the charges of
a conductor is similar to that of an electric field directed
oppositely to the Coulomb field. This field is created not by
Coulomb forces, but by forces of a magnetic origin—Lorentz

Fig. III.5.4

forces. For this reason, an electric field whose characteristic


is the induced e.m.f. is an extraneous electric field (111.2.2.4°)
(i.e. a field produced by non-electromagnetic processes).
By definition, the magnitude of the intensity of this extra­
neous electric field (111.2.2.4°) is
Eex = = Bv i

where FL = qBv± is the magnitude of the Lorentz force.


3°. The direction of the intensity of the induced extra­
neous electric field in a straight conductor moving in a
magnetic field is determined by the right-hand rule: if
the palm of the right hand is arranged so that the mag­
netic induction vector B enters it, and the thumb made
to form an angle of 90 degrees with the other fingers coin­
cides with the direction of the conductor velocity compo­
nent at right angles to the conductor, then the stretched-
out four fingers will show the direction of the induced
extraneous electric field intensity set up in the conduc­
tor (Fig. III.5.4).
20-0211
306 Fundamentals of Electrodynamics 111.5.2

4°. In a flat rectangular loop rotating in a homogeneous


magnetic field with the angular velocity co (1.1.9.4°) so
that its axis of rotation is in the plane of the loop and
is at right angles to the magnetic induction vector B of
the external field, the induced e.m.f. is

<£{ = BS co sin (Dt

where S is the area of the loop.


Problem 1. The span of the wings of an airplane is
15 m. Its horizontal velocity is 830 km/h. Find the po­
tential difference appearing between the tips of the wings.
The vertical component of the Earth’s magnetic field in­
duction is 50 pT.
Given: Z = 15m ,y = 830 km/h = 230 m/s, B = 50 pT =
= 5.0 X 10-5 T.
Required: (cp2 — cp2).
Solution: When a conductor moves at right angles to
the magnetic field lines, the potential difference across
the ends of the conductor is

q)i —cp2= SZy = 5.0 X 10 6 x 15 x 230 « 0.17 V

Problem 2. A loop 400 cm2 in area having 100 turns


rotates in a homogeneous magnetic field having an induction
of 0.01 T. The period of revolution of the loop is 0.1 s.
Find the maximum value of the induced e.m.f. in the loop.
The axis of rotation is perpendicular to the magnetic field
lines.
Given: S = 400 cm2 = 0.04 m2, N = 100, B = 0.01 T,
T = 0.1 s.
Required: max.
Solution: When a loop rotates in a magnetic field, the
induced e.m.f. is = (a>BS sin cot) N.
It is obvious that

» 1, max = <»BSN = ^ BSN =


2X3.14
0.1 X 0.01 x 0.04 x 100 2 V
111.5.3 Electromagnetic Induction 307

3. Induced Electric Field


1°. If a bar magnet approaches a stationary conducting
loop with the velocity y, then an induced current I i appears
in the loop (Fig. III.5.5). The cause of the ordered motion
of the charges is the induced electric field in which forces
(111.5.2.1°) act on the
positive and negative
charges of the conduc­
tion loop.
2°. Properties of an
induced electric field:
(a) An induced elec­
tric field is not a Cou­
lomb field. It is due to
a varying magnetic field
instead of to charges dis­
tributed in space.
(b) An induced electric field, similar to a magnetic
one, is a vortex field (111.4.1.7°). Its field lines close on
themselves, and do not begin and terminate on charges
as in a Coulomb field.
(c) An induced electric field is a non-potential one.
The work done in this field upon thejnotion of a unit pos­
itive charge along a closed circuit does not equal zero
and is the induced e.m.f. in a conducting loop in a varying
magnetic field.
3°. A varying magnetic field produces an induced vortex
electric field. This fundamental law of electrodynamics
was established by J. Maxwell as a generalization of Far­
aday’s law of electromagnetic induction (111.5.1.2°).
A vortex electric field is detected by its action on the free
charges of an electric conducting loop placed in this field.
The direction of the intensity vector of a vortex electric
Held is established according to Faraday’s law of electro­
magnetic induction (111.5.1.2°) and Lenz’s law (111.5.1.3°)
(see also IV.4.1.20).
A vortex electric field with closed field lines is used
lor accelerating electrons in a betatron (VI.4.16.3°).
308 Fundamentals of Electrodynamics 1II.5A

4. Induced Currents in Solid Conductors


1°. Induced currents appearing in solid conductors are
called eddy currents, or Foucault currents. Many closed
lines of such currents appear inside solid conductors. Eddy
currents liberate a great amount of heat in a unit time, and
are directly proportional to the square of the frequency
of variation of the magnetic field. High-frequency alternat­
ing currents are used in induction furnaces for heating.
2°. The liberation of heat in many kinds of electrical
equipment due to eddy currents results in losses of energy.
To reduce them, the cores of transformers (111.5.6.3°), the
magnetic circuits of electrical machines, and other pieces
of equipment are made of separate insulated plates whose
surfaces are arranged parallel to the magnetic field lines
instead of being made solid. Eddy currents are formed in
the planes perpendicular to the magnetic field lines. Such
an arrangement of the plates therefore diminishes the
energy losses due to eddy currents.
3°. The eddy currents appearing in solid conductors
moving in a magnetic field interact with the latter accord­
ing to Lenz’s law (111.5.1.3°), which results inretardation
of the moving conductors. This phenomenon is used for
retarding the movable systems of electrical measuring
instruments (the retarding action of eddy currents).

5. Self-Induction
1°. The setting up of an induced field in a circuit as
a result of a change in the current in the circuit is called
self-induction. The change in the current causes its own
magnetic field to change. The phenomenon of electro­
magnetic induction is set up in a current-carrying conductor
whose own magnetic field is varying. This phenomenon
is characterized by the e.m.f. of self-induction.
2°. The internal magnetic field of the current^in^a loop
sets up the magnetic flux Os through the area of the sur­
face confined by the loop itself (111.4.1.8°). The magnetic
flux Os is called the flux of self-induction of the loop. If
the latter is not in a ferro-magnetic medium (111.6.5.1°),
i n . 5.a Electromagnetic Induction 309

then is proportional to the current I in the loop: Os =


= LI. The quantity L is called the inductance of the loop
and is an electrical characterise of it similar to the re­
sistance R of the loop and other characteristics.
The inductance of a loop numerically equals the mag­
netic flux of self-induction of the loop at a current equal
to unity. The value of L depends on the dimensions of
the loop, its geometrical shape, and the relative permea­
bility (111.4.3.2°) of the medium containing the loop. For
example, for a sufficiently long solenoid (111.4.3.6°, 7°)
of length I and turn cross-sectional area S with a total
of N turns, the inductance is

L= = nofm 2F

where |m0 = magnetic constant in the SI system (VII.5.3°)


\x = relative permeability of the medium
n1= N /l = number of turns per unit length
fV = SI = volume of the solenoid.
3°. According to Faraday’s law of electromagnetic
induction (111.5.1.2°), the e.m.f. of self-induction Sfsi is
<£ _
- w
If a current-carrying loop is not deformed and the re­
lative permeability of the medium is constant, then the
inductance of the loop is constant. Consequently, £ sl is
proportional only to the rate of change of the current:

The inductance of a loop numerically equals the e.m.f.


of self-induction in it when the current in the loop changes
by unity in one second (| ^si | = L when A//AJ = 1).
4°. An induced current 7S appears in a loop under the
action of J?st. According to Lenz’s law (111.5.1.3°), it counter­
acts the'* change in the current in the circuit that gave
rise to self-induction. The current of self-induction im­
posed onto the main current retards its growth or pre­
vents its reduction. According to the formula in Par. 3°,
310 Fundamentals of Electrodynamics 111.5.6

$sl and, consequently, 7S, other conditions being equal,


are directly proportional to the inductance of the loop.
The inductance of a loop is a measure of its “inertia”
with respect to a change in the current in the loop. In
this sense, the inductance L of a loop in electrodynamics
plays the same part as the mass m of a body in mechanics
(1.2.3.1°).
5°. The increase (decrease) in a current with time
when the circuit is closed (open) has the form shown in

Fig. i l l . 5.6. The nature of the curves is explained by the


part of self-induction when a circuit is closed or open
(Par. 4°).

6. Mutual Induction. Transformer


1°. The phenomenon of mutual induction consists in the
appearance of an induced field (111.5.3.1°) in conductors
that are near other conductors with time-varying currents.
Thus, if the current 7X in the loop 1 changes, then in the
loop 2, which contains no current source, an induced field
will be set up characterized by the e.m.f. of mutual in­
duction 2i- An induced current is produced that is de­
tected by the galvanometer G (Fig. III.5.7).
2°. According to Faraday’s law of electromagnetic in­
duction (111.5.1.2°), we have
A<D21
1IL5.6 Electromagnetic Induction 311

where Ozl is the magnetic induction flux produced by the


magnetic field of the current I x and passing through the
area of the surface confined b y the loop 2. The magnetic
flux 0 21 is proportional to the current I x in the loop 1:
^21 = L 2i l \
where L 21 is the mutual inductance of the second and first
loops. It depends on*the dimensions, geometrical shape,

and mutual arrangement of the loops 1 and 2 and, in ad­


dition, on the relative permeability of the medium containing
the loops.
3°. The functioning of a transformer used to step up
or down the voltage of an alternating current is based on
mutual induction. Two coils
or windings, the primary Wx
and the secondary W2 with N x
and N 2 turns, respectively,
are wound on a closed core
(111.6.5.6°) consisting of sepa­
rate plates (Fig. III.5.8). The
alternating current I x pro­
duces a varying magnetic field
in the primary, and it is
exactly this field that produces an e.m.f. of mutual in­
duction in the secondary.
When no load is connected to the secondary of a trans­
former (I 2 = 0), the ratio of the absolute values of the
voltages U2 and across the ends of the secondary and
primary windings is called the voltage ratio:

For a step-up transformer, N 2 > N ly and for a step-down


one, N 2 < lN x.
312 Fundamentals of Electrodynamics III.5.7

7. Energy of a Magnetic Field


1°. To produce the current I in a loop of inductance L ,
work must be done to overcome the e.m.f. of self-induction.
The intrinsic, or proper, energy Wm of the current I is the
quantity numerically equal to this work. If the medium
containing the loop is not ferromagnetic (III.6.5.1°), then

2°. In accordance with the theory of short-range force


action (111.1.3.2°), the intrinsic energy of a current is
concentrated in the magnetic field produced by the current-
carrying conductor. This is why we speak of the energy
of a magnetic field and consider that the intrinsic energy
of a current is distributed over the entire space containing
a magnetic field. The energy of the latter equals the in­
trinsic energy of the current. In a long solenoid (111.4.3.6°),
the energy of the magnetic field is concentrated mainly
in the volume V of the solenoid and is

Wm= — n0^ I W

where n = number of turns per unit length of the solenoid


Ix0 = magnetic constant in the SI system (VII.5.3°)
\i = relative permeability of the medium.
3°. The energy of a homogeneous magnetic field con­
centrated in the volume V of an isotropic and non-ferro­
magnetic medium is
B2
Wm V
2 H 'oH'

where B is the magnetic field induction (111.4.1.6°). The


remaining notation is explained in Par. 2°.
4°. The volume density of the energy of a magnetic field
wm is defined as the energy in a unit volume of the field:
III.5.7 Electromagnetic Induction 313

For a magnetic field in an isotropic and non-ferromagnetic


medium, we have
1 &
2 lioH

This expression holds not only for a homogeneous field, but


also for arbitrary ones including time-varying magnetic
fields.
5°. If an electric and a magnetic fields exist simul­
taneously, then the volume density of the energy of the electro­
magnetic field w in an isotropic medium having neither
ferromagnetic (111.6.5.1°) nor ferroelectric (111.1.6.9°) pro­
perties equals the sum of the volume densities of the energies
of the electric (111.1.12.7°) and magnetic fields:
e 0e E 2 1 B*
w = w0+ wm 2 p0[i
~ ~

Problem. Determine the inductance of a coil with a


non-ferromagnetic core having 800 turns. The length of
the coil is 0.25 m, and the diameter of the turns is 4 cm.
A current of 1 AJflows through the coil. What is the magnetic
flux through the*cross section of the coil? What is the energy
of the coil’s magnetic field?
Given: I = 0.25 m, d = 4 cm = 0.04 m, N = 800,
I = 1 A, p = 1, p 0 = 1.26 X 10- 6 H/m.
Required: ®, W.
Solution: The inductance of a coil with a non-ferro­
magnetic core is
r /v2 c , Ofiw 8002 3.14X 0.042 , „
L = ii0— S = 1.26 X 10 6 X -q-25~X------ 4------ « 4 mH
The magnetic flux is

(D = ^ - = 4X18°w8 x 1 « 5 x 10~6 Wl,

The energy of the magnetic field is


x,7 ZJ2 4 x to -2 x l 2
W m —
2 2
2 x 10~3 J
314 Fundamentals of Electrodynamics III.6.1

Chapter 6
M agnetic Properties
of a Substance

1. Magnetic Moments of Electrons


and Atoms. Electron Spin
1°. Every electron travelling in an atom about the nu­
cleus in a closed orbit* (VI.2.2.1°) is an electron current
flowing in a direction opposite to that of the electron. Fi­
gure 111.6.1 shows the direction of the velocity vector
v of an electron and the direction of the current flowing in
the orbit. The electron current I (111.2.1.3°) is

where e = absolute value of the charge of an electron


T = period of revolution of an electron in its orbit.
2°. The magnetic moment pm of the electric current pro­
duced by the motion of an electron in orbit, by definition
(111.4.1.4°), is
Pm — I S n 0

where S = area of the electron’s orbit


n0 = unit vector of the normal setting the direction
of the vector pm (111.4.1.4°) (see Fig. III.6.1).
The magnetic moment pm of an electron current is
called the orbital magnetic moment of the electron.
3°. The vector of the orbital magnetic moment of an atom
Pm is defined as the vector sum of the orbital magnetic mo­
ments of all its Z electrons:
Pm = P m > + P m ) +
where Z is the atomic number of the atom in Mendeleev’s
periodic table (VI.2.9.1°) equal to the total number of
* For a more precise explanation of the concept of an electron
orbit, see V I.1.6.2° and VI.2.6.3°.
III.6.1 Magnetic Properties of a Substance 315

electrons in the atom. If a substance consists of molecules,


then the magnetic moment of a molecule is the vector sum
of the orbital magnetic moments of its atoms.
4°. Every atom or molecule having a magnetic moment
can be considered similar to a current flowing through
a loop confining the surface area S (an atomic or molecular
current). A molecular current, like any current-carrying loop,
has a magnetic moment and produces a magnetic (field

Fig. III.6.1 Fig. III.6.2


(111.4.3.5°). According to Ampere's hypothesis, the magnetic
properties of substances are determined by molecular cur­
rents.
5°. An electron, regardless of the system of particles it
is in (an atom, molecule, crystal), has the intrinsic mechani­
cal angular momentum Ls (1.3.2.2°) called spin. The elemen­
tary model notion of spin is associated with rotation of an
electron about its own axis. This notion, however, contra­
dicts the special theory of relativity (VI.2.8.3°). It is proved
in modern physics that spin is just as important a characte­
ristic of an electron as its electric charge e and rest mass
m0 (V.4.10.4°) (see also VI.2.8.1°).
6°. A very important feature of the spin of an electron
is that in a magnetic field* the spin can be oriented so that
its projection on the direction of the magnetic field induc­
tion vector B will take on only two values (Fig. III.6.2):
h_ h
L sb — zb 2 4tc
* This field can be set up either by current-carrying conductors
(an “external field”) or by electron, and also atomic and molecular cur­
rents (an “internal field”).
316 Fundamentals of Electro dynamic 111.62

where h is the Planck constant (V.3.2.3°), and h = h/2n


(see also VI.2.8.2°).
If a system of electrons (an atom, crystal) has an even num­
ber of them, then the spins of each pair of electrons direc­
ted oppositely produce a summary spin equal to zero.
Such a system is called compensated with respect to spin.
A system with an odd number of electrons has an uncompen­
sated spin different from zero.
The presence of spin in an electron and certain other ele­
mentary particles (VI.5.1.1°) explains many important laws
in modern physics. For example, the spin of an electron ex­
plains the magnetic properties of ferromagnetics (111.6.5.8°).
It determines the distribution of electrons by energy states
and, in this connection, by shells in *atoms [(VI.2.8.6°).

2. Classification of Magnetic Substances


1°. All substances capable of becoming magnetized in an
external magnetic field, i.e. of producing an internal
magnetic field of the substance itself are called magne­
tic substances, or magnetics. The latter are divided with res­
pect to their magnetic properties into weakly magnetic and
strongly magnetic substances. The weakly magnetic substan­
ces include paramagnetics and diamagnetics. The basic
strongly magnetic substances are ferromagnetics. Weakly
and strongly magnetic substances differ in the magnitude of
their relative permeability p (111.4.3.2°). For weakly magne­
tic substances, p differs insignificantly from unity: for para-
magnetics p ^ 1, for diamagnetics p 1. In addition, for
weakly magnetic substances, p does not depend on the induc­
tion B 0 of the magnetic field in which the substances are
magnetized. For strongly magnetic substances, p 1 and
depends on B 0 (111.6.5.3°).*
2°. Paramagnetics include oxygen, nitrogen oxide, alu­
minium, platinum, the rare-earth elements, alkali and
alkaline-earth metals, and other substances.

* For ferromagnetics, the'relativ e permeability'is defined in’a way


that differs fronfthe one in 111.4.3.2°. This, however, is beyond the
scope of elementary physics.
111.6.3 Magnetic Properties of a Substance 317

The permeability p, for paramagnetics depends on the


temperature and diminishes with its elevation according to
the law p = 1 + CIT, where T is the absolute temperature,
and C is the Curie constant characteristic of a given sub­
stance.
3°. Diamagnetics include the inert gases (helium, argon,
etc.), many metals (gold, zinc, copper, mercury, silver),
water, glass, marble, and many organic compounds. For
these substances, the relative permeability is independent
of the temperature.
4°. Ferromagnetics include a comparatively small group
of solid crystalline bodies—the so-called transition metals
(iron, nickel, cobalt), and a number of alloys. For the mag­
netic properties of ferromagnetics, see III.6.5.10-7°.

3. Diamagnetism
1°. Diamagnetics are defined as substances in which the
atoms or molecules have no magnetic moments (111.6.1.3°)
in the absence of an external magnetic field. The atoms of
such substances are called diamagnetic atoms. An example is
the helium atom. The nucleus of this atom has the charge
q = -\-2e (VI.2.1.1°), where e is the absolute value of the
charge of an electron. Let us assume that both electrons of
a helium atom are travelling about the nucleus with an
identical velocity in identical orbits (VI.2.2.1°), but in
opposite directions (Fig. III.6.3). Hence, their orbital mag­
netic moments (111.6.1.2°) will be equal in magnitude, but
opposite in sign, and the summary magnetic moment of an
atom Pm = p< m) + p(m* will equal zero.
2°. When a diamagnetic substance is brought into a mag­
netic field, an additional atomic (or molecular) induced cur­
rent 1 1 (111.6.1.4°) of magnetic moment APm>1 is set up
in each of its atoms (or molecules). The vector APm,i is
directed oppositely to the magnetic induction vector B0
of the external magnetic field (Fig. III.6.4). The vector
APm,i and the induced current / 4, according to Lenz’s law
(111.5.1.3°), should be directed so that the magnetic field
produced by the induced currents is opposite to the
318 Fundamentals of Electrodynamics tll.6.4

extenral magnetizing field. The total magnetic field set up


by the currents induced in all the atoms (molecules) is an
internal magnetic field. The magnetic induction vector of
an internal field is directed oppositely to the induction vec­
tor of the external magnetizing field. This is the essence of
the magnetization of a diamagnetic substance.
If the action of the magnetizing field stops, then the in­
duced currents vanish in the atoms (molecules), and the dia­
magnetic properties also vanish (demagnetization of a dia­

magnetic). The appearance of induced currents in atoms


(molecules) is not affected by the chaotic thermal motion of
the atoms (molecules). Hence, the diamagnetic properties
of a substance do not depend on the temperature.
3°. Diamagnetism is a universal property of all substances
because induced currents are set up in the atoms (mole­
cules) of any substances placed in a magnetic field. Diamag­
netism, however, is a very small effect. Diamagnetic pro­
perties are therefore observed only in substances in which these
properties are the only ones and are not masked by other
stronger magnetic properties.

4. Paramagnetism
1°. Atoms (or molecules) having a certain magnetic mo­
ment Pm (111.6.1.3°) are called paramagnetic, and the sub­
stances consisting of them —paramagnetics. The magnetic
moments of the atoms (molecules) of a paramagnetic depend
on the structure of the atoms (molecules), are constant for
III.6.4 Magnetic Properties of a Substance 319

a given substance, and do not depend on the external mag­


netic field.
2°. In the absence of a magnetite field, the thermal motion
of the atoms (molecules) of a paramagnetic and their col­
lisions with one another prevent the appearance of an
ordered arrangement of the magnetic moment vectors Pm
of separate atoms (molecules). Consequently, in a paramag­
netic substance in the absence of an external magnetic field,
the atomic (molecular) currents
do not produce a total magnetic Atomic current
field. The substance is not mag­
netized—no internal magnetic
field is produced in it.
3°. When a paramagnetic
is brought into an external homo -
geneous magnetic field, every
atomic (molecular) current tends
to position itself so that its mag­
netic moment vector is orient­
ed parallel to the induction vec­
tor B0of the external field. This is hindered by the thermal mo­
tion of the atoms (molecules). The joint action of the magne­
tic field and the thermal motion results in the orientation of
the magnetic moments of the majority of the atoms (mole­
cules) in the direction of the external magnetic field (Fig.
111.6.5). In a paramagnetic substance, a total magnetic
field of all the atomic (molecular) currents is set up, and
the substance ismagnetized—an internal magnetic field appears
in it. The induction vector of this field and the induction
vector of the external magnetizing field have the same di­
rection.
4°. Elevation of the temperature of a paramagnetic is
attended by an increase in the chaotic thermal motion of
the atoms (molecules). It prevents the orientation of their
magnetic moments and reduces the magnetization of the
substance. This is why the relative permeability of paramag-
netics diminishes when they are heated (111.6.2.2°).
320 Fundamentals of Electrodynamics III.6.5

5. Ferromagnetism
1°- Ferromagnetics are a group of solid crystalline sub­
stances having a combination of magnetic properties due to
the specific interaction of the atomic carriers of magnetism.
In ferromagnetic substances, the internal magnetic field
(111.6.2.1°) has an induction that is hundreds and thousands
of times greater than the induction of the external magne­

tic field producing the magnetization, i.e. the formation of


the internal field.
2°. The quantity M called the magnetization of a sub­
stance* is introduced to characterize the phenomenon of
its magnetization.
The magnetization in the SI system is determined by the
formula
M = B — B0= \xB0— B0= {\i — i) B0
where \i = relative permeability of the substance
B 0 = induction of the magnetic field in a vacuum
B = induction of the magnetic field in the substance:
B = \iB p (111.4.3.2°).
The magnetization M for para- and diamagnetics is
directly proportional to the induction B 0 of the magnetic
field in a vacuum (Fig. III.6.6).
For ferromagnetics, the magnetization M is a complicated
non-linear function of B 0. The dependence of M on the quan-
* T h e quantity o f m a g n e t iz a t io n , w h ic h i s a m e a s u r e o f t h e in t e r ­
n a l fie ld o f a s u b s t a n c e , s h o u ld n o t b e c o n fu s e d w it h t h e phenomenon
o f m a g n e t iz a t io n o f t h e s u b s t a n c e — t h e s e t t i n g u p o f t h is fie ld .
III.6.5 Magnetic Properties of a Substance 321

tity B 0l\i0 is called a technical magnetization curve (Fig.


I l l .6.7). The curve depicts the phenomenon of magnetic
saturation: beginning with a certain value of B J |x0 = B0,J
p0, the magnetization remains virtually constant, equal to
Ms (the magnetization of saturation). The quantity p0 is
the magnetic constant in the SI system (VII.5.3°).
3°. The relative permeability \i of ferromagnetics, un­
like para- and diamagnetics, has very great values and de­
pends on the induction B 0 of the magnetic field in which the
substance is placed (Fig. I l l .6.8). For example for iron
Pmax = 5000, for Permalloy (78% Ni and 22% Fe) |LLmax =
= 100 000 (see the footnote on p. 316).
4°. The magnetic hysteresis of a ferromagnetic is defined
as the lagging of the change in the degree of magnetization

of a ferromagnetic substance behind the change in the exter­


nal magnetic field in which the substance is placed. A very
important cause of magnetic hysteresis in a ferromagnetic is
the dependence of its magnetic characteristics (|i, M) not
only on the state of a substance at the given moment, but
also on the values of the quantities \i and M at previous mo­
ments (i.e. on the past history of the substance). Thus, the
magnetic properties depend on the past history of magnetiza­
tion of a substance.
5°. A hysteresis loop is the loop formed by the curves
showing the change in magnetization of a ferromagnetic
body placed in an external magnetic field whose induction
changes from + 5 0,s/p,0 to — B 0JS/\i0 and back. The value
322 Fundamentals of Electrodynamics 111.6.5

+ 5 0,s/(>t0 corresponds to magnetization of saturation M s


(Par. 2°) (Fig. III.6.9). If the process of magnetization of a
ferromagnetic up to saturation (the point a in Fig. III.6.9)
occurs along the curve Oa, then upon a further reduction of
B 0/\i q the magnetization changes along the curve aMr.
When B 0 = 0, the ferromagnetic retains a certain residual,
or permanent, magnetization MT called the remanence.
This means that the ferromagnetic body has an internal mag­
netic field in the absence of an external one.
To completely demagnetize the ferromagnetic body, the
direction of the external field must be reversed. At a certain
value of the magnetic induction h Q,c, which the quantity
Z?oC/p 0 called the coercive force corresponds to, the magne­
tization M of the body will vanish. Upon a further increase
in the induction of the external magnetic field in the direction
opposite to the original one, the value of the magnetization
again reaches saturation (the point b). The reduction of the
external magnetic field to zero and its further increase to
the value of the quantity B 0l\i0 = Z^g/po leads to a loop
symmetrical with respect to the point 0 —a hysteresis
loop.
6°. The coercive force and the shape of the hysteresis
loop characterize the property of a ferromagnetic to retain
its remanence and determine the use of ferromagnetics for
different purposes. Ferromagnetics with a broad hysteresis
loop are called hard-magnetic materials (carbon, tungsten,
chromium, aluminium-nickel, and other steels). They have
a great coercive force and are used for producing permanent
magnets of various shapes (bar and horseshoe magnets, mag­
netic needles). Soft-magnetic materials with a small coercive
force and narrow hysteresis loop include iron and its alloys
with nickel. These materials are used for making the cores
of transformers, generators, and other equipment whose
operating conditions require magnetic reversal in varying
magnetic fields. The magnetic reversal of a ferromagnetic
is associated with turning of the regions of spontaneous mag­
netization (Par. 8°). The work needed for this is done at the
expense of the energy of the external magnetic
field (111.5.7.2°). The quantity of heat liberated in
magnetic reversal is proportional to the area of the hys­
teresis loop.
Ul.6.5_ _ _ _ _ _ _ _ _ Magnetic Properties of a Substance_ _ _ _ _ _ _ _ _ _ _ _ 3 2 3

7°. The specific properties of a ferromagnetic manifest


themselves only at temperatures lower than a certain so-
called Curie point 0 (). When the* temperature is above this
point, the ferromagnetic properties vanish, and a substance
becomes a paramagnetic (111.6.4.1°). The Curie point 0c
for iron is 770 °C, for nickel is 360 °C, and for Permalloy is
70 °C.
8°. At temperatures below the Curie point, a ferromagnet­
ic body consists of domains—small regions with linear di­
mensions of the order of magnitude of from 10“2 to 10 “3 cm
within which the highest degree of magnetization equal to
M s—the magnetization of saturation (Par. 2°)—is encoun­
tered. Domains are also called regions of spontaneous magne­
tization.
The formation of the domains is explained as follows.
The atoms of the transition metals have uncompleted elec­
tron orbits (VI.2.8.6°) in which the spins of the electrons
are not compensated completely (111.6.1.6°). If in a crys­
tal lattice of such atoms the condition dla^ 1.5 is observed,
where d is the diameter of an atom, and a the diameter of
the uncompleted orbit, a specific interaction appears be­
tween the uncompensated spins. (The quantum-mechanical
nature of this interaction is not considered in an elementary
course of physics.) As a result of this interaction, the spins
of electrons are oriented parallel to one another within small
regions called domains. A very strong magnetic field ap­
pears inside a domain, so that the latter is magnetized to
saturation. Every domain, in addition, is characterized by
a definite value and direction of the magnetic moment
vector Pm,d of the entire domain* (Fig. III.6.10).
9°. In the absence of an external magnetic field, the
magnetic moment vectors of the separate domains are orient­
ed absolutely chaotically inside a ferromagnetic so that the
total magnetic moment of the entire body equals zero (Fig.
III.6.11). Under the influence of an external magnetic field,
the reorientation of the magnetic moments of whole regions
of spontaneous magnetization—domains—along the field
occurs in ferromagnetics instead of the reorientation of the

* T h e m a g n e t ic m o m e n t o f a d o m a in m u s t n o t b e c o n fu s e d w it h
th a t o f a s e p a r a te a t o m o r m o le c u le (1 1 1 .6 .1 .3 ° ) .
324 Fundamentals of Electrodynamics 111.6/)

magnetic moments of separate atoms or molecules as in


paramagnetics (111.6.4.3°). The vectors Pm,d turn along the
field first of all in the domains in which the direction of
Pm,d is closest to the direction of the induction vector B0
of the external field. Consequently, the magnetization M
grows gradually with an increase in B0 (see Fig. III.6.7).

F ig. I I I . 6 .1 0 F ig . I I I . 6.11

Upon a growth in the external field, the dimensions of the


domains magnetized along the external field grow at the
expense of a reduction in the dimensions of the domains with
other orientations of the vectors Pmid. When the external
magnetic field is sufiiciently strong, the entire ferromagnet­
ic body is magnetized. The magnetization reaches its maxi­
mum value Ms—magnetic saturation sets in (Par. 2°).
10°. Diminishing of the induction B0 of the external
field is attended by a gradual disorientation of the regions
of spontaneous magnetization in a magnetized ferromagnetic.
Even in the absence of an external field, however, part of
the magnetic moments of the domains remain oriented, and
this explains the existence of remanence (Par. 5°) and the
possibility of manufacturing permanent magnets.
11°. The thermal motion of the atoms of ferromagnetic
substances facilitates the diminishing of the remanence.
This is why the remanence decreases with elevation of the
temperature. When the latter reaches the Curie point (Par.
7°), the remanence M r (see Fig. I l l .6.9) completely van­
ishes—the regions of spontaneous magnetization break up,
and the substance loses its ferromagnetic properties.
OSCILLATIONS
PART
AND WAVES

Chapter 1

Mechanical O scillations

1. Basic Concepts and Definitions


of Oscillatory Processes
1°. Oscillations, or oscillatory motion, are defined as
motion or changes in state having a certain degree of recur­
rence in time. Oscillations (vibrations) are very diverse in
their physical nature: the mechanical oscillations of a body
suspended on a spring (IV.1.3.1°) (a spring pendulum) (Fig.
IV. 1.1), the oscillations of pendulums (IV.1.4.1°) (Fig.
IV.1.2), the vibrations of strings, the vibrations of building
foundations, electromagnetic oscillations in an oscillator
circuit (IV.2.1.1°), etc. While differing in their nature, os­
cillations have common laws and are described by mathe­
matical methods of the same type.
2°. Oscillations are called periodic if the values of the
physical quantities changing in the course of the oscilla­
tions repeat at equal intervals of time. For example, the
positions of the pendulum in a clock and the magnitude of
the current in alternating-current mains (IV.2.2.3°) repeat.
326 Oscillations and Waves IV.1.1

3°. The period of oscillations T is the minimum time in­


terval after which the values of all the quantities charac­
terizing the oscillatory motion repeat. During this time, one
complete oscillation is performed.
The frequency of periodic oscillations v is the number of
complete oscillations performed in a unit time:

The cyclic {circular) frequency of periodic oscillations


co is the number of complete oscillations performed during
2ji time units:
co = 2jiv = ^ ^ whence T =

4°. A particular case of periodic oscillations are har­


monic oscillations in which the oscillating physical quantity
x changes with time according to the law
x = A sin (co£ + <p0) (*)
where A , co, and cp0 are constants, with A > 0 and co > 0.
The quantity A equal to the greatest absolute value of the

Fig. IV. 1.1 Fig. IV .1.2

oscillating physical quantity x is called the amplitude


of the oscillation. The expression co£ + cp0 = cp determines
the value of x at a given moment and is called the phase
of the oscillation. At the initial moment (£ = 0), the phase
equals the initial phase cpq.
IV.LI Mechanical Oscillations 327

Sometimes, the expression (*) is replaced with the equa­


tion x = A cos ((tit + <Pi) that differs from Eq. (*) in the
initial phase <px = <p0 — Jt/2. •
A very simple example of harmonic oscillations is the
displacement x along the OX axis of the projection of the
end of the position vector be­
longing to a point moving
around a circle of radius A .
When t = 0, the position vector
OB forms the angle (p0 with
the OY axis, while during the
time t it describes the angle
a t so that at an arbitrary mo­
ment we have x = A sin((o£-f
+ <Po) (Fig. IV.1.3).
5°. Free oscillations are de­
fined as oscillations that appear
in a system not subjected to
the action of varying external
forces as a result of a single
initial deviation of the system
from its state of stable equilibrium. Examples of free oscil­
lations are the oscillations of a body suspended on a spring
and brought out once from its position of equilibrium 0 0 '
(see Fig. IV.l.lfc), and the oscillations of a pendulum de­
flected once through the angle a (see Fig. IV. 1.2). When there
are free oscillations in a system, forces always act in it that
tend to return the system to its equilibrium position. If
the system is conservative (1.5.2.6°), then no energy is dis­
sipated during the oscillations. In this case, the free oscil­
lations are called undamped. Undamped free oscillations are
possible in a system only in the absence of friction and other
resistance forces. It is quite obvious that undamped oscil­
lations are an idealized case of oscillations. Real free oscil­
lations in mechanics are damped ones (IV.1.7.10). The ampli­
tude of undamped oscillations is independent of time and
remains constant.
328 Oscillations and Waves IV.1.2

2. Velocity and Acceleration


of Harmonic Oscillation
1°. By the magnitude u of the velocity of the harmonic
oscillation of a particle, in accordance with the definition of
velocity( 1.1.3.2°), is meant the change Ax in the absolute
value of the displacement x during a sufficiently small in­
terval of time:

,. Ax
v = lim-T—
At-0 At

The velocity of the harmonic oscillation described by


Eq. (*) (IV.1.1.4°) is:

v = toA cos (a t + <p0) = v0 cos (<*>£+ <Po) =

= VoSin (© i + -^- + <po)

where v0 = toA is the velocity amplitude proportional to the


cyclic frequency and the displacement amplitude A
(IV.1.1.4°).
The velocity v changes according to a sinusoidal law
with the same period T as the displacement x. The phase
of the velocity is in advance of the displacement phase
(IV.1.1.4°) by jt/2. For example, the velocity of a spring
pendulum is maximum and in absolute value equals the
velocity amplitude at the moment when the pendulum passes
through its equilibrium position (x = 0) (see Fig. IV.1.1a).
At the maximum displacements of the spring pendulum
(x = ± A ), the velocity vanishes (see Fig. IV. 1.16).
2°. The magnitude of the acceleration a of the harmonic
oscillation of a particle, in accordance with the definition of
acceleration (1.1.4.2°), is the magnitude Av of the change in
the velocity of the harmonic oscillation during a sufficiently
small time interval:
IV .1.2 Mechanical Oscillations 329

The acceleration of the harmonic oscillation described by


Eq. (*) (IV.1.1.40) is:
a = —<&2A sin (co£ + cp0) = — ft0 sin (co* + q>0) = — co2#
or
a = a0sin (coJ + n -f <p0)
where a0 = co2 A is the acceleration amplitude proportional
to the square of the cyclic frequency and the displacement
amplitude A (IV.1.1.40).
The acceleration a changes according to a sinusoidal law
with the same period T as the displacement x. The phase of
yX;Via

the acceleration is in advance of the displacement phase by


n (Fig. IV.1.4). For example, the acceleration of a spring
pendulum (see Fig. IV. 1.1) vanishes when the pendulum
passes through its equilibrium position and reaches maxi-
ximum values equal to the acceleration amplitude at the ma-
mum displacements of the spring pendulum (x = zk A).
The acceleration of a spring pendulum is always dire­
cted toward its equilibrium position: the pendulum slows
down when moving away from this position and speeds up
when approaching it. Figure IV. 1.4 shows the time dependence
of the displacement, velocity, and acceleration assuming
that the initial phase (p0 = 0.
3°. If the displacement x changes with time according
to the law of harmonic oscillation (IV.1.1.40), then the mag-
330 Oscillations and Waves IV.1.2

nitude of the acceleration a is always directly proportional


to the absolute value of x, and the direction of the accelera­
tion is always opposite to the direction of the change in x.
The formula a = —(o2# holds for any harmonic oscillations
and can be used to define such oscillations.
Problem 1. A particle performs harmonic oscillations
with a period of 2.0 s. The amplitude of the oscillations is
10 cm. Find the displacement, velocity, and acceleration of
the particle in 0.20 s after it passes through its equilibrium
position. The beginning of the oscillations coincides with
the equilibrium position.
Given: T = 2.0 s, A = 10 cm = 0.10 m, t = 0.20 s, (p0 = 0.
Required: x, v, a.
Solution: The displacement x of an oscillating particle
is
2ji
x —A sin (cof + cpo), x = A sin 1=

= 0.10 sin 2 * * 80 x 0.20 = 0.10 sin 36° m 0.059 m

The velocity of a harmonically oscillating particle is

v=- (i>A cos (co/ + cp0), v = A cos 1=

x 0 -10xcos 36° ^ °-25 m/s

The acceleration at the moment t is

a = — co2A sin (co* + cp0) = — (o2#, a = ---- x

|a| = 4x430142 x 0.059 « 0.57 m/s2

Problem 2. The middle of a vibrating string has a max­


imum acceleration of 2.02 X 103 m/s2. Find the frequency
of the vibrations if their amplitude is 2.00 mm.
Given: a = 2.02 XlO3 m/s2, A = 2.00 mm = 2.0 X
X 10"3 m.
Required: v.
IV .1.3 Mechanical Oscillations 331

Solution: The maximum acceleration corresponds to the


maximum displacement of a point from its equilibrium
position: | amSLX | = co2A = 4jt2vM, whence

| /~ | a m ax I __ | / " 2.02 X 103


V 4jt 2 A ~ V 4 X 3 .1 4 2 X 2.00 X 10-3
160 s"1

3, Harmonic Oscillations
of a Spring Pendulum

1°. In the state of equilibrium of a spring pendulum


(IV. 1.1.1°), the force of gravity P acting on it is balanced
by the elastic force F (11.7.2.2°) of the stretched spring (see
Fig. IV .l.la). If we bring
the body A out of its equi­
librium position by moving
it downward along the OX
axis over the distance x,
then the pendulum will
begin to perform free oscil­
lations (IV.1.1.5°) under the
action of the elastic force
F of the spring directed op­
positely to the displacement
of the pendulum (Fig. IV.
1.5a, b). According to
Hooke’s law (1.2.9.4°), the Fig. IV.1.5
force F is directly propor­
tional to the absolute value
of the displacement x of the pendulum and is always
directed toward the equilibrium position. Such forces in
oscillatory motion are called restoring forces.
2°. If the origin of coordinates coincides with the equi­
librium position of a spring pendulum, and the OX axis is
directed downward (see Fig. IV.1.5), then according to
Hooke’s law

F — —kx
332 O sc illa tio n s and W aves IV .1.3

where F = acting force


x = absolute value of the displacement of the pen­
dulum
k = spring constant (1.2.9.4°).
3°. According to Newton’s second law (1.2.4.1°)

F = ma
where m = mass of the body of the spring pendulum
a = its acceleration.
From the above two equations, we get ma = —kx, or
a = —kxlm. In the preceding section, we found that a =
-c o fc (IV. 1.2.2°). Hence,

and

According to IV. 1.2.2°, a spring pendulum performs free


harmonic oscillations with a cyclic frequency of co0 (the
natural cyclic frequency of free oscillations):

The period of oscillations is

Problem. A helical spring under the action of a weight


suspended on it is stretched by 6.5 cm. If the weight is pulled
downward and then released, it begins to oscillate along
a vertical line. Find the period of oscillation of the weight.
Given: x = 6.5 cm = 6.5 X 10"2 m.
Required: T.
Solution: We assume that the weight performs harmonic
oscillations with the period T: T = 2zi]f mlk, where m
is the mass of the weight, and k is the spring constant.
lV.1.4 M e c h a n ic a l O s c illa tio n s 333

The magnitude of the force of gravity in the equilibrium


position of the weight will equal the magnitude of the
elastic force: mg = kx, whence* k = mglx, and

7’ = 2 j t | / y = 2 x 3 . 1 4 ) / 65 9 g° ~2 = 0-51 s

4. H a r m o n ic O s c illa tio n s
of a M a th e m a tic a l P e n d u lu m
1°. A mathematical, or simple, pendulum is defined as
a point particle M suspended from a fixed point by a mas­
sless unstretchable string and moving in a vertical plane

Fig. IV.1.6 Fig. IV. 1.7

under the action of the force of gravity P. In the equilibri­


um position, the two forces acting on the particle—the force
of gravity P = mg (1.2.8.3°) and the tensile force Ft of
the string—balance each other (Fig. IV. 1.6). If we deflect the
pendulum from its equilibrium position through the small
angle a, then the force of gravity P and the tensile force
Ft will be at an angle to each other, and they will not be
balanced.
2°. The restoring force (IV.1.3.10) for a mathematical
pendulum is the component F of its force of gravity mg
equal to (Fig. IV. 1.7)
F = mg sin a
334 O s c illa tio n s and W aves 1 V .1 A

At small displacement angles, we have sin a ^ a = x ll.


Taking into account that the directions of the displacement
and the restoring force are opposite, we get

y-| X
F = — rag —

where x is the absolute value of the displacement of the


pendulum from its equilibrium position.
3°. Examination of the last equation in Par. 2° shows
that the acceleration of a pendulum is

a= — y x= —

where coq = gll.


By comparison with IV. 1.2.2°, we see that small oscil­
lations of a mathematical pendulum are free harmonic
oscillations with a natural cyclic frequency of co0 = ]/ gll.
The period of small oscillations of a mathematical pen­
dulum T = 2jt/co0 = 2nYIIg does not depend on the mass
of the pendulum and the amplitude of its oscillations.
Observations of the oscillations of pendulums are used to
determine the acceleration of free fall g (1.2.8.4°).
Problem. A mathematical pendulum with a period of
oscillations of 2.000 s has a length of 0.9973 m at the lati­
tude of Arkhangelsk and 0.9952 m at the latitude of London.
Find the accelerations of free fall corresponding to these
latitudes.
Given: T = 2.000 s, Zx = 0.9973 m, Z2 =: 0.9952 m.
Required: g1? g2.
Solution: A mathematical pendulum has a period of
oscillations of T = 2nYUg- Hence,

4x3.1422x0.9973 = 9.882 m/s2


4
and g2
4X3.1422X0.9952 9.812 m/s2
4
1V.1.5 M e c h a n ic a l O sc illa tio n s 335

5. E n e r g y o f H a r m o n ic O s c illa to r y
M o tio n
1°. In the harmonic oscillations of a spring pendulum
(see Fig. IV. 1.1), the potential energy of the elastically de­
formed body Ev = kx2!2 (1.5.3.6°) transforms into its ki­
netic energy E k = mi;2/2 (1.5.3.3°), where k is the spring
constant (1.2.9.4°), x is the absolute value of the displace­
ment of the pendulum from its equilibrium position, m is
the mass of the pendulum, and v is its velocity. In accordance
with IV.1.1.40 and IV.1.2.10,
jp kx2 kA2 . 9 / , x
E p — -g ” — ~~2~ Sm 2 9°)

Ey = —= — — cos (©*+ <Po) 2

2°. The total energy E of a spring pendulum is


p mt>2 . kx2 kA2
E= — + — =—

Energy is transformed in the oscillations of a spring pen­


dulum in accordance with the law of conservation of me­
chanical energy in a conservative system (1.5.4.1°). When
a pendulum moves downward or upward from its equili­
brium position, its potential energy grows, and its kinetic
energy diminishes. When the pendulum passes through its
equilibrium position (x = 0), its potential energy vanishes,
while the kinetic energy of the pendulum has its maximum
value equal to its total energy.
3°. The total energy of the harmonic oscillations of a
spring pendulum is proportional to the square of the am­
plitude of the oscillations:
„ kA2 m(i)2A 2
E ~ 2 ~ 2
Figure IV. 1.8 shows a graph of the potential energy of
the elastic oscillations of a spring pendulum with the value
E of its total energy laid off. A glance at the figure shows that
the value of the amplitude of the oscillations x = ± A
336 Oscillations and Waves IV.1.5

equals the displacement of the pendulum at the “turning


points” B and C — the extreme points of deviation of the
pendulum from its equilibrium position. The amplitude of
oscillations of a pendulum of
a given mass and spring con­
stant is determined by its
store of total energy:

A V k (0 V m

4°. The information on the


energy of the oscillations of
a spring pendulum is of a gen­
eral nature and holds for free
harmonic undamped oscilla­
tions in any oscillating system where oscillations of the
kind indicated are being performed.
Problem. A spring pendulum performs harmonic oscil­
lations with a displacement amplitude of 0.04 m. At a dis­
placement of 0.03 m, the elastic force is 9 X 10"5 N. Find
the potential and kinetic energies corresponding to the
given displacement, and the total energy of the pendulum.
Given: x = 0.03 m, A = 0.04 m, F — 9 X 10“5 N.
Required: Ev, i?k, E.
Solution: The total energy of the pendulum is E =
= 2?k + jEp, where E k is the kinetic energy, and Ep is the po­
tential energy. But E = kA2l2, where k is the spring constant
determined according to the elastic force F, i.e. k = Fix,
where x is the absolute value of the displacement.
Hence,
P _FA2 9 X 10~5 X 16 X 10~4
2.4 xlO"6 J
E 2x 2 X 3 X IQ"2

The potential energy is

Ev = •— = = 9 x 10~6^ 3 x 10 2 = 1.35 x 10~6 J

The kinetic energy is


£ fc= £ - £ p = 1.05 X IQ'6 J
1 V .1 .6 Mechanical Oscillations 337

6. Addition of Harmonic
Identically Directed Oscillations
1°. If a point particle simultaneously participates in
two harmonic oscillations of the same cyclic frequency
(IV. 1.1.3°), then addition of harmonic oscillations occurs.
2°. In the simplest case when adding two identically
directed harmonic oscillations with the displacements x±
and x2 of identical cyclic frequency co and differing only
in their displacement amplitudes (A± and A 2) and initial
phases (cpj and cp2) (IV. 1.1.4°):
x 1= i41sin(co^ + q)1) and x2 = A2 sin («£ + ^2)

the resultant harmonic oscillation has the displacement


x = xx + occurs in the same direction, and is a harmonic
oscillation of the same frequency:
x = A sin (co£ + cp)

where A = amplitude of the displacement of the resultant


oscillation
cp = its initial phase.
The quantities A and cp are calculated by the formulas

A = V A \ + A \ + 2AtAz cos (<p2— q>i)

<P arc tan A} sin ^ + A/ 8ip


Ax cos cpx+ A2 cos q>2

3°. The quantity cos (cp2 — cpd cannot be greater than


+ 1 and smaller than —1. Hence, the possible values of the
amplitude A are within the limits:
Ai + A 2 ^ A ^ \ A 2 — A 1 |

We take into consideration here that by definition of the


amplitude (IV.1.1.40), it cannot be negative.
Particular cases of the addition of oscillations:
(1) cp2 — cpx = 2nn, where n — 0, 1, 2, . . . . Here
cos (cp2 — cpi) = 1, and A = A x + A 2.
22-0211
338 Oscillations and Waves 1V.L7

Figure 1V.1.9 shows the addition of two such harmonic


oscillations.

(2) <p2 — <Pi = (2n + 1) Jt, where n = 0, 1, 2, . . . .


Here cos ((p2 — cpx) = —1, and A = | A 2 — A x | =
I Ai (A 2 | .

Fig. IV.1.10

The time dependences of the displacements of the oscil­


lations being added having opposite initial phases and of
the displacement of the resultant oscillation are shown in
Fig. IV.1.10.

7. Damped Oscillations
1°. Oscillations are called damped when their energy
diminishes with time. The damping of free harmonic oscil­
lations (IV. 1.1.5°) is associated with the decrease in the
mechanical energy of the oscillating system because of the
action of forces of friction and other forces of resistance.
1 V .1 .7 Mechanical Oscillations 339

2°. The amplitude of damped oscillations decreases with


time according to the law A (t) = A 0e~6t, where A 0 is the
initial amplitude of the oscillaljons at the moment t = 0
determined by the initial store of total energy of the oscil­
lating body (IV.l.fi.S0), e is the base of natural logarithms,
and 8 is the damping factor, it characterizes the rate of dimin­
ishing of the amplitude and depends on the forces of fric­
tion and the mass of the oscillating body. If the force of
friction is proportional to the velocity v of the oscillations,
i.e. Ffr = —fv, where / is the coefficient of friction, then
6 = f/2m (here m is the mass of the body). Diminishing of
the amplitude of damped oscillations according to the law
A = A$e~bt is observed only with small damping. The
values of the amplitudes for the moments t, t + At, t -f 2At,
etc. in this case form a decreasing geometric progression
whose common ratio is e~&At.
3°. Damped oscillations are not periodic oscillations
(IV. 1.1.2°) because the values of the physical quantities
characterizing such oscillations (for example the displace­
ment, velocity, and acceleration) are never repeated in them.
This is why the concepts of period and frequency in­
troduced for periodic oscillations (IV. 1.1.3°) cannot be ap­
plied to damped oscillations.
The conditional period (period) T of damped oscillations
is defined as the time interval between two consecutive states
of an oscillating system in which the physical quantities
characterizing the oscillations take on similar values while
changing in the same direction, diminishing or growing.
The period of damped oscillations is calculated by the for­
mula
rp __ 2JI

where co0 = natural frequency of free undamped oscilla­


tions (IV.1.3.30)
8 = damping factor.
The quantity cod = V — 82 is called the cyclic fre­
quency of damped oscillations. It shows how many times an
oscillating body passes through its equilibrium position
during n seconds.
22 *
340 Oscillations and Waves IV .1.8

4°. Provided that 6 < co0, damped oscillations are de­


scribed by the equation

x = A0e~6t sin (o)d£ + (p0)

where cp0 is the initial phase of the oscillations determined


by the initial conditions of appearance of the oscillations.
Figure IV. 1.11 shows how
x depends on t.
5°. With great friction
(8 > a)0), no damped oscil­
lations occur. A system

Fig. IV.1.12

brought out of its equilibrium position by external forces


returns to its equilibrium position aperiodically (not period­
ically) (Fig. IV. 1.12) after the action of these forces ceases.
The store of mechanical energy of the body is used
up to overcome friction by the moment it returns to its
equilibrium position.

8. Forced Oscillations

1°. Undamped oscillations of a system produced by the


action on it of external forces F (t) periodically changing
with time are called forced oscillations. Examples of the lat­
ter are the oscillations of the current in alternating-current
mains (IV.2.2.30), and the vibrations of screw propellers,
turbine blades, and shafts under the action of periodically
varying external forces. The force F (t) setting up forced
oscillations is called the driving (disturbing) force.
I V . 1. 8 Mechanical Oscillations 341

2°. If the driving force F (t) changes harmonically


(IV. 1.1.4°) according to the law
F (t) = F0 cos cot
where F0 is the amplitude of the driving force, and co is
its cyclic frequency (IV.1.1.30), then forced oscillations may
be produced in the system on which such a force is exerted.
These oscillations are also harmonic, occur with a cyclic
frequency equal to the frequency co of the driving force,
and are described by the equation
x = A cos (co£ + cpx)
Here A = amplitude of the forced oscillations of the rel­
evant quantity (for example the displacement)
(Pi = phase difference between the forced oscillations
x and the force F (t).
For instance, a spring pendulum (IV. 1.3.1°), which is
pushed upward periodically, after a definite time begins to

Fig. IV.1.13

oscillate with a definite amplitude (Fig. IV.1.13). Inspection


of Fig. IV.1.13 shows that initially, in the course of setting
up the forced oscillations, their nature is quite intricate. Free
damped oscillations (IV.1.7.1°) and forced oscillations are
superposed. After the free oscillations terminate, only the
forced oscillations remain.
3°. The amplitude A of stationary forced oscillations
is determined by the formula

my (0)2 _ 0)2)2+ 462(02


342 Oscillations and Waves IV.1.8

where F0 = amplitude of the driving force (Par. 2°)


m = mass of the oscillating system
co0 = cyclic frequency of the free undamped
oscillations of the system (IV. 1.3.3°)
o) = cyclic frequency of the external force
8 = damping factor (IV. 1.7.2°).
At constant values of F0, m, and 6, the amplitude of
forced oscillations depends on the ratio of the cyclic frequen­

cies of the driving force (co) and of the free undamped oscil­
lations ((00).
4°. The curves showing how the amplitude A depends
on co at different damping factors depicted in Fig. IV. 1.14
have been plotted according to the formula for A (Par. 3°).
Corollaries from the formula of Par. 3°:
(a) If the cyclic frequency of the driving force equals zero
(co = 0), then

Here no oscillations occur, and the deviation of the system


from its equilibrium position is called a static deflection.
IV .1.9 Mechanical Oscillations 343

(b) In the absence of damping (6 = //2m = 0), the am­


plitude of forced oscillations grows with increasing (o, and
when (o = (o0, and the denominator in the formula for A
vanishes, the amplitude of the oscillations becomes equal
to infinity. A further growth of the frequency is attended by
diminishing of A , and lim A = 0.
(D—>-oo
(c) If damping exists (6 0), then the amplitude of the
forced oscillations reaches its maximum value at a frequen­
cy (Ores of the driving force that does not coincide with the
frequency of free undamped oscillations co0:

(o = cores= / c D ’ - 2 5 2 = c o 0j / 1 — ^

5°. The phenomenon of the growth in the amplitude of


forced oscillations when the cyclic frequency of the driving
force approaches the value cores is called resonance. The quan­
tity (Ores is called, accordingly, the resonance cyclic frequency,
and the curves showing how A depends on co (see Fig.
IV. 1.14), resonance curves. When friction is present (6 0),
the resonance cyclic frequency (ores is somewhat lower than
the natural cyclic frequency of free damped oscillations
(cod = y (Oq — 82) (IV.1.7.30) and lower than co0—the natu­
ral frequency of free undamped oscillations (IV. 1.3.3°).
6°. The shape of the resonance curves in Fig. IV. 1.14
depends on the value of the damping factor 8. With increas­
ing 8, the resonance curves become less steep, the “sharp­
ness” of the curves diminishes, and also the value of A maX
when (o = (ores-
7°. The phenomenon of resonance is used in acoustics
for analysing sounds, their amplification, etc.
Resonance phenomena may appear in machines and var­
ious structures under the action of periodically varying loads
that are sometimes dangerous for operation of the machines
or for service of the structures.

9. Auto-Oscillations
1°. An oscillating system performing undamped oscil­
lations at the expense of the action of an energy source hav­
ing no oscillating properties is called an auto-oscillating
344 Oscillations and Waves IV.1.9

system (auto-oscillator). An example of such a system is a clock


or watch with an anchor (recoil) escapement (Par. 5°).
2°. An auto-oscillating system consists of four parts
(Fig. IV.1.15):
(a) an oscillating system;
(b) an energy source compensating the losses of energy
due to damping of the oscillations as a result of friction or
other resistance forces;
(c) a valve—a device that controls the supply of energy
to the oscillating system in definite portions; and
(d) a feedback—a device for reverse action of the auto-
oscillating system on the valve, i.e. for control of valve
Feedback

Fig. IV.1.15

operation at the expense of the processes in the oscillating


system itself.
3°. The feedback is called positive (negative) if during the
time of action of the energy source on the oscillating system
the source does positive (negative) work on the system and
transmits to it (removes from it) a certain amount of energy.
Positive feedback is employed for initiating auto-oscilla­
tions. Negative feedback increases damping, and the auto-
oscillations are suppressed.
4°. Examples of auto-oscillating systems are clocks and
watches (Par. 5°), steam and internal-combustion engines,
pick hammers, and electric bells. Auto-oscillations are pro­
duced by strings under the action of a bow in a violin, by
the columns of air in wind instruments, by the reeds in ac­
cordions and other reed instruments, and by the vocal cords
when talking or singing. A v?rlve generator of undamped elec­
tric oscillations (IV.2.9.10) is an electrical auto-oscillating
system. It is sometimes difficult to reveal the feedback
mechanism in an auto-oscillating system and divide the
latter into its basic parts (Par. 2°).
IV.1.9 Mechanical Oscillations 345

5°. A clock with an anchor escapement (Fig. IV. 1.16)


is an auto-oscillating system. The escape wheel with spi­
ral teeth is rigidly fastened to a cogged drum over which
a chain with a weight passes. One end of the pendulum
has an anchor with two pallets
rigidly fastened to it. The pal­
lets are fiat pieces of a hard ma­
terial bent to the shape of the
arc of a circle with its centre
at the pendulum axis. In
watches, the weight is replaced
with a spring, and the pendu­
lum with a balance—a small fly­
wheel fastened to a helical spring.
The balance performs torsional
oscillations about its axis. The
pendulum or balance is the os­
cillating system in clocks and
watches. The lifted weight or the
wound-up spring is the energy
source. The anchor, which per­
mits the escape wheel to turn
by one tooth during a half pe­
riod, is the valve. Feedback is
accomplished by the interaction
between the anchor and the
escape wheel. When the pendu­
lum passes through its equilibrium position and has the
greatest velocity, a tooth of the escape wheel briefly comes
into contact with an end of a pallet and pushes the pendu­
lum. In a clock, the potential energy of the weight is gradual­
ly, in separate portions, transmitted to the pendulum and
compensates the losses of energy on friction.
346 Oscillations and Waves IV.2.1

Chapter 2
Electromagnetic O scillations

1. Free Electromagnetic Oscillations


in an Oscillator Circuit
1°. An oscillator circuit is defined as an electric circuit
consisting of a resistor R (111.2.4.1°), a coil of inductance
L (111.5.5.2°), and a capacitor of capacitance C (111.1.10.1°)
connected in series (Fig. IV. 2.1). Such a circuit is often
called an RLC circuit. In a very simple idealized case when we

Fig. IV.2.1 Fig. IV.2.2 Fig.IV.2.3

can disregard the resistance of the circuit (i?->- 0), an oscil­


lator circuit consists of a capacitor of capacitance C and a
coil of inductance L connected in series (an LC circuit)
(Fig. IV.2.2).
2°. Periodic variations of the charge q, the potential
difference Acp across the capacitor plates, and the electric
current I in the circuit may occur in an oscillator circuit.
If these variations are due to the capacitor plates being
charged only once, then free electromagnetic oscillations of q,
/, and A<p appear in the oscillator circuit. Figure IV.2.3
shows how electromagnetic oscillations are set up and pro­
ceed in an oscillator circuit when R -v 0. If the switch S
is in its position a at the initial moment t = 0, then the
capacitor is charged, receiving the charge q0. The capacitor is
thus supplied with the energy W = q\!2C (111.1.12.2°),
and the potential difference Acp across its plates becomes equal
I V . 2 .1 Electrom agnetic Oscillations 347

to the maximum value Acp0. There is no current I in the


oscillator circuit (Fig. IV.2.4a). Turning of the switch to
its position b is attended by discharge of the capacitor.
Owing to self-induction (111.5.5.1°), the current in the os­
cillator circuit gradually increases, and it reaches its maxi­
mum value I = 70 at the moment t = 774 when q and Acp
vanish (Fig. IV.2.4&). Next, the current in the circuit
(a) (b) (c) (d) (e)

(a') ( b') (cf) (d') (e')


Fig. IV.2.4

while retaining its direction gradually diminishes and va­


nishes at t = 772. Here the charge of the capacitor and the
potential difference across its plates again reach their maxi­
mum values, but the signs of the charges on the plates and
the direction of the intensity of the electric field between them
are opposite to those at the moment t = 0 (Fig. IV.2.4c).
As a result, owing to self-induction, the capacitor is re­
charged. After this in the time intervals from 772 to 3774 and
from 3774 to 71, the processes occur in the reverse direction
(Fig. IV.2.4d, e).
3°. It is convenient to study electromagnetic oscilla­
tions by taking advantage of the circumstance that oscilla­
tions of a various nature—mechanical and electromagnetic—
obey similar laws. For a mathematical pendulum (IV. 1.4.1°),
the energy transformations shown in Fig. lY.2A a'-ef cor­
respond to the processes in the oscillator circuit shown in
Fig. IV.2.4a-e. This makes it possible to apply the results of
348 Oscillations and Waves IV.2.1

investigations obtained for mechanical oscillations to os­


cillator processes in a circuit (see Fig. IV.2.1). Here use is
made of the analogies existing between the physical quan­
tities characterizing mechanical systems and electric circuits
(Table IV.2.1).
T a b le I V . 2 . 1

Mechanical system Electric circuit

Mass m Inductance L
Spring constant k Reciprocal of capacitance, i/C
Coefficient of friction / Resistance R
Force F E.m.f. %
Displacement x Charge q
Velocity y= A x/A t Current / = Ag/A*
Acceleration a = k v /& t Rate of current change A//AJ

4°. The process considered in Par. 2° is characterized by


the periodic transition of the energy of the electric field
of the capacitor into the energy of the magnetic field of the
electric current (111.5.7.1°). At the moments t = 0, 772, 71,
etc., the energy of the electric field is maximum and equals
q\!2C, while the energy of the magnetic field equals zero
because there is no current in the circuit. At the moments
t = 774, 3774, etc., the energy of the magnetic field is maxi­
mum and equals LPJ2, while the energy of the electric
field equals zero because the capacitor is completely
discharged.
Similarly, in free undamped oscillations, a periodic tran­
sition of potential energy into kinetic energy and vice versa
occurs. Figure IV.2.4a'-e' shows the conversions of the po­
tential (W) and the kinetic (K ) energy upon the undamped
oscillations of a mathematical pendulum corresponding to
the processes in an oscillator circuit. The varying electromag­
netic field (111.1.3.1°) that appears in the oscillator circuit
of Fig. IV.2.1 is concentrated (localized) in the region of the
space about the circuit. Hence, such a circuit is called a
closed one and cannot be used for emitting electromagnetic
waves (IV.4.4.6°).
rv. 2 .i Electromagnetic Oscillations 340

5°. Free electromagnetic oscillations are damped in a


real (R =^=0) oscillator circuit. For example, the change in
the charge q on the capacitor plates is described by a formu­
la similar to the equation of damped mechanical oscilla­
tions (IV.1.7.40):
Q= Qoe~6t sin (cod/ + ( p 0)

where q0 = amplitude value of the charge at the moment


t= 0
6 = RI2L is called the damping factor (R is the
resistance and L the inductance of the circuit)
q)0 = initial phase of oscillations of the charge.
Figure IV.2.5 shows q against t for such an oscillator
circuit.

6°. The quantity (Dd is defined as the cyclic frequency


of free electromagnetic oscillations in a circuit: cod =
= y \! L C — i?2/4L2. For mechanical oscillations, the for­
mula of the frequency of damped oscillations corresponds
to this formula.
For free undamped oscillations (R = 0), the cyclic fre­
quency (Da = <d0 = j/"1!LC.
The period T of free undamped oscillations is expressed
by the Thomson formula

T = — = 2nV LC
co0 r

Undamped electromagnetic oscillations can be obtained,


for example, with the aid of a valve generator (IV.2.9.10).
350 Oscillations and Waves IV.2.2

2. Forced Electromagnetic Oscillations.


Alternating Current
1°. Forced electromagnetic oscillations are defined as the
undamped oscillations of the charge q, the potential dif­
ference Acp across the capacitor plates, the current / , and
other physical quantities in an oscillator circuit due to a pe­
riodically varying sinusoidal (alternating) e.m.f.:
%= %0 sin cd£
where g 0 = amplitude value of the e.m.f.
co = cyclic frequency of the alternating e.m.f.
Energy is supplied to the circuit that is needed to make
up for the losses of energy in the circuit (111.2.7.4°) because
of the presence of the resistance i?.
2°. A sinusoidal e.m.f. is set up in a loop that rotates
with the angular velocity co in a stationary homogeneous
magnetic field of induc­
tion B (Fig. IV.2.6).
The magnetic flux O
(111.4.1.8°) through a loop
of area S is
O = BS cos a = BS cos co£
where a = (at is the angle
between a normal n to the
loop^and the magnetic in­
duction vector, and is directly proportional to the time
t. According to Faraday’s law of electromagnetic induction
(111.5.1.2°), the induced e.m.f. is
op AO

where A0/A£ is the rate of change of the magnetic induc­


tion flux.
A harmonically varying magnetic flux results in a si­
nusoidal induced e.m.f.

g{= — BS cocos ^co/ + -^-) = £ 0sin cat


IV.2.3 Electromagnetic Oscillations 351

where %0 = BSco is the amplitude value of the induced


e.m.f.
3°. A current varying according to a harmonic law is
called an alternating current.
An alternating current is produced by forced oscilla­
tions of the current in an electric circuit occurring with
the frequency co that coincides with the frequency of the
driving e.m.f. (cf. IV. 1.8.2°):

/ = / 0 sin (co/-|-(p)

where I 0 = amplitude value of the current


cp = phase shift between the oscillations of the cur­
rent and the e.m.f.

3. Alternating-Current Circuit.
Resistance

1°. An alternating-current (a.c.) circuit in the general


case is an oscillator circuit to which an external sinusoidal
e.m.f. is applied (Fig. IV.2.7).
To set it up, it is necessary to
connect an oscillator circuit to
the terminals of an a.c. genera­
tor. In the given circuit, apart
from the external sinusoidal
e.m.f. <£, there is an e.m.f. of S
self-induction $sl = — L A ll At —0 0—
(111.5.5.3°), and there is the
potential difference (p2 — cpx Fig. IV.2.7
across the capacitor plates.
2 • According to Ohm s law for a circuit containing an
e.m.f. (111.2.4.2°), we have

I R = %+ &si + cp2 — cpi


or
%= IR + (9l - «p2) - gsl = UB + Uc + UL
352 Oscillations and Waves IV.2.4

where UR = IR = voltage across the resistance R


C/c = cpi — q)2 = g/C = voltage across the capaci­
tive reactance xc
UL = —£si = L AH At = voltage across the induc­
tive reactance xL.
The external e.m.f. % equals the sum of the voltages
across the resistance i?, the capacitive reactance xCl and the
inductive reactance xL.
3°. The circuit depicted in Fig. IV.2.7, in which UR >
> Uc and UR > UL (Par. 2°),
(J is called an a.c. circuit with a
resistance. Here
% = UR = IR
or, on the basis of IV.2.2.10,
IR = <S0 sin (d£
The current oscillates according
to the law

Fig. IV.2.8 / = ——sin (i)t —70 sin octf
H

where I 0 = %JR is the amplitude value of the current.


In a circuit with a resistance, the current oscillates harmo­
nically with the frequency and phase of oscillations of the
external sinusoidal e.m.f. (Fig. IV.2.8).

4. Inductive Reactance

1°. The circuit depicted in Fig. IV.2.7 in which the ca­


pacitor is short-circuited ( Uc = 0 ) and U l ^ UR is called
an a.c. circuit with an inductive reactance xL. The oscil­
lations of the current in such a circuit lag in phase by
j i /2 behind the oscillations of the e.m.f. (Fig. IV.2.9):

^= sin (0/ and I = I 0sin^a)t—


IV.2.5 E lectro m a g n etic O scilla tio n s 353

2°. The relationship between the amplitude values oi


the current I 0 and the e.m.f. <£0 is
T __ _1 %
0 Lto ~ xL

where xL = coL is the inductive reactance (L is the induct­


ance of the circuit). According to Lenz’s law (111.5.1.3°),
the e.m.f. of self-induction in a circuit prevents the changes
of the current in it. This results in the existence of the induc­
tive reactance xL that retards
the changes of the current in
the circuit in comparison with
the changes of the e.m.f.

5. Capacitive Reactance
1°. If inductance is absent
( UL = 0 ) in the circuit depicted
in Fig. IV.2.7, and Uc > UR,
then we have an a.c. circuit with
a capacitive reactance xc. The
oscillations of the current in such a circuit are in advance
of the oscillations of the external e.m.f. in phase by j t /2
(Fig. IV.2.10):
%= ^0sin tot

and / = 70sin + -y j

2°. The relationship between


the amplitude values of the
current I 0 and the e.m.f. 5f0 *s

Fig. I V . 2.10
coC

where xc = l/coC is the capacitive reactance (C is the capaci­


tance of the capacitor).
A capacitor in a circuit has an infinitely great resistance
for a steady current—the latter does not pass through it.
It has a finite resistance inversely proportional to its capac-
23 — 0211
354 Oscillations and Waves tV .H

itance for an alternating current. The voltage Uc across


the capacitor plates gradually grows as the capacitor is
charged, and the oscillations of the current in a circuit with
a capacitor are in advance of the e.m.f. oscillations.

6, Ohm’s Law for


an Alternating-Current Circuit
1°. In the a.c. circuit depicted in Fig. IV.2.7, the cur­
rent and the e.m.f. oscillate according to a sinusoidal law
with the identical frequency o) and the phase shift <p:
/ = / 0sin(L>£ and %= <£0 sin (cd£ + (p)

2°. The relationship between the amplitude values of


the current 70 and the e.m.f. <£0 in an a.c. circuit is:

where Z = j/^ R2 + ^Z/Co — is called the impedance


of the a.c. circuit.
Ohm's law for an a.c. circuit: the amplitude of an alter­
nating current is directly proportional to the amplitude of
the e.m.f. and inversely proportional to the impedance of
the circuit.
Ohm’s law in this form also holds for the effective values
of the current and the e.m.f. (IV.2.7.3°):

3°. The phase shift between the oscillations of the cur­


rent and the e.m.f. is determined by the relationship
I 0R R_
cos <p= —— Z
©o
where R — resistance
Z = impedance of the a.c. circuit (see also Par. 2°).
tV ± 7 E le c tro m a g n e tic O scilla tio n s 355

7. Power of Alternating Current.


Effective Values
of Current and Voltage
1°. The irreversible conversion of the energy of the elec­
tric current into the internal energy of the conductor
occurs in a circuit with a resistance—Joule heat (111.2.7.4°)
is liberated.
The instantaneous power of an alternating current, i.e.
its power at a certain moment t, is determined by the pro­
duct of the instantaneous values of the current / and the e.m.f.
t: (provided that there is no phase shift between them):

P = I(g = poR sin2 (ot = P0 sin2 (ot

where P 0 = I\R is the amplitude value of the power.


2°. The mean power P of an alternating current is de­
fined as the work A done by the current during the time T
and related to a unit time; there T is the period of the alter­
nating current:

3°. The effective values of the current ( /eff), e.m.f.


(&eff), and voltage (f/eff) of an alternating current are de­
fined as the values of these quantities for such a steady cur­
rent which in a circuit of the same resistance releases a po­
wer that is the same as the mean power P of the alternating
current:
I eff i and C7„ = £o_
V*
where 70, and t/0 are the amplitude values of the cur­
rent, e.m.f., and voltage.
4°.yrhe active power of an alternating current P a is the
mean power of the irreversible transformations of energy in
the a.c. circuit:
P a = leff&eff COS(p
23*
356 O sc illa tio n s a n d W a v e s 1V .2.8

where Iv[f and ? eff = effective values of the alternating


current and the e.m.f., respectively
cos cp = RIZ — power factor determined by the phase
shift between the oscillations of the
current and the e.m.f. (IV.2.6.3°).
At a low power factor, the load consumes a small active
power from the generator, i.e. only a small part of the po­
wer produced by the generator. The remaining part of the
power produced by the generator is periodically pumped
from the generator to the consumer and back and is dis­
sipated in the transmission lines.

8. Resonance in an
Alternating-Current Circuit
1°. The amplitude of the current / 0 in an a.c. circuit
reaches its maximum value 70,max at the smallest value of
the impedance Z of the circuit (IV.2.6.20), i.e. provided
that

The cyclic frequency co of the oscillations of the current and


the e.m.f. here is

and coincides with the cyclic frequency of the free undamped


electromagnetic oscillations in the electric circuit (IV.2.1.6°).
2°. The phenomenon of a sharp growth in the amplitu­
de of forced oscillations of the current in an oscillator cir­
cuit when the cyclic frequency co of the external alternating
e.m.f. approaches the frequency co0 of free undamped oscil­
lations in the circuit is called resonance in an ax. circuit.
The frequency co = co0 is called the resonance cyclic frequency.
The latter does not depend on the resistance R (cf. IV. 1.8.5°).
The plot of / 0 against co is a resonance curve (Fig. IV.2.11).
The sharpness of the peak of a resonance curve grows with
a decreasing resistance R.
I V . 2 .8 E lectro m a g n etic O scilla tio n s 357

3°. Upon resonance in the electric circuit shown in


Fig. IV.2.7, the phase shift between the oscillations of the
external e.m.f. and the current vanishes. The active power
(1V.2.7.40) coincides with the total power produced by the
generator, i.e. the most favourable conditions are ensured
for the supply of energy from the a.c. source to a consumer.
The amplitudes of the voltage across the inductance UL
and across the capacitance Uc (IV.2.3.20) are the same:

while the phases are opposite: UL is in advance of Uc in


phase by jt s o that UL -j- Uc = 0. The total voltage drop

'o
Fig. IV.2.11 Fig. IV.2.12

in the circuit (see Fig. IV.2.7) equals the voltage drop across
the resistance UR. This phenomenon is called voltage reso­
nance.
4°. In an electric circuit consisting of a capacitor of ca­
pacitance C and a coil of inductance L connected in paral­
lel (Fig. IV.2.12), with low resistances of the parallel
branches (Rx and R 2 tend to zero), the amplitude of the current
/ 0 in the external (unbranched) circuit is

where I 0tl and / 0>2 = amplitude values of the currents in


the parallel branches
r 0 — amplitude value of the external
e.m.f.
When co = co0 = 1 we have
I , i —A), 2
q find I —0 q
358 O s c illa tio n s and W aves I V . 2.9

The sharp reduction in the current amplitude in the


external circuit feeding a capacitive reactance xc and an
inductive reactance xL (IV.2.3.20) connected in parallel,
provided that (o w0 = i/y L C , is called current reso­
nance.
5°. The phenomenon of resonance in an electric circuit
makes radio communication possible and is used when tun­
ing radio receivers to the frequency of a selected radio
station.
Harmful effects resulting from resonance may occur in
electric circuits if excessive currents or voltages appear
in a circuit not designed for operation in resonance condi­
tions (the melting of wires, the breakdown of insulation,
etc.).

9. V a lv e G en er a to r

1 . A value generator is a device intended for the pro­


duction of undamped high-frequency electric oscillations.
It is an electrical auto-oscillating system (IV.1.9.1°). If
we connect the electric circuit of the valve generator shown
in Fig. IV.2.13 to the input of
an electron oscillograph (EO),
then when the switch is closed,
undamped electromagne­
tic oscillations can be seen
on the oscillograph screen.
2°. The components of the
auto-oscillating system in a
valve generator are (IV. 1.9.2°):
the anode battery BA—the
energy source, and the loop
in the anode circuit—the
oscillating system. The function of a valve is performed by
the grid of the triode (111.3.9.1°), which controls the anode
current. The feedback coil, whose ends are connected to the
cathode and the triode grid, is inductively coupled to the
loop coil and is responsible for feedback control of the
oscillating system to the valve.
IV.3.1 E la stic W aves. Sound 359

3°. For the energy of the source to be supplied to the loop


and compensate the damping losses, the oscillations of the
current in the anode circuit of the triode must be coherent
with the free electromagnetic oscillations in the loop
(IV.3.9.8°). The anode current is changed periodically by
changing the potential of the triode grid. This is what en­
sures the part of the triode grid as a valve periodically
opening and closing the access of energy to the loop.

Chapter 3
E lastic W aves. Sound

1. P r e lim in a r y C o n ce p ts
1°. A medium is defined as elastic if forces of interaction
exist between its particles that prevent any deformation of
the medium (11.7.2.1°). For example, the pressure of gases
on the wall of a vessel ensures the ability of the gases to
resist a change in their volume (the cubic, or volume, elas­
ticity of gases). Gases change their form unhindered, i.e.
do not have elasticity of form. Liquids have the same proper­
ties. The forces of interaction between the particles of solids
are so great that the latter have both volume elasticity and
elasticity of form.
2°. If a body oscillates in an elastic medium, then it
acts on the adjoining particles of the medium and makes
them perform forced oscillations (IV.1.8.10). The medium
near the oscillating body is deformed, and elastic forces
(1.2.9.1°) appear in it. These forces act on particles of the
medium that are farther and farther away from the body,
bringing them out of their equilibrium position. Gradually,
all the particles of the medium are involved in oscillating
motion.
The presence of an elastic medium is not an essential
condition for the propagation of any oscillations. For exam-
360 O s c illa tio n s and W aves IV.3.1

pie, electromagnetic oscillations can propagate in a vacuum


(IV.4.1.1°).
3°. Waves are defined as disturbances of the state of
a substance or field that propagate in space with time.
For instance, sound waves (IV.3.7.1°) in gases or liquids are
pressure oscillations propagating in these media.
Electromagnetic waves are oscillations of the intensity
E and induction B of an electromagnetic field (111.1.3.1°)
propagating in space.
4°. By elastic waves are meant mechanical disturbances
(deformations)Tthat propagate in an elastic medium. The
Wavefront

bodies producing these disturbances in the medium are


called wave sources (oscillating tuning forks, strings of musi­
cal instruments, etc.). Elastic waves are sound or acoustic
ones if weak disturbances propagate in an elastic medium,
i.e. the mechanical deformations of the medium cor­
responding to them have small amplitudes (see also
lV.3.7.1°-5°).
5°. A wave surface (or wavefront) is defined as the locus
of points of the medium oscillating in identical phases
(IV. 1.1.4°). The phases of the oscillations of different points
at the moment of time under consideration have indentical
values on a wavefront.
The direction line is defined as a line whose tangent at
every point coincides with the direction of propagation of
the wave. The direction line in a homogeneous isotropic
medium is a straight line perpendicular to the wavefront and
coincides with the direction of transfer of the wave energy
(IV.3.6.2°).
In a plane wave, the wavefronts are planes perpendicular
to the direction of propagation of the wave. The direction
/ V .3 .2 Elastic Waves. Sound 361

lines are parallel lines coinciding with the direction of the


velocity of propagation of the wave (IV.3.3.10). Such waves
can be obtained on the surface of water with the aid of oscil­
lations of a flat bar (Fig. IV.3. la).
Figure IV.3.16 shows the wave-
fronts and direction lines of a
plane wave.
In a spherical wave, the wave-
fronts are spheres. Such waves
are produced by a point
source. The direction lines of a
spherical wave coincide with the
radii of the spheres from the
centre accommodating the wave
source (Fig. IV.3.2) (see also
IV.3.6.30).
6°. Elastic waves in a medium differ from any other or­
dered motion of its particles in that the propagation of the
waves is not associated with transfer of the substance of the
medium from one place to another over great distances.

2. Transverse
and Longitudinal Waves
1°. A wave is transverse if the particles of the medium os­
cillate in directions at right angles to the direction of pro­
pagation of the wave. For example, a transverse wave pro­
pagates along a stretched rubber cord, one end of which is
fixed and the other is made to oscillate (Fig. IV.3.3a). Every
portion of the cord oscillates relative to its unchanged posi­
tion of equilibrium in a direction at right angles to the di­
rection of propagation of the wave (Fig. IV.3.36).
2°. A wave is longitudinal if the particles of the medium
oscillate in the direction of propagation of the wave. The
harmonic oscillations of a piston in a tube filled with a fluid
under the action of elastic forces are transmitted to the fluid
particles, and an elastic longitudinal wave propagates
along the tube (Fig. IV.3.4). It is a system of regions of com­
pression and rarefaction of the medium periodically chang-
362 Oscillations and Waves IV.3.2

ing their states: if at a certain moment there is a rarefac­


tion at one place of the medium and compression at the neigh­
bouring place, then after the time 7V2, where T is the pe-

(«)
Direction of oscillations
__________ D ire c tio n o f w av e ^
------- 1 n Z p r o p a g a tio n
k____
(b)

Fig. IV.3.3

riod of oscillations of the piston, there will be compression


in the first region and rarefaction in the second, etc. The
particles of the medium oscillate in the same direction in

Fig. IV.3.4

which the oscillations are transmitted from one layer to


another, i.e. along the direction of propagation of the wave.

A longitudinal wave appears in a long helical spring (a


slinky) if one end of it is subjected to a periodic external ac­
tion (Fig. IV.3.5). The elastic wave comprises consecutive
I V , 3, 3 Elastic Waves. Sound 363

Compressions and tensions of the spring propagating along


it that periodically, through intervals of 772, replace one
another (T is the period of external action on the spring).
3°. The propagation of transverse waves is impossible
in fluids that have no elasticity of form (IV.3.1.1°). In
solids, the propagation of both longitudinal and transverse
waves associated with the presence of elasticity of form is
possible (for example, the waves propagating along the
strings of musical instruments).

3*. Wave Velocity


1°. The wave velocity (phase velocity) is the physical quan­
tity numerically equal to the distance which any point of
a wavefront (IV.3.1.5°) travels in a unit time. The velocity
vector v is directed along a normal to the wavefront in the
direction of propagation of the wave and in a homogeneous
isotropic medium coincides with the direction line (IV.3.1.5°).
2°. The velocity with which the energy of waves of any
physical nature propagates is finite and cannot exceed the
speed c of light in a vacuum. This follows from the fun­
damental statements of the special^theory of relativity
(V.4.4.40) and conforms to the theory of short-range force
action (111.1.3.2°). These restrictions do not relate to the
phase velocity.
3°. The velocity of elastic sound waves in gases depends
on the absolute temperature of the gas. The velocity of
sound for ideal gases (11.2.1.1°) is

o-V'Tp
where R =
molar gas constant (11.3.3.7°)
T =
absolute temperature
\i =
molar mass (11.1.1.7°)
y =
constant for a given gas depending on the
structure of its molecules.
For air, for example, y = 1.4, and v = 20]/"T. For T =
273 K, we have v = 330 m/s, and for T = 293 K, we have
v = 343 m/s.
3fi4 Oscillations and Waves IV.3.4

The velocity of elastic waves in liquids and of longitu­


dinal waves in solids exceeds the velocity of sound in gases
and depends on the compressibility (elasticity) and
density of the medium:

where K = bulk (compression) modulus (11.7.2.6°)


p = density of the medium.
For example, for water v = 1430 m/s, for copper u =
3910 m/s, for aluminium v = 4880 m/s.

4. Wavelength
1°. A wavefront (IV.3.1.5°) propagates from the source
of the wave during the time At over a certain distance. For
a wave in an isotropic medium, it equals
Ax = v At
where v is the velocity of the wave. This signifies that the
oscillations of the particles of the medium at a distance of
A# from the source occur with a time delay of At, and with
a phase delay of Acp (IV. 1.1.4°), and
Acp __"2ji

because during the period T of the oscillations at the source


the phase changes by 2jt.
2°. The lag in time At and in phase Acp of the oscilla­
tions of points of the medium at the distance x from the
source is

where the quantity % = vT is the wavelength. If Ax =


then Acp = 2jt. The wavelength \ is the distance between the
two closest points oscillating in the same phase, i.e. with
a phase shift of'Acp = 2jt. In other words, the wavelength is
the distance travelled by a wavefront during the time T
equal to the period of oscillations at the wave source.
IV.3.5 Elastic Waves. Sound 365

3°. Tlie wavelength is related to the frequency of oscil­


lations of the wave source as follows:
2nv
X ^vT = —
v O)
where v = wave velocity
_jv = HT = frequency of oscillations at the source
^ " (o = cyclic frequency (IV. 1.1.3°).
The frequency of oscillations depends only on the prop­
erties of the wave source (see, for example, IV. 1.3.3°).
The wave velocity and, consequently, the wavelength, de­
pend on the properties of the medium.

5*. Equation of a Plane Wave


1°. If in a wave source an oscillating quantity changes
according to the law s = A cos (cot + cp) with the ampli­
tude A , the cyclic frequency co, and the initial phase cp, then
the oscillations of the particles of a wavefront of a plane
wave at a point at the distance x from the source lag in
time by At:
sx = A cos [co (t — At) + cp]
It is assumed that the wave propagates without damping.
2°. The equation of a plane (IV.3.1.5°) sinusoidal wave
propagating along the axis OX is
sa: = 4 co s[co (* — v)+<J>]
The quantity
^ co 2jiT 2ji
V V k

is called the wave number. It shows how many wavelengths


can be accommodated on 2ji units of length (cf. IV. 1.1.3°).
The equation of a plane wave has another form:
sx = A cos (gU — kx + <p)
It follows from this equation that:
(a) The amplitude of a plane undamped wave at a given
point of a medium remains constant.
366 Oscillations and Waves tV.3.6

(b) Any point of a medium^# = x0 = const) performs


tlie harmonic oscillations sx = A cos (coJ + a) whose phase
depends on the distance x0 to the given point from the source
of oscillations:
a = <p— kxQ
(c) At a certain moment
(t = ^0 = const), the positions
of the oscillating points of
the medium are described
by the expression
sx = A cos (kx + p)
where |3 = — (a)t0 + <p)-
Figure IV.3.6 shows a plot of this function for t0 = 0
and (p = 0, which is, as it were, an “instantaneous photo­
graph” of a wave.

6*. Energy and Intensity of a Wave.


Equation of a Spherical Wave
1°. An oscillating source of waves has energy
(IV.1.5.30, 4°). When a wave propagates, every particle of
the medium which the wave reaches also oscillates and has
energy. The volume V of an elastic medium in which a wave
propagates with the amplitude A and the cyclic frequency
a) contains the mean energy W equal to
W = — mo)2A2

where m is the mass of the medium in the volume V (cf.


IV.1.5.30).
The mean density (mean volume density) of energy w of
a wave is the energy of a wave concentrated in a unit volume
of the medium:
— W 1 0 ,0
W== — = Y A
where p is the density of the medium.
tV.8.f Elastic Waves. Sound 36 1

2°. The intensity J of a wave is the quantity equal to


the energy transferred on an average by the wave in a unit
time through a unit area of a surface perpendicular to the
direction of propagation of the wave:
_ \
J = wv = -^ pvu>2A2

where v is the velocity of the wave.


The energy and the intensity of a wave are directly pro­
portional to the square of its amplitude.
The power P (mean power) of a wave is defined as the mean
total energy transferred by the wave in a unit time through
a surface of area S. The power P is related to the intensity
/ of a wave as follows:
P = JS
3°. In a spherical wave (IV.3.1.5°), the surface area of
a wavefront grows in direct proportion to r2, where r is
the distance to the source. Hence, the intensity of a spheri­
cal wave diminishes in inverse proportion to r2, i.e. / oc
oc 1/r2. Since J & A 2, it follows that the amplitude of a sphe­
rical wave does not remain constant, but diminishes in
inverse proportion to the radius r of the wavefront, i.e.
A oc 1/r.
The equation of a spherical wave is written in the form
sr = cos (at — kr + <p)

where A 0 is the amplitude of the wave at points of the medi­


um at a unit length from the wave source. For the other sym­
bols, see IV.3.5.1°, 2°.

7. Some Characteristics
of Sound Waves
1°. The branch of physics dealing with the properties
of sound waves (IV.3.1.4°), the laws of their generation,
propagation, and action on obstacles in their way is called
acoustics.
368 Oscillations and Waves 1V.3.7

Sound waves with frequencies from 16 to 2 X 104 Hz


act on a person’s ears, call forth feelings of hearing, and are
called audible sounds. Sound waves with frequencies below
16 Ilz are called infrasounds, and above 2 X 104 Hz, ultra­
sounds.
2°. The perception of sound by the ear^depends on the
frequencies which the sound wave is composed of. Sounds
forming a set of frequencies continuously filling a certain
interval (a continuous spectrum of frequencies) are called
noise.
Musical sounds (tones) have a line spectrum]of frequencies:
the frequencies v* in musical sounds form a number of dis­
crete (intermittent) values. Periodic or almost periodic
oscillations correspond to musical sounds.
3°. Every sinusoidal sound wave is called a tone (a
simple, or pure, tone).
The pitch of a tone depends on its frequency: the greater
the frequency, the higher is the tone.
4°. The tone corresponding to the lowest frequency in
a set of frequencies of a given sound is called the fundamen­
tal tone. The tones corresponding to the remaining frequen­
cies in the sound are called overtones. If the frequencies of
the overtones are multiples of the frequency v0 of the funda­
mental tone, then the overtones are called harmonics. The
fundamental tone of frequency v0 is called the first harmonic,
the overtone with the next frequency 2v0 the second harmonic,
etc.
Musical sounds with the same fundamental tone are
distinguished by their timbre (quality). The latter is deter­
mined by the presence of overtones—their frequencies and
amplitudes, the nature of the growth of the amplitudes
when the sound begins, and their attenuation when it termi­
nates.
5°. The loudness of sound depends on its intensity,
i.e. is determined by the amplitude of the oscillations in
a sound wave (IV.3.6.2°). The human ear has the maximum
sensitivity to sounds of frequencies ranging from 700 to
6000 Hz. In this range, the ear can perceive sounds having
an intensity of about 10~12-10-11 W/m2.
The audibility thereshold is the lowest intensity of
a sound wave that can be perceived by the ear. The standard
I V . 3 .8 Elastic Waves. Sound 369

audibility threshold j^is taken equal to J 0 = 10“12 W/m2


at a frequency of v0 = 1 kHz.
The pain threshold is defined aS the maximum intensity
of a sound wave at which perception of the sound does not
cause a feeling of pain. The pain threshold depends on the
frequency of the sound and varies from 0.1 W/m2 at
6000 Hz to 10 W/m2 at low and high frequencies.
6°. A measure of the sensitivity of the ear to the percep­
tion of sound waves of a given intensity J is the sound inten­
sity level L :

L = 101og-^-
J0
where J 0 is the standard audibility threshold (Par. 5°).

8. Ultrasounds
1°. Sound waves of frequencies ranging from 2 X 104
to 1013 Hz are called ultrasounds. Ultrasounds of frequencies
of 109 Hz and above are also called hypersounds. Ultrasounds
are generated by mechanical and electromechanical emit­
ters (oscillators). A siren is a mechanical emitter of low-
frequency ultrasonic waves (with a frequency ranging from
about 20 to 200 kHz) of a high intensity. It emits sound owing
to the periodic cutting off of a powerful stream of compressed
air or steam passing through openings in two coaxial disks,
of which one is stationary and the other rotates.
2°. Magnetostriction and piezoelectric electromechani­
cal oscillators are employed most often.
Magnetostriction oscillators are used for generating ultra­
sounds of frequencies of up to 200 kHz. The design of these
oscillators is based on the phenomenon of magnetostriction—
the change in the shape and volume of ferromagnetics
(111.6.5.1°) placed in a varying magnetic field. If a ferro­
magnetic is magnetized in a periodically varying magnetic
field, then forced mechanical oscillations are set up in it
that are a source of ultrasound. The ferromagnetic bar used
as the core of a high-frequency transformer is a very simple
ultrasonic magnetostriction oscillator.
24-0211
370 Oscillations and Waves iv.$.d

3°. Piezoelectric oscillators emit ultrasounds of frequen­


cies of up to 50 MHz. The action of these oscillators is based
on a phenomenon consisting in that certain crystals, for
instance quartz, change their linear dimensions under the
action of an electric field. A plate of such a piezoelectric*
in a varying electric field performs forced mechanical os­
cillations that generate ultrasounds.
4°. Ultrasounds are used in engineering for control and
measuring purposes, and also for carrying out and accel­
erating some production processes.
Sonar [see radar (IV.4.5.6°)] is based on determining the
distance to a body in water by measuring the time between
the sending of an ultrasonic signal and the reception of an
echo-signal produced as a result of reflection of the ultrasound
from a body. Sonar takes advantage of the absorption of
ultrasound by liquids. In air, this absorption is a thousand
times greater than in water.
Ultrasonic flaw detection is the tracing of internal defects
(cracks, cavities, structural heterogeneities) in solids with
the aid of ultrasound. Such flaw detection is based on the
different reflection of ultrasound from damaged and un­
damaged parts of a body.
5°. Ultrasonic waves of a sufficient intensity have a com­
minuting action on bodies. This is used in various produc­
tion processes such as the preparation of emulsions, the re­
moval of oxide films, the degreasing of component surfaces,
and the comminution of photographic emulsion grains.
Ultrasounds accelerate diffusion processes and some chemi­
cal reactions.

9*. Interference
1°. If two point sources produce spherical waves
(IV.3.6.3°) in a homogeneous and isotropic medium, then
at an arbitrary point of space M superposition of the waves
* The word “piezo” is from the Greek piezein meaning to squeeze
press. Piezoelectrics are crystals in which the direct and inverse piezoelec­
tric effects are possible. The direct effect consists in the appearance of
electric charges on the boundaries of certain crystals when they are
compressed or stretched. The inverse effect consists in the appearance
of deformations when such crystals are introduced into an electric field.
IV.3.9 Elastic Waves. Sound371

occurs in accordance with the superposition principle: ev­


ery point of a medium at which two or more waves arrive
participates in the oscillations produced by each wave se­
parately. The waves do not interact and propagate inde-
dendently of one another. For example, the waves propaga­
ting in water from the two point sources depicted in

Fig. IV.3.8

Fig. IV.3.7 reach the point Af, and each of them, inde­
pendently of the other, causes it to oscillate.
2°. Two simultaneously propagating sinusoidal spher­
ical waves s1 and s2 set up by the point sources Bx and B 2
(Fig. IV.3.8) produce an oscillation at the point M that
in accordance with the superposition principle is described
by the formula s = s± -f s2. According to IV.3.6.30, we
have
= sin (o)^ — k{ri -f- o&i) = ■^ sin
A A
$2= — sin (g)2£— Jt2 r2 + a 2) = — sin 0>2
*2 r2
where Oj = co^ — k^ + a x and = (o2t — k2r2 -f a 2
are the phases of the propagating waves. For the other sym­
bols, see IV.3.5.20.
In the resultant wave, s = s± + s2 = (Air) sin O, the
amplitude A ir and the phase O are determined by the for­
mulas

— - s in d > H -------S i n 02
O = arc tan — ^ ---------
— — c o s <DX~\— cos 0 2
ri r*%
24*
372 Oscillations and Waves IV.3.9

3°. Waves and the sources producing them are called


coherent if the phase difference of the waves 0 2 — does
not depend on the time. Waves and the sources producing
them are called incoherent if the phase difference of the
waves <P2 — Oj changes with time. The formula for the
phase difference is
$2 - 0 1 = (o)2—co4) t — (k2r 2 — kiT i) + (a2—a {)
where A1= a)1/i;, k2 = (&2/v, v is the wave velocity (IV.3.5.2°),
which is identical for both waves in a given medium. Only
the first term is time dependent in the above expression.
Two sinusoidal waves are coherent if their frequencies are
the same (cOi = cd2), and incoherent if their frequencies are
different.
4°. For coherent waves (a)! = co2 = co), provided that
a 2 — a i = 0,
O z — d > i = -------^ - ( r 2 — r j ) = — k ( r 2— r t)

4=
The amplitude of the resulting oscillations at any point of
the medium is independent of the time. The cosine equals
unity, and the amplitude of the oscillations in the resultant
wave is maximum (Air = A J rx + A 2lr2) at all the points
M of the medium for which k (r2 — rx) = 2mn, where
m = 0, ± 1 , ± 2, . . ., or, since k = 2nIK (IV.3.5.20),
r2 — rx = m%
The quantity r2 — rx = A is called the geometrical
path difference of the waves from their sources Bx and B z
to the point of the medium being considered (see Fig. IV.3.8).
The amplitude of the oscillations in the resultant wave
is minimum (Air = | Ajr-^ — A 2lr2 | ) at all the points
of the medium for which
k (r2 — rx) = (2m — 1) Jt
where m = 1 ,2 , . . ., or
X
IV.3.10 Elastic Waves. Sound 373

Upon the superposition of coherent waves, the square


of the amplitude and the energy of the resultant wave dif­
fer from the sum of the squares «of the relevant quantities
for the waves being superposed.
5°. The interference of waves is defined as the phenomenon
of wave superposition when they amplify one another at
some points of space and weaken one another at other points.
The result of interference depends on the phase difference
of the superposed waves.
Only coherent waves (Par. 3°) with oscillations in the
same direction can interfere. For example, two spherical
waves on the surface of water propagating from two coherent
point sources (see Fig. IV.3.7) produce a resultant wave upon
interference. The wavefront (IV.3.1.50) of the resultant wave
will be a sphere. (Figure IV.3.7 shows plane sections of
hemispherical wavefronts.)
When waves interfere, their energies are not summated.
The interference of waves leads to the redistribution of the
energy of oscillations between different close particles of
the medium. This does not contradict the law of conserva­
tion of energy because, on an average, for a large region of
space, the energy of the resultant wave equals the sum
of the energies of the interfering ones.
6°. When incoherent waves are superposed, the mean
value of the square of the amplitude of the resultant wave
equals the sum of the squares of the amplitudes of the super­
posed waves. The energy of the resultant oscillations of
each point of space equals the sum of the energies of its
oscillations due to all the incoherent waves taken separa­
tely.

10*n Standing Waves


1°. Standing waves are a particular case of wave inter­
ference. A standing wave in the simplest case is formed by
l lie superposition of two waves propagating in mutually
opposite directions if the interfering waves comply with the
following conditions: their frequencies, amplitudes, and di­
rections of oscillations must be the same. The interfering
waves, unlike the standing one, are called running waves.
374 Oscillations and Waves IV.3.10

A standing wave is formed in a cord that is fixed at one


end when the other end receives periodic oscillations

(Fig. IV.3.9). A standing wave may also be formed in


a column of a gas in a tube of a definite length.
2°. The oscillations of the arbitrary point M
(Fig. IV.3.10) at the distance x from the unfixed end of

M 0
X I
" 1 X
c
L ---------- >■
--------- -----
Fig. IV.3.10

a cord of length I are described by the equation of a plane


standing wave:
s = 2A cos £ft(Z —#) + y ] sin [ a t — k l—

The equations of the incident sx and reflected s2 waves have


the form (IV.3.5.20)
sf = A sin (g)£ — kx) and s2 = A sin [cd2 + k (x — 21) — cp]
In the reflected wave, the displacement s2 of the point M
lags in phase behind by the amount a = 2k (I — x) +
where cp is the additional lag in phase that may appear in
reflection (Par. 3°).
3°. The amplitude Ast of a standing wave is independent
of the time and is a periodic function of the distance x
from the points of the cord to the source of the waves:

Asi = 2A cos [/c(Z — *) + - y ] |


IV. 3.10 Elastic Waves. Sound 375

The points at which the amplitude A st vanishes are called


the nodes of the standing wave (the points D, D lT D 2, etc.
in Fig. IV.3.9). At these poinfe

k (l — x) -f- -y- = (2m + 1) - y

where m = 0, ± 1 , ± 2 , . . . .
The points at which the amplitude of oscillations is
maximum and equals 2A are called the antinodes of the stand­
ing wave (the points C, C2, etc. in Fig. IV.3.9). At
these points
k ( l - x ) + ^ -== 2m ^ -

When a wave is reflected from the boundary with a dens­


er medium (see Fig. IV.3.9), the phase changes by jt,
and the “loss of a half-wave” occurs. A phase shift of n
corresponds to a change in the phase of the oscillations dur­
ing the time 772 in which a running wave covers a distance
of XI2.
4°. The length of a standing wave Xst is defined as the
distance between two adjacent nodes or antinodes:

equal to half the length of a running wave. The distance


between a neighbouring node and antinode of a standing
wave is XS\I2 = XIA (see Fig. IV.3.9).
5°. All the points of a standing wave between two adja­
cent nodes oscillate with different amplitudes, but in the
same phase, whereas in a running wave, on the contrary,
all the points oscillate with the same amplitudes, but in
different phases.
6°. Unlike a running wave, there is no energy transfer
in a standing wave—the energy of the oscillations of every
(dement in the volume of the medium confined by a neigh­
bouring node and antinode does not depend on the time.
It periodically transforms from kinetic energy into the
potential energy of the elastically deformed medium and
vice versa. The absence of energy transfer in a standing wave
is explained by the fact that energy is transferred in equal
376 Oscillations and Waves IV.4.1

amounts in opposite directions in the falling and the re­


flected waves forming it.
7°. Standing waves are formed in a cord, rod, or in other
media of a restricted length only at definite frequencies
called the natural frequencies of oscillations of the relevant
bodies. Depending on how the ends of bodies are fixed,
either nodes or antinodes of a standing wave are formed on
their ends. Hence, one of two conditions is observed:

I = mhst
where m = 1, 2, 3.......... Since A,st =XI2 (Par. 4°), or Xsi =
= v/2v where v is the velocity of elastic waves of frequency v,
then the natural frequencies of standing waves are found
from the conditions
v = (2m — or v = 2m ^-

To change the natural frequencies of oscillations of


strings or columns of a gas, their length I must be changed.
This is taken advantage of when playing string and wind
instruments.

Chapter 4
Electrom agnetic W aves

1. Relationship Between Varying


Electric and Magnetic Fields
1°. A varying electromagnetic field (111.1.3.1°) propagat­
ing in space is called an electromagnetic wave. An electro­
magnetic wave is characterized by the electric field inten­
sity vector E (111.1.3.3°) and the magnetic field induc­
tion vector B (111.4.1.6°), the two fields composing a single
I V .4.2 Electromagnetic Waves 377

electromagnetic field. Electromagnetic waves are divided


into radio waves (IV.4.5.1°) and light waves (V.1.1.1°).
2°. Electromagnetic waves gan exist due to the fact
that there is a relationship between a varying electric and
a varying magnetic fields. A
varying magnetic field sets
up a vortex electric field
(111.5.3.3°). The reverse phe­
nomenon is also observed: an
electric field in a vacuum or £“<L
a dielectric varying in time
sets up a vortex magnetic m
field. Figure IV.4.1 shows the
appearance of an electric E
and magnetic B vortex fields (a) (b)
when the magnetic and the
electric fields change, respec­ Fig. IV.4.1
tively.
3°. A varying electric field in a vacuum or a dielectric
is characterized by the vector of the displacement current
density ;dls. In a homogeneous isotropic medium
. _ e0e AE
3dls~ ~
where e0 = electric constant in the SI system (VI1.5.4°)
e = relative permittivity of the medium (111.1.2.6°).
The name “displacement current”* stresses that, simi­
lar to a conduction current (111.2.1.1°), a displacement cur­
rent sets up a vortex magnetic field.

2. Velocity and Some Basic Properties


of Electromagnetic Waves
1°. The velocity (phase velocity) v (IV.3.3.1°) of an
electromagnetic wave in a medium is determined from
Maxwell’s formula

__________________ Y*V
* The word “displacement” in the term “displacement current”
is associated with the fact that the current density j is also determined
as jdis = AD/A*, where the vector quantity D = e0eE is called the
electric displacement vector. The vector D is not treated in an elementary
course of physics. *
378 Oscillations and Waves IV.4.2

where e = relative permittivity of the medium (111.1.2.6°)


|i = relative permeability of the medium (111.4.3.2°)
c = speed of light in a vacuum.
The velocity of electromagnetic waves in a given medium
coincides with that of light in this medium. This coinci­
dence is not by chance and is one of the substantiations of
the electromagnetic nature of light (V.2.1.1°).
2°. The relative permeabilities of all non-ferromagnetic
media, i.e. dia- and paramagnetics (111.6.2.1°), differ only
slightly from unity; therefore in such media

and is a function of the refractive index n because e = ri*


(V.1.2.1°).
3°. The relative permittivity of a substance in a varying
electric field depends on the frequency v of oscillations of

the field: e = / (v). The phenomenon of dispersion of elec­


tromagnetic waves—the dependence of the velocity v of
these waves on the frequency of oscillations of a varying
electromagnetic field, i.e. v = cp (v)—exists in all substances.
Dispersion is absent only in a vacuum for which e = 1
and does not depend on the frequency (see also V.2.6.1°).
4°. Electromagnetic waves are transverse ones (IV.3.2.1°).
The oscillations of the intensity vector E of the varying elec­
tric field and the induction vector B of the varying magne­
tic field in an electromagnetic wave are mutually perpen­
dicular and are in a plane perpendicular to the wave velo­
city vector v. The vectors v, E, and B form a right-handed
IV.4.2 Electromagnetic Waves 379

system: from the tip of the vector v the rotation from E to


B through the smallest angle can be seen to occur counter­
clockwise (Fig. IV.4.2). •
5°. An electromagnetic wave is called monochromatic*
if its vectors E and B perform harmonic oscillations of a
constant identical frequency called the wave frequency.
The equations of a plane (IV.3.5.2°) monochromatic wave

propagating along the axis OX (the oscillations of the vectors


E and B occur correspondingly along the directions of the
axes OZ and OY) are
Ex = 0, Ey = 0, Ez = E0 cos (a t — fcr + <p)
Bx = 0, Bz = 0, By = B0cos (a>t — kx-\~q>)

where E 0 and B 0 = amplitudes of the electric field inten­


sity and the magnetic field induction in
the electromagnetic wave, respectively
co = cyclic frequency of the wave
k = colv = wave number
cp = initial phase of the oscillations of E
and B at points of the coordinate plane
ZOY (x = 0) (Fig. IV.4.3).
6°. The magnitudes of the vectors E and B in a plane
electromagnetic wave (IV.3.1.5°) are related by the expres­
sion
e0eE 2 = ——
VoV

* From the Latin monochromatos meaning consisting of one colour.


For the non-monochromatic nature of light waves in principle, see
VI.1.5.20*
380 Oscillations and Waves IV.4.3

In a vacuum, © = jm = 1, and B = \i0 E = Elc, where


c = i / y e 0 m is the speed of light in a vacuum, e0 and p0
are the electric and the magnetic constants in the SI system
(VII.5.1°,3°).
The mutually perpendicular vectors E and B in an elec­
tromagnetic wave propagating in free space oscillate in an
identical phase—they simultaneously vanish and simulta­
neously reach their maximum values (E 0 == | Z?max | and
B0 = Ve0|J.0e fi# 0).
3. Energy and Intensity
of Electromagnetic Waves
1°. The propagation of electromagnetic waves is asso­
ciated with the transfer of the energy of the electromagnetic
field of a wave like the propagation of elastic waves in a sub-
stance is associated with the transfer of energy (IV.3.6.2°).
Energy is transferred in the direction of wave propagation,
i.e. in the direction of the vector v. The volume density of
the energy of the electromagnetic field of a wave (111.5.7.5°)
is
e0e £ 2 1 B2
w = we+ w m 2 2 p0p
But, according to IV.4.2.6°, we have we = wm, and
_ EB

~"PoP
where v = c /Y ep = velocity of the electromagnetic wave
(IV.4.2.1°)
c = i / Y e0p0
e0 and p0 = electric and magnetic constants in the SI
system (VII.5.1°, 3°).
2°. The intensity of an electromagnetic wave (not to be
confused with the electric field intensity) is determined
similar to the intensity of an elastic wave (IV.3.6.2°) and
is expressed by the formula
EB
J = wv
1V.4.4 Electromagnetic Waves 381

The intensity of an electromagnetic wave is proportional


to the mean value of the product of the magnitudes of the
electromagnetic field vectors E and B , i.e. is proportional
to the square of the electric field intensity E : J <x E2.

4*. Emission
of Electromagnetic Waves
1°. Electromagnetic waves are produced by electric
currents varying in time, and also by separate electrically
charged particles moving with acceleration.*
The process of the production of electromagnetic waves
by a source is called the radiation, or emission, of electromag­
netic waves, and the source itself is called the emitting system,
or emitter. The electromagnetic field of a wave is called a
radiation field.
2°. The emission of electromagnetic waves by a source
is accompanied by its emitting energy into the surrounding
space.
The mean radiated power P is defined as the mean energy
which a source of electromagnetic waves emits in a unit
time in all directions. The mean power P is related to the
intensity of electromagnetic waves as follows: P = JS
(cf. IV.3.6.2°).
A wave zone is a region of space whose distance from the
source of radiation considerably exceeds the linear dimen­
sions of the source and the wavelength % of the waves it
emits.
3°. An electric charge q that is at the origin of coordi­
nates at t = 0 and moves in a vacuum along the axis OZ with
the acceleration a is a source of an electromagnetic wave.
The magnitudes of the vectors of the intensity E and the
induction B of the fields of this wave at the point M

* Electromagnetic waves may be produced in a substance by


charged particles moving uniformly with the speed v exceeding the
speed of light c!n in the given substance, where n is the refractive in­
dex of the substance (V.1.2.1°). The consideration of this radiation
(Vavilov-Cerenkov radiation) is beyond the scope of a course in elemen­
tary physics.
3S2 Oscillations and Waves t V . 4 .4

(Fig. IV.4.4) of the wave zone are determined by the for­


mulas
j? _ M a sin 0
JLwaye~ ~ W r
E_ |x0qg sin 6
^wave — Y e oM'0® c 4ncr
where c = \ l Y e0p,0. The directions of the vectors Ewave
and Bwayg, the angle 0, and the distance r are shown in
Fig. IV.4.4.
The mean radiated power P of a charge moving with
acceleration is
77 _ M 2*2
6nc
The mean radiated power is directly proportional to the
square of the acceleration of the charge (see Fig. IV.4.4).
4°. The electric charge
q performing harmonic os­
cillations (IV. 1.1.4°) along
the axis OZ relative to the ‘
origin of coordinates (see
Fig. IV.4.4) according to
the law z = A cos cot,
where A is the amplitude
of the oscillations of the
charge, and co is their cyc­
lic frequency, is a source
of a spherical electromag-
Fig. iv.4.4 netic wave (IV.3.6.3°). The
magnitudes of the field
vectors E and B in the wave zone are described by the
equations
E = M ^ — -? C 0 S ( v t - k r + n)
B = ^ Af s* * cos( M - k r + n)
The mean radiated power of an oscillating charge is
77 _ VoQ2*2®*
12 jic
The mean radiated power is directly proportional to the
fourth power of the frequency of oscillations of the charge,
I V .4 .4 Electromagnetic Waves 383

hence the radiated power increases very rapidly with a grow­


ing frequency of oscillations.
5°. An example of an emitter #(Par. 1°) is an electric dipole
whose electric moment p e = ql (111.1.4.6°) performs har­
monic oscillations according to the law p e = p 0 cos cat,
where p 0 is the amplitude of oscillations of the dipole moment,
and to is the cyclic frequency of oscillations of the dipole.
The fields of the electromagnetic wave emitted by the dipole
and the mean radiated power are described by the formulas
of Par. 4° in which p 0 should be introduced instead of qA.
6°. A Hertz oscillator is an example of an emitter. It
consists of two metal rods with identical spheres A and B
on their ends and a small spark gap C between them
(Fig. IV.4.5). The capacitance
of the oscillator is determined
by the capacitances of the
spheres, and the inductance by
the inductances of both halves
o CJ
of the rods. Electromagnetic
oscillations are produced in the T ~T
oscillator by the induction
coil IC comprising a high-freq­ IC
uency transformer (111.5.6.3°).
The secondary of the coil IC Fig. IV.4.5
is connected to the spark gap.
When the varying voltage in the secondary of the coil
reaches the value of the breakdown voltage (111.1.6.1°), a
spark passes through the spark gap, and damped high-fre­
quency electromagnetic oscillations (IV.2.1.5°) attended by
the emission of electromagnetic waves are set up in the os­
cillator. Unlike a conventional oscillator circuit (IV.2.1.1°)
in which a varying electromagnetic field is concentrated in
a small region of space (IV.2.1.4°), the varying electromagnet­
ic field propagates from a Hertz oscillator through the en­
tire space surrounding it. Therefore, this oscillator is an
open oscillator circuit.
Electromagnetic waves are registered (received) by a
similar oscillator—a resonator, in which the action of the
electromagnetic field of a wave induces forced electromagnet­
ic oscillations (IV.2.2.1°). If the frequencies of the oscilla­
tions in the vibrator and the resonator are the same, then
384 Oscillations and Waves IV .4.4

electrical resonance (IV.2.8.2°) sets in, and the forced oscil­


lations in the resonator are detected either according to the
passage of a spark through its spark gap or according to the
glowing of a small gas-discharge tube connected to the spark
gap (Fig. IV.4.6). H. Hertz used such a system to conduct a
series of experiments in which he discovered the existence of
electromagnetic waves, their transverse nature (IV.4.2.4°),
and observed the phenomenon of interference of electromag­
netic waves. Hertz obtained standing electromagnetic waves
(IV.3.10.1°) and determined the speed of propagation of
electromagnetic waves (IV.4.2.1°)
with their aid.
7°. A ^Hertz oscillator whose
linear dimensions I are small in
comparison with the wavelength
it emits (I is called a Hertz
dipole. The emission of a Hertz
dipole is similar to that of the di­
pole treated in Par. 5°, the only dif­
ference being that the varying
Fig. IV.4.6 electric moment p e of the dipole
is set up by oscillations of the
charge q of the oppositely charged spheres A and B (see
Fig. IV.4.5), and not by periodic changes in the distance
between them. The current I = I 0 sin cot is supplied to a
Hertz dipole, and it is considered identical at a given
moment in the entire circuit (a quasi-stationary current).
The mean radiated power
of a Hertz dipole is
p _Voiwil
12nc
The emission of a dipole
is not the same in different
directions. Figure IV.4.7
shows a diagram of the
intensity of the electro­
magnetic waves emitted by a dipole at various angles
0 to its axis. The position vector J (0) in the figure character­
izes the intensity of radiation at the angle 0. A glance at
the figure shows that the dipole emits the greatest amount
TV.4.5 E le c tro m a g n e tic W a v e s 385

of energy at the angles 0 = jt/2 and 3jt/2, i.e. in a plane pass­


ing through its middle at right angles to its axis. A dipole
emits no electromagnetic waves «along its axis (0 = 0, jt).

5. Concept of Radio Communication,


Television, Radar,
and Radio Astronomy
1°. Radio, or wireless, communication is the transmission
of information with the aid of radio waves—electromagnetic
waves whose frequencies cover a broad range from 3 X 104
to 3 X 1011 Hz.|j Radio waves are divided into the groups
given in Table IV.4.1.
Table IV.4.1
C la s s if ic a t io n o f R a d io W a v e s

Frequency range, Range of wave­


Name Hz lengths (in
vacuum), m

Ultralong <3X 10* > 10 000


Long 3 X 104 -3 X 105 10 000-1000
Medium 3 X 105 - 3 X 106 1 000-100
Short 3 X 106 -3 X 10? 100-10
( metre 3 X 10’ .3 x 108 10-1
J decimetre 3 X 108 _3 x 109 1-0.1
Ultrashort ] centimetre 3 X 10» -3 X 10‘0 0.1-0.01
I* millimetre 3 X 1 0 * 0 -3 X 1 0 1 1 0.01-0.001

Radio broadcasting is used to transmit speech and music,


and television to transmit images.
2°. Monochromatic waves (IV.4.2.5°) are not suitable
for transmitting definite signals by radio. Radio communi­
cation is accomplished with the aid of modulated radio
waves. By the modulation of an electromagnetic wave is meant
a change in its parameters (amplitude, frequency, initial
phase) with frequencies considerably lower than that of the
electromagnetic wave itself. The frequency of the initial
386 Oscillations and Waves IV .4.5

(unmodulated) wave is called the carrier frequency, and the


frequency of the change in the wave parameters in modula­
tion, the modulation frequency.
3°. A diagram of a modern radio transmitter is shown
in Fig. IV.4.8a. The generator of undamped oscillations
(the driving oscillator) (IV.2.9.1°) produces high-frequency
oscillations of the carrier frequency shown in Fig. IV.4.9a.
The sound oscillations (Fig. IV.4.96) enter the microphone
and are converted into electric oscillations. In the modula­
tor, the undamped sinusoidal oscillations are converted into

Driving Modu­ -> Ampli­ Detec­ Ampli­


oscillator lator fier tor fier

Microphone D ynam ic
lo u d s p e a k e r

(a) (b)
Fig. IV.4.8

modulated oscillations (Fig. IV.4.9c). After amplification,


the modulated oscillations are fed to the transmitting ae­
rial (antenna), which emits electromagnetic waves.
4°. A radio receiver is shown schematically in Fig. IV.4.86.
The electromagnetic waves are picked up by the aerial of the
receiver and set up electromagnetic oscillations in the reso­
nance circuit RC. The weak high-frequency oscillations pass
through an amplifier into a detector—& conductor with uni­
polar conduction (for example a two-electrode valve, see
111.3.8.1°). The detector is responsible for demodulation—
separation of the low-frequency component of the oscillations
from the carrier-frequency oscillations. The low-frequency
(sound) component (Fig. IV.4.9e) is separated from the de­
tected oscillations (Fig. IV.4.9d), and is then amplified and
fed to a sound reproducer (loudspeaker, earphone, etc.).
The resonance circuit of the receiver consists of a coil
and a variable capacitor (111.1.11.4°). This makes it possible
to achieve coincidence of the frequencies of oscillations of
the circuit with the frequency of the waves emitted by a
I V . 4 .5 Electromagnetic Waves 387

definite radio broadcasting station. High-quality^reproduc-


l ion in the receiver of the signals transmitted by a radio sta­
tion requires that the modulation frequency be from one-
fifth to one-tenth of the carrier frequency. For the transmis­
sion of speech and music, the modulation is performed with
sound frequencies usually not exceeding from 10 to 13 kHz.
All the ranges of radio waves beginning from the long ones
can be used for radio broadcasting. In practice, radio broad­
casting stations use the ranges of long, medium, and short
radio waves (see Table IV.4.1).
5°. The block diagram of tele­
vision is basically similar to
that of radio broadcasting (see
Fig. IV.4.8a and b). In the
transmitter, the oscillations of
the carrier frequency are modu­
lated not only with the sound
signal, but also with the preli­
minarily amplified image signals
supplied from transmitting tubes
(iconoscopes or superorthicons).
The volume of modulation also
includes signals for synchroniza­
tion of the scanning of the elec­
tron beam in the cathode-ray
tube (111.3.10.3°)—the icono­
scope on whose screen the image
appears. In a television receiver,
the high-frequency signal is divid­
ed into three: the image signal,
the accompanying sound, and
the control signal.
These signals after amplifica­
tion are fed to their relevant
channels and are used in accor­
dance with their purpose. The
Fig. IV.4.9
control signals synchronize the
operation of the oscillators re­
sponsible for scanning of the electron beam horizontally
(111.3.10.4°) — along the lines — and for transferring it from
one line to the next one. During only l/25th of a second, the
388 Oscillations and Waves I V . 4 .5

electron beam registers 625 lines forming one frame. If


there is no image signal, the screen is illuminated uniformly.
The amplified image signal is fed to the grid of the electron
gun (111.3.10.3°). The intensity of the electron beam changes,
and together with it the brightness of the given point of the
screen where the image appears. A television signal carries
a large volume of information and occupies a frequency band
of the order of magnitude of from 4 to 5 MHz (in a radio
receiver about 10 kHz). High frequencies, from 50 to 900 MHz
(which correspond to wavelengths from 6 m to 30 cm),
are used as the carrier frequencies of the electromagnetic
waves (compare with Table IV.4.1).
6°. Radar is the detection and determination of the loca­
tion of various objects with the aid of radio waves (Par. 1°).
Radar is based on the reflection and scattering of radio
waves by bodies.
Radar equipment is a combination of an ultrashort wave
(see Table IV.4.1) radio transmitter and a radio receiver
having a common receiving and transmitting aerial that
produces sharply directed radiation (a radio beam). The
radiation is in short pulses each lasting about 10“6 s. In
the interval between two consecutive pulses of radiation,
the aerial is automatically switched over to reception of
the signal reflected from an object or target. The distance to
the object, its location, is determined according to the time
interval At between the emission of a signal and the recep­
tion of the reflected signal. Radar is most effective when
d ^ X, where d signifies the linear dimensions of the bodies
being located. Therefore, radar uses ultrashort radio waves
of the decimetre, centimetre, and millimetre ranges (see
Table IV.4.1). In radar astronomy, the radar methods are
used for determining more precisely the motion of the plan­
ets of the solar system and their satellites, artificial satel­
lites of the Earth, spaceships, etc.
7°. Radio astronomy is the branch of physics and astro­
nomy that studies cosmic objects by registering their own
ultrashort-wave radio emission [mainly in the range of cen­
timetre and decimetre waves (see Table IV.4.1), which are
absorbed only slightly on their way from the objects to the
Earth].
Special radio telescopes are used to receive and study the
I V.4.5 Electromagnetic Waves 389

radio emission of cosmic objects. Their sensitivity, owing


Io the large effective areas of their aerials, considerably ex­
ceeds that of the largest modern optical telescopes (V. 1.7.11°).
The methods of radio astronomy make it possible to study
Ihe physical properties of the surface layers of the solar-
system planets and their temperatures. Investigation of the
radio emission of the Sun makes it possible to predict changes
in solar activity and other important optical properties.
Methods of radio astronomy are the only possible means for
sludying the nucleus of the Galaxy, and also radiogalaxies—
parts of the metagalaxy that are very far from the Earth and
cannot be observed in optical telescopes.
OPTICS

Chapter 1
Geometrical (Ray) Optics

1. The Rectilinear Propagation


of Light
1°. Optics is the branch of physics that studies the phe­
nomena and laws associated with the generation, propaga­
tion, and interaction with a substance of electromagnetic
light waves (IV.4.1.10). Light waves occupy an enormous
range on the scale of electromagnetic waves (V.3.7.10). It
is after the ultrashort millimetre radio waves and extends
up to the shortest electromagnetic waves known at present—
gamma rays with a wavelength X less than one angstrom
(A) (1A = lO"10 m).
2°. Geometrical {ray) optics treats the laws of propagation
of light in transparent media on the basis of notions of light
as a combination of light rays (IV.3.1.5°)—lines along which
the energy of light electromagnetic waves propagates. Geo­
metrical optics does not take into consideration the wave
properties of light and the diffraction phenomena (V.2.3.10)
associated with them. For instance, when light passes
through a lens (V. 1.5.1°) with a mount diameter of X,
V .J .2 Geometrical (Ray) Optics 391

where X is the length of the light wave, we may ignore the


diffraction at the edges of the lens.
3°. A medium is defined as Optically homogeneous if its
refractive index (V .l.2.20) is everywhere the same. In an
optically homogeneous me­
dium, the rays are straight: in
such a medium light propa­
gates rectilinearly. This is con­
firmed by the formation of sha­
dows. If S is a point source
of light, and A is a body in
the path of the light from the
source, then a cone of a shadow Fig. V.1.1
is formed beyond the body
K (Fig. V.1.1). No point inside this cone receives light from
the source. A well-defined shadow of the body K is obtained
on a screen placed at right angles to the axis of the cone.
Beams of light rays upon intersecting do not interfere
(V.2.2.10) and propagate independently of each other after
intersection.

2. L a w s o f R e fle c tio n
an d R e fr a c tio n of L ig h t.
T o ta l R e fle c tio n
1°. The ratio of the speed of light in a vacuum to its
speed v in a given medium
n= = Y^s\i y^e

is called the absolute refractive index of this medium. Here


e and p are the relative permittivity (111.1.2.6°) and the
relative permeability (111.4.3.2°) of the medium. For non­
ferromagnetic media, p « 1. For any medium except a
vacuum, we have n > 1. The refractive index n depends on
the frequency of the light (V.2.6.10) and the state of the
medium (its density and temperature). For gases in standard
conditions, n « 1. In anisotropic media (11.7.1.1°), n de­
pends on the direction of propagation of the light and its
polarization (V.2.5.1°).
392 O p tics V.1.2

2°. The relative refractive index n21 of a second medium


relative to a first one is defined as the ratio of the speeds of
light and v2 in the first and second media, respectively:

where nx and n2 are the absolute refractive indices of the


first and second media, respectively.
If n21 > 1, then the second medium is called optically
more dense than the first one.
3°. When light rays strike an ideally flat interface be­
tween two media whose dimensions considerably exceed
the wavelength, reflection and refraction of the light occur.

(b)
Fig. V.1.2

The direction of propagation of the light changes when it


passes into the second medium, except when the rays strike
the interface at right angles to it. The angle of incidence i
is the angle between the incident ray and a normal to the
interface erected at the point of incidence. The angle of
reflection i' is the angle between the reflected ray and the
same normal (Fig. V.1.2). The angle of refraction r is the
angle between the refracted ray and the same normal.
4°. The laws of reflection of light: *
(a) The incident ray, the reflected ray, and the normal
to the interface of two media erected at the point of incidence
of the ray all lie in the same plane.
(b) The angle of incidence equals the angle of reflection:
i = V (Fig. V.1.2a).
V .1 .2 Geometrical (Ray) Optics 393

The laws of reflection hold in the reverse direction of


the path of the light rays. A ray travelling along the path
of the reflected ray is reflected almig the path of the incident
one (the reversibility of the path of light rays) (Fig. V.1.26).
5°. The reflection of light obeying these laws is called
regular, or specular, reflection. If the condition of specular
reflection is not observed, then the laws of reflection do not
hold, and the reflection of light is called irregular, or diffuse.
6°. The laws of refraction of light:
(a) The incident ray, the refracted ray, and the normal
to the interface of two media erected at the point of incidence
of the ray all lie in the same plane.
(b) The ratio of the sines of the angles of incidence and
refraction is a constant quantity equal to the relative refrac­
tive index of the two given media (Par. 2°):
sin i
■n2l

The incident and refracted rays are mutually reversible:


if the incident ray is made to travel in][the direction of the
refracted one, then the refract­
ed ray will travel in the direc­
tion of the incident one.
7°. The laws of reflection
and refraction of light hold
for homogeneous isotropic me­
dia in the absence of absorption
of light.
8°. If light rays from ;an
optically more dense medium 1
strike the interface with an op­
tically less dense medium 2, Fig. V.1.3
for example if they pass from
glass into water f|(Fig. V.1.3),
then at angles; of incidence of i ^ iCT, where sin ic = n2 1 ?
no refraction of light occurs. When i = iCT, the angle of refrac­
tion r = j i / 2, and when i > iCT, no light passes into the sec­
ond medium. This; phenomenon is called total reflection.
The angle iCT is"called the critical angle of total reflection. If
light passes from a substance with the absolute refractive
394 O p tics V.1.2

index nx = n into air where n2 « 1, then the condition of


total reflection acquires the form
. . 1
sin iCT= —

Problem. An observer’s eye is arranged so that the wall


of a flat-bottom bowl completely closes the bottom. If a

liquid is poured into the bowl up to its edges, then the observ­
er will see the picture at the centre of the bowl’s bottom.
The depth of the bowl is 8.1 cm and its diameter is 14 cm
(Fig. V.1.4).
Determine the relative refractive index of the liquid
and the apparent depth h of the bowl.
Given: CA = 8 . 1 cm, CB = 14 cm.
Required: n21, h.
Solution: In the empty bowl, a light ray travels in the
direction B A , and the eye does not see the bottom and the
picture. In the liquid, the ray from the picture travels in
the direction;£M, is then refracted, and passes on to the observ­
er’s eye. The relative refractive index of the liquid is n21 =
= sin ilsin r.lThe angles i and r are determined from the
triangles ACB and A CO:

tan i = = — = 1.7, the angle £= 60°

tan r = ~ 2 x 8 1 = 0-86, the angle r = 40°


V.I.3 G e o m etric a l (R ay) O p tics 395

Hence,
sin 60° _ CL86
sin 40° 0.65
1.33

Water has such a relative refractive index.


Inspection of the figure shows that the apparent depth
of the bowl is
AC 8.1 , n
2
— « 4.0 cm

3. Plane Mirror.
P lan e-P arallel Plate. Prism
1°. Every point S of a source of light* in geometrical
optics is considered to be the centre of a divergent beam of
rays called homocentric. If after reflections and refractions
in various media a beam remains homocentric, then its
centre S' is called the image of the point S in the optical
device.
The image S' is called realii the rays of a beam themselves
intersect at the point S', and virtual if the continuations
of these rays intersect at it. The energy of the light rays is
concentrated at the point where a real image is produced, and
this can be detected, for example, by a photocell (V.5.4.10)
or light-sensitive paper. This does not occur with a virtual
image—the light rays appear to emerge from a point at
which no energy can be detected. It is significant, however,
that even with a virtual image of a point, its real image al­
ways appears on the retina of the eye (V. 1.7.3°).
2°. A homocentric light beam emerging from the point
S' of a source remains homocentric after reflection in the
plane mirror AC (Fig. V.1.5). The point Sx is the virtual
image of the point S. Its position is determined by the
intersection of the continuations of any two rays reaching
the eye, for example A B and CD. The line SSj is perpendic­
ular to the plane of the mirror, and SO = OSv To find the

* In Chapter 1 of Part V, we consider only monochromatic light


(IV.4.2.5°).
396 O p tics V .1 .3

image of a point in a plane mirror, it is sufficient to lay off


on a perpendicular line dropped from the point onto the
mirror the same distance behind the mirror. The geometrical

Fig. V.1.5

dimensions of an extended light source and its virtual


image in a plane mirror are identical (Fig. V.1.6).
3°. After passing through a plane-parallel (parallel­
sided) plate (Fig. V.1.7),Uhe rays.leavejt atjthe same angle
S

y///////////sZ'r

Fig. V.1.6

i at which they enter it. The plate displaces a ray of light


parallel to itself over the distance

A= d s i n i ( l - ] /
V .1 .4 Geometrical (Ray) Optics 397

where d = thickness of the plate


i = angle of incidence of the rays
n = refractive index of the plate material relative
to the surrounding medium.
The luminous point S of a source or an illuminated object
seems to be brought closer to the surface of the plate by the
distance
A' = d V( i - ri / -n 12-2—
~ -sm
sii1V
2 i -/ )
With normal incidence of the rays (i = 0), we have A = 0,
and A '= d (n —l)/w.
4°. A light ray falling on a prism in its section with a
planejperpendicular to its edges shown in Fig. V.1.8 is deflect­
ed toward its base after refraction
on the faces AC and CB. The
angle of deflection of the ray is
cp = + r2 — a, where is
the angle of incidence of the
ray on the face AC, r2 is the
angle of refraction on the face
BC, and a is the angle between
the faces AC and CB called the
prism angle.
When r2 = ily the angle
of deflection of the rays cp is
minimum (cp = cpmln). With such an arrangement of a prism
relative to the light source (arrangement at the angle of the
smallest deflection),
sin ^ — = n sin -y
where n is the refractive index of the prism material relative
to the surrounding medium.

4. Spherical Mirrors
1°. If two media have a spherical interface, then a homo-
centric beam emerging from the point source S (V. 1.3.1°)
remains the same after refraction only provided that SA
« SO and AS' « S '0 ±1 i.e. when the points 0 and Ox vir­
tually coincide (Fig. V.1.9).
398 Optics V.1.4

The optical axis of a spherical surface is defined as the


straight line passing through the point light source S and
the centre of curvature C of the surface. The preceding con­
ditions hold only for a narrow cone of light rays with their
axis at right angles to the spherical interface. Only such
beams of light rays called paraxial beams remain homocen­
tric after refraction and produce an image of the luminous
point S in the form of the point S' (see Fig. V.1.9).

In this figure, ax and a2 are the distances from


the source S to the spherical surface and from this
surface to the image S'. These distances are measured
from the point O of intersection of the spherical surface with
the optical axis. According to the sign convention, they are
considered positive in the direction of propagation of the
light, and negative in the opposite direction: a2 > 0, and
a1 < 0.
2°. A spherical mirror is a carefully polished surface of
a body having the shape of a spherical segment (Fig. V.1.10).
Such a mirror reflects light specularly (V .l.2.50). The cen­
tre C of the spherical surface which the segment has been cut
out from is called the optical centre of the mirror. The point
O on the spherical segment is its vertex. Any straight line
passing through the optical centre of the mirror C is called
an optical axis of the mirror. The optical axis CO passing
through the optical centre of the mirror and its vertex is the
chief, or principal, optical axis.
3°. The rays of a paraxial beam (Par. 1°) parallel to the
principal optical axis intersect at one point F after being
reflected from the mirror. The point F is called the focal
point of the mirror (the principal focal point). The distance
OF = f from the vertex to the focal point of the mirror is its
V.1.4 Geometrical (Ray) Optics 399

focal length: / = R!2, where R is the radius of curvature of


the mirror. The plane passing through the focal point per­
pendicular to the principal optical axis is defined as the
focal plane.
4°. The equation of a spherical mirror for paraxial light
beams (Par. 1°) is
1 1 _ 2 _ 1
«i a2 ~ R ~~ f
where R = radius of curvature of the mirror
ax = distance from the mirror to a luminous point
on its principal optical axis
a2 = distance from the mirror to the image.

Fig. V.1.11

The sign convention for i?, a1, and a2 is indicated in


Par. 1°.
The image in a spherical mirror is real if a2 < 0, i.e.
both the source and its image are on the same side of the
mirror (Fig. V.1.11). When a
luminous point and its image
are on different sides of the mir­
ror (a2 > 0) (Fig. V.1.12), then
the image in the mirror is vir­
tual.
A mirror is convex if with
account taken of the sign con­
vention, R > 0. The image is
always virtual in such a mir­
ror. The image in a concave mirror (R < 0) is real when
a jf > 1, and is virtual when a jf < 1.
5°. The image of a luminous object of linear dimension
ftobj at right angles to the principal optical axis has the lin-
400 Optics V .1 .5

ear dimension /^m and is at right angles to the same axis.


The image should be formed with the aid of paraxial rays
(Par. 1°), and it is essential that hohj | ax |. Figures
V.1.11 and V.1.12 are not drawn to scale. When using the
equation of a spherical mirror for constructing the images
of objects, ax and a2 must be understood as the distances from
the mirror to the object and to its image (with the observance
of the sign convention, Par. 1°). The ratio of the linear di­
mensions of the image him and the object h0bj at right angles
to the principal optical axis is called the linear (lateral) magni­
fication:
M = ±
^obj
The plus sign (M > 0) corresponds to an upright image (see
Fig. V.1.12), and the minus sign (M <C 0) to the inverted
image (see Fig. V.1.11).

5. L e n s e s
1°. A transparent body having two curved surfaces is
called a lens. In a particular case, one of the surfaces may be
flat. In the majority of cases important in practice, both lens

surfaces are spherical. A lens is considered to be thin if its


thickness is much smaller than the radii of curvature R 1
and i?2 °f both surfaces. (Thick lenses for which this condi­
tion is not observed are not treated in an elementary course
of physics.)
2°. The straight line drawn through the centres C1 and
C2 of curvature of both surfaces is the principal optical axis
of the lens (Fig. V.1.13). In a thin lens, the points Ox and 0 2
V .1 .5 Geometrical (Ray) Optics 401

of intersection of the principal optical axis with both sur­


faces can be considered to merge into a single point 0 called
the optical centre of the lens. Straight lines passing through
the optical centre of a lens
and not coinciding with
the principal optical axis
are called auxiliary optical Fz P rin c ip a l a x is
axes. A light ray propagat­ Focal p o in t
ing along an optical axis Optical centre
Focal point
passes through a lens
without refraction.
3°. The rays of a para­ Fig. V.1.14
xial light beam propagating
parallel to the principal optical axis intersect at a point on
this axis called the focal point of the lens (the principal focal
point). Every lens has two
focal points at both sides of
it (Fig. V.1.14).
The plane M N drawn
through a focal point of a lens
at right angles to the principal
optical axis is called a focal
plane (Fig. V.1.15). The inci­
dent rays parallel to an au­
xiliary optical axis after refrac­
tion in a lens intersect at a
point in the focal plane. A lens
has two focal planes at both
sides of it. The points of intersection of the auxiliary opti­
cal planes with the focal planes of a lens are the auxiliary
focal points of the lens (the point F' in Fig. V.1.15).
4°. The distance OF = / from the optical centre of a
lens to its focal points is called the focal length of the lens:

(i!r -x r )
The thin-lens equation holding for paraxial rays (see
also Par. 6°) is
402 Optics V .U

In these equations, n21 = n2ln^ n2 and nx are the absolute


refractive indices (V. 1.2.1°) for the lens material and the
surrounding medium, respectively, Rx and R 2 are the radii
of curvature of the front and rear (relative to the object)
surfaces of the lens, and ax and a2 are the distances to the
object and to its image that are measured from the optical
centre of the lens along its principal optical axis. For the
sigh convention for a±1 a2, R±, and i?2, see V .1.4.1°. If the
light source is at the front principal focal point F x (see
Fig. V.1.14), then its image will be at infinity (a2 = °°)«

(a) (b) (c) (d ) (e) (f)

Fig. V.1.16

after refraction in the lens, the rays will be parallel to one


another. If the light source is at infinity (ax = —-oo), i.e.
if a beam of light parallel to the principal optical axis
falls on the lens, then its image will be at the point F 2 (see
Fig. V.1.14).
5°. A lens is defined as positive (converging) if its focal
length is positive ( / > 0), and negative (diverging) if its
focal length* / < 0. Converging glass lenses in air (n2 > n2)
include double-convex (biconvex) (Fig. V.1.16a), plano-convex
(Fig. V. 1.166), and converging meniscus (concavo-convex)
(Fig. V. 1.16c) ones. Diverging glass lenses include
double-concave (biconcave) (Fig. V.1.16d), plano-concave
(Fig. V. 1.16c), and diverging meniscus (convexo-concave)
(*Fig. V:1.16/) ones. The focal points of converging lenses are
real, and those of diverging ones virtual.
6°. Another form of the thin-lens equation for paraxial
rays (see also Par. 4°) is:
or
/
where s = | ax |, d = a2, and / is the focal length of the lens.
The plus sign is for a converging lens, and the minus sign
V .1 .5 Geometrical (Ray) Optics 403

for a diverging one. The quantity P = 1// is defined as the


lens power. For a converging lens P > 0, and for a diverging
one P < 0 . The lens power R equals one diopter when
/ = 1 m.
7°. The image of a point of an object in a lens is at the
point of intersection of two rays (or their continuations)
emerging from the point and passing through the lens.
Generally two of the following three rays are used to con­
struct an image: a ray passing without refraction through
the optical centre of a lens; a ray parallel to the principal
optical axis [after refraction in the lens this ray (or its con­
tinuation) passes through the rear principal focal point
(rear relative to the object)!; and a ray (or its continuation)
that passes through the front principal focal point and after
refraction in the lens is parallel to the principal optical axis.
8°. The linear magnification M of a thin lens is defined
as the ratio M = aja^ with a view to the sign convention
for a2 and ax (V. 1.4.1°). For real images, we have M < 0,
i.e. they are inverted; for virtual images, M > 0, they
are upright.
Problem. How will the focal lengths and powers of
double-convex and double-concave glass lenses having radii
of curvature R1 and R 2 change if the lenses are immersed in
a medium (for example nut oil or cinnamon oil) with a re­
fractive index n\ greater than that of glass n2?
Given: n2 = 1.5, nx = 1.0, n[ = 1.58.
Required: /', /;, P'v P'2.
Solution: If a light ray passes from a lens into air, then
the focal length of the lens is found by the lens equation

where the relative refractive index of glass n21 = n2!nly


and R x and R 2 are the radii of curvature. When the lens is
immersed in oil, the focal length is determined by the same
equation:

26*
404 O p tits V.1.6
The ratio of the focal lengths is

i'i " a i-i 1 .5 - 1


/i « ii — 1 1-5 .

A converging lens when immersed in an optically more


dense medium becomes a diverging one (/' < 0), and its
power diminishes to one-tenth of the original value:

For a double-concave lens, we get

- I H 0' “ d n ~ i0h

A diverging lens when immersed in an optically more


dense medium becomes converging, and its power diminishes
to one-tenth of the original value:

6. Concept of Photometry
1°. Photometry is the branch of optics dealing with mea­
surements of the energy transferred by electromagnetic
light waves (V. 1.1.1°), i.e. of the quantity of light. Photo­
metry usually considers the action of electromagnetic waves
of the visible optical range on the eye and other optical in­
struments (V. 1.7.1°). This action is described by introducing
the following physical quantities characterizing light from
the viewpoint of its quantity: the luminous flux, luminous
intensity, and illumination.
2°. The luminous flux O is defined as the power of visible
radiation (IV.4.4.1°), which is assessed according to its
V.1.6 Geometrical (Ray) Optics 405

action on a normal eye. In other words, O is the quantity


of light carried in a unit time through a certain surface area
and assessed by visible perception. For monochromatic
light (IV.4.2.5°) corresponding to the maximum of the spec­
tral sensitivity of the eye
(X = 5500 A), the luminous
flux equals 683 lumens (lm)
(VII.6.2°) if the radiated power
equals one watt.
3°. A point source of light
is a source whose linear di­ Fig. V.1.17
mensions are considerably
smaller than the distance from it to the point of observa­
tion. Such a source emits spherical electromagnetic waves
(IV.3.1.50).
The luminous intensity I of a point source is the quantity
numerically equal to the luminous flux produced by this
source in a unit solid angle* (Fig. V.1.17).
If a point source emits light uniformly in all directions,
then
r _ Qtot
1 ~~ 4n

where Otot is the total luminous flux of the source, i.e. the
power radiated by it in all directions—-the quantity of light
carried in a unit time through an arbitrary closed surface
surrounding the light source.
4°. If a light source is extended instead of being a point
one, then the luminous intensity I of a small section AS of
its surface in a given direction is

where AO is the luminous flux produced by the section AS


of the surface of the source in the given direction inside the
solid angle AQ.

* The solid angle AQ is determined as the ratio of the area AS of


the surface of a spherical segment to the square of the radius of the
sphere: AQ = AS/r2 (see Table VIIt2)t
406 Optics V.1.6

The mean spherical luminous intensity for an arbitrary


light source is defined as the quantity

7 — ^ tot
4 ji

where Otot is the total luminous flux of the light source.


When an extended light source radiates uniformly in all
directions, 1 = 1. The unit of luminous intensity in the
SI system—the candela (cd)—
is a fundamental unit (see Tab­
le VII.2).
5°. The illumination (illumi­
nance) E of a surface is defined
as the ratio of the luminous flux
AO striking the area AS of the
surface to the value of this area:
A(D
E = AS
The illumination E at each
point of a screen on which light
falls is proportional to the intensity of the electromagnetic
light wave at this point (IV.4.3.20).
The illumination produced by a point source of luminous
intensity / on a surface at the distance r from the source is
I cos i
E-

where i is the angle of incidence (V. 1.2.3°) (the law of illu­


mination by a point source).
Problem. A bulb considered as a point source of light
uniformly radiating in all directions hangs above a round
table 1.6 m in diameter at a height of 0.6 m. The luminous
flux falling on the table is 201 lm. Find the luminous inten­
sity of the bulb, the total luminous flux emitted by it, the
illumination at the centre of the table, and at its edge
(Fig. V.1.18).
Given: AO = 201 lm, d = 1.6 m, h = 0.6 pi,
Required: /, d>tot, E, E '•
V .1 .7 Geometrical (Ray) Optics 407

Solution: The intensity of the source I = AO/AQ, where


AO is the luminous flux radiated within the solid angle
AQ. #
The solid angle at which the surface of the table is seen
from the source is
AQ = 2ji (1—cos i)
.Si
where i is the angle of incidence of a ray.
It follows from Fig. V.1.18 that

cos i = -y-, r = j / rfe2 + ^ ) 2= 1^-0.36 + 0.64 = 1 m

Hence,
1= ^ 201,;- -^ v = 8 0 Cd
2 X 3.14 X (1 — 0.6)
'

The total luminous flux emitted by the point source is


(Dtot = AnI = 4 x 3.14 x 80 = 1000 lm -

The illumination at the centre of the table is 9


.'i

The illumination at the edge of the table is

Ji cos i i —
r Ih 80 X 0.6
E' = r3 13 = 48 lx

7., Optical Instruments


1°. Optical instruments are devices intended for produc
ing the images of various objects on screens, light-sensitive
plates, photographic films, and in the eye. Optical instru­
ments generally produce a plane (two-dimensional) image
of three-dimensional objects. Examples of optical instru­
ments are a camera and a projector, the eye, eyeglasses, $
magnifying glass and a microscope, a telescope.
408 Optics V.1.7

2°. A camera is used to obtain a real reduced (sometimes


magnified) image of an object that in the majority of cases
is beyond the double focal length from the objective (a
system of lenses or, in a simple camera, a single converging
lens). The objective is in the front wall of the camera. Its
back wall accommodates a photographic plate or film coated
with a thin layer of a light-sensitive photoemulsion usually
containing silver bromide or chloride as the active ingredient.
The objective produces the image A 1B1 of the object AB
(Fig. V.1.19) on the emulsion.
The luminous fluxes (V. 1.6.2°)
striking the emulsion from dif­
ferent points of the object cause
different decomposition of
the silver bromide at different
Fig. V.1.19 places (V.5.6.20). The photo­
graph obtained is then proces­
sed with special substances (is developed, fixed, and washed
with water) to obtain a visible image with differing degrees
of blackness. The luminous flux striking the light-sensitive
layer is controlled by a shutter that opens during the time
called the exposure time t. The optimal value of t depends on
the illumination of the photographic plate or film (V .l.6.50)
and its sensitivity. The illumination depends on the lumi­
nous flux passing through the objective of diameter d. When
the object is far enough from the objective, the film is near
its focal plane (V .l.5.30) and is at the distance / (V. 1.5.4°)
from it. The illumination of the film is proportional to the
quantity d2//2—the rapidity of the objective. The value of the
relative aperture dlf shows the quality of an objective or lens.
The mounts of camera lenses carry an indication of their
focal lengths / and relative apertures in the form 1: fid
(sometimes the so-called f-number equal to fid is indicated).
Projectors are intended for obtaining real magnified im­
ages of objects on a screen. Diascopic projectors (diascopes)
are used to project transparent objects (slides, etc.), and
episcopic projectors (episcopes) are used for opaque objects
(illustrations, photographs, etc.). Figure V.1.20 shows the
design of a diascope. The objective Ob—a system of lenses
functioning like a single converging lens—produces a magni­
fied image of the slide, or diapositive, Z), which is near the
V .1 .7 Geometrical (Ray) Optics 409

focal plane of the objective, on the screen MN. A short-


focus system of lenses—the condenser C—is used to direct
all the light from the source through the diapositive. The
condenser is arranged so as to produce the image of the source

S on the objective. An object to be projected by an epi-


scope is illuminated strongly from a side with the aid of a
lamp and mirrors, and is projected by the objective onto
a screen. Projectors in which means for both dia- and epi-
scope projection are combined
are called epidiascopes.
3°. The eye is our organ of
vision. Figure V.1.21 shows
the structure of the human
eye. The sclera 1 forms the
outer covering of the eyeball
and protects the interior of
the eye, retaining its rigidity.
The sclera at the front of the
eye changes into the cornea 2
through which light enters the
Fig. V.1.21
eye. Behind the cornea is the
iris 5, a circular diaphragm
whose central hole is the pupil 4. The iris is a ring of muscles
controlling the area of the pupil and the amount of light
entering the eye. The crystalline lens 5 is an elastic lens-like
body. The action of special ciliary muscle 6 changes the
shape and curvature of the crystalline lens, its power
(V.1.5.6°)? and focal length. The space between the cornea
410 Optics V.1.7

and the crystalline lens is filled with a fluid called the aqueous
humour. The crystalline lens is followed by the vitreous body
7. The optical system of the eye, similar to a lens of power
P « 5 8 . 5 diopters (V .l.5.60), is formed by the cornea,
aqueous humour, crystalline lens, and the vitreous body.
The optical centre of this system (V. 1.5.1°) with the prin­
cipal optical axis (V .l.5.20) AB is at a distance of about 5 mm
from the cornea. A real inverted image of the object which
the eye is accommodated to (Par. 5°) is always formed on
the retina 9—a hemisphere consisting of special light-sen­
sitive cells having the shape of rods and cones. The cells are
on the back surface of the retina, which lies on the choroid 8.
The nerve cells of the retina combine to form the optic nerve
10 that leaves the eye at a -spot where there are no light-
sensitive cells (the blind spot 11). The centre of the retina on
the optical axis is the region of most distinct vision called the
fovea centralis, or yellow spot, 12 at which the cones are con­
centrated. They permit the eye to perceive colour. The
rods are arranged mainly on the other parts of the retina.
4°. The action of light on the cells highly sensitive to
light (every rod can react to one photon—V.5.1.20) causes
complicated physicochemical processes to occur in them. As
a result, a nerve impulse is generated in each cell which is
transmitted by the optic nerve to the brain. The joint action
of the rods and cones is responsible for the process of vision.
The rods detect the dimensions and shape of an object. Col­
our vision is the duty of the cones when the image of an
object reaches the yellow spot.
5°. The focal length of the optical system in an eye must
change to create a sharp image of objects at different distances
from the eye on the retina. This is achieved by a change
in the curvature of the crystalline lens surfaces. The property
of an eye to adjust itself to the distance at which objects
being viewed are is called accommodation. The latter occurs
spontaneously as a result of contraction or stretching of the
ciliary muscle (Par. 3°). The point which an eye sees when
the ciliary muscle is relaxed is called the far point. The
point distinctly seen by the eye at the greatest tension of the
muscle is called the near point. For a normal eye, the far
point is at infinity, and the near one at a distance of from
15 to 20 cm,
V .1 .7 Geometrical (Ray) Optics 411

6°. Nearsightedness (,shortsightedness) is a defect of the


eye when the far point is at a finite distance, and the image
of a remote object is obtained not *>n the retina, but in front
of it (Fig. V. 1.22a) when the eye is not strained. This is
corrected by glasses (eyeglasses) with diverging lenses
(Fig. V. 1.226).
Farsightedness (longsightedness) is a defect of the eye when
the near point is at a greater distance from the eye than in
normal vision. [At present farsightedness exceeding 1.0
diopter (V. 1.5.6°) is considered to be a defect of the eye.]

Fig. V.1.22 Fig. V.1.2a


The image of an object is obtained behind the retina when
the eye is not strained (Fig. V. 1.22c). Glasses with converg­
ing lenses remedy this defect and return the image to the
retina (Fig. V.1.22d).
7°. The magnitude of the image S'S[ of the object S S 1
on the retina is determined by the angle of vision (p = hlf
(Fig. V.1.23), whose apex is at the optical centre of the eye.
The rays forming the angle are directed toward the extreme
points of the object.
The distance of normal vision D is defined as the distance
from an object to the eye at which <p is maximum provided
that the accommodation strain is not great, and the eye does
not tire. For a normal eye, D « 25 cm. An eye is considered
to be normal if it retains a good ability of accommodation.
This ability gradually diminishes as a person gets older.
8°. Two points of the image of an object are considered
to be resolved, if they fall onto two different light-sensitive
cells on the retina and are perceived separately by the eye.
The resolving power of the eye is assessed according to the
minimum angle of vision (pmln at which two points of an
object are seen separately in conditions of good illumination,
412 Optics V.1.7

9°. A magnifying glass is an optical instrument making


it possible to increase the angle of vision in the simplest
way. A magnifying glass is a lens of short focal length that
is positioned for looking at the object AB = h so that the
latter is closer than the principal focal point F of the lens.
The path of the rays in a mag­
nifying glass is shown in Fig.
V.1.24.
A normal naked eye sees the
object AB at an angle of vision
(p0 such that tan cp0 = hlD,
where D is the distance of normal
vision. An eye looking through
Fig. V.1.24 a magnifying glass sees the
object at the angle cp whose tan­
gent is tan cp = hlf, where / is the focal length of the lens.
The image ab of the object on the retina is the same as if
the eye were considering the object A XBX that is the virtual
image of the object AB
in the glass. A^%
The angular magni­
fication y of an optical
instrument is defined as
the quantity
_ tan cp _ D
^ ~~ tan cp0 ~ T

10°. A microscope is
an instrument that per­
mits us to obtain con­
Fig. V.1.25
siderable angular magni­
fication of close minute
objects. It is a combination of two lenses of short focal
length—an objective and an ocular, or eyepiece. The focal
points of both lenses and the path of the rays in a micro­
scope are shown in Fig. V.1.25. The object AB = h is placed
beyond the focal point F1 of the objective. A real magnified
image H is obtained behind the objective, before the focal
point F2 of the eyepiece. This image is what is seen by the
eye looking through the eyepiece as through a magnifying
V.1.7 Geometrical (Ray) Optics 413

glass, and a greatly magnified virtual image A 1B 1 of the


object is formed.
The angular magnification of » microscope is
Z>A
Y fobfe
All the distances are indicated in Fig. V.1.25, D is the dis­
tance of normal vision.
11°. A telescope is used to view the details of very remote
objects. A very simple telescope is the Keplerian telescope.
When considering two points A and B of a remote object

of which one (B) is on the optical axis of the system and the
other 04) is higher than the axis, then the image A 1B 1 of
the object is obtained in the focal plane of the objective
(Fig. V.1.26). The eyepiece functioning as a magnifying
glass is arranged so that its front focal point will coincide
with the rear focal point of the objective. The eye sees a
beam of parallel rays at the angle of vision cp > (p0, where cp0
is the angle of vision at which the object is seen by the naked
eye. The final image is formed as shown in Fig. V.1.26.
The angular magnification of a telescope is
__ tan (p __ /pb
7 tan <p0 /e
Objectives of a long focal length and eyepieces of a short
focal length are used for considerable angular magnifica­
tions.
414 Optics V.2.1

Chapter 2

W ave Optics (Light W aves)

1. The Speed of Light


1°. Wave (physical) optics is a branch of the science of
light in which light waves (V. 1.1.1°) are considered as elec­
tromagnetic waves (IV.4.1.1°) occupying a definite interval
on the electromagnetic spectrum (V.3.7.10). Wave optics
deals with the classical laws of the radiation (IV.4.4.10),
propagation, and interaction of
Io light waves with a substance.
2°. The speed of light is deter­
mined in the same way as the ve­
locity of propagation of a wave
of any nature (IV.3.3.10). The
speed can be measured by astronom­
ical and laboratory methods. One
of the astronomical methods, the
Roemer method, is based on observ­
ing the time intervals T between
two consecutive eclipses of Jupi­
ter^ satellite Io (Fig. V.2.1). The
delay AT7 of the eclipse at the
moment when the Earth is at its
gr atest distance from Jupiter
(the points J' and E' in Fig. V.2.1) in comparison with the
moment when the two planets are the closest distance apart
(the points J and E) is associated with the fact that light,
travelling at the finite speed c, passes a distance equal to the
diameter D of the Earth’s orbit {D = 2.99 X 1011 m) during
the time AT, i.e. AT = Die. Modern data for AT7 = 16.5
-min result in a value of c close to c = 3 X 108 m/s.
3°. A setup of a laboratory method of measuring the
speed of light (the Fizeau method) is shown in Fig. V.2.2.
The light from the source S is directed by the half-silvered
mirror A onto the toothed wheel TW rotating with a speed
of z about its axis O. After passing through a gap between
V .2 .1 Wave Optics (Light Waves) 415

the teeth, the light reaches the plane mirror M . The latter
reflects the light back to the toothed wheel, which may
either let it pass through and reach the observer’s eye, or
retain it. In the latter case during the time At spent by the
light for travelling over the distance 2Z, the wheel turns
through half the width of a tooth, and the passage of the

according to this method is determined by the formula


c = 2----
zZ(on
-
ji
where co0 is the smallest angular velocity (1.1.9.5°) of the
wheel at which no light reaches the observer.

Fig. V.2.3
4°. In Michelson's method (Fig. V.2.3), the toothed wheel
is replaced by an octahedral prism. It rotates with such a
velocity that the image of the slit D is continuously seen in
416 O ptics V.2.2

the telescope C. For this to occur, the light must travel the
distance 21 when the prism rotates through one-eighth of
its length.
5°. Measurements of the speed of light v in various sub­
stances (for example in water and glass) confirmed the fact
that it diminishes compared with the speed c in a vacuum in
accordance with the formula v = cln (V. 1.2.1°), where
n = j/ep, is the absolute refractive index of the substance.

2. Interference
1°. The interference (IV.3.9.50) of electromagnetic waves
of the optical range (V. 1.1.1°) when the oscillations occur
in the same planes is called the interference of light. The
result of superposition of coherent (IV.3.9.30) light waves
observed on a screen, photographic plate, etc. is called an
interference pattern. Upon the superposition of incoherent
light waves (IV.3.9.50), only an increase in the amount of
light is observed, and no interference is noted.
2°. The length X of a light wave in a substance of refrac­
tive index n diminishes in comparison with the wavelength
A,0 in a vacuum: X = X0/n. This is associated with the fact
that the speed v of propagation of an electromagnetic wave
in a substance decreases to 1/rc-th the speed of light: v =
= cln (V.l.2.10). Consequently, X = vT — {cln) T = X0lny
where T is the period of oscillations. If an electromagnetic
wave travels the distance d in a substance, this distance w ill
accommodate n times more wavelengths than in a vacuum.
The concept of the optical path length nd is introduced in
optics. Here n is the refractive index, and d is the geometri­
cal path length of the wave. The optical path length character­
izes the number of wavelengths that can be accommodated
in a given medium over the length of the geometrical path
of a wave. The conditions for the constructive interference
(amplification) and destructive interference (weakening) of
two coherent light waves propagating from the sources S 1
and S 2 to the point M (Fig. V.2.4) are written differently
than in IV.3.9.40. The light waves emitted by the coherent
sources Sx and S 2 can travel in different substances with the
refractive indices nx and n2. The difference 8 between the
V.2.2 Wave Optics (Light Waves) 417

optical path lengths of two rays is called the optical path


difference:
8 —n2r2—'nir i
The condition for constructive interference is
8 == n2r2— nir i = mX
where m = 0, ± 1 , ± 2 , etc.
The condition for destructive interference is
X
8 = n2r2— nir i = (2m — 1) —

The interference of light can be observed if one of two


rays emitted by the source S at the focal point of a lens trav­
els the distance d, for example,
in glass, and another ray trav­
els in air (n « 1) (Fig. V.2.5).
At the point M , the rays have
the optical path difference 8 =
— nd — d = d (n — 1) whose
value determines the interference
pattern at the point M.
When interference of mono­
chromatic light of a definite wave-
length is observed, the inter­
ference pattern on a screen con­
sists of alternating dark and
Fig. V.2.4
bright spots. An interference
pattern in white light (V.2.6.20) is coloured because each
component of the white light having the wavelength X

produces constructive and destructive interference at its


relevant place on the screen.
27 — 0211
-418 Optics V.2.2

3°. For light to interfere, the light waves must be coher­


ent (IV.3.9.30). A way of achieving the interference of
light is to divide a wave emitted by a single light source into
two or several waves. After travelling different optical path
lengths (Par. 2°), these waves become superposed at the
points of observation and, having a certain optical path
difference, produce an inter­
ference pattern.
4°. In Fresnel's double mir­
ror (Fig. V.2.6), the beam of
light from the point source S
is divided into two beams
by the two mirrors I and II
that make an angle close to
180 degrees with each other.
The direct rays do not reach
the screen A A. They are re­
tained by the partition KK.
The virtual images Sx and S 2
of the source S in the mirrors
I and II produce two systems
of coherent waves SB1OC1C1
and SOB2C2C2, and their in­
terference is observed on the
screen AA.
5°. Interference in thin
F ig . V .2 .6 films is observed in the sim­
plest case when a beam of
rays falls on a plane-parallel layer of thickness d
(Fig. V.2.7). At the point C, the reflected ray 2' and the
ray 1" refracted in the glass have an optical path difference
of 8 = (AD + DC) n - BC.
An interference pattern of this kind is observed when
light is reflected from a thin soap film, and from oil films
on the surface of water that produce a very fantastically
coloured interference pattern as a result of their varying
thickness.
6°. The interference fringes formed in a very simple
case on a plano-convex lens contacting a plane-parallel
plate at the point O (Fig. V.2.8) are called Newton's rings.
The ray 1 that has passed twice through the air gap in-
V .2 .2 Wave Optics (Light Waves) 419

terferes with the ray 2 at the point C. The interference


pattern has the form of light and dark fringes because
all the points of a ring of radius r have the same optical
path difference and produce either constructive or destruc­
tive interference of the light.
Problem 1. When yellow light passes from a vacuum
into a liquid, its wavelength X0 decreases by 0.147 pm.

X !
N^ i C*r
0'/ 'A 'A 'A
/, A // %

Fig. V.2.8

Determine the absolute refractive index of the liquid and


the speed of light in it.
Given: X0 = 0.589 pm, AX = 0.147 pm.
Required: n, v.
Solution: The length of a light wave in the given medium
diminishes in comparison with the wavelength in a vacuum:
X = X0/n, where n is the absolute refractive index of the
medium. Hence,
Xq _ XQ 0.589 i nn
n = T = X0— AX = 0.589—0.147 =*= 1,M

The speed of light in a vacuum is c = A,0v, and that


in a liquid is v = Xv. The frequency v does not depend
on the properties of the medium. Therefore, c/X0 = v/X.
Ilence,
cX c 3X108
V~‘ n 1.33
2.26 x 108 m/s

Problem 2. In a Rayleigh interferometer (Fig. V.2.9),


Ilie rays from two coherent sources pass through closed
glass tubes 10.0 cm long and, after being gathered by a
420 Optics V.2.B

lens, produce an interference spectrum on a screen. If one


of the tubes is filled with a gas whose refractive index
differs from that of the air filling the other tube, the spec­
trum will be displaced by several orders. Determine the
dn]
S,*
dn0
s2*

Fig. V.2.9

refractive index of chlorine if the yellow band of the spec­


trum was displaced by 82 orders.
Given: m = 82, n± = 1.000 292, X = 5880 A = 5.88 X
X 10"7 m, d = 10.0 cm = 0.100 m.
Required: n2.
Solution: The optical path difference of the rays is 6 =
= dn2 — dnv An optical path difference of the rays of
6 = mX corresponds to displacement of the spectrum by
m orders. Hence, mX = d (n2 — n j, and
n2 = = 82 x 5 -8^ 10"7 + 1 .0 0 0 292 = 1.000 773
z d 1 1 O.lOt)

3. D iffr a c tio n
1°. The diffraction of light is the property of light waves
of bending around obstacles in their way.
In a broader sense, the diffraction of light is defined
as the combination of phenomena due to the wave properties
of light and observed when it propagates in a medium
with sharply expressed inhomogeneities (holes in opaque
screens, edges of opaque bodies, etc.). The phenomenon
of diffraction points to violation of the laws of geometrical
optics.
Diffraction is observed at distances I from an obstacle
equal to / « D 2/4X, where D signifies the linear dimen­
sions of the obstacle, and X is the wavelength (the condi­
tion of observance of diffraction).
i '.2.3 W ave O ptics (Light W aves) 421

Diffraction problems—the finding of the distribution


on a screen of the intensity of a light wave propagating
in a medium with obstacles—afe solved with the aid of
approximate methods based on the Huygens and the Huy-
gens-Fresnel principles.
2°. The Huygens principle: every point Sx, S 2, . . ., Sn
of the wavefront AB (IV. 3.1.5°) is a source of new secondary

waves. The new position of the wavefront A 1B1 after the


Iime At is the envelope of the secondary waves (Fig. V.2.10).
The Huygens principle is a purely geometrical one. It
permits us, for example, to explain the equality of the
iingles of incidence i and reflec­
tion i on the surface M N of a
medium upon the reflection of
light. The path difference CB of
(lie rays A XA and BXB produces
such a front DB of the reflected
wave that the equality i = i'
follows from the right triangles
ABB and ACB (AD = CB)
(Fig. V.2.11).
3°. The Huy gens-Fresnel prin­
ciple: all the secondary sources
.V,, S 2, etc. on a wavefront are
inherent (IV.3.9.3°). The amplitude (IV.1.1.4°) and the
phase (IV. 1.1.4°) of a wave at any point of space M are
the result of the interference of the waves emitted by
the secondary sources (Fig. V.2.12).
422 Optics V.2.4

4°. The rectilinear propagation of the ray SM emitted


by the source S in a homogeneous medium is explained
by the Huygens-Fresnel principle. The interference of all
the secondary waves emitted by the secondary sources on
the surface of the wavefront AB is destructive except for
the waves from the sources on the small part of the spherical
segment ab at right angles to SM (see Fig. V.2.12). The
light propagates along a narrow cone with a very small
base, i.e. virtually rectilinearly.

4. Diffraction by a Slit.
Diffraction Grating
1°. Assume that a beam of parallel rays of monochromat­
ic light falls on the barrier AE having a thin slit BC of
constant width b = BC and length L > b (Fig. V.2.13).
The phenomenon of diffrac-
b tion will be observed on the
screen Sc at the distance I
from the slit. If this phenome­
non were absent, then J we
would get an image of the
light source on the screen Sc
placed in the focal plane of
the converging lens ML
(V. 1.5.3°) at the point $F0—
the principal focal point of
the lens (V .l.5.30). When
light is diffracted by a nar­
row slit, an interference pat­
Fig. V.2.13 tern is observed on the
screen: a sequence of blurred
images of the light source
separated by dark spaces. All the parallel rays gather at
the point F<p on the screen that fall on the lens at the angle
cp (the diffraction angle) to the optical axis OF0 of the lens
(V. 1.5.2°) at right angles to the wavefront.
2°. The intensification of light (diffraction peaks) upon
diffraction by a narrow slit is observed at the angles cp
V .2 .4 Wave Optics (Light Waves) 423

complying with the condition

b sin cp= (2m -f-1) ^

where m is a positive integer.


The condition for the weakening of light (diffraction
minima) is

ft sin cp= 2m-—

The integer m is called the order of a diffraction maximum


(peak) or minimum. The quantity 6 = CD = b sin <p is the
optical path difference between the extreme rays CN and
BM travelling from the slit at the angle cp (see Fig. V.2.13).
The most intensive central zero-order peak is observed in
the direction cp = 0. Intensification of the light is always
observed at the point F0 regardless of its wavelength X.
3°. When diffraction by a slit is observed in white
light (V.2.6.20), the interference pattern on the screen
is coloured. Diffraction peaks with smaller wavelengths
are closer to the central uncoloured peak in the diffrac­
tion peak of each order (m = const).
4°. A diffraction grating in optics is defined as a combi-
nation of a large number of obstacles and holes concentrated
in a restricted space on which light is diffracted.
The simplest diffraction grating is a system of N iden­
tical parallel slits in a plane barrier each of width b with
equal opaque portions a between them (Fig. V.2.14). The
quantity d = b + a is called the grating spacing. The
number of slits per unit length of the grating is called
the grating constant.
According to the Huygens-Fresnel principle (V.2.3.30),
every slit is a source of coherent secondary waves capable
of interfering with one another. If a beam of parallel light
rays strikes a diffraction grating at right angles to it, then
a system of diffraction maxima and minima obtained as
a result of the interference of light from different slits
will be observed at the diffraction angle cp (Par. 2°) on
the screen Sc in the focal plane of the lens.
424 Optics V.2.4

5°. The principal maxima produced by a diffraction


grating are observed at the angles (p complying with the
condition
d sin (p = nX
where n = 0, 1, 2, 3 is the order of a principal maximum.
The quantity 6 = D K = d sin (p is the optical path dif­
ference between the similar rays BM and DN travelling
from neighbouring slits (Fig. V.2.14).
The principal minima produced by a diffraction grating
are observed at diffraction angles cp for which completely
destructive interference of
I iflj I Icl 111. the light from different
parts of each slit occurs.
The condition for the prin­
cipal minima coincides
with the condition for weak­
ening by one slit (Par. 3°)
b sin (p = mX
where m is a positive in­
teger.
6°. When diffraction by
a slit is observed in non-
Fig. V.2.14 monochromatic light, all
the principal maxima except
for the central zero-order one are coloured. With an increase
in the wavelength, the principal maxima within a given
order are at greater angles from the central one. The rain­
bow strip containing seven colours, from violet to red
(counting from the central maximum), is called a diffrac­
tion spectrum (cf. V.2.6.20). A diffraction grating is one
of the simplest sufficiently accurate means for measuring
wavelengths.
Problem 1. A beam of monochromatic light rays whose
wavelength is 5890 A falls on a flat slit 0.010 00 mm wide
at right angles to the slit. Find the angles at which the
first three diffraction minima will be arranged on a screen.
Given: b = 0.010 00 mm = 1.000 X 10”5 m, X = 5890 A =
= 5.890 X 10-7 m.
V.2.5 Wave Optics (Light Waves) 425

Required: q>t, cp2, cp3.


Solution: The condition for the minimum illumination
upon diffraction by a flat slit is b sin <p = 2mk!2 = ink,
where m is the order of the minimum.
For the first order, m = 1, whence
5.890 XIO'7
sin cpt = 1.000 x i o - &
=0.0589 and cpj = 3°22'

For the second order,


5.890 x 10-7 X 2
sin (p2 — 1 0 0 ) Xl()-5 11.87 x 10“2 and q>2 = 6°46'

For the third order,


.
sin <p3= 5.890
L()0X,)x10”1()_t
7X 3
- 17.67 x 10 2o andj <p3= 10
Art rtrj 4 r\o
10 \ A

Problem 2. Light from a gas-discharge tube falls nor­


mally onto a diffraction grating with 600 lines per mil­
limetre. The diffraction spectrum is examined through
a telescope mounted on a graduated circle. The red line
in the first-order spectrum is seen at an angle of 23°, and
the green line at an angle of 19°8'. Find the wavelengths
of these lines
Given: N = 600 lines/mm, = 23°, q)2 = 19°8'.
Required: k±, k2.
Solution: The condition for the maximum illumination
in diffraction by a grating is (a + b) sin q) = 2nkl2 = nk =
= (1 IN) sin q), where a + b = i/N is the grating spacing,
N is the grating constant, and n is the order of the diffrac-
tion spectrum. Hence,
A,1= -^'sinq)1= g^sin23° = 6.5 x 10‘4 mm = 6500 A

^ ^ ^ s i n 19°8' — 5.46
OOU
x 10“4 mm =5460 A

5. Polarization
1°. By the polarization of light is meant the combination
of phenomena of wave optics (V.2.1.10) in which the trans­
verse nature of electromagnetic light waves (IV.4.2.4°)
426 Optics V.2,5

manifests itself. An electromagnetic light wave is defined


as plane-polarized (linearly polarized) if the directions of
oscillations of the vectors E and B in this wave are strictly
fixed and are in definite planes. Figure V.2.15 shows the
planes A and C in which the vectors E and B oscillate
in a plane-polarized wave travelling with the velocity

Fig. V.2.15 Fig. V.2.16

v in a homogeneous isotropic medium. A plane-polarized


(linearly polarized) light wave is called plane-polarized
(.linearly polarized) light.
2°. An electromagnetic light wave is called natural
(;unpolarized) if the directions of oscillations of the vectors
f
E
\ e 1 p

'E
E
Fig. V.2.17

E and B in this wave can be in any planes perpendicular


to the vector of the velocity of propagation of the wave.
Figure V.2.16 shows possible directions of oscillations
of the vector E in natural, unpolarized light. The vector
can oscillate in any of the directions shown in Fig. V.2.16.
The directions of oscillations of the vector B are always
perpendicular to the directions of oscillations of the vec­
tor E and are not shown in Fig. V.2.16. In other words,
V .2 .6 Wave Optics (Light Waves) 427

natural, unpolarized light is the name given to light waves


in which the directions of oscillations of the vectors E
and B change randomly so that all the directions of oscilla­
tions in planes perpendicular to a ray are equally probable.
3°. Devices used to transform natural light into po­
larized light are called polarizers. A crystal of tourmaline
is a very simple polarizer. It can transmit light waves
with oscillations of the vector E (and correspondingly of
B) in one definite plane. Figure V.2.17 shows that if natural
light is incident on a plate made
from a tourmaline crystal cut out
so that one of its sides coincides
with the axis of the crystal, then
the tourmaline transmits only a
light wave in which the vector E
oscillates in a definite plane P.
4°. We can convince ourselves
that the light passing through a
tourmaline crystal is plane-polar­
ized with the aid of a second tour­
maline crystal playing the part of
an analyser—a device permitting
us to detect the position of the
plane in which the vector E oscil­
lates in linearly polarized light. .
Figure V.2.18 shows that when Fig. V.2.18
light is passed through two con­
secutive tourmaline crystals, it passes entirely through
both of them when their axes are parallel to each other,
and is gradually absorbed to an increasing extent when
the angle between the axes approaches 90 degrees. If the
two crystals are placed with their axes at right angles to
each other (Fig. V.2.18c), then the light is entirely absorbed
in the analyser.

6. Dispersion
1°. The dispersion of light is the phenomenon of the
dependence of the absolute refractive index of a sub­
stance n (V. 1.2.1°) on the frequency v of the light falling
428 Optics V.2.6

on a substance (or on the wavelength in a vacuum


= c/v, where c is the. speed of light in a vacuum):
rc = / (v) = cp(lo)
It follows from the definition n = civ that the disper­
sion of light can be characterized as the phenomenon of
the dependence of the speed
of the propagation of a
light wave in a substance
on its frequency: v = f 1(\).
B* The dispersion of light
is called normal if the re­
fractive index grows mono­
tonously with increasing
frequency (diminishes with
increasing wavelength);
otherwise dispersion is called
abnormal, or anomalous.
Figure V.2.19 shows the function n = /(v) for normal and
anomalous dispersion.
2°. Newton experimentally discovered the dependence of
the absolute refractive index on the frequency of light.

red
orange
yellow
green
light blue
dark blue
violet

Fig. V.2.20

He showed in a series of experiments that when non-mono-


chromatic white light is passed through a prism, a visible
rainbow strip M N is observed on a screen set up behind
the prism (Fig. V.2.20). The strip consists of seven colours
called a prismatic, or dispersive, spectrum. Thus, the dis­
persion of light results in the decomposition of white,
V.3J Radiation and Spectra 429

noil-monochromatic light into its monochromatic compo­


nents each of which has a definite frequency (or wavelength)
(the spectral decomposition of wtiite light).
The dispersive spectrum of visible light occupies the
range from X = 7.5 X 10~5 cm (red rays) to 3.9 X 10~5 cm
(iviolet rays) on the scale of electromagnetic waves.
3°. The dispersion of light, like interference and dif­
fraction in non-monochromatic light, proves that a mono­
chromatic electromagnetic wave with a definite frequency
(or wavelength in a vacuum), belonging to the range of
visible light waves (IV.4.1.1°), results in the perception
by the eye of a definite colour with the aid of the cones
(V. 1.7.3°). Strictly monochromatic light cannot exist in
principle. This is associated with the processes of the emis­
sion of light (VI.1.5.2°).

Chapter 3
Radiation and Spectra

1. Thermal Radiation. Blackbody


1°. All bodies emit electromagnetic waves (IV.4.4.1°)
as a result of the transformation of the energy of the chaotic
thermal motion of their particles into the energy of radia­
tion. The electromagnetic waves emitted by a body—
a source of thermal radiation—in the state of thermody­
namic equilibrium (11.3.1.3°) are called isothermal radia­
tion. The latter is produced by a source at a constant tem­
perature. Isothermal radiation is achieved if the source
S of this radiation is within a closed space o with opaque
walls whose temperature equals that of the source (Fig. V.3.1).
The intensity of the electromagnetic waves (IV.4.3.20)
emitted by the source will equal the energy absorbed by
the walls of the space in a unit time (Par. 3°). Thermody­
namic equilibrium will exist between the radiation field
430 Optics V.3.1

(IV.4.4.1°) and the walls. The Sun is a source of isothermal


radiation. Its temperature is kept constant by the liberation
of energy in thermonuclear reactions (VI.4.15.1°).
Non-isothermal radiation (or simply thermal radiation)
occurs when the source of radiation is heated. For example,
in incandescent lamps, a small portion of the heat liberated
when an electric current (111.2.7.4°) is passed through
their filaments transforms into
the energy of electromagnetic
waves.
2°. The emissivity of a body
E v%
t is the physical quantity
numerically equal to the energy
of thermal radiation of a given
frequency v that is emitted at
the temperature T by a unit sur­
face area of the body in a unit
time. The emissivity of a body
Ey,%
\t shows the fraction formed
by thermal radiation of the given
frequency v in the total ther­
mal radiation of a source. Apart from the frequency v
and temperature 71, the quantity EV%T depends on the
material of a body and the state of its surface.
3°. The absorptivity of a body AV|T is the physical quan­
tity showing the part of the energy of an electromagnetic
wave of a given frequency v incident in a unit time on
a unit surface area of the body absorbed by the body. Apart
from the frequency v and temperature T, the quantity
A VtT depends on the material of a body and the state of
its surface.
4°. A blackbody is defined as a body which at any tem­
perature, regardless of the material of the body and the
state of its surface, completely absorbs electromagnetic
waves of any frequencies v, i.e. all the rays incident on
it: A VtT = 1 for a blackbody. Carbon black, black velvet,
and platinum black are close in their properties to a black­
body. The Sun is close to a blackbody in its optical pro­
perties.
A small opening O made in the opaque wall of a box B
(Fig. V.3.2) is a model of a blackbody. A ray entering
V .3 .2 Radiation and Spectra 431

the box through the opening after repeated reflection back


and forth by the internal walls of the box is virtually com­
pletely absorbed regardless of th<* material of the walls—
the opening in the wall seems to be black.
5°. Kirchhoff's law for thermal radiation: for an arbitrary
frequency v and temperature T, the ratio of the emissi-
vity EVtT of an opaque body to
its absorptivity A VtT is the same.
It equals the emissivity eVfr of
a blackbody:

The value of ev,r depends only on


the frequency and the temperature
of a blackbody. It follows from
Kirchhofi’s law that EVtT is di­
rectly proportional to A VtT for
any opaque body. If at a given
temperature T a body does not
absorb a certain frequency v, then it also cannot emit
equilibrium radiation of the same frequency (if A VtT = 0,
then = 0).
The significance of Kirchhoff’s law is that it permits
us to express the emissivity £ ViT of any opaque body through
the emissivity ev>T of a blackbody if we know the value
of A VfT for it.

2. Energy Distribution
in the Spectrum
of a Blackbody
1°. A blackbody consists of a tremendous number of
ntoms. Each of them as regards its properties of emitting
electromagnetic waves is similar to a miniature oscillator—a
dipole (IV.4.4.50 = 7°). Every oscillator atom oscillates
with many frequencies and emits energy of the correspond­
ing frequencies. Therefore, a blackbody radiates electro­
magnetic waves of all possible frequencies. The dependence
432 Optics V.3.2

of the emissivity ev,r °f a blackbody on the frequency v


of radiation at a constant temperature T is called a curve
of energy distribution in the spectrum of the blackbody.
This relationship ev,r = /(v) has the nature of a conti­
nuous curve. A curve obtained experimentally is shown
in Fig. V.3.3 by a solid line. Figure V.3.4 shows curves
of 8V|t versus v for different temperatures. Examination

of these curves shows that with elevation of the tempera-


ture the greatest power emitted by a unit surface area
of a blackbody falls to constantly growing frequencies
v m4 ^ V ma ^ 'V m , V TOt.
2°. Theoretical investigation of the shape of the curve
8v,t = / (v) with T = const played an outstanding part in
the development of physics. The shape of this curve obtained
by physicists at the end of the nineteenth century is shown
by a dash line in Fig. V.3.3. The theoretical curve can
be seen to differ appreciably from the experimental one.
A very low value of ev,r falls to the part of high frequen­
cies in the experimental curve (the point N on the curve
in Fig. V.3.3).
The theoretical curve ev,T = /(v) with T = const has
no maximum (the point M on the experimental curve) and
depends on the frequency according to the law

ev, T= Cv2
VJJ Radiation and Spectra 433

where C is a constant. The unlimited growth of ev,7’ with


an increasing frequency v contradicts not only the experi­
mental curve^ but also the law of conservation of energy.
A blackbody having a definite finite store of internal energy
at a certain constant temperature would have to radiate
an unlimitedly great power from a unit area: cVfr -*■ 00
at v — 00 . This conclusion was called the “ultraviolet
catastrophe”, “ultraviolet” because it involves high (ultra-
violet) frequencies, and “catastrophe” because it predicts
an apparent violation of the law of conservation of energy.
The conditional nature of this name is obvious. Actually,
the frequencies (and wavelengths) of ultraviolet rays are
finite and occupy a definite range on the electromagnetic
spectrum (V.3.7.10). The term “catastrophe” stresses that
it corresponds to very high frequencies v that increase
unlimitedly.
3°. The cause of the difficulties arising in the theory
of the thermal radiation of a blackbody was associated
with one of the fundamental postulates of physics that
appeared as a result of its prolonged development. This pos­
tulate states that the energy of any body (or system) can
change continuously, i.e. can take on any infinitely close
values. The energy E of a system is a combination of con­
tinuous possible values.
According to M. Planck’s hypothesis, the energy of an
oscillator atom (Par. 1°) of a blackbody changes only in
definite portions (quanta) that are multiples of a certain
energy e, i.e. can take on values of e, 2e, 3e, . . ., ne.
The value of an elementary portion of energy e is called
a quantum of energy:
e = hv
where v = frequency of oscillations of an oscillator atom
h = Planck's constant (a quantum of action), h —
= 6.625 X 10-34 J-s.
The name “quantum of action” used for h is associated
with the dimension of this quantity. A physical quantity
having the dimension of the product of energy and time is
called an action.*
* The value of Planck’s constant is obtained from the laws of
radiation of a blackbody, which are not treated in elementary physics
(see also V.5.3.4°).
4°. It follows from Planck’s hypothesis that an oscil­
lator atom of a blackbody oscillating with the frequency
{E
v0 can have only the values
of the energy indicated in
1 3/iVn
”*« Fig- V.3.5.
According to Planck’s hy-
2/?vn pothesis, the energy of elec­
tromagnetic waves can be
■hVn emitted and absorbed in sep­
‘h arate quanta that are mul­
tiples of the quantity fcv. The
Fig. V.3.5 consistent development of
Planck’s hypothesis led to
the appearance of notions of the quantum properties
of light (V.5.1.10).

3. Luminescence
1°. Apart from thermal radiation (V.3.1.10), bodies at
the same temperature can have another form of radiation
in excess of thermal radiation—luminescence* that is not
associated with the transfer of the energy of thermal motion
of the molecules into the energy of electromagnetic waves.
Luminescence consists in the emission of light by sources
at the expense of the supply of energy to them owing to
various processes.
2°. Cathodoluminescence is defined as the glow of bodies
due to the bombardment of a substance by electrons or
other charged particles (for example ions).
Electroluminescence is caused by the passage of an elec­
tric current through a substance or the action of an electric
field on it. In these kinds of luminescence, the kinetic energy
of the charged particles of the energy of the electric field
is partly transmitted to the atoms (molecules) of the sub­
stance that emit electromagnetic waves.
The glow of a gas discharge (111.3.5.2°) in the tubes
of advertisements is an example of these kinds of lumines­
cence. Streams of charged particles from the Sun cause

* From the Latin lumin meaning light.


V.3.4 Radiation and Spectra 435

the glow of atoms of gases in the upper layers of the at­


mosphere that is visible from the Earth.
Some chemical reactions in substances attended by the
liberation of energy are the cause of a glow called chemilu­
minescence. The glow of many living organisms such as
bacteria, insects, and fish is due to chemical reactions.
Photoluminescence is the glow of bodies as a result of
their irradiation by visible or ultraviolet light, X-rays
(V.3.6.10) or gamma rays (VI.4.4.2°). In this case, the
energy of the incident radiation is partly transformed
into the self-radiation of the substance. In luminescent
sources of light—daylight lamps—the internal surfaces
of a discharge tube are coated with luminophors. These
are substances that under the action of ultraviolet light
or some other short-wave radiation with a high frequency
begin to emit visible light of a smaller frequency.
3°. The cause of all luminescent phenomena is that the
centres of luminescence—the atoms, molecules, or ions of
a substance that is a source of luminescent glow—are trans­
ferred into excited states (VI.2.5.3°) at the expense of
(he energy supplied by an extraneous source. The transi­
tion of the excited centres of luminescence into the "normal
or less excited states (VI.2.4.3°) is attended by the emission
of light, which is luminescent glow of the substance itself.

4. Kinds of Spectra
1°. The combination of frequencies (or wavelengths)
contained in the radiation of a substance is called the
emission spectrum of the substance. The combination of
frequencies (or wavelengths) absorbed by a given substance
is called its absorption spectrum.
2°. Glowing gases (rarefied) in the atomic state pro­
duce line emission spectra consisting of separate narrow
spectral lines. The latter have a definite intensity (IV.4.3.20)
aud are separated from one another by dark spaces. Iso-
Inted atoms of a chemical element emit a quite definite
combination of spectral lines belonging only to this ele­
ment. For example, the vapour of sodium in a vacuum
has two bright yellow lines of wavelengths 5890 and 5896 A
in its spectrum in addition to other lines.
4a6______________________ dpttct____________________

3°. Emitting molecules produce band emission spectra


in which a multitude of close spectral lines form groups—
bands separated by dark spaces. For the origin of band
spectra of molecules see VI.3.4.4°.
4°. Molten solids and glowing liquids create continuous
emission spectra that are a continuous sequence of frequencies
(or wavelengths) smoothly changing into one another.
An example of a continuous spectrum is the emission spectrum
of a blackbody (V.3.2.10) with continuous distribution of
the energy in it (see Fig. V.3.3). The glowing surface of
the Sun—the photosphere—produces a continuous spectrum.
5°. Reversal of emission and absorption spectral lines:
the atoms of a given chemical element absorb the same
spectral lines (more exactly, frequencies) which they them­
selves emit. The gas envelope surrounding the Sun—the
chromosphere—and the Earth’s atmosphere absorb a number
of lines in the continuous spectrum emitted by the Sun,
and numerous dark Fraunhofer lines (named after J. Fraun­
hofer who discovered them) are observed in the continuous
spectrum of the Sun.
6°. Spectral analysis is defined as the studying of the
chemical composition and concentration of atoms (and
molecules) in a substance according to its spectrum. To
determine concentrations in quantitative spectral analysis,
we consider not only the positions of the spectral lines
and bands (their frequencies or wavelengths), but also their
intensities (iV.4.3.2°). The methods of quantitative spec­
tral analysis allow us to determine very minute amounts
of a given element in the composition of a compound sub­
stance (of the order of magnitude of 10“13 kg). The chemical
composition of the Sun and stars is determined in this
way.

5., Infrared and Ultraviolet


Radiation
1°. The radiation directly following the red part of a
prismatic spectrum (V.2.6.20) is called infrared radiation.
Infrared rays with wavelengths X exceeding 7.5 X 10“5 cm
fill the interval between ultrashort radio waves (IV.4.5.10)
V.3.6 Radiation and Spectra 437

with wavelengths X ranging from 1 to 2 mm and visible


light (V.2.6.20) on the electromagnetic spectrum (V.3.7.10).
All heated bodies are sources «of infrared radiation.
2°. Any absorbed radiation heats bodies to some extent
or other. Infrared radiation, however, is detected only
according to its thermal action—the heating of the bodies
that absorb it. All heating appliances are sources of in­
frared waves and heat the bodies that absorb these waves.
Infrared radiation acts on special photographic plates
sensitive to such rays. This is the basis of photography
(V.5.6.20) in infrared rays, which is possible at any time
of the day.
3°. The radiation directly following the violet part of
a prismatic spectrum (V.2.6.20) is called ultraviolet radia­
tion. Ultraviolet rays with wavelengths X less than 3.9 X
X 10~5 cm occupy a position on the electromagnetic spec­
trum (V.3.7.10) between violet rays with wavelengths
of 3.9 x 10"5 cm and X-rays (V.3.6.1°) with wavelengths of
about 8 X 10“6 cm.
4°. The invisible ultraviolet light is detected accord­
ing to its active chemical and biological action. Special
protective glasses absorbing or reflecting ultraviolet ra­
diation do not let the eye’s retina (V .l.7.30) be harmed by
the destructive action of ultraviolet rays. Various doses
(VI.4.14.1°) of ultraviolet light when acting on the tissues
of the skin result in the formation of a protective pigment,
tan, of the vitamin D2, and have a bactericidal action—de­
stroy pathogenic bacteria.
All the features of the action of ultraviolet radiation
are explained by the energy of its quanta hv (V.3.2.40)
having a greater value than that of quanta of visible light.

6. X-Rays
1°. X-radiation is defined as electromagnetic waves hav­
ing a shorter wavelength than that of ultraviolet rays
(V.3.5.30). X-rays occupy a broad range of wavelengths
from 8 X 10"({ to 10~10 cm on the electromagnetic spectrum
(V.3.7.10). X-radiation is produced when the kinetic energy
438 Optics V.3.6

of fast electrons is transformed into the energy of electro­


magnetic waves.
2°. Two kinds of X-radiation are distinguished: white
(continuous) and characteristic.
White X-radiation appears when fast electrons are retard­
ed during their motion in a substance, particularly in
metals. According to IV.4.4.30, when an electric charge is
retarded, it emits electromagnetic waves. The white ra­
diation of electrons has a con­
tinuous spectrum (V.3.4.40)
(an X-ray continuous spec­
trum). The latter appreciably
differs from the continuous
radiation spectra emitted by
solids or liquids. First, it is
in the far short-wave region,
and, second, the continuous
X-ray spectrum is limited at
the side of small wavelengths
by a certain boundary Xmln
called the boundary of a con­
tinuous X-ray spectrum. Fig­
ure V.3.6 shows typical curves
of the distribution of X-ray
Fig. V.3.6 intensity by wavelengths for
different potential differences
across the X-ray tube electrodes (Par. 4°). The existence
of ^min can be explained on the basis of quantum notions
of the nature of light (V.5.1.20). The maximum energy
hvmax a quantum of X-rays produced at the expense
of the energy of an electron E cannot exceed this energy,
i.e.
h v m a x == ^ 1=3ecpo

where cp0 is the potential difference at whose expense the


energy E was imparted to the electron (111.1.8.5°), and e
is the absolute value of the charge of an electron. If we
pass over from the frequency vmnx to the wavelength,
then
« c __ ch ch

^lln vmax *<Po Q


V.3.6 Radiation and Spectra 439

This formula can be verified experimentally by calculating


Planck’s constant (V.3.2.30) from it. The value of h ob­
tained in this way is one of the mo$t accurate and authentical
ones.
3°. Characteristic X-radiation has a line spectrum
(V.3.4.2°). This name is due to the fact that the frequencies
of the spectrum lines in this radiation are characteristic
of the substance in which the fast electrons are retarded.
Characteristic radiation is the result of an external fast
electron, when retarded in a substance, tearing out from

X-ray photons
Fig. V.3.7 Fig. V.3.8

an atom of the substance an electron on one of the inner


electron shells (VI.2.8.6°). Another electron of the atom
passes over to the vacancy from a shell farther from the
nucleus. The result is the appearance of an X-ray photon
(V.5.1.20) with a definite frequency v characteristic of
Ilie given atom and having the charge Ze (VI.2.1.1°). Fig­
ure V.3.7 shows how an X-ray photon is formed when an
electron is knocked out of the /ST-shell of an atom and an
electron from the L-shell passes over to the vacancy.
4°. An X-ray tube is a device for producing X-rays.
Figure V.3.8 shows the design of an electronic X-ray tube.
The hot tungsten filament Ca heated by a special battery
nr a special heater transformer is a source of electrons.
The stream of electrons is accelerated in the strong electric
held set up by a high-voltage source between the anode
mikI the cathode. The accelerated stream of electrons is
retarded in the substance of the anode A and causes X-rays
In appear. (In powerful X-ray tubes, tha anode, or anti-
cathode, is cooled to prevent overheating when the electrons
are retarded.)
440 Optics V.3.6

5°. The small wavelength of X-rays and their great


“hardness” underlie the fundamental properties of X-rays:
a high penetrating ability, action on photographic emulsions,
and an ability of causing ionization in substances through
which these rays pass.
The absorption of X-rays attended by the transition
of their energy into the internal energy (11.4.1.2°) of the
substance depends very greatly on the atomic number of
the substance (it is proportional to Z4). The different ab­
sorption of X-rays when they pass through non-uniform
substances is taken advantage of in medicine, science and

Fig. V.3.9 fFig. V.3.10


engineering. When a human body is X-rayed, the absorption
in the bones consisting mainly of calcium phosphate is
about 150 times greater than the absorption in the soft
tissues of the body where water is responsible for the main
absorption. In X-raying, a sharp shadow is formed by
the bones. X-ray diagnostics—the identification of various
new growths in an organism—is based on the principle
of the different absorption of X-rays by substances having
different densities. X-ray therapy is based on the possi­
bility of destroying new growths by means of X-ray pho­
tons. X-ray flaw detection is based on the dependence of
the absorption of X-rays on the atomic number Z in opti­
cally opaque solid bodies. The various flaws in a body
(for example impurities, etc.) are detected on the screen
when X-rayed. If Zx < Z2, where Zx and Z2 are the atomic
numbers of a flaw and the substance of the body, respec-
V.3.7 Radiation and Spectra 441

tively, then the region occupied by flaws will be lighter


on the screen than the remaining field of vision. In the
opposite case (Z2 < Z2), the defective region will be darker.
6°. The condition for observing diffraction (V.2.3.10)
is obeyed for X-ray radiation upon diffraction by the crys­
tal lattices of solids because the periods of the crystal
lattices (11.1.6.5°) are commensurable with the wavelengths
of the X-rays (diffraction by crystalline structures ). In the
diffraction of X-rays by monocrystals (11.1.6.4°), the dif­
fraction pattern has the form of separate spots regularly
arranged around a central spot (Fig. V.3.9). In diffraction
by polycrystals (11.1.6.4°), the diffraction pattern has the
form of symmetrically arranged concentric rings (Fig. V.3.10).
The X -ray diffraction analysis of crystals—investigating
the structure of the crystal lattices of solids—is carried
out by studying the arrangement of the diffraction spots
and rings.

7. The Electromagnetic Spectrum


1°. Figure V.3.11 shows to a logarithmic scale the
electromagnetic spectrum —a continuous sequence of frequen-
log v (v is in Hz)
4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 if) 90 91
1 R a d io w a v e s 1 1

o
.i
>
*8
-E -S
% r r
l•s i
i -§
i
___ 1___ » i
3 2 1 0 -1 -2 -3 -4 -5 -6 -7 -8 -9 -10 -11 -12
log A (X is in m)

Fig. V.3.11

cies and wavelengths of electromagnetic radiation consisting


of a varying electromagnetic field (IV.4.1.1°) propagating
in space.
442 Optics V.4.1

2°. A glance at the electromagnetic spectrum shows


that the boundaries between the frequencies v (or the wave­
lengths in a vacuum X0 = c/v) of different kinds of electro­
magnetic radiation are very conditional—adjacent por­
tions of the spectrum continuously transform into one
another. Electromagnetic waves whose frequencies differ
by many orders of magnitude (for example radio waves
and X-rays) have qualitatively different properties. These
differences are determined by the general law of the electro­
magnetic spectrum: as we pass from longer waves (low
frequencies) to shorter ones (high frequencies), the wave
properties of light (interference, diffraction, polarization)
manifest themselves to a smaller extent, while the quantum
properties (V.5.1.10) in which the decisive part is played
by the magnitude of a quantum of energy hv (V.3.2.30)
manifest themselves to a greater extent.
3°. The electromagnetic nature of light waves in the
spectrum of Fig. V.3.11 is expressed in that visible light
occupies a quite definite, although narrow, section of it.

Chapter 4*
Fundamentals
of the Special Theory
of R elativity
1. Laws of Electrodynamics and
the Classical Principle
of Relativity
1°. The velocity of light c in a vacuum, according
to the classical addition law for velocities (1.2.7.2°), should
vary in the different inertial reference frames (1.2.1.3°)
shown in Fig. V.4.1. If it is c in the frame XYZ associated
with the light source (Fig. V.4.1a), then in the frame X'Y'Z'
V .4 .1 The Special Theory of Relativity 443

Z '\
Zk

~r '
c"'= £z^ ?

id)
fig. V.44
444 Optics V.4.2

travelling relative to XYZ as shown in Fig. V.4.16, the


velocity of light in a vacuum should be c = c — v. Cor­
respondingly in the frames X"Y"Z" and X WY " ' Z trav­
elling relative to XYZ as shown in Figs. V.4.1c and V.4.1d,
the velocity of light in a vacuum ought to be c" = c + v
and c" = Y 7 + 7 " .
2°. The conclusion on the possibility of there being
different values of the velocity of light in a vacuum in
various inertial reference frames signifies that a very im­
portant theorem of Newtonian mechanics—the classical
principle of relativity (1.2.7.5°)—must be violated in
electrodynamics. Another assumption that the classical
principle of relativity has a universal applicability, but
that the entire system of laws of electrodynamics and
optics (see Parts IV and V of this book) is not correct,
sharply contradicts the enormous number of experiments
that have been explained on the basis of these laws.
3°. A great number of experiments set up with a very
high accuracy were aimed at detecting the dependence of
the velocity of light c on the motion of the reference frame
relative to the source. They all gave negative results.
The velocity of light was the same in all the inertial ref­
erence frames regardless of the magnitude and direction
of their velocity, and was equal to the value of this velo­
city in the reference frame associated with the source
(Par. 1°):
c = c" = c'" = c

2. Postulates
of the Special Theory
of Relativity
1°. The1! special* theory of relativity also called the re­
lativistic theory is based on two postulates.
The first postulate is the principle of relativity: all phys­
ical phenomena (mechanical, electromagnetic, etc.) pro-
* Apart from his special theory of relativity, A. Einstein also
developed a general theory of relativity, information on which is far
beyond the scope of an elementary course in physics and of the present
hookT
V.4.2 th e Special theory of Relativity 445

ceed identically in the same conditions in any inertial


reference frames. No experiments conducted in a closed
system of bodies can show whether the system is at rest
or moves uniformly and rectilinearly.
The principle of relativity is a generalization of the
classical principle of relativity for all physical phenomena,
particularly for electromagnetism.
The second postulate is the principle of the constancy
of the velocity of light: in all inertial reference frames, the
velocity (speed) of light in a vacuum is the same and does
not depend on the velocity of the light source.
According to the first postulate of the special theory of
relativity, the laws of electrodynamics and optics hold in
all inertial frames of reference.
Therefore, all experiments K K'
set up with the object of de­
tecting how the Earth’s mo­
tion along its orbit affects the
laws of electromagnetic phe­
nomena always had negative
results.
2°. The postulates of the
special theory of relativity
obviously contradict the no­
tions of time and space formed
in Newtonian mechanics. This
can be seen from the exam­
ple shown in Fig. V.4.2. The inertial] framed of reference
K' (X', Y', Z') with the centre O' that coincides when
t = 0 with the centre O of the frame K (X, Y, Z) travels rela­
tive to the stationary frame K as shown in the figure. At
the moment t = 0, a light flashes at the centre 0 in a point
source of light (V .l.6.30). By the moment t > 0, the sphe­
rical wavefront (IV.3.1.5°) will reach the sphere A of radius
r = ct in the frame K. But we can consider with the same
right that the flash occurred in the source at the point
O' of the frame K' when the points 0 and O' coincided. Ac­
cording to the second postulate, the light wave should trav­
el relative to the frame K' in exactly the same way as rel­
ative to the frame K because the velocity of light c does not
depend on the velocity v of the source—point O'. According
446_______________________________ d p t t c i _____________________________ V A M

to the first postulate, the frames K and K ' are absolutely


equivalent, and the light must propagate in the moving
frame K' in the same way as in the stationary one. Hence, by
the time t, the front of the wave emitted from the point O'
should be a sphere of the same radius ct with its centre at the
point O'. But the centre O' of this sphere during the time t
moves away from the centre 0 of the frame K over the di­
stance vt. A situation arises that contradicts common sense:
during the time t the spherical wave reaches a sphere of defi­
nite radius r = ct whose centre is simultaneously at two
different points 0 and O'.

3. Concept of the Length


of a Body

1°. Einstein found that the contradiction noted in


V.4.2.20 had been in the limitation of the Newtonian con­
cepts of space and time, which were taken for granted. In
the special theory of relativity, these notions were revised
quite seriously.
When using the two basic concepts of length and time
in physics, the ways of unambiguously measuring the length
and time intervals are indicated. The length l0 of a bar is
measured by comparing it with the length of a standard body
which, by definition, is considered to equal a unit length.
It is simple to use ^this method of measuring length if the bar
and the standard rule are fixed in the frame K where the
measurements are being made. If the length is being mea­
sured in the reference frame K' moving together with the
rule that is the standard of length in this frame, then the
length of the bar Iq will coincide with l0: Z'0 = Z0. If matters
were not so, if, for example, it was found that Zq > Z0,
then owing to the equal rights of the frames K and K' we could
consider that the frame K' is stationary and the frame K
moves relative to K' with the velocity —v. Consequently,
owing to our above assumption, we would have Z0 > Z'0.
The two inequalities Iq > Z0 and Z0 > Zq are contradictory,
and therefore Zq = Z0.
V.4.i The Special Theory of Relativity 447

2°. The length of a bar moving together with the refer-


ence frame K' and arranged along the axis O'X' (Fig. V.4.3)
can be determined with the aid o f a measuring rule in the
stationary frame K . For
this purpose, the coordin­ ^A .A
ates xx and x2 of the be­
! M ,/V
ri PZZZZZZZA".
ginning M and end N of the i

moving bar have to be 01 0 1 * B l ___ **_____ 9


~X'
measured at an arbitrary,
but the same, moment of
time (V.4.4.20). The length
of the bar will be I = x2— Fig. V.4.3
(Fig. V.4.3). In the
moving frame K ', the coordinates of the ends of the bar will
be x[ and x%, respectively, and the length of the bar, which
is stationary in the frame K \ will be l0 = x 2 — x\.
In Newtonian mechanics, it follows from the Galilean
transformations (1.2.7.1°) that x^ — x[ = x2 — xx, and
therefore I = lQ. In the special theory of relativity, this mat­
ter is solved in a different way (V.4.7.20).

4. Simultaneity of Events.
Synchronization of Timepieces
1°. Time is measured by means of a standard whose
capacity is filled by a periodic process (the rocking of a pen­
dulum, the motion of the hands of timepieces over a dial,
etc.). The measurement of time is associated with the concept
of the simultaneity of two events.
By an event is meant any phenomenon occurring at a
given place having the coordinates x, y, z, at a certain mo­
ment of time t. Two or more events are called simultaneous
if they occur at the same moment t. The simultaneity of two
events occurring at the same place is established according
to the reading of a clock installed at this place.
In Newtonian mechanics, time is treated as absolute.
According to I. Newton, time flows identically without
regard to any external thing. The age of substances remains
the same regardless of whether their motions are fast or slow,
or are in general absent. The simultaneity of two events oc-
Us Optics V.4.4

curring at the same and at different points of space is consid­


ered to be an obvious concept in Newtonian mechanics,
not requiring any indication as to how simultaneity must
be detected.
2°. To determine the simultaneity of two events occur­
ring at the different points A and B of space, it is necessary
to install clocks running at the same rate {synchronously)
at them. The clocks installed at A and B will be synchronous
if at the moment when the clock at the point A shows the
time t±1 the one at the point B shows the same time tx and
runs at the same rate as the clock at the point A. That the
rates of the clocks are identical can be checked experimental­
ly if we send a signal from the point A at definite equal in­
tervals of time and note on the second clock installed at B
the intervals between the moments when the signals arrive
at the point B.
3°. It would be very simple to check the identical run­
ning of the clocks at A and B with the aid of a signal travel­
ling instantaneously between these points. The absence of
such signals in nature signifies that the question of the iden­
tical readings of two clocks at different points can be solved
by synchronization of the clocks. Assume that a light signal
is dispatched according to the clock at the point A at the
moment tx and after reflection at the point B returns to A at
the moment t2. The clock at B is considered to be synchronous
with the one at A if they run at the same rate, and when the
signal arrives at B the clock at B shows the time t =
= -(- t2)/2.
4°. The choice of a light signal in a vacuum as the phys­
ical process for the synchronization of clocks is determined
by the fact that the speed of any signal in nature cannot
exceed that of light c in a vacuum. The limiting nature of the
speed of light c is not a postulate of the special theory of
relativity, but it plays the same part as the postulates of
this theory and is confirmed by numerous experiments.
The synchronization of clocks at different points of space
makes it possible to time reference frames: a quite definite mo­
ment t corresponds to every event regardless of the place
where the event occurred.
V.4.5 The Special Theory of Relativity 449

5. Relative Nature
of Simultaneity of Events
1°. It follows from the Galilean transformations
(1.2.7.1°) that if two events occur in the frame K at the mo­
ments tx and t2, and in the frame K' (see Fig. V.4.2) at the
moments t[ and t2, respectively, then since t = t', the in­
terval of time between the two events is the same in both
reference frames:
At = t2— t i = t ’2—t\ = At'
The simultaneity of two events in Newtonian mechanics is
absolute and does not depend on the reference frame
(1.1.1.3°): if At = 0, then we also have At' = 0.
2°. In the special theory of relativity, the simultaneity
of two events occurring at different points of space is rela­

tive: events that are simultaneous in one inertial reference


frame are not simultaneous in other inertial frames moving
relative to the first one.
Figure V.4.4 shows schematically an experiment illus­
trating this. The reference frame K is associated with the
Earth, and the frame K ' with a car moving rectilinearly
and uniformly relative to the Earth with the velocity v.
The points A, M, and B are marked on the Earth, and the
points A \ M ', and B' in the car correspond to them; here
AM = MB and A'M ' = M 'B '. At the moment when the
points marked on the Earth and in the car coincide, two
events occur at A and B —there are two flashes of lightning
at them. In the frame K , the signals from the two flashes
will arrive at M simultaneously because AM = M B, and
Ihe speed of light is the same in all directions. Hence, the
450 Optics V.4.6

events at A and H occurred simultaneously. In the frame


K ' associated with :he car, the signal from B' will arrive at
M' before that from A' because the speed of light is the same
in all directions, but M ' is moving toward the signal emit­
ted from B' and away from the signal emitted from A'.
Consequently, the events at A' and B' are not simultaneous:
the event at B r occurred before that at A'. If the car were
moving in the other direction, the result would be opposite:
the event at B r would occur later than at A'.
3°. The concept of the simultaneity of events separated
in space is relative. If follows from the postulates of the
theory of relativity and the existence of a finite velocity
of propagation of the signals that time flows differently in
different inertial reference frames. This solves the contradic­
tion in the arrangement of the fronts of the spherical waves
noted in V.4.2.20. Light simultaneously reaches points
of the spherical surface with its centre at the point O from
the viewpoint of an observer who is stationary in the frame
K. From the viewpoint of an observer associated with the
frame K \ light reaches points of the spherical surface at
different moments, and there is no paradox with the centres
of the spheres.

6. Lorentz Transformations
1°. In accordance with the two postulates of the special
theory of relativity (V.4.2.10), relationships called Lorentz
transformations exist between the coordinates and time in
two inertial reference frames.
2°. In the simplest case, when the frame K' moves with
the velocity v relative to the frame K as shown in Fig. V.4.3,
the Lorentz transformations for the coordinates and time
have the following form:

X —vt
Y i —y2/c2*
x —
(—v t
X
Y 1 — v 2J c 2 ’
V .4 .7 The Special Theory of Relativity 451

3°. The close relationship between the space and time coor­
dinates in the special theory of relativity follows from the
Lorentz transformations. Not onTy do the space coordinates
depend on the time [as in the Galilean transformation
(1.2.7.1°)I, but also the time in both reference frames de­
pends on the space coordinates, and on the velocity v of the
reference frame K '.
4°. The Lorentz transformations are converted into the
Galilean ones provided that vie 1. In this case,

x' = x — vt, y' = y, z r = z, t' = t


x = x' -f vt, y = y ', z = z', t = t'

The transition of the formulas of the special theory of rela­


tivity into the formulas of kinematics provided that vie ^ 1
is a verification of the formulas of the special theory of re­
lativity.
5°. The principle of relativity (V.4.2.10) can be formu­
lated with account of the Lorentz transformations as follows.
All the laws of physics describing any physical phenomena
must have the same form in all inertial reference frames.
This signifies that when passing over from one inertial re­
ference frame K to another frame K' with the aid of Lorentz
transformations, the laws of physics must retain their form.
This conclusion is called the relativistic invariance (the
Lorentz invariance) of the laws of physics.

7. Relativity of Lengths
1°. It follows from the Lorentz transformations for the
coordinates x and x' and the time t and t' that vie ^ 1.
Otherwise, these coordinates and times will be imaginary.
The velocity v of the relative motion of two inertial refe­
rence frames cannot exceed the speed of light in a vacuum.
2°. Assume that the bar MN moves together with the
reference frame K f relative to the frame K as shown in
Kig. V.4.3. The length of the bar in the frame K ' is (V.4.3.20)
Iq= X2 — X\
452 Optics V.4.7

The length of a body in a reference frame where it is at


rest (Z0) is called its proper length. To find the length I
of the moving bar in the frame K, we must determine the
coordinates x2 and xl of the points N and M of the end and
beginning of the bar at the same moment t according to a
clock in the frame K :

I = x2 (t) — xx (t)

It follows from the Lorentz transformations that

x2 (t) — x { (t) = (x '2 — x\) ]A — v2/c2> or l = l0 Y l — v2/c2

The length of a body depends on its velocity. The proper


length of a body is its maximum length. The linear dimensi­
on of a body moving relative to an inertial reference frame
diminishes in the direction of motion Y 1 — ^2/c2 times
(the Lorentz length contraction). It also follows from the
Lorentz transformations that

y2 — y i = y 2 — y i and 4 —4 = z2 — z t

i.e. the transverse dimensions of a body do not depend on


its velocity and are the same in all inertial reference frames.
3°. The Lorentz length contraction is not a seeming one.
It stresses that the length of a body is relative and depends
on the velocity of the body in the given reference frame K.
This dependence, however, can tell only at velocities when
v2/c2 appreciably affects the value of ]/ 1 — v2/c2, i.e. at
velocities v close to the speed of light. Such velocities are
not achievable in practice for macroscopic bodies, and the
Lorentz contraction is not detected experimentally.
4°. The Lorentz length contraction is a kinematic effect
of the special theory of relativity and is not associated with
the action of any forces “compressing” the bar along its length.
It follows from the Lorentz contraction that no body can
travel in space with a velocity v ^ c. If it could, this would
mean that the length of the body is an imaginary quantity
or vanishes.
V.4.8 The Special Theory of Relativity 453

8. Relativity of Time Intervals


1°. The time interval between two events is defined as
the time that has elapsed between these events and measured
by a clock in the given reference frame. Assume that two
events occurred at the moments t\ and t'2 at the point A
stationary relative to the frame K '. For example, an oscil­
lating pendulum passed twice through its position of equi­
librium. The time interval t 0 between these events will be
t 0 = t 2 t i.
2°. The time measured in the reference frame where
the point A is stationary is called the proper time. The pro­
per time is measured by a clock moving together with the
reference frame. In the frame K relative to which the frame
K' moves (Fig. V.4.3), the time interval x between events
occurring at A will be x = t2 — where the time is meas­
ured by a clock in the frame K . From the Lorentz transfor­
mation for time (V.4.6.20)
V
t2—-^rx2 t x1
t ’2 = —7= and t\
Y i — v 2I c 2 y T — v2/c2
it follows that
v / \
(t2 fj) (x2 X\) T— ^5-(*2 — *l)
To — t 2 — 11—
Y 1 — v 2/ c 2 Y 1— v 21 c 2

But x2 — xx is the displacement of the point A along the


axis OX of the frame K during the interval x between events.
Consequently,
x 2 — x { = vt and t 0 = t [ A — v2/c2
i.e.
t0< t , or x= T0
Y 1 — V2 / c 2

3°. The duration of a phenomenon occurring at a cer­


tain point of space will be minimum in the inertial reference
frame relative to which the point is stationary. This means
454 Optics V .4 .8

that a clock moving relative to an inertial reference frame


runs slower than a stationary clock and shows a smaller
time interval between events (relativistic dilation of time).
The kinematic effect of the slowing down of a clock becomes
noticeable only at velocities v close to the speed of light c
in a vacuum, when the term v2/c2 in the formulas of Par.
2° has an appreciable influence.
4°. The relativistic dilation of time has been proved in
experiments with the decay of muons (VI.5.4.3°)—unstable
spontaneously decaying elementary particles. The mean
proper lifetime t 0 of a muon, i.e. its lifetime measured by
a clock travelling together with it, is t 0 = 2.2 X 10~6 s.
If there were no time dilation, then a muon created in the
upper layers of the atmosphere and moving toward the
Earth with a velocity v close to c would travel a small di­
stance s & cx0 = 660 m in the atmosphere. The muon would
never reach the Earth’s surface. Instruments, however, re­
gister such particles. The formulas of Par. 2° explain this
paradox. The lifetime x of a muon for an observer on the
Earth (in the fixed reference frame K) will be considerably
greater than t 0, namely t = t 0/]/ 1 — v2/c2. Estimates show
that x » 10t0, and a muon travels a distance in the atmo­
sphere that permits it to reach the Earth and be registered
by an instrument.
5°. The relativistic dilation of time makes possible, in
principle, a “trip to the future”. Suppose that a spaceship
travelling with the velocity v relative to the Earth completes
a flight from the Earth to a star and back. If light travels the
distance from the star to the Earth during the time tQy
then Z0 = ct0. The duration of the spaceship’s trip for an
observer on the Earth will be
T __ 2t0

where P = vie. This time will be added to people’s ages on


the Earth by the moment when the astronauts return. A
clock in the spaceship will show that the flight took less
time: t 0 ~ 2 toy I — P2/P. According to the principle of
relativity (V.4.2.10), all processes on the spaceship includ­
ing ageing of the astronauts occur in the same way as on the
V .4 .8 The Special Theory of Relativity 455

Earth, but according to the spaceship’s clock instead of one


on the Earth. Hence, by the moment they return to the Earth,
the astronauts will become older* only by the time t 0. If,
for example, t0 = 500 years and P = 0.9999, then the for­
mulas for t and t 0 give x = 1000.1 years and x0 = 14.1
years. The astronauts will return to the Earth according to
a clock on the Earth in 10 centuries after starting off, and
will become only 14.1 years older.
6°. This result is the basis of the clock paradox in the
special theory of relativity. The Earth can be regarded as
moving relative to the spaceship with the velocity v. Hence,
a clock on the Earth ought to lag behind the spaceship’s
clock, and the duration of the flight ought to be greater for
the astronauts than for the people remaining on the Earth.
It follows that after landing of the spaceship its clock should
simultaneously lag behind the clock at the cosmodrome
and be ahead of it, which is absolutely meaningless. The re­
sult of comparing the passage of time according to two cor­
rectly running mutually stationary clocks at the same point
of space should be unambiguous.
7°. Actually, there is no clock paradox. It appears owing
l.o the wrong application of the principle of relativity in
Par. 6°. The principle of relativity (V.4.2.10) postulates the
physical equality of only inertial reference frames, and not
of any ones. The clock at the cosmodrome always remains at
rest relative to the same inertial reference frame. The space­
ship’s clock, on the other hand, is stationary relative to
the ship, which is not always an inertial reference frame.
When starting, flying around the star, and landing, the ship’s
speed changes, and an accelerating reference frame is a non-
inertial one (1.2.12.1°). The Earth’s and the spaceship’s re­
ference frames are not equivalent within the special theory
of relativity. The calculation of the durations of the flight
t and t 0 given in Par. 5° with the use of an inertial (the
Earth’s) reference frame must not be compared with the
reasoning of Par. 6°. Here it is necessary to use the general
theory of relativity (see the footnote on p. 444), where it
is proved that from the astronaut’s viewpoint x is also
greater than t 0.
456 Optics V .4 .9

9. The Relativistic Law


of Velocity Addition
1°. The law of velocity addition in Newtonian mecha­
nics (1.2.7.2°) contradicts the postulates of the special theory
of relativity (V.4.2.10) and is replaced in the latter with
a new relativistic law of ve­
locity addition. The law of ve­
2 ,\ zkI locity addition following from
K
\Kr
the Lorentz transformations
(V.4.6.20) is called relativistic.
0 o'L This law satisfies the postu-
lates of the special theory of
relativity and the limiting
Y'J nature of the speed of light in
F ig . V . 4 . 5
a vacuum (V.4.4.40).
2°. If a point particle or
body M moves along the axes
OX and O 'X' in the inertial frames K and K ' and has vel­
ocities of v and v \ respectively, in these frames, then
v '+ V
i + v'V/c*

where V is the velocity of the frame K f relative to the axis


OX of the frame K (Fig. V.4.5) (the law of velocity addition
in the special theory of relativity).
3°. At VIc < 1 and v'lc < 1, and at the arbitrary velocity
v, the relativistic law of velocity addition transforms into
the law of velocity addition of Newtonian mechanics:

v = v' + V

4°. It follows from the relativistic law of velocity ad­


dition that the sum of two velocities less than or equal to
c is a velocity not exceeding c. Particularly, if vr = c,
then v = c at any velocity V. When v' = V = c, it follows
that v also equals c (see also V.4.7.40).
V .4 .1 0 The Special Theory of R ela tivity 457

10. Relativistic Dynamics.


Dependence#of Mass
on Velocity
1°. Dynamics based on the postulates of the special theo­
ry of relativity (V.4.2.10) and invariant relative to the
Lorentz transformations (V.4.6.20) is called relativistic
dynamics*.
2°. The fundamental law of dynamics—Newton’s second
law (1.2.4.2°)—has the following form for a point particle
or a body in relativistic dynamics:
F = AP
At
The vector of the resultant force F applied to a particle
(a body) equals the change in the momentum vector p of
the particle (or the body) in a unit time.
3°. The vector p of the mechanical momentum in the
special theory of relativity (the relativistic momentum),
as in Newtonian mechanics, is proportional to the velocity
vector v:
WqV m\
P
Y 1— V2 / c 2
where

is called the relativistic mass of a particle (a body).


4°. A glance at the expression for the relativistic mass
shows that the mass of a body depends on its velocity.
The mass mQ of a body that is stationary in the given ref­
erence frame is called its rest mass (proper mass). Figure
V.4.6 shows a plot of m/m0 against vtc. It can be seen from
the plot and the formulas of Par. 3° that if m0 = 0, then
v c when m — oo and p - > o o . Hence it follows that no

* A sim ila r d e fin itio n can be ap p lied to r e l a t i v i s t i c k i n e m a t i c s ,


in form ation on w h ich is g iv e n in the p reced in g se c tio n s of the present
chapter.
458 Optics V .4 .1 1

particle (or body) with a rest mass differing from zero can
travel with a velocity equal to the speed of light in a vacuum.
Particles with a rest mass not equal to zero (mQ=7^ 0)
and moving with such high velocities v that the term v2lc2
in the [formulas of Par. 3° can­
not be ignored are called relativ­
istic particles. A velocity v great­
er than c leads for conventional
particles to an imaginary mass
and an imaginary momentum,
which is physically meaningless.
The dependence of the mass on
the velocity begins to tell only
at velocities very close to c
(Fig. V.4.6). The formulas of
Par. 3° cannot be applied to a
photon (V.5.1.20), which has no
rest mass (m0 = 0 ). A photon
always travels with a velocity
equal to the speed of light in a vacuum, and is an ultra-
relativistic particle. This does not mean, however, that the
speed of light is constant in all substances (V.2.1.50).
5°. When v!c < 1, the expression for the momentum be­
comes the same as that used in Newtonian mechanics
(1.2.3.5°):
p = mv

where m signifies the rest mass (m = m0) because when


v!c < 1 the difference between m and m0 is of no importance.

11. The Mass-Energy Relation


1°. The total energy E of a body (or a particle) is pro­
portional to the relativistic mass m (V.4.10.3°) (the mass-
energy relation):
E = me2

where c is the speed of light in a vacuum. The relativistic


mass depends on the velocity v with which a body (a par-
V . 4 .1 1 The Special Theory of R ela tivity 459

tide) travels in the given reference frame. Consequently, the


total energy is different in different reference frames.
2°. A particle (a body) ha^ the minimum energy E 0
in the reference frame relative to which it is at rest (v = 0).
The energy E0 is called the rest energy or proper energy of
a particle (a body):
E0= m0c2
The rest energy of a body is its internal energy (11.4.1.2°).
It consists of the sum of the rest energies of all the particles
of the body 2 modc2i the kinetic energy of all the particles
i
relative to the common centre of mass {1.2.3.4°), and the
potential energy of their interaction. Hence,

m0c2 ^ 2 moac2 and mo¥= 2 m°’i


i i
where m0%i is the rest mass of the i-th particle.
The law of conservation of the rest mass does not hold
in relativistic mechanics. For example, the rest mass m0
of an atomic nucleus is smaller than the sum of the proper
masses of the particles in the nucleus (see also VI.4.2.3°).
Conversely, the rest mass m0 of a particle capable of sponta­
neous decay is greater than the sum of/the proper masses of
the decay products m0fl and m0)2:

m0A> m 0A + m0t2
[see the decay of a neutron (VI.4.7.7°)I.
3°. The non-conservation of the rest mass does not sig­
nify violation of the law of conservation of mass in general.
The law of conservation of the relativistic mass (V.4.10.30)
holds in the theory of relativity. It follows from the formula
of the mass-energy relation E = me2 (Par. 1°). The total ener­
gy (1.5.4.1°) is retained in an isolated system of bodies
(1.2.2.5°). Consequently, the relativistic mass is also re­
tained. In the theory of relativity, the laws of conservation of
energy and relativistic mass are interrelated and form a
single law of conservation of mass and energy. The possibil­
ity of the conversion of mass into energy and vice versa,
however, does not at all follow from this law. Mass and energy
460 Optics V .4.11

are two qualitatively different properties of matter that by


no means are “equivalent” to each other. No experimental
evidence gives grounds for the conclusion on the “conver­
sion of mass into energy”. The conversion of the energy of
a system from one form to another is attended by transfor­
mation of the mass. For example, in the birth and destruc­
tion of an electron-positron pair (VI.5.5.5°), their mass does
not transform into energy, in complete accordance with the
law of conservation of relativistic mass and energy. The
rest mass of the particles (electron and positron) is trans­
formed into the mass of photons (V.5.1.20), i.e. into the mass
of an electromagnetic field.
4°. The kinetic energy Ek of a free particle (a body)
(1.5.3.3°) is the difference between the total energy of the
body E and the rest energy E0:
Ek = E — E0 — (m — m0) c2 = me2 ^1 — — j =

= me2 (1 — V l — v2/c2) = (l — V i - v 2/c2)


V V ' |/T _ i; 2 /c2 v r 1

or, in a different form:


p ________m v2_______ p2________
l + j / " l — v2/c2 m ( 1 + Y 1 — v2/c2)
where p = mv is the relativistic momentum (V.4.10.3°).
When v2lc2 1, we get the formula used for calculating the
kinetic energy in Newtonian mechanics (1.5.3.3°)
r mv2 p2

At velocities much smaller than the speed of light in


a vacuum, the kinetic energy Ek of a body is considerably
smaller than the rest energy E0:

E0 2c2

For example, at the velocity v = 3 X 106 m/s, which is ten


times greater than the Earth’s velocity on its orbit, we have
Ek/E 0 = 5 X lO"7 - 5 X 10-5%.
The kinetic energy of relativistic particles (V.4.10.4°)
considerably exceeds their rest energy. For example, in mod-
V .4 .1 1 The Special Theory of Relativity 461

ern synchrophasotrons (VI.4.16.7°), protons are accelerated


up to velocities differing from c by 0.05%, i.e. v = 0.9995c.
In these conditions, •

4 £ = (m~'^o) _ _ 1 ——---------1 ^ 30

For relativistic particles, the rest energy may be ignored


in comparison with their kinetic energy Ek (E 0 <€ Ek):

E — E0-\- Ek « E k, i.e. Ek t t E = mc2

5°. There is a relativistic energy-momentum relationship


between the total energy £ of a particle (a body), its rest
energy E0, an its momentum p:

E2 = E q-\- pk2

The following relationship holds for relativistic particles


for which E « Ekl i.e. E 0 Ek:

Ek ~ pc

Problem. At what velocity does the kinetic energy of


a particle equal its rest energy?
Given: Ek = E0.
Required: v.
Solution: In relativistic mechanics, the kinetic energy of
a particle Ek = E — E 0, where E is the total energy, and
E0 is the rest energy.
Since E 0 = E — E 0J then E = 2E0, or me2 = 2m0c2,
where m = m j']f1 — v2/c2. Hence, m0c2/Y ^ — v2/c2 =
= 2m0c2, or Y \. — v2/c2 = 1/2, whence y=c'|/r3/4, v = 0.866c,
where c = 3 X 108 m/s is the speed of light in a vacuum,

v = 2.6 X 108 m/s


462 O p tic s t'.5.1

Chapter 5
Quantum Optics

1. Basic Concepts
1°. Quantum optics is the branch of the science of light
studying the discrete nature of the emission, propagation
of light, and its interaction with a substance.
2°. Light in quantum optics is considered as a stream
of special particles—photons, having no rest mass (m0 = 0)
(V.4.10.3°) and travelling with the velocity c equal to
the speed of light in a vacuum. The fundamental charac­
teristics of a photon are its energy Ek and momentum p:
h v _ h
and p
c ~~ X o

where v = frequency of an electromagnetic light wave


(IV.3.4.30, V .l.1.10)
X0 = wavelength in a vacuum (V.2.2.20)
h — Planck’s constant (V.3.2.30).
The momentum vector p of a photon has a direction coin­
ciding with that of the wave vector k:

The vector k has a magnitude equal to the wave number k


(IV.3.5.2°), k = 2n/k, and a direction coinciding with that
of the wave velocity.
A photon has The mass m = E^/c2 = hvlc2, which is the
mass of an electromagnetic field and is not associated with
a rest mass because photons at rest do not exist. In any iner­
tial reference frame, the speed of light in a vacuum equals
c(V.4.2.1°). In monochromatic light of frequency v, all
photons have the same energy, momentum, and mass.
3°. In any substance of absolute refractive index n
(V .l.2.10), photons always travel with the speed of light in
a vacuum although the velocity v of a light wave in a sub­
stance is n times smaller: v = d n . One should never confuse
V .5 .1 Quantum Optics 463

the velocity v of the propagation of the front of an electro­


magnetic wave in a substance (IV.3.3.1) with the velocity
of photons in the substance. Platons in a substance trav­
el from one of its particles (an atom or molecule) to another
in a vacuum, as it were, and when they “strike” a particle,
are absorbed in it and again appear.*
4°. Photons appear (are emitted) when atoms, mole­
cules, ions, and atomic nuclei pass over from excited energy
states (VI.2.4.3°) to states with a smaller energy. Photons
are also emitted upon the acceleration and retardation of
charged particles (1V.4.4.30, V.3.6.20), in the decay of
some particles (VI.5.4.4°), and in the destruction of an elec­
tron-positron pair (VI.5.5.5°). The process of absorption
of light by a substance consists in that the photons transmit
all their energy to the particles of the substance. The process
of absorption of light in quantum optics is regarded as being
intermittent both in space and in time.
5°. In the formulas of Par. 2°, the properties of a photon—
its energy, momentum, and mass—are expressed through
the characteristics of an electromagnetic wave: the fre­
quency or wavelength in a vacuum. Here the wave-particle
duality of the properties of light manifests itself. On the
one hand, light has wave properties that are detected in the
phenomena of interference, diffraction, and polarization;
on the other hand, light is a stream of photons. At low
frequencies v, the predominating part is played by the wave
properties of light, at high frequencies, by the quantum prop­
erties of light.
6°. There is a relation between the corpuscular (parti­
cle) and wave properties of light that is detected when light
propagates in a non-uniform medium. When, for example,
light passes through a slit and diffraction maxima and minima
(V.2.4.30) are observed on a screen, photons interact
with the substance (slit), and the result is the redistribution
of the photons in space. At the spots on the screen where
a greater illumination is observed, the summary energy of
the photons striking these points should be greater. But the
illumination at a given point of the screen is proportional to
* The analysis of the complicated phenomena occurring upon the
motion of photons in a substance is far beyond the scope of elementary
physics.
464 O ptics V.5.2

the intensity J of the light at this point: E <x J* (V. 1.6.5°).


The intensity of an electromagnetic light wave, in turn,
is proportional to the square of its amplitude: J ex A 2
(IV.4.3.2°). Thus, the square of the amplitude of a light
wave at a point of space is a measure of the number of pho­
tons striking this point.
7°. The quantum and wave properties of light mutually
supplement each other and reflect the mutually related laws
of propagation of light and its interaction with a substance.
The quantum properties of light are due to the fact that the
energy, momentum, and mass of electromagnetic radiation
are concentrated in special particles—photons. The wave
properties of light describe the laws of distribution of pho­
tons in space. The number of photons at a definite point of
space is determined by the wave properties of light.
8°. The energy of a photon of visible light hv consider­
ably exceeds that of the molecules of a substance. Let us
compare, for example, this energy at a frequency of v «
101B Hz with the mean kinetic energy of thermal motion
belonging to a molecule of an ideal gas at the temperature
T (11.2.4.4°). It follows from the equation hv = 3kT/2
that T = 2hv/3k « 105 K. Only at this temperature, which
is practically unattainable for a gas, would its molecule have
a kinetic energy comparable with the energy of a photon of
visible light.

2. The Photoelectric Effect


1°. The photoelectric effect (photoeffect) is the phenomenon
of the interaction of light with a substance as a result of
which the energy of the photons is transmitted to the ele­
ctrons of the substance. The photoemissiue and photoconduc-
tive effects (V.5.4.30) are distinguished for solids and liquids.
In the photoemissive effect, the absorption of photons is
attended by the emission of electrons out of the body. In
the photoconductive effect, the electrons emitted by the
atoms, molecules, or ions remain inside the substance, but

* Do not confuse the intensity of light with the intensity I of a


light source (V.l.6.30).
V.5.2 Quantum Optics 465

the energies of the electrons change. In gases, the photoelec­


tric effect consists in photoionization—the emission of elec­
trons from atoms and molecules ef a gas under the action
of light (111*3*3.2°).
2°. The photoemissive effect is detected by experiments
involving the ejection of electrons from the surface of metals
onto which short-wave light falls.
Figure V.5.1 is a diagram of Stole- A Ca
ton's experiment. The anode A —a
thin metal screen—is illuminated
with light from an electric arc. The
beam of light falls on the cathode
Ca—a solid zinc plate. The galva­
nometer G connected to the circuit
shows a current. Electrons are emit­
ted from the illuminated nega­
tively charged zinc plate, and the
electric circuit is closed. If the
screen A is used as the cathode and
the zinc plate as the anode, there
will be no current.
3°. The electrons escaping from the surface of a body in
the photoemissive effect are called photoelectrons. The latter,
accelerated by the electric field between the cathode and the
anode (see Fig. V.5.1) pro­
duce a photoelectric current
(photocurrent) I. Figure
V.5.2 shows how the pho­
tocurrent I depends on the
R e ta rd in g Accelerating U voltage U between the
v o lta g e vo lta g e cathode and the anode that
can be changed with the
Fig. V.5.2 aid of a potentiometer
(111.2.6.7°, Problem 3),
both in magnitude and in sign (the potentiometer is not
shown in Fig. V.5.1). The curves in Fig. V.5.2 cor­
respond to two different illuminations E of the cath­
ode (V. 1.6.5°). It follows from the curves that at certain
values of U = Us the photocurrent at the given illumi­
nation of the cathode reaches its maximum value I =
= Is called the saturation photocurrent. When U = C/s, all
466 Optics i7.5.3

ilie electrons ejected from the cathode upon its illumina­


tion reach the anode: / s ---cw, where n is the total number of
photoelectrons flying out from the cathode in a unit time,
and e is the magnitude of the charge of an electron [cf.
the saturation current in a diode (111.3.8.2°)].
4°. The existence of a photocurrent at negative volt­
ages from 0 to — £/r (see Fig. V.5.2) is explained by the fact
that the photoelectrons ejected by the light from the cath­
ode have an initial kinetic energy whose maximum value
is rni;max/2 (1.5.3.3°). This energy allows the electrons to do
work against the forces of the retarding electric held bet­
ween the cathode and the anode and reach the anode. Ac­
cording to the law of conservation of energy

where UT is the absolute value of the retarding voltage


U — UT at which the photocurrent stops, and e and m
are the magnitude of the charge of an electron and its mass,
respectively. When U f/r, there is no photocurrent ( / =
= 0 ).

3. Laws of the Photoemissive Effect.


Einstein’s Photoelectric Equation
1°. Laws of the photoemissive effect:
I. The maximum initial velocity vmax of photoelectrons
depends on the frequency of the light and the properties of
the surface of the metal. It does not depend on the illumi­
nation of the cathode.
II. The total number of photoelectrons n ejected by light
from a cathode in a unit time and the photocurrent of sat­
uration / s are directly proportional to the illumination of
the cathode.
III. For every substance there is a threshold frequency
(the photoelectric threshold)—such a minimum frequency
v min (or maximum, threshold wavelength ^max) at which
the photoemissive effect is still possible. There can be
no photoelectric effect at vmln (or X ^ ^max).
V.5.3 Quantum Optics 467

The virtually non-inertial nature of the photoelectric


effect has been established: it appears immediately when
the surface of a body is illuminatecfif the frequency v ^ vmin.
2°. The photoemissive effect in metals is explained in
quantum optics (V.5.1.10). To escape from a metal, an
electron must have a certain amount of energy called the
work function A (111.3.7.3°). As a result of the absorption
of a photon by the metal, the energy of the photon hv may
be entirely transmitted to an electron.* If h v ^ A , then
it is energetically possible for an electron to leave the
surface of a metal. The maximum kinetic energy mu2maJ 2
of an ejected photoelectron according to the law of con­
servation of energy is found from Einstein's photoelectric
equation:
2
m v max mi;2
2
h v - A , or hv = —~ + A

3°. Einstein’s equation explains all the laws of the


photoemissive effect (Par. 1°). Thus, ymax depends only
on the frequency of light v and the workfunction.
The photoemissive effect is possible only provided
that hvmln ^ A . The threshold frequency vmin = A/h,
or Jimax = ch/A depends only on the workfunction, i.e.
on the chemical nature of a metal and the state of its sur­
face. The total number of photoelectrons n emitted from
a cathode in a unit time is directly proportional to the
number of photons n' falling during this time on the sur­
face of the cathode. For a cathode uniformly illuminated
by monochromatic light, in turn, n' is directly propor­
tional to the illumination E.
4°. On the basis of V.5.2.4° and the photoelectric thresh­
old, Einstein’s equation acquires the form
mv2
m ax
2
eUT= h(v — vmin)

The determination of Planck’s constant from the last


equation was an experimental verification of the correctness

* T h e a b s o r p t io n o f a p h o t o n b y a p o s it iv e io n in th e c r y s t a l la t -
lic e o f a m e t a l ( 1 1 .1 .6 .5 ° ) r e s u lt s in o th e r p h e n o m e n a n o t c o n s id e r e d
h ere.
468 Optics ir.5.d

of Einstein’s equation and the quantum explanation of the


photoelectric effect. The linear relationship between the
quantity eUr and the frequen-
e^r, eV cy v of light follows from
'Zn this equation. Figure V.5.3
y . contains experimental straight
7 lines showing how eUTdepends
\ t on the frequency v for the
photoelectric effect on selec­
-^4 & % ted metals. The angle between
1.0 1.4 1.8
v*10~15, Hz the straight lines and the axis
of abscissas permits us to
Fig. V.5.3 find the value of h, namely,
h = k tan a , where k is the
ratio of the dimensions taken as scale units along the axes
eUT and v (1.1.3.5°).
The value of h obtained from photoelectric effect ex­
periments coincides with the one obtained for this quantity
from the laws of thermal radiation (V.3.2.3°): h = 6.625 X
X 10~34 J.s.

4. Applications
of the Photoelectric Effect
1°. Photocells are devices based on the utilization of
the photoelectric effect. A photocell is shown schemati­
cally in Fig. V.5.4. A light-sensitive layer is applied to
the greater part of the internal surface of a glass bulb that
is evacuated or filled with an inert gas and serves as the
cathode Ca. Light enters the bulb through the window W.
A wire ring or disk is the anode A. The latter is connected
to the positive pole of the battery 5 , and the light-sen­
sitive layer through the galvanometer G to its negative
pole. A current appears in the circuit when the cathode
is illuminated by a source of light whose radiation spectrum
contains frequencies above the threshold one (V.5.3.3°).
2°. A photorelay is a device for the automatic control
of various installations employing the ability of a photo­
cell to virtually instantaneously react to the action of
light or its change. A photorelay operates either when
V.5.4 Quantum Optics 469

light falls on a photocell or when illumination of the latter


is stopped.
3°. The photoconductive effect in semiconductors results
in the appearance in them of electrons or holes (111.3.11.5°)
and is called photoconduction. The
design of devices called photoresis­
tors and of photocells with a pho­
toconductive effect is based on
photoconduction.
4°. The simplest photoresistor is
a plate of an insulator onto which
a thin layer of a semiconductor is
applied. Photoconduction appears
when the plate is illuminated,
and a current flows through the
circuit. Photoresistors find appli­
cation in sound cinematography,
for signalling, in television, auto­
matic and remote control.
Photoresistors make it possible
lo automatically detect violations
of the normal course of various production processes
at a distance and stop the processes in such cases. Upon
violations of the normal course of a process, the lu­
minous flux (V .l.6.20) striking a photocell may change.
This changes the photo­
Light current and, consequent­
Transparent ly, the course of the entire
Layer process.
Metal oxide Photoresistors are em­
(semiconductor)
ployed for sorting large
\ Metal
batches of articles by dimen­
sions and colour. A beam
of light falls on a photo­
Fig. V.5.5
cell after being reflected
from the articles being
sorted, which are continuously fed onto a conveyor. The
colour of an article or its dimensions determine the lu-
mi nous flux falling on the photocell and the photocurrent.
Tim articles are sorted automatically depending on the
photocurrent-
470 O ptics V.5M

5°. Figure V.5.5 shows the design of a photocell with


a barrier layer. Two plates made of a metal and its oxide
(a semiconductor) are coated on top with a thin transparent
layer of metal. The boundary layer between the metal
and its oxide has unipolar conduction—the electrons can
pass only in the direction from the metal oxide to the metal.
The flow of electrons in this direction is produced under
the action of light without any external voltage. A bar­
rier-layer photocell directly transforms the energy of a
light wave into the energy of an electric current, i.e. is
a source of current. The functioning of the solar cells in­
stalled on spaceships is based on this principle.

5. Light Pressure
1°. By light pressure is meant the pressure exerted
by electromagnetic light waves falling on the surface of
a body. The existence of light pressure is predicted in
the electromagnetic theory of light. If, for example, an
electromagnetic wave falls on the metal M, then under
the action of the electric field of
the wave of intensity E, the elec­
trons of the metal will travel in
a direction opposite to the vector
E with the velocity v (Fig. V.5.6).
The magnetic field of the wave of
induction B acts on the travelling
electrons with the Lorentz force
F l (111.4.5.1°) in a direction at
right angles to the surface of the
metal. Thus, a light wave exerts
F ig . V . 5 . 6 a pressure on the surface of a
metal.
2°. In quantum optics, light pressure is the result of
the fact that a photon has the momentum p (V.5.1.20).
When a photon collides with the surface of a body, this
momentum is transferred to the atoms or molecules of the
substance. Similarly, the pressure of a gas is the result
of the transmission of momentum by molecules of the gas
to particles on the surface of the walls of a vessel.
V.5.5 Quantum Optics 471

Light pressure is determined by Maxwell’s formula


P = (1 +*r) we

where r = coefficient of reflection of light by the surface


of a body
we = volume density of the energy of the electro­
magnetic field of a wave (111.5.7.5°).
The light pressure on a perfectly reflecting (mirror)
surface (r = 1), i.e. p = 2we, is twice the light pressure
p = we for a blackbody for which r = 0. The different
light pressures in these two cases are ex­
plained by the fact that the momentum of
a photon hv/c (Par. 2°) is transmitted to
the atoms or molecules of the body if its
surface is an absorbing one. Upon reflec­
tion from a mirror surface, the momentum
of a photon hv/c changes its sign and be­
comes —hv/c, so that the momentum trans­
mitted to the particles of the substance
is 2hv/c.
3°. The existence of light pressure and
(he correctness of Maxwell’s formula were
experimentally proved by experiments con­
ducted by P. Lebedev. His instrument
was a lightweight frame with thin “wings”
fastened to it. The wings were light and
dark disks from 0.01 to 0.1 mm thick.
The disks were arranged symmetrically relative to the axis
of the suspension about which the frame could turn
(big. V.5.7). The light falling on the wings exerted dif­
ferent pressures on the light and dark disks. As a result,
a torque (1.3.1.4°) was exerted on the frame suspended
on a thin glass thread and twisted it. The light pressure
was determined according to the angle through which the
thread was twisted. Lebedev measured the volume density
of the energy of an electromagnetic field with the aid of
a specially designed magnetic calorimeter. A beam of light
was directed onto it for a definite time, and the elevation
of the temperature was registered. The value of we* was
determined according to the quantity of absorbed heat.
472 Optics V.5.6

After cleverly overcoming a number of difficulties ap­


pearing in the measurement of light pressure, Lebedev
proved the correctness of Maxwell’s formula (Par. 2) ensu­
ing from the electromagnetic
theory of light.
Figure V.5.8 shows a sim­
plified diagram of Lebedev’s set­
up for measuring the light pres­
sure on gases. The light passing
through the glass wall A acts
A
on the gas contained in the cy­
Fig. V.5.8
lindrical channel B . Under the
action of the light pressure, the
gas from the channel B flows over into the channel C
communicating with it. The channel C contains a light­
weight movable piston D suspended on a thin elastic
thread E at right angles to the plane of the drawing.
The light pressure was measured according to the angle
of twisting of the thread.

6. Chemical Action of Light.


The Photographic Process
1 °. The action of light is defined as chemical if chemical
transformations—photochemical reactions—occur in substan­
ces that absorb light. For example, when light acts on
bromine vapour, a molecule of Br2 dissociates into two
bromine atoms; a molecule of silver bromide AgBr de­
composes under the action of light into atoms of bromine
and silver. (The process of the transformation of the ions
Ag+ and Br~ into atoms is not considered here.) The green
parts of plants absorb carbon dioxide and under the action
of sunlight decompose its molecules into their constituent
parts—carbon and oxygen. The energy of sunlight results
in the creation of very complicated molecules of proteins,
fats, and carbohydrates on the basis of carbon and atoms
of other chemical elements.
2°. The photochemical reactions occurring in the photo­
sensitive layers of photographic plates and films are the
J)asis of photography. A photosensitive layer consists of
V .5 .6 Quantum Optics 473

minute crystals of silver bromide (or silver chloride) em­


bedded in gelatin. The action of light leads to a photochemi­
cal reaction of decomposition of* AgBr (Par. 1°). A neglig­
ible amount of metallic silver is liberated that is differ­
ent in different parts of the film, and a latent image of
the object is formed.
To process a film, it is immersed in a solution of de­
veloping substances. It causes a greater and greater number
of silver atoms to precipitate from the silver bromide crys­
tals on which light acted. They group around the silver
atoms that appeared before developing. As a result of
the precipitation of silver, the exposed parts of the film
become the blacker, the brighter was the light falling
on the corresponding part of the film. The non-exposed
parts of the film remain unchanged—light-coloured silver
bromide is concentrated at them. To prevent the negative
image obtained from changing under the action of light,
it is fixed by immersing the film in a solution of sodium
thiosulphate (hyposulphite) (fixer). Here the silver bro­
mide not reduced by the developer dissolves. After washing
and drying, the negative can be stored and used for various
purposes. For example, X-ray photographs of a person’s
organs are stored as negatives (V.3.6.50). To obtain a pos­
itive image from a negative, the latter is projected onto
photographic paper with a light-sensitive layer. After
developing, a reverse image of the negative is obtained,
i.e. a positive, in which the light spots of the photographed
object are light, and the dark ones are dark.
Objects from which very weak light reaches the camera
produce an image upon a large exposure, i.e. a prolonged
time of illuminating the film. The photochemical process
in special films or plates makes it possible to photograph
objects in infrared rays in the dark.
ATOMIC
PART AND NUCLEAR
PHYSICS

Chapter 1*

Elements of Quantum Mechanics

1. De B roglie’s Ideas
on the W ave Properties
of Particles
1°. Quantum mechanics is one of the main directions of
development of modern physics. It studies the laws of
phenomena occurring in the microcosm—within the limits
of distances of about 10-15 to 10“10 m. Atoms, molecules,
crystals, and also atomic nuclei and elementary particles
are what is studied by quantum mechanics.
2°. The physical fundamentals of quantum mechanics
are:
(a) M. Planck’s ideas on quanta of energy (V.3.2.30);
(b) A. Einstein’s ideas on photons (V.5.1.20); and
(c) L. de Broglie’s ideas on the wave properties of the
particles of a substance.
According to de Broglie, the particle-wave duality of
the properties of light (V.5.1.50) is characteristic not only
of light particles, photons, but also of the particles of a
substance having a rest mass (V.4.10.40), electrons, pro­
tons, neutrons, and their ensembles, atoms, molecules,
V I .1 .1 Elements of Quantum Mechanics 475

and atomic nuclei. De Broglie’s hypothesis signifies that


the particle-wave duality of properties characteristic of
an electromagnetic field has a •universal nature.
3°. The propagation of a de Broglie wave is associated
with any particle of mass m travelling with the velocity
v. The de Broglie wavelength X is calculated by de Brog­
lie's formula

X= - = -
mv p

where h = Planck’s constant (V.3.2.30)


p == mv = magnitude of the momentum of the moving
particle.
At a velocity v of a particle comparable with the speed
c of light in a vacuum, the momentum p is considered as
relativistic (V.4.10.3°). When v < c, the momentum p of
a particle is computed as is usual practice in Newtonian
mechanics (1.2.3.5°). Another form of de Broglie’s formula
is

P= hk

where k = 2jiiiA = wave vector (V.5.1.20)


n = unit vector in the direction of wave propagation
h = h/2n = 1.05 X 10-34 J -s.
De Broglie waves are not electromagnetic ones and have
no analogy among all the kinds of waves studied in clas­
sical physics (VI.3.1.3°). De Broglie’s formula is one of
the fundamental relationships of quantum mechanics. The
de Broglie wavelength for an electron after it has been
subjected to the accelerating potential difference A<p is
« h 12.25 4 a_j/|
X= — X 10 10 m
| / 2 m e Acp y Acp

Ilore m = mass of an electron


e = absolute value of the charge of an electron
Acp = potential difference in volts
h = Planck’s constant (V.3.2.30)
476 Atomic and Nuclear Physics VI.1.2

2. Wave Properties of Electrons,


Neutrons, Atoms, and Molecules
1°. De Broglie’s hypothesis was confirmed in numerous
experiments showing the appearance of diffraction (V.2.3.10)
when beams of particles (electrons, neutrons, atoms, and
molecules) interact with a substance.
2°. C. Davisson and L. Germer investigated the scattering
of electrons on a monocrystal of nickel (11.1.6.4°). The
experiments are shown schematically in Fig. VI. 1.1. A beam

of electrons was produced in the electron source A (an


electron gun). The energy and velocity of the electrons
were controlled by the accelerating electric field produced
in the gun. A narrow beam of electrons having a definite
velocity was directed onto the earthed nickel crystal B
and was reflected from it. The target B could rotate about
an axis perpendicular to the plane of the drawing. The
movable receiver of electrons C rotated about the same
axis and registered the electrons reflected by the nickel
crystal in different directions in the plane of the drawing.
The result was expected to be a mirror reflection of the
electrons with the angle of incidence equalling the angle
of reflection (V. 1.2.4°). The experiments showed that the
laws of "geometrical optics are violated in reflection. With
a given angle of incidence, the electrons are reflected from
the surface of the crystal at different angles. Maxima of
the number of reflected electrons are observed in some
directions, and minima in others. Figure VI. 1.2 contains
a diagram showing the distribution by directions of the
number of electrons scattered by the target B for a given
angle of incidence of the primary beam of electrons N.
V l± 2 Elements of Quantum Mechanics 47?

The length of the radius r conducted from the centre of


the target is proportional to the number of electrons re­
flected at a given angle. Using# the methods of observing
the diffraction of X-rays on monocrystals (V.3.6.60), Da­
visson and Germer determined the de Broglie wavelength
experimentally for the scattered electrons and confirmed
the correctness of de Broglie’s formula.
3°. Diffraction is detected when a beam of electrons
is passed through thin layers of metals (with a thickness

Fig. V I.1.3/

of about 10“7 m) having a polycrystalline structure (11.1.6.4°).


Experiments have shown that the diffraction of electrons
on the polycrystals is observed similar to the diffraction
of X-rays on polycrystalline powders (V.3.6.60). Figure
VI. 1.3 shows photographs of the diffraction patterns ob­
served when X-rays (Fig. VI. 1.3a) and a beam of electrons
(Fig. VI. 1.36) passed through thin films of the same poly­
crystal. The radii of the diffraction fringes were used to
determine the de Broglie wavelength and verify the cor­
rectness of de Broglie’s formula. The wave properties of
electrons are observed only provided that the length of
a de Broglie wave has the same order of magnitude as the
interatomic distances in the crystals on which diffraction
478 Atomic and Nuclear Physics VI.1.2

occurs. Hence, the structure of a substance can be studied


by electron diffraction. It is similar in essence to X-ray
structural analysis (V.3.G.00).
4°. Experiments revealed the diffraction of neutrons—
particles in the nuclei of atoms (VI.4.1.1°). The experiments
were set up as shown in Fig. VI. 1.4. The neutrons are
formed in a nuclear reactor (VI.4.12.6°) and are retarded
in a layer of graphite—a graphite thermal column. The
thermal neutrons (VI.4.9.2°) pass through a narrow slit,
and the beam of them (TV) falls on the crystal B , where

diffraction occurs. The receiver C registers the neutrons


reflected from the target B at different angles of diffrac­
tion (V.2.4.10). Neutron diffraction is used to study the
structure of solids.
5°. Beams of neutral atoms and molecules at room
temperatures (T « 300 K) travel with such velocities v
that the de Broglie wavelengths corresponding to these
particles have the order of magnitude of %« 10"10 m
(VI. 1.1.3°). This makes it possible to observe the wave
properties of atoms and molecules when beams of particles
are reflected from the surface of crystals. Apart from mir­
ror reflection (V. 1.2.5°), in some directions additional
diffraction maxima of the number of reflected atoms (mo­
lecules) are observed. Figure VI. 1.5 shows the first dif­
fraction maxima in addition to the main maximum of the
number of mirror reflected particles.
6°. De Broglie waves are associated with any moving
particle regardless of whether it is electrically charged
VI.1.3 Elements of Quantum Mechanics 479

or neutral. W aves of th is kind ap p reciably differ from all


the w aves known in cla ssica l p h y sics in th at th ey are not
em itted by any w ave sources. •

3. The Physical Meaning


of de Broglie Waves
1°. The wave properties of particles manifest themselves
when we can detect diffraction on the particles experimental­
ly. For this, it is essential that the condition of observation
of diffraction (V.2.3.10) be satisfied.
2°. The de Broglie wavelengths of macroscopic bodies at
their usual velocities are so small that it is impossible
to detect such a wave in an experiment. For example, when
a body of mass m = 1 g travels with a velocity v = 1 cm/s,
we have X = 6.62 X 10~27 cm. Such a wavelength cannot
be detected because there are no periodic structures with
a lattice constant (11.1.6.5°) of the order of magnitude
of 10“27 cm. Wave properties are detected only in moving
microscopic particles of a substance having a mass com­
parable, within the limits of several orders of magnitude,
with the mass of elementary particles (VI.5.2.1°).
3°. The physical meaning of de Broglie waves is estab­
lished from an analysis of the relationship existing between
the particle (corpuscular) and the wave properties of light
(V.5.1.60). A similar relationship exists between the cor­
puscular and the wave properties of the particles of a sub­
stance. The square of the amplitude of a de Broglie wave
at a given point of space determines the probability of
the fact that a certain number of particles will get to this
point. The diffraction maxima (VI.2.2.1°-5°) where the
amplitudes of the waves have the greatest values corres­
pond to the points of space reached by the greatest number
of particles. The diffraction minima correspond to the
points of space reached by the minimum number of par­
ticles.
4°. Wave properties are characteristic not only of a
beam of moving particles, but also of a separate moving
particle. V. Fabrikant, L. Biberman, and N. Sushkin
discovered the diffraction of single electrons flying one
480 Atomic and Nuclear Physics Vl.1.4

by one onto a crystal experimentally. In this experiment,


only one electron at a time fell onto a thin metal film of
a polycrystal. After multifold “bombardment’’ of the film
with single electrons, the same diffraction pattern was
observed as in the simultaneous passage of a beam of elec*
trons through a crystal (see Fig. VI. 1.36). This signifies
that for a single particle the square of the amplitude of
the de Broglie wave at a given point of space is a measure
of the probability of detecting a particle at this point.
De Broglie waves associated with moving particles have
no relation to the propagation of a field, for example an
electromagnetic or some other one.

4. Linear Harmonic Oscillator.


Motion of an Electron
in a Limited Region of Space
1°. A body (a particle) may move under the action of
forces so that it is retained in a definite region of space.
For example, in the harmonic oscillations of a body of
mass m suspended on a spring
Ev (x) (VI. 1.3.1°) under the action of
elastic forces, the body cannot
move away from its equilibrium
position by more than the dis­
tance equal to the displacement
amplitude A. The value of the
displacement amplitude is deter­
mined by the total energy E
(Fig. VI. 1.6). The potential ener­
gy of a body Ev (x) = kx*l2 =
m<£>\x2!2, where coj = klm is the
natural cyclic frequency of the
oscillations (IV. 1.3.3°), and k
is the spring constant.
A particle oscillating along the axis O X under the
action of quasi-elastic forces and having the potential
energy Ep (x) = mu>\x2l2 is called a linear harmonic os­
cillator.
VL1A Elements of Quantum Mechanics 481

2°. In Newtonian mechanics, a linear harmonic oscillat­


or can have any value of the potential energy Ep (x) not
exceeding the value Ev (A) = mcw*Aa/2 at the points B and
C (B ' and C') where the potential energy equals the total
energy. The velocity of harmonic oscillations (IV. 1.2.1°)
of an oscillator can also take on any values limited by
the store of total energy E of the oscillator. No restrictions
are imposed in classical mechanics on the nature of the
possible change^ of the total energy
E. In quantum mechanics, the total
energy of an oscillator changes dif­ \ ep
ferently (Par. 6°).
3°. The one-dimensional motion of
a particle along the axis O X (Fig.
VI. 1.7) can be restricted as follows.
The particle moves freely in the re­
gion O x L. It cannot pass out
of the limits of the region OL. At the
L X
boundaries of the region 0L, at the
points x = 0 and x = L, the poten­ Fig. VI.1.7
tial energy Ev of the particle becomes
equal to infinity. Such motion of a
particle is called motion in a one-dimensional square poten­
tial well. This motion can be illustrated by the following
model: a particle moves along the bottom of a flat box
with perfectly reflecting infinitely high walls (see also a
potential barrier, VI.4.7.2°).
4°. The motion of an electron inside a potential well is
attended by the propagation of a de Broglie wave (VI. 1.1.3°).
The wave is reflected on the walls of the well, and standing
de Broglie waves (IV.3.10.1°) are formed as a result of
the superposition of the incident and reflected waves.
The condition for the formation of standing waves along
the length L of a potential well is similar to that for the
formation of standing waves in a string fixed at both ends
(IV.3.10.7°). A whole number of half-waves should be
contained on the length L :

§ 1-0211
482 Atomic and Nuclear Physics V i.i.i

where n is an integer. Hence,

The lengths of the de Broglie waves of an electron mov­


ing in a potential well can take on only definite values
inversely proportional to a series of integral numbers n
0discrete* values of the wavelengths). The velocity vn of an
electron in a potential well according to de Broglie’s for­
mula (VI.1.1.30) is
_ h _ nh
Vfl mXn 2 mL
The velocity vn takes on discrete values directly proportion­
al to the integers n.
5°. The momentum p n = mvn of an electron in a po­
tential well has discrete values:

The energy En of an electron “imprisoned” inside a


square potential well of infinite depth is
jp _ mv% _ n 2 h?
n~ 2 “ 8mL*
The energy En can have only discrete values directly
proportional to squares of the integers n. Physical quan­
tities (for example, energy, momentum, etc.) that can
take on only discrete (quantized) values are called quan­
tized physical quantities (the quantization of physical quan­
tities).
6°. The potential well for a linear harmonic oscillator
is the region of the axis O X limited by the potential energy
curve Ev (x) = &r2/2, and the store of total energy E of
the oscillator. At the two values of the total energy E1
and E 2 (see Fig. VI. 1.6), the oscillator can oscillate with
the displacement amplitudes equal to A 1 and A 2l respec­
tively, i.e. it is shut in on the segments BC and B'C' of
straight lines parallel to the axis O X .
* From the Latin discretus meaning separated, distinguished be­
tween.
vi.i.4 Elements of Quantum Mechanics 483

The wave properties of a linear harmonic oscillator


lead to the fact that the possible quantized values of its
total energy (the energy levels ef the oscillator) have the
form
E n = ( » + y ) h v o = ( « + y ) ft0)o

where n = integer
h = Planck’s constant
v0 = natural frequency of oscillations of the oscil­
lator (IV. 1.3.3°)
G)0 = 2 jiv 0 = cyclic natural frequency
h = hl2n.
Figure VI. 1.8 shows the energy levels of a linear har­
monic oscillator that are directly proportional to a series
of half-integral numbers. The
energy levels are at identical
“energy distances” from one
another. The minimum value
Eo of the energy of a linear
harmonic oscillator (at n = 0)
E0 = h v j 2 = h g)0/2 is called
the zero-point energy. It
cannot be diminished by any
external actions. The energy
Eo never becomes equal to ze­
ro at any ultralow tempera­
tures, including the absolute zero of temperature (T =
= OK = — 273.15 °C) (11.4.9.4°). The existence of a zero-
point energy in a particle is a purely quantum effect. (For
the nature of the zero-point energy see V I.1.7.3°.)
7°. The quantization of physical quantities in definite
conditions is an important result of quantum mechanics
that is new in principle. In classical mechanics and in all
of classical physics, the physical quantities characterizing
any physical phenomena change, as a rule,* continuously.
Planck’s idea that the energy of an atom-emitter can take
on only definite values (V.3.2.3°) was consistently developed
in quantum mechanics.
* An exception here, for example, are the natural frequencies of
oscillations of fixed strings and plates (IV.3.10.7°).
4s4 Atomic and Nuclear Physics Vl.1.5

8°. The results given in Paragraphs 4° and 5° have


a general significance. The energies of the electrons in
atoms, molecules, and their ensembles have quantized
values called energy levels (electron energy levels). The num­
ber n determining the quantized value of the energy of
an electron is called the quantum number. For electrons
in atoms or molecules, the number n is called the prin­
cipal quantum number (VI.2.5.3°). This name is associated
with the fact that, apart from the quantum number n, there
are other quantum numbers on which the physical quan­
tities characterizing the electrons in atoms and molecules
depend (VI.2.8.40).
A stationary quantum state of an electron (or of any
other particle) is defined as its state with a definite quantized
value of the energy E n.* A stationary quantum state of
a particle does not change with time in the absence of
external action on it.

5. Uncertainty Relation
1°. An excited atom (VI.2.5.3°) has surplus energy in
comparison with an atom in the normal energy state. The
transition of an atom from an excited state to the normal

one lasts a finite time of t « 10”8 s and is attended by


the emission of light (VI.2.4.3°). From the viewpoint of
wave optics, this signifies that the emission of an electro­
magnetic wave by an atom continues for the finite time t .
During this time, the atom emits a “broken sinusoid” called

* For a complete characteristic of the quantum state of an elec­


tron in an atom and in other ensembles of particles, see VI.2.8.40.
V I.L 5 Elements of Quantum Mechanics 485

a wave train. The length of a wave train in a vacuum


is Ax = cx = 3 m (Fig. VI. 1.9V, The length of a light
wave is X « 10”6 m. Several million wavelengths can be
accommodated on the length of a wave train emitted by an
atom during one act of radiation.
2°. A wave train cannot be strictly monochromatic,
i.e. cannot have a definite frequency v or wavelength X in a
vacuum (IV.4.2.5°). A wave train inevitably contains a
certain range of frequencies Av.* The range of cyclic fre­
quencies Aco = 2jt Av contained in a wave train is in­
versely proportional to the duration t of the train:
Aco « —,
x 1
or A c o t« 1

The duration of the emission of light by an excited


atom is generally t « 10“8 s, therefore Aco ^ 1/t « 108 s*"1.
This quantity is called the natural width of a spectral line.
This name underlines the fact that the spectral lines in
a line spectrum of atoms (V.3.4.20) cannot be narrower
Ilian Aco.
3°. A certain interval of wavelengths AX or a definite
range Ak of wave numbers corresponds to the range of
frequencies that is inevitably contained in a wave train.
Considering the propagation of a wave train along the
axis OX, we see that the length Ax of a wave train (Par. 1°)
is inversely proportional to the range Ak of the wave num­
bers in the train:
Ax Ak & 1 and Ax & «■
?—
Ak
The relationships of Paragraphs 2° and 3° hold for waves
of any physical nature: electromagnetic, sound, and de
Itroglie waves associated with moving particles.
4°. For a wave of any nature, the idea that it has cer­
tain coordinates, that is, has a definite place in space,
is deprived of a physical meaning. For example, if a wave
propagating over the surface of water reaches a boat, there
is no sense in stating that the wave is only at the place

* The substantiation of this important conclusion is beyond the


‘M’ope of the present course.
486 Atomic and Nuclear Physics V I.1.5

where it encounters the boat. It follows from the physical


meaning of the de Broglie waves that the length Ax of
a wave train of a de Broglie wave is associated with the
position of the particle in space (VI. 1.3.3°). For micro­
particles having wave properties, the concept of the coordi­
nate of a particle must be used in a different sense than
in classical mechanics. When a ball is rolling along a hori­
zontal path in bowling, the position (coordinate) of the
ball is absolutely accurately determined by the distance
to its centre of mass (1.2.3.4°) from the beginning of the

A ,
0------------ C T _________ ^
0 M X

Fig. V I.1.10

alley. In any problem of classical mechanics, a point par­


ticle (of a body) has definite coordinates characterizing
its position in space at each moment of time. When the
particle M having wave properties moves along the axis
OX (Fig. VI. 1.10), its coordinates on this axis can be de­
termined only with an accuracy to the quantity Ax called
the uncertainty in the coordinate of the particle. The un­
certainty in a coordinate Ax is assessed by the linear di­
mensions of the region of space containing the wave train
associated with the moving particle.
5°. The uncertainty Ax in the position of a particle
on the axis OX determines the range Ak of the de Broglie
wave numbers associated with a moving particle (Par. 3°):
Ak « \! Ax. The quantity Ap determined by a relation­
ship following from de Broglie’s formula (VI. 1.1.3°):
Ap = h Ak
is called the uncertainty in the momentum of a particle.
The concept of the momentum of a particle must be
employed for particles having wave properties in a different
way than in classical mechanics. In the latter, a definite
value of the velocity v of a particle or its momentum p = mv,
where m is the mass of the particle, corresponds to each
definite value of its coordinate. In quantum mechanics
VI.1.5 Elements of Quantum Mechanics 487

in connection with the fact that particles have wave pro­


perties, the coordinate £ of a particle (we are considering
the motion of a particle only ia one direction, along the
axis OX) is determined with an accuracy up to the quan­
tity Ax, and the momentum p of the particle also has no
accurate value. The momentum p is determined only with
an accuracy up to the value Ap of the uncertainty in the
momentum, and
ft
Ap ==% Ak ^ , or Ax Ap ^ h

The product of the uncertainties in the coordinate of a par­


ticle and its momentum has the order of magnitude of
the Planck constant (Heisenberg's uncertainty relation).
When a particle moves in an arbitrary direction, the
following uncertainty relations hold:

AxApx t t h , A yA p y ttT i%and A zA pz t t h


where Ax, Ay, and Az are the uncertainties in the coordi­
nates of the particle along the axes OX, OY and OZ, and
Ap x, Apy, and Ap z are the uncertainties in the projections
of the momentum of the particle onto the same axes.
6°. The uncertainty relation shows that the coordinate
x of the particle M and its momentum p (we are considering
the propagation of a wave train along the axis OX) cannot
simultaneously have values exactly equal to x and p. The
uncertainty in the coordinate Ax and in the momentum
Ap cannot vanish simultaneously because if they do, i.e.
if Ax = 0 and Ap = 0, the uncertainty relation loses its
meaning. The uncertainty relation permits the uncertainty
in only one of these quantities to vanish, for example Ap=0.
This means that the particle has a strictly definite value
of the momentum p (or the velocity v), but now Ax « h/Ap =
= oo. Consequently, the position of the particle on the
axis OX (its coordinate) become absolutely indeterminate:
the particle may be detected at any place on the axis OX
(see Fig. VI. 1.10) within the limits from 0 to oo.
7°. In connection with the fact that no wave properties
are detected in macroscopic bodies (VI.1.3.2°), the un­
certainty relation imposes no restrictions on the possibil-
488 Atomic and Nuclear Physics VI.1.6

ity of determining the coordinates and momenta of such


bodies. A macroscopic body moving along the axis OX
(see Fig. VI. 1.10) can simultaneously have exact values
of its coordinate and momentum.

6. Part Played
by Uncertainty Relations
in Studying the Motion
of Microparticles
1°. If a particle having wave properties moves in a
region whose linear dimensions are much greater than the
dimensions of an atom (VI.2.1.1°), then the uncertainty
relation does not virtually limit the possibility of the
particle simultaneously having an exact value of the co­
ordinate and the momentum. This can be explained by
the following problem.
Let us assume that an electron travels in a betatron
(VI.4.16.3°) along a circle of radius 2.5 m with a velocity
v = 2.97 X 108 m/s. Find the uncertainty in the velocity
Av if the radius of the trajectory has been determined
with the uncertainty Ar forming 0.002% of the radius
of the trajectory.
The uncertainty in the radius of the trajectory Ar —
= r X 0.002% = 0.05 mm, i.e. the trajectory has been
determined very accurately. According to the formulas
given in (VI. 1.5.5°), the uncertainty in the velocity will
be
Av = m- Ar
- « 0.3 m/s

Here the mass of the electron must be taken with a view


to the dependence of the mass on the velocity (V.4.10.30):
m = m J Y 1 — y2/c2 = 7 . 1 m0. The rest mass of the elec­
tron m0 = 9.1 X 10"31 kg.
At a velocity of the electron almost equal to the speed
of light, the uncertainty in the velocity Av = 30 cm/s
may be disregarded, and we may consider that the velocity
of the electron has been determined absolutely exactly.
vr.i.6 Elements of Quantum Mechanics 489

The electron moves along a circle of a definite radius and


has an exact value of the velocity.
When an electron travels in a cathode-ray tube
(111.3.10.3°), we have to do with a situation similar to
the preceding problem. The electron travels along a quite
definite trajectory. At each moment of time, it has an
exact value of the coordinate and an exact value of the
velocity.
2°. When an electron moves in an atom, the uncertainty
relation introduces serious changes into the classical no­
tions of the trajectory of an electron—its orbit (VI.2.2.1°).
The radius of an atom is approximately r « 5 X 10“u m.
The velocity of an electron in its orbit is v » 106 m/s.
If we assume that the uncertainty in the radius of the or­
bit is one per cent of the radius, i.e. Ar = O.Olr =
= 5 X 10"ls m, then the uncertainty in the velocity Ay is

1 0 -3 4
= 2.2 X 10® m/s
9.1 X 10“31 X 5 X 10"13

i.e. is 220 times greater than the value of the velocity itself.
The value of Au almost equals the speed of light. The veloc­
ity of an electron moving in an orbit of a definite radius
becomes absolutely indeterminate, and there is no sense
in speaking of the motion of an electron in an atom along
a definite trajectory—an orbit.* If, conversely, we take
a reasonable value of the uncertainty in the velocity Ay
of an electron, for example, if we assume that Ay = O.Oly =
= 104 m/s, then the uncertainty in the radius Ar is Ar =
= him Ay = 10"34/9.1 X 10~31 X 104 = 1.1 X 10~8 m and
is 220 times greater than the radius of the orbit. Thus,
the radius of the orbit becomes absolutely indeterminate,
and, consequently, we cannot consider that an electron
moves in an orbit as understood in the meaning of the
classical trajectory in mechanics (1.1.1.7°).

* For what is understood by an electron orbit in an atom in quant­


um mechanics, see VI.2.6t3°,
490 Atomic and Nuclear Physics VI.1.7

7. Zero-Point Energy
of a Linear Harmonic Oscillator
1°. The zero-point energy E 0 of a linear harmonic os­
cillator (VI. 1.4.6°) is associated with the quantum prop­
erties of the oscillator and the uncertainty relation. If
a particle of mass m oscillates with the amplitude A along
the axis OX (an oscillator) (see Fig. VI. 1.6), then its total
energy E 0 at the moment of reaching the turning points
B and C is
m®lA2
Z?0= 2
2°. If a particle of mass m has wave properties (a quantum
linear harmonic oscillator), then the de Broglie wave as­
sociated with the particle (VI. 1.1.3°) is “shut in” in the
region with the linear dimensions A , where A is the dis­
placement amplitude of the oscillator. The uncertainty
in the coordinate of the particle Ax (VI. 1.5.4°) will be
A# « A. According to the uncertainty relation (VI. 1.5.5°),
the uncertainty in the momentum of the particle Ap is
h A
tip Ax A
The momentum p of a particle cannot be smaller than
the uncertainty in the momentum Ap*: p ^ Ap. The
momentum p equal in magnitude to the uncertainty in
the momentum tip is called the momentum of a localized
particle: p = Ap.
A particle with wave properties always has a certain
zero-point energy (VI. 1.4.6°) that is the energy of a lo­
calized particle E 0. This is the minimum energy determined
by the momentum of a localized particle p « hi Ax:
F P2 - h2
0 2m ~ 2m ( Ax)2
3°. For a quantum linear harmonic oscillator, Ax « A,
and E 0 « h2/2mA2. On the other hand, according to Par. 1°,

* The substantiation of this conclusion is beyond the scope of


the present course.
V I.1.8 Elements of Quantum Mechanics 491

we have E 0 = m(oJ^42/2. Multiplying the two expressions


for E 0 and extracting a square root, we get E 0 = K&J2
(VI.1.4.6°).
The zero-point energy of a quantum harmonic oscillator
is characterized by “zero-point oscillations” of a particle
occurring at temperatures that are as close as desired to
absolute zero (T = 0 K) (11.4.9.4°). The existence of zero-
point energy has been confirmed experimentally in the
phenomenon of the scattering of light by crystals of solids
at ultralow temperatures. The light is scattered on the
oscillating atoms, molecules, or ions at the points of a
crystal lattice (11.1.6.5°). From the classical viewpoint,
at T 0 K, the oscillations of lattice points should ter­
minate, and scattering of light should stop accordingly.
Experiments have shown that when the temperature of
a body drops, the intensity of the scattered light does not
drop below a certain limit and retains its value upon further
cooling. This occurs because “zero-point oscillations” of
the lattice points continue at ultralow temperatures close
to absolute zero, and light is scattered.

8. The Degeneracy of Gases


1°. The degeneracy of gases is defined as the deviation
of their properties from those of ideal gases (11.2.1.1°)
due to the quantum properties of both the particles of the
gases themselves and of their ensembles. The degeneracy
of gases becomes appreciable at very low temperatures and
great densities.
The degeneracy temperature r deg is the temperature
below which a given gas behaves like a degenerate one.
At T > r deg, a gas is not degenerate, and its properties
are described by the Mendeleev-Clapeyron equation for
ideal gases (11.3.3.7°). The condition for the degeneracy
of gases is T < T^tg .
Figure VI. 1.11 shows how the mean kinetic energy
e of a gas particle depends on the temperature T. At T >
> Tdeg» the straight line AB expresses the direct proportion­
ality between e and T characteristic of ideal gases. At
492 Atomic and Nuclear Physics VI.1.8

T < 7deg> in the degeneracy region BC, there is a non­


linear dependence of e on T. In this region, the absolute
temperature T cannot be determined as a quantity directly
proportional to the mean kinetic energy of a gas molecule
(11.2.4.4°). The ordinate OC char­
acterizes the zero-point energy
of a particle.
2°. The energy of a localized
particle of mass m (VI.1.7.2) is
E0 = h2l2mL2, where L signifies
the linear dimensions of the re­
gion in which the particle is lo­
calized. The energy E 0 can be
associated with the degeneracy
temperature. If n is the number
of particles in 1 cm3 of a gas, then
L = n~ 1/3 because of the condition that L3= 1. The mean ki­
netic energy of a particle at the degeneracy temperature
77(ieg is e = 3/cTdeg/2 (11.2.4.4°), where k is the Boltzmann
constant (11.2.4.4°). From the equality of the two expres­
sions for the energy E 0 = e, we get
^ _ h2nV3
2m ~ 2 KL d(* ’ or ide*~~3*ST
3°. If we can disregard the finite value of Planck’s
constant, i.e. consider that h -> 0, then we can assume that
r deg -> 0. In this case, the degeneracy of gases may be dis­
regarded. In a real case, for instance for hydrogen (m &
« 2 X 10"27 kg) in standard conditions (T = 300 K and
n & 3 X 1025 m~3), we have Tdeg « 1 K. For heavier
gases than hydrogen, r deg is still lower. Atomic and molecu­
lar gases at standard pressures and temperatures are never
degenerate. The point B in Fig. V I.1.11 is near the absolute
zero of temperature. The degeneracy due to the quantum
properties of gases tells to a considerably smaller extent
than the deviations of gases from the ideal state associated
with the forces of interaction between molecules of real
gases (11.1.4.1°).
4°. For an electron gas in metals (11.7.1.2°), we have
1029 m"3, m = 9 X lO"31kg, and = 1.84 X104K.
V l.2.1 th e Structure of Atoms 493

Owing to the small mass of electrons and the great density


of the particles, an electron gas is virtually always dege­
nerate. Only at very high temperatures, above several
tens of thousands of kelvins, would an electron gas obey
the laws of an ideal gas. But metals cannot exist in the con­
densed state at such temperatures. Owing to the degenera­
cy of an electron gas, the conclusions on its properties ob­
tained with the aid of the molecular-kinetic theory of ideal
gases—Ohm’s law for the current density j (111.2.4.7°) —
are in sharp contradiction with experimental results. The
methods of quantum mechanics are used for a correct de­
scription of the electrical conductance of metals.*
5°. With an increase in the concentration n of the par­
ticles in a degenerate gas, the potential energy of inter­
action of its particles Ev « e*n1/3** grows at a slower rate
than the kinetic energy of the particles E 0 = h2n2l3l2m
(E0 is proportional to rc2/3, and Ep to ra1/3, therefore E0
grows at a faster rate than 2?p with increasing concentra­
tion). With a growth in the concentration of the particles,
the properties of a degenerate gas become closer to those of
an ideal gas. The opposite holds for conventional non­
degenerate gases: with a reduction in the concentration of
the molecules (atoms) of a gas, its properties become closer
to those of an ideal gas.

Chapter 2
The Structure of Atoms
1. Rutherford’s Nuclear Model
of an Atom
1°. By a nuclear (planetary) model of an atom is meant
such a model of its structure in which the entire positive
charge of the atom is considered to be concentrated in its
* Information on such a description is beyond the scope of the
I)resent course.
** Ep ^ ePlr, but r oc n1/3 because r*n ^ 1 (see Par. 2°).
494 Atomic and Nuclear Physics Vt.2.1

nucleus (VI.4.1.1°)—a region occupying a very small vol­


ume in comparison with the entire volume of the atom.
The linear dimensions of a nucleus are about 10-15 to 10"14 m.
The remaining part of the atom, whose linear dimensions
are about 10“10m, is occupied by a cloud of negatively charged
electrons. The absolute value of the total negative charge
of the electrons equals the positive charge of the nucleus.
The number of protons in the nucleus equals that of the elec­
trons in the negatively charged cloud and coincides with
the atomic number Z (also called the Z number) of an atom
of a given element in Mendeleev’s periodic table (VI.2.9.1°).
1 9
L.

F ig . V L 2 . l i

The entire mass of an atom is virtually concentrated in its


nucleus. The mass of the electrons travelling near the nu­
cleus is considerably smaller than that of the nucleons
(VI.4.1.1°) contained in the nucleus.
2°. The nuclear model of an atom is the result of Ruth­
erford's experiments set up to study the passage of alpha
particles (VI.4.4.1°) through gold and platinum foil. The
alpha particles emitted by a uranium nucleus (VI.4.7 2°)
have an energy of 4.05 MeV. Rutherford and his collaborat­
ors “bombarded” metal foil with such particles and stud­
ied the scattering of the alpha particles in the substance.
The setup of these experiments is shown schematically in
Fig. V I.2.1. The alpha particles were emitted by the source
1 placed inside a lead chamber with the channel 2. All
the alpha particles except those flying along the channel were
absorbed by the lead. A narrow beam of the alpha particles
was directed onto the gold foil 3 perpendicular to its surface.
The alpha particles that passed through the foil and were
scattered by it caused flashes (the so-called scintillations*)
* F r o m th e L a t in scintillatio m e a n in g a s p a r k , fla s h o f l i g h t .
VI.2.1 The Structure of Atoms 495

on the screen 4 coated with a substance capable of emitting


light when particles strike it (a fluorescent* substance).
A sufficient vacuum (III.3.7.1°)#was ensured in the space
between the foil and the screen to prevent additional scat­
tering of the alpha particles in the air. The design of the
setup permitted alpha particles scattered at angles up to
150 degrees to be observed.
3°. Rutherford’s experiments showed that almost all
the alpha particles which passed through the foil retained
their previous direction of motion or were deflected through
very small angles. Only a few particles were deflected through
large angles of about 135 to 150 degrees.
The results of these experiments were explained very
simply from the viewpoint of the nuclear model of an atom
(Par. 1°). When an alpha particle passes through the elec­
tron shell of an atom, it should not deviate noticeably from
its original direction. The mass of an electron is much small­
er than that of an alpha particle, while the negative charge
of all the electrons is distributed over the entire volume
of the electron shell. The alpha particles that encounter
electrons along their path in a substance are not virtually
scattered on them. Only the few alpha particles that pass
near the nucleus are deflected sharply. At small distances r
from the nucleus, a positively charged alpha particle hav­
ing the charge +2^ (VI.4.4.2°) experiences a considerable
repulsive force F from the nucleus**
p 1 2 ePe
4 jie 0 r2
Here P = number of protons in a nucleus
e0 = electric constant in the SI system (VII.5.1°)
e = 1.6 X 10~19 C = absolute value of an elementary
electric charge.
4°. Figure V I.2.2 shows the trajectory of an alpha par­
ticle near an atomic nucleus. Rutherford succeeded in de­
riving a formula that related the number of alpha particles
scattered through a definite angle 0 to the energy of the al­
pha particles and the number of protons P in a nucleus.

* F lu o r e s c e n c e is o n e o f t h e k in d s o f lu m in e s c e n c e ( V .3 .3 .1 ° ) ?
* * A ll th e fo r m u la s in C h a p . 2 a re g iv e n in t h e S I s y s t e m o f u n it s .
49G A to m ic and N u c le a r P h ysic s VL2.2

Experimental verification of Rutherford’s formula confirmed


its correctness and showed that P = Z, where Z is the
atomic number of a chemical element (VI.2.9.1°). This coin­
cidence is an important proof of the fact that the nuclear
model of an atom corresponds to its actual structure.
5°. If an alpha particle of mass m strikes a nucleus of
charge Ze along the straight line BA (a central collision)

B A
+Ze

Oncoming
alpha particle S c a tte r e d
a lp h a p a rtic le

F ig . V I .2 .2

(see Fig. VI.2.2), then the minimum distance d between the


particle and the nucleus is determined from the condition
that
mv* _ 1 (2g) (Z e)

2 4 jte 0 d
Over the distance d, the entire kinetic energy of the parti­
cle transforms into the potential energy of electrostatic re­
pulsion of the nucleus and the particle. The linear dimension
of the region occupied by the nucleus of an atom is deter­
mined from this formula, namely, d « 10“1B to 10“14 m.

2. Predicam ents
in the C lassical E xplanation
of the N uclear M odel
of an A tom
1°. The electrons of an atom according to the nuclear
model must move relative to the nucleus (the latter is con­
sidered to be stationary throughout the present chapter).
Otherwise, as a result of Coulomb forces of attraction to the
nucleus, the electrons would fall immediately onto the
nucleus. The dynamic stability characteristic of an atom is
explained by the high velocities of its electrons (v « 10® m/s).
V L 2 .2 The Structure of Atoms 497

Let us consider the simplest atom—the hydrogen atom.


It consists of a nucleus—a proton, and a single electron.
The closed trajectory of motion of«an electron relative to the
nucleus is called its orbit in the classical meaning of this
term (the orbit of an electron is treated in greater detail
in V I.1.6.2° and VI.2.6.3°). The velocity v of an electron
travelling along a circle of radius r is determined from the
condition that the centripetal force (1.2.4.5°) retaining the
electron on its orbit is the Coulomb force of its attraction to
the nucleus (111.1.2.5°):
m v2 e2
r 4Jie0ra
where m =mass of an electron
e = its charge
8q = electric constant in the SI system (VI1.5.1°).
An electron has the velocity v « 106 m/s on an orbit
of radius r « 10~10 m. The centripetal acceleration of the
electron here, a = v2/r (1.1.9.1°), is about 1022 m/s2.
2°. The accelerated motion of an electric charge in an
atom should be attended by the emission of electromagnetic
waves of a frequency equal to that of rotation of the electron
about the nucleus (IV.4.4.3°). The energy of an electron in
an atom should continuously diminish at the expense of the
radiation, and the atom cannot be stable. The electron will
not be able to stay on its orbit. It should approach the nu­
cleus along a spiral with a continuously changing frequency
and fall onto it. The spectrum of a hydrogen atom should
contain all kinds of frequencies, i.e. a hydrogen atom should
emit light with a continuous spectrum of frequencies
(V.3.4.40).
3°. All the results given in Paragraphs 1° and 2° have
been obtained with the aid of classical mechanics and elec­
trodynamics. They sharply contradict experimental data
and signify that the laws of classical physics cannot be ap­
plied to the electrons in atoms. The modern theory of the
atom is based on quantum mechanics (VI.1.1.1°).
498 Atomic and Nuclear Physics VI.2.3

3. Line Spectrum of a Hydrogen Atom


1°. The radiation spectrum of a hydrogen atom is a line
one (V.3.4.20). The frequencies vmn of the lines of this

eV
v*,cnr
13.53 o-
13

12

n
10.15 - eg %
2-
10 HgHphfHsHelkHqHeh^
B alm er sen
9
40000-
8
7

6 .8 60000-

5 S

4
80000-
3

2
100000-
1
0

Fig. VI.2.3

spectrum are described by the Balmer-Rydberg formula:

Vmn — ^ ~n* nfi )


V I.2.4 The Structure of Atoms 499

where/? = 3.288 X 1015 s"1 is called the Rydberg constant*.


The integers n and m are called the principal quantum
numbers (VI.2.5.3°), and m = A -f- 1, rt -f 2, etc.
2°. A group of spectral lines with identical values of
n is called a series of spectral lines. The greatest frequency
for each series with the principal quantum number n corre­
sponds to the value m = oo and is called the boundary of
the series, or the spectral term Tn = R/n2.
The frequency vmn of a line equals the term difference:

When n = 1, we get a series of lines in the far ultravio­


let part of the spectrum (the Lyman series):

where m = 2, 3, . . . .
When n = 2, the Balmer series in the visible part of the
spectrum is observed:

where m — 3, 4, 5, . . . .
The infrared part of the spectrum contains other series
of spectral lines. Figure VI.2.3 shows the series of spectral
lines of a hydrogen atom. The right-hand scale shows the
values of v* = 1IX = v/c, which are measured in cm"1.
The left-hand scale shows the values of the energy levels of
a hydrogen atom in eV (VI.2.5.3°).
3°. The line spectrum of a hydrogen atom contradicts the
classical interpretation of the nuclear model of an atom
(VI.2.2.2°).

4. Bohr’s Postulates
1°. The quantum theory of the structure of an atom de­
veloped by N. Bohr (the Bohr model of an atom) is based on
the idea of combining into a single entirety:
* The quantity Rlc, where c is the speed of light in a vacuum, is
also called the Rydberg constant. It is measured in m-1 (VI1.8).
:j2*
500 Atomic and Nuclear Physics V1.2A

(a) the laws of the line spectrum of a hydrogen atom ex­


pressed in the Balmer-Rydberg formula;
(b) Rutherford’s nuclear model of an atom not permit­
ting a classical explanation; and
(c) the quantum nature of the radiation and absorption
of light (V.5.1.40).
2°. To realize this idea, Bohr, while retaining the clas­
sical approach to describing the behaviour of an electron in
an atom, advanced three postulates called Bohr's postu­
lates,,*
Bohr's first postulate (the postulate of stationary states):
stationary quantum states exist in an atom that do not change
with time if no external forces act on the atom.
An atom in these states does not radiate electromagnetic
waves. A definite energy En of an atom corresponds to each
stationary state. Stationary orbits in which electrons trav­
el correspond to the stationary states of an atom. Electrons
travelling in the stationary orbits do not radiate electro­
magnetic waves, notwithstanding their accelerated motion.
Bohr’s first postulate negates the conclusions of electrody­
namics that an electric charge travelling with acceleration
always radiates electromagnetic waves (IV.4.4.3°).
3°. Bohr's second postulate (the frequency condition):
when an atom passes over from one stationary state to an­
other, one photon (V.5.1.20) is emitted or absorbed. An atom
emits (absorbs) one quantum of electromagnetic energy when
an electron makes a transition from an orbit with a great­
er (smaller) principal quantum number to an orbit with
a smaller (greater) one. The energy of a photon equals the
difference between the energies of an atom in two of its
stationary states:
h^mn = En
If E m > En, a photon is emitted, and if E m < En, a photon
is absorbed.

* Bohr’s model is sometimes called the semiclassical model of an


atom. This name is associated with the fact that Bohr, using the laws
of mechanics and electrodynamics, introduced postulates into the de­
scription of the behaviour of an electron that contradicted classical
physics.
VI.2.5 The Structure of A to m s 501

The frequency vmn emitted (absorbed) by an atom is

On the other hand, on the basis of VI.2.3.2°,


Vmn ~ T n 'I'm
where Tn and T m are the spectral terms corresponding to
the principal quantum numbers n and m.
Bohr’s second postulate points to the inversion of spec­
tral lines: atoms absorb only the spectral lines (frequencies)
which they themselves can emit (V.3.4.50).
Bohr’s second postulate was a further development of
the idea of the quantum nature of the radiation and absorp­
tion of light.
4°. Bohr's third postulate (the rule of orbit quantization):
in a stationary state of an atom, an electron travelling in
a circular orbit must have discrete quantized values of the
angular momentum (1.3.2.2°):
Ln = mvnrn = nh
Here n = positive integer
m = mass of an electron
rn = radius of the n-th orbit
vn = velocity of an electron in this orbit
k = h/2n.
(See also VI.2.7.1° for the quantized angular momentum.)
Bohr’s third postulate can be interpreted quite simply
if we take into account the wave properties of an electron
(VI. 1.1.3°). Like the de Broglie wave of an electron travel­
ling in a square potential well (VI.1.4.4°), a whole number of
de Broglie wavelengths Xn should be accommodated on the
length 2jirn of a circular orbit of an electron in an atom:
2nrn = rikn. But Xn = hlmvn, and therefore 2nrn = nhlmvn,
or munrn = nh/2jt = nh.

5. Bohr’s Model of the Hydrogen Atom


1°. A hydrogen atom consists of a nucleus—a proton,
and one electron. Assuming that the electron travels in a
circular orbit, Bohr’s postulates permit us to find the radii
502 Atomic and Nuclear Physics VI.2.5

rn of the allowable stationary orbits of the electron*:


_ _ 2 h2 (4ne0) 2 hHa
I. -- o lb «

Here e is the charge of an electron, and e0 is the electric con­


stant in the SI system (VII.5.1°). The remaining notation is
given in VI.2.4.4°. The radii of the electron orbits in the hy­
drogen atom are directly propor­
tional to the squares of the prin­
cipal quantum number n.
The radius of the first orbit
in a hydrogen atom when n = 1
is called the first Bohr radius:
(4Jte0)
ri = a0--

= 0.528 x 10“10 m = 0.528 A

The first Bohr radius is a unit of length in atomic physics.


2°. The total energy E of the electron in the hydrogen
atom consists of the kinetic energy Ek of the electron and the
potential energy Ev of attraction of the electron to the nu­
cleus:

E=*E* + EV- - 2(4‘Eo)r

The potential energy Ev is negative (Fig. VI.2.4):


e2
~ (4ne0) r

The nucleus is at the origin of coordinates O.


3°. The energy levels (VI. 1.4.8°) En of the electron in
the hydrogen atom
jp me4 1 __ me4 1
“ 2/*2 (4jce0)2 ~ ~ ShHl
* Wherever it seems expedient, the factor 4jie0 characterizing
units of the SI system in electricity and magnetism has been separated
in the formulas.
V I .2 .5 The Structure of Atoms 503

From the relationship between the energy level En and the


spectral term Tn (VI.2.4.3° and VI.2.3.2°)

E ft rp R
h ” l n ~ n2

it follows that
? Rh R2nh
'n ~ n2 ~ n2

where n is a positive integer.


The energy En of the electron in the hydrogen atom de­
pends only on one principal quantum number n. The prin­
cipal quantum number is an integer that determines the ener­
gy levels of the electron in the hydrogen atom. The energy
level at n = 1 is called the ground energy state (the normal
state of the atom). Energy levels at n > 1 are called ex-
cited energy states (excited states of the atom).
4°. The Rydberg constant R (VI.2.3.1°) calculated from
Bohr’s postulate is
„ me4 _ m e4
4jih* (4ne0)2 “ 8hH l

The value of R calculated according to this formula coin­


cides very accurately with the experimental value of this
constant obtained from observations of the frequencies
of the line spectrum of the hydrogen atom. The coincidence
of the experimental and theoretical values of the Rydberg
constant confirms the correctness of Bohr’s model of the
hydrogen atom.
Apart from the hydrogen atom, Bohr’s model can be ap­
plied to hydrogen-like ions. This is the name given to ion­
ized atoms with a nucleus charge of Ze and one electron (for
example, Li2+, Be3+, B4+, etc.).
5°. For atoms with two and more electrons (helium,
lithium, etc.), the model based on Bohr’s postulates does not
el low us to calculate the energy levels of the electrons and
(lie frequencies of the line spectra. For complex atoms, the
methods of quantum mechanics are used for this purpose.
504 Atomic and Nuclear Physics Vl.2.6

6*. Substantiation of Bohr’s Postulates


and the Physical Meaning
of the Electron Orbit
in Quantum Mechanics
1°. Bohr’s first and second postulates were theoretically
substantiated in quantum mechanics. For the substantiation
of his third postulate, see VI.2.4.4°. The postulate of sta­
tionary states (VI.2.4.2°) is a corollary of the fact that in
the stationary state of an electron of energy En the square of
the amplitude of the de Broglie wave (VI. 1.3.3°) does not
depend on the time. The energy of an electron in the station­
ary state remains constant. This signifies (VI. 1.3.4°)
that the probability of an electron being in a state with the
energy En does not change with time. But if the energy En
of the electron does not change, then no radiation will occur.
2°. The second postulate—the frequency condition—
is substantiated in quantum mechanics as follows. Let us
assume that the stationary state of an atom changes under
the influence of external actions, and the atom passes from
the state m to the state n. If a quantum transition between
the two states occurs, then an electron in the atom is in one
state part of the time, as it were, and part of the time in the
other. It is proved in quantum mechanics that an electron in
an atom behaves here like an oscillating charge (IV.4.4.4°)
that emits light. The frequency of the oscillations of the charge
coincides with the frequency of the photon emitted upon
the transition of the electron (and the entire atom) from
the state m to the state n:

3°. When an electron moves in an atom, the uncertainty


relations change the classical concepts of an electron orbit
(VI.2.2.1°). An electron, which has wave properties
(VI.1.1.3°), can be discovered in different places in an atom.
An electron cloud is a definite distribution of the de Brog­
lie wave of an electron in an atom. The electron cloud has
a different density p at different points of an atom (VI.2.7.3°).
VI.2.7 The Structure of Atoms 505

An electron orbit in an atom is defined as the locus of the


points where the electron may be detected with the greatest
probability. In other words, iff
is the locus of the points where
the density of the electron cloud
is greater. In the hydrogen
atom, the probability P (r) of
finding the electron in the ground
energy state at the distance
r from the nucleus has the
form! shown in Fig. VI.2.5.
The probability P (r) has the
greatest value at the distance
from the nucleus that coincides
with the radius of the first Bohr orbit a0 in the hydrogen
atom (VI.2.5.1°).

7*. Quantization of the Angular Momentum


of an Electron and of Its Projection
1°. Quantum mechanics has introduced appreciable
refinements into Bohr’s third postulate on the quantization
of the angular momentum L, (1.3.2.2°) of an electron in an
atom (VI.2.4.4°).
The orbital quantum number I of an electron in an atom
is an integer determining the possible values of Lx of the
electron:
Lx = V l ( l + 1) h
where h = h!2n. The orbital quantum number I does not
coincide with the principal quantum number n (VI.2.5.3°).
At a given value of n, the orbital quantum number can have
Ihe following values:
I = 0, 1, . . ., (» - 1)
The possible values of Lx in quantum mechanics differ
from the quantized values of Ln according to Bohr’s postu­
late in that the expression for Lx includes (/ + 1) in­
stead of the principal quantum number n (VI.2.4.4°).
2°. The formula for Lx when I > 1 and I + I I
gives Li = lh and is similar to Bohr’s postulate Ln = nh.
506 Atomic and Nuclear Physics VI.2.7

The possible values of the orbital quantum number Z,


however, differ from those of the principal quantum number
n. It is especially important that Z can equal zero. There
are such states of an electron in an atom in which the elec­
tron has no angular momentum
p(r) (Lt = 0). In Bohr’s model,
A an “orbit” of ' an electron pas­
sing through the ^nucleus of the
atom corresponds tof-such states.
Experiments show that such
^ states do exist, and this is another
X" r proof of the limited nature of
N Bohr’s model (VI.2.5.50).
3°. The different values of the
orbital quantum number Zof an
electron are the basis of syste­
matizing electron states in atoms
Fig. V I.2.6
and molecules in atomic phys­
ics and modern chemistry. The
following notation isused: if I = 0, then the state of
an electron is called the s-state, if I = 1 , the state of
an electron is called the p-state; a state with Z = 2, 5.
etc. is called the d-, / -state, respectively, with the letters
after / continuing in alphabetical or­
der.
In the Bohr model of the atom, dif­
ferent “shapes” of an electron orbit in
an atom correspond to different values
of the orbital quantum number Z* (ex­
cept for Z = 0).lln quantum mechanics,
different distributions of the density p
of the electron cloud around the nucleus
(VI.2.6.3°) correspond to different values
of Z. For the 5-state of an electron in any
atom, the distribution of the electron
cloud around the nucleus has the shape of a sphere. The
density of the electron cloud is greatest at the distance

* The introduction of the quantum number playing the part of


the orbital one in the Bohr model is not considered in the present
course.
VI.2.7 The Structure of Atoms 507

from the nucleus equal to a0 in the hydrogen atom


(Fig. VI.2.6).
4°. The list of the characteristics of an electron in an
atom that can have only discrete, quantized values is not
exhausted by the energy of the electron and its angular mo­
mentum. The vector L/ of the angular momentum of an
electron cannot have an arbitrary orientation in space. The
orientation of the vector L* in an external magnetic field
of induction B is characterized by the projection LitB
of the vector L* onto the direction of the vector B
(Fig. VI.2.7):

LitB = Li cos a

The absence of arbitrary values of the projection LitB


is called spatial quantization. The vector L; can have only
such orientations in space at which the projection L i% b
will take on integral values
that are multiples of h=h!2n:
LitB = m%
The integer m determining
the possible values of LitB
is called the magnetic quantum
number. It can have the fol­
lowing values:
m = 0, ± 1 , ± 2 , ..., ± 1 d -s ta te

where I is the orbital quan-


turn number (VI.2.7.1°). The
magnetic quantum number can have (21 -f 1) possible val­
ues. The vector Li can have (21 + 1) orientations in space
in accordance with the number of its possible projections
onto the direction of the external magnetic field. Figure
VI.2.8 shows the possible orientations of the vectors L/
Tor an electron in the p- and d-states, i.e. for I = 1 and
/ - 2 (VI.2.7.3°).
508 Atomic and Nuclear Physics VI.2.8

8*. Electron Spin.


The Pauli Exclusion Principle
1°. The spin of an electron or another elementary par­
ticle (VI.5.1.1°) is defined as the intrinsic angular momentum
of the particle due to its quantum nature. Many elementary
particles (VI.5.2.1°) have spin. Among them are the proton,
neutron, antineutrino, and also the atomic nucleus (VI.4.1.4°).
Spin is a property of elementary particles that characterizes
them to the same extent as the rest mass (V.4.10.3°) or the
electric charge (111.1.1.2°) (see also 111.6.1.5°).
2°. A feature of the spin of an electron (and also that of
a proton, neutron, antineutrino, and other particles) is its
quantization (VI.1.4.5°). The spin of an electron (and
other particles) can have only two orientations in an exter­
nal magnetic field. The projections of the spin on the direc­
tion of induction B of the external magnetic field can take
on only two values (111.6.1.5°, see Fig. III.6.2):

If we introduce the magnetic spin number ma = ± 1 /2 , then


L S,B= msh

3°. An illustrative notion of spin is associated with the


rotation of an electron about its axis. This notion supposed­
ly “deepens” the analogy between the structure of an atom
and that of the solar system, where the planets revolve around
the Sun and spin on their axes. The “graphic” notion of spin
contradicts the special theory of relativity (V.4.4.4°).
The velocities u with which points on the “diameter” of an
electron-sphere (VI.5.1.6°) should spin on their axes exceed
the speed of light c in a vacuum.
4°. The stationary quantum state of an electron in an
atom or molecule is characterized by a complete set of the
four quantum numbers: the principal one n, the orbital one
Z, the magnetic one m, and the magnetic spin quantum num­
ber ms. Each of them characterizes quantization: of energy
(ft), of the angular momentum (Z), of its projection on the
V I . 2 .8 The Structure of Atoms 509

direction of the external magnetic field (m), and of the pro­


jection of the spin (ms).
Electrons, protons, neutrons, and other elementary par­
ticles having a spin equal to HI2 obey the Pauli exclusion prin­
ciple: in any system of particles with a spin of H!2 there can
never be more than one particle in a stationary state deter­
mined by a given complete set of the four quantum numbers.
If Zx (n, Z, m, ms) is the number of electrons in an atom
in the state given by a set of the four quantum numbers,
then
Z1 (n, Z, m, ms) = 0 or 1
5°. The maximum number Z2 (n, Z, m) of electrons in an
atom that are in states determined by a set of the three quan­
tum numbers, n, Z, and m is:
Z2 (n, Z, m) = 2

Such electrons differ only in the orientation of their spins.


The maximum number Z3 (rc, Z) of electrons in an atom
that are in states determined by the principal n and orbital
I quantum numbers is:
Z3 (n, I) = 2 (21 + 1)

These electrons differ in the possible values of the magnetic


quantum number m (VI.2.7.4°) and in the orientation of
their spins. Table V I.2.1 gives the values of Z3 (n, I) for
different values of Z.
Table V I .2.1

O rb ital q u an tu m n u m b er I 0 1 2 3 4
S ta te o f e le ctr o n s P d / g
M axim u m nu m b er o f e le c tr o n s
Z 3 ( n , I) 2 6 10 14 18

The maximum number Z (/z) of electrons in an atom that


are in states determined by the value of the principal quan­
tum number n is:
Z (») = 2n2
516 Atomic and N u clea r P h y s ic s Vl.2.9

This number includes the electrons differing in the possible


values of I from 0 to n — 1 (VI.2.7.1°), in the possible values
of the magnetic quantum number m (VI.2.7.4°), and in the
orientation of their spins.
6°. The electrons in an atom occupying a combination of
states with identical values of the principal quantum
number n form an electron shell. The following electron shells
are distinguished: the if-shell at n — 1, the L-shell at
n = 2, the M-shell at n = 3, the TV-shell at n = 4, the
O-shell at n = 5, etc.
The electrons in each electron shell of an atom are dis­
tributed among orbitals, or subshells. A subshell corresponds
to a definite value of the orbital quantum number I
Table V I .2.2
Number of electrons in states
Electron Maximum
n shell p d / 8
number of
(1—0) (i—l) (1=2) (1 = 3) 0=4) electrons

1 K 2 _ _ _ _ 2
2 L 2 6 — — — 8
3 M 2 6 10 — — 18
4 N 2 6 10 14 — 32
5 0 2 6 10 14 18 50

(VI.2.7.1°). Table V I.2.2 gives the maximum numbers of


electrons in the various electron shells ,'and subshells. This
table, which^follows from the Pauli exclusion principle,
explains the sequence of filling the electron states in the atoms
of chemical elements.

9. Mendeleev’s Periodic System


of the Elements
1°. The periodic system (or table) of the elements originally
devised by D. Mendeleev is the name given to the law of the
periodic change in the chemical and physical properties of
elements depending on their atomic number Z (VI.2.1.4°).
V 1 .2 .9 The S tru c tu re of A to m s 511

At iniervals called periods, the elements in one column of the


periodic table (a group of elements) have similar physical and
chemical properties.
2°. The chemical and physical properties of the atoms
of the chemical elements are explained mainly by the be­
haviour of the electrons on the outer shell and outer subshell
of the atoms. These electrons are called peripheral, or valence,
electrons. The periodic nature of the properties of the chemi­
cal elements is associated with the periodic nature in the ar­
rangement of the valence electrons of the atoms of the dif­
ferent elements in a given group.
3°. The theory of the periodic table is based on four pro­
positions:
{a) the total number of electrons in an atom of a given chem­
ical element equals the atomic number Z of this element;
(b) the state of an electron in an atom is determined by
the complete set of the four quantum numbers: n, Z, m, and
(VI.2.8.4°);
(c) the distribution of the electrons in an atom by energy
states (levels) must comply with the principle of the mini­
mum energy: with a growth in the number of electrons, each
following electronAmust occupy an allowable energy state
with the minimum energy; and
(d) the electrons fill the energy states (levels) in an atom
in accordance with the Pauli exclusion principle (VI.2.8.4°).
4°. The electrons fill the states in the different shells
(VI.2.8.6°), and in the subshells within a single shell in ac­
cordance with the requirement of Par. 3° (c): first the states
with the minimum allowable energy are filled, and then the
states with a greater and greater energy. This sequence for
many atoms corresponds to the fact that first the shells with
a smaller value of n are filled, and then the shells with a great­
er value of the principal quantum number. Within the lim­
its of one shell, first the subshells are filled for which I =
^0, and then those with greater values of Zup to I = n — 1.
Filling of the shells and subshells according to this principle
must correspond to Table VI.2.2.
5°. In the actual periodic table (see the back inside cover),
the distribution of electrons in atoms by states differs from
that corresponding to Table VI.2.2. As a result of the inter­
action between electrons at quantum numbers of n =
Table VI.2.3

Distribution of Electrons in Atoms

K L M N 0 P Q
E le­
z ment
Is 2s 2p 3s 3p 3d 4s 4p 4d 4 / 5 s 5p 5d 5 / 6s 6 p 6d 6 / 7s

1 H 1
2 He 2
3 Li 2 1
4 Be 2 2

5 B 2 2 1
6 G 2 2 2
7 N 2 2 3
8 0 2 2 4
9 F 2 2 5
10 Ne 2 2 6

11 Na 2 2 6 1
12 Mg 2 2 6 2

13 A1 2 2 6 2 1
14 Si 2 2 6 2 2
15 P 2 2 6 2 3
16 S 2 2 6 2 4
17 Cl 2 2 6 2 5
18 Ar 2 2 6 2 6

19 K 2 2 6 2 6 1
20 Ca 2 2 6 2 6 2

21 Sc 2 2 6 2 6 1 2
22 Ti 2 2 6 2 6 2 2
23 V 2 2 6 2 6 3 2
24 Cr 2 2 6 2 6 5 1
25 Mn 2 2 6 2 6 5 2
26 Fe 2 2 6 2 6 6 2
27 Co 2 2 6 2 6 7 2
28 Ni 2 2 6 2 6 8 2
29 Cu 2 2 6 2 6 10 1
30 Zn 2 2 6 2 6 10 2

31 Ga 2 2 6 2 6 10 2 1
32 Ge 2 2 6 2 6 10 2 2
33 As 2 2 6 2 6 10 2 3
34 Se 2 2 6 2 6 10 2 4
35 Br 2 2 6 2 6 10 2 5
36 K 2 2 6 2 6 10 2 6
VI.2.10 The Structure of Atoms 513

= 3 (Af-shell), n = 4 (TV-shell), and so on, states with greater


n's, and smaller /’shave a smaller energy and are more advan­
tageous from the energy viewpoint* than states with smaller
rVs, but greater Z’s. The real filling of the energy states by
the electrons in atoms with Z from 1 to 36 is shown in Table
VI.2.3.
Inspection of this table shows, for example, that vio­
lation of ideal filling begins from potassium (Z = 19).
The nineteenth electron of potassium ought to occupy the
state in the Af-shell with n = 3 and 1 = 2. But the chemical
and optical properties of potassium are similar to those of
lithium and sodium in which the valence electron occupies
the states n = 2, 1 = 0, and n = 3, 1 = 0, respectively.
Therefore the nineteenth valence electron of potassium must
be in the 5-state (/ = 0). It occupies this state in the follow­
ing iV-shell (n = 4). From potassium (Z = 19), “build­
ing up” of the iV-shell begins with the c/-subshell unfilled
(/ = 2) in the Af-shell. This continues up to scandium (Z =
21). Beginning from the latter, filling of the d-subshell in
the Af-shell is renewed, and this terminates in copper (Z =
=29). Further up to krypton (Z=36), normal filling of the
Af-shell occurs.*
The periodic table of elements contains chemical elements
with a nucleus charge (atomic number) from Z = 1 (hydro­
gen) to Z = 104 (kurchatovium). The chemical elements
with the atomic numbers Z = 105 and Z = 106 have not
yet been named.

10*. Lasers and Masers


1°. Lasers (an acronym for Zight amplification by stim-
ulated emission of radiation) and masers (/mcrowave
amplification by stimulated emission of radiation) are de­
vices used for amplifying the relevant waves—lasers in the
optical range, and masers in the range of ultrashort radio
waves (microwaves).

* A d e ta ile d d iscu ssio n of a ll the v io la tio n s in the id e a l fillin g


nf the a to m s b y ele ctr o n s in th e p eriod ic tab le is b eyond the scope c l
the present course.
514 A to m ic a n d N u c le a r P h y s ic s VI.2.10

2°. Stimulated (induced) radiation is the radiation of


excited atoms (molecules, Lions) of a substance due to the
action of the light falling on it. An'atom in an excited energy
state (VI.2.5.3°) can drop to a lower (usually ground) ener­
gy state under the action of an electromagnetic field. The
latter, as it were, “throws” the atom from the excited energy
level down to the ground or a less excited level.
3°. The stimulated radiation has absolutely the same prop­
erties as the radiation that caused it to appear. The new
photon appearing as a result
Before in te ra c tio n A fte r of the fact that an atom (mol­
in te ra c tio n ecule jjion^of a substance drops
/7V from a higher level E 2 to
a lower level Ex under the ac­
(a) A b so rp tio n tion of light does not differ in
any way from the photon
h v , that caused its appearance.
2hvz
From the viewpoint of wave
(b) S tim u la te d em ission
optics, the phenomenon of
stimulated radiation consists
F ig. V I .2.9
in an increase in the intensity
of the electromagnetic wave
passing through a substance.
The frequency of the wave, the direction of its propagation,
the phase and polarization remain unchanged. Stimulated
radiation is strictly coherent with the transmitted light
that induced it (IV.3.9.3°).
4°. The new photon that appears as a result of the stimu­
lated radiation amplifies the light passing through the me­
dium. The absorption of light occurs simultaneously with
the stimulated radiation. A photon may be absorbed by an
atom in the ground state Ex. The photon vanishes, and the
atom ascends to the excited state E 2. The absorption of
photons diminishes the intensity of the light passing through
the medium. Figure VI.2.9shows the two competing process­
es: absorption and stimulated emission. The former process
reduces the number of photons passing through the medium.
The latter process increases their number.
5°. A medium is called amplifying (an active medium)
if the processes of stimulated emission predominate over the
processes of light absorption. Otherwise the medium weakens
VL2.10 T he S tru c tu re of A to m s 515

(lie light passing through it instead of amplifying it. An


amplifying medium is also called a medium with negative
absorption of light. The intensity0J of the transmitted light
rapidly grows in such a medium with an increase in its thick-

hv^

ft
hv
hv^ — *~hv
hv — *»hv
hv^ — *»hv
hv —
/?v> — *~hv
— *~hv

Fig. V I.2.10 Fig. V I.2.11

ness (Fig. VI.2.10) as a result of the avalanche-like growth


in the number of photons. The two photons formed in one
act of stimulated emission (Fig. VI.2.11) when they encounter
two atoms at an excited state cause the latter to descend to
Ihe ground state, after which
lour identical photons will fly £
along, etc. (see Fig. V I.2.11). 3 Laser
radiation
6°. To obtain a medium
with negative absorption of
light, an unusual non-equilib­
rium state (an inverted state)
must be created in it. The po- _
filiation of atoms (molecules,
ions) at the excited state must
Fig. V I.2.12
lx* greater than that in the
ground state. Such a distribu­
tio n of the population of atoms by states is inverted in com­
parison with the conventional one. There are usually less
a to m s at higher levels than at lower ones.
7°. The process of transferring a medium into the in­
verted state is called pumping of the amplifying medium.
In practice, pumping is conducted according to a three-
level scheme of a laser. In one kind of gas lasers, the plasma
5i6 Atomic and Nuclear Physics VI.2.10

(111.3.6.1°) of a high-frequency gas discharge (111.3.3.1°)


obtained in a mixture of helium and neon is used as the
amplifying medium. Figure VI.2.12 shows a simplified
three-level energy diagram of such a laser. The helium atoms
are excited by collisions of the electrons and ascend to the
excited state E3. When the excited helium atoms collide
with the neon atoms, the latter are also excited and ascend
to one of the upper levels of neon. The transition of the neon
atoms from this level to one of the lower levels E2 is attended
by laser radiation.
8°. The light amplification effect in lasers is increased
as a result of the multifold passage of the light being am-

Mirror Half-silvered
mirror/

Active medium

Fig. VI.2.13

plified through the same layer of the active medium. This


can be achieved if we place a layer of a medium with neg­
ative absorption (a tube with a gas or a crystal) between
two mirrors parallel to each other (Fig. VI.2.13). A schema­
tic diagram of laser operation is shown in Fig. VI.2.14.
The photon A , which travels parallel to the axis of the tube
or crystal, gives birth to an avalanche of photons flying in
the same direction (Fig. VI.2.14a). Part of this avalanche pass­
es out through the half-silvered (partially reflecting) mir­
ror 3, and part is reflected and grows in the active medium
1 (Fig. VI.2.146). When the avalanche of photons reaches the
mirror 2, it is partly absorbed, and after reflection from the
mirror 2 the amplified stream of photons will travel in the
same way as the initial “seed” photon (Fig. VI.2.14c). The
stream of photons amplified many times and leaving the
laser through the half-silvered mirror produces a beam of
light rays of an enormous intensity with a small angular
divergence, i.e. a sharply directed beam. The photons B
V I . 3. 1 Structure and Spectra of Molecules 517

and C flying to a side, at an angle to the axis of the tube or


crystal, set up avalanches that after a small number of
reflections leave the active meflium (see Fig. VI.2.14a)
and do not participate in
amplification of the light.
9°. The high coherence
and sharp directivity of
laser rays permit lasers to
be used successfully for
communication, radar
(IV.4.5.6°), etc. It is theo­
retically possible to trans­
mit 10000 radio programmes
when the radiation
band has a width of 1 A
and the wavelength is 1 p,m.
Lasers are used for com­
munication over tremen­
dous distances of an astro­ 7
nomical magnitude.
Laser rays pierce very ( c)

minute holes in such hard


substances as diamond, and Fig. V I.2.14
arc used in the welding of
microcomponents. Laser rays are used in microsurgery for
“welding” torn and detached retinas in the human eye to the
underlying tissue (V .l.7.30). Laser radiation is finding
greater and greater application with every passing year.

Chapter 3
Structure and Spectra of Molecules
1, General Characteristics
of Chemical Bonds
1°. Molecules (11.1.1.3°) consist of atoms or ions joined
into a single whole by chemical bonds (11.1.1.3°). The stabi­
lity of molecules is an indication that the chemical bonds are
518 Atomic and Nuclear Physics VI.3.1

due to the forces of interaction (11.1.4.2°) binding the atoms


into molecules.
2°. To decompose a molecule into its constituent atoms
or ions, work has to be done (VI.3.1.5°). The formation of a
molecule is attended by the liberation of energy. Two atoms
of hydrogen (H) in the free state have a greater energy than
the same atoms forming a molecule (H2). The energy liber­
ated in the formation of a molecule is a measure of the work
done by the forces of interaction that bind atoms (ions) into
a molecule.
3°. All the kinds of chemical bonds are due to the inter­
action between the valence electrons of the atoms (VI.2.9.2°).
This is confirmed by the sharp change in the optical spectra
of atoms when molecules are formed. The line spectra of
atoms are determined by the state of the external valence elec­
trons (VI.2.3.1°). The changes in these spectra when mole­
cules are formed signify that the states of the valence elec­
trons change. At the same time, the characteristic X-ray
spectra (V.3.6.30) that depend on the electrons on inner shells
of the atoms (VI.2.9.4°) do not change when atoms combine
chemically. The electrons whose states are easily changed
with the outlay of a small amount of energy participate in
the formation of chemical bonds. The outer valence electrons
are such electrons.
4°. The general characteristic of the forces of interatomic
interaction is similar to the same characteristic of the forces
of intermolecular interaction (II.1.4.1°-5°). The simulta­
neous action of oppositely directed forces—attraction and
repulsion—results in the two forces balancing each other at
a certain distance r0 between the atoms. When r = r0, the
geometrical sum of the forces of attraction and repulsion
equals zero. The minimum potential energy Ev (r0) of inter­
action of the atoms in a molecule corresponds to this distance.
Figure VI.3.1 gives three curves: the force of attraction
F 2, the force of repulsion Fx, and the resultant force F of
interaction of the atoms in a diatomic molecule depending
on the distance r between the atoms. Forces of repulsion are
considered to be positive (11.1.4.4°). Figure VI.3.2 shows
the dependence on r of the potential energy Ev (r) of inte­
raction of the atoms in a diatomic molecule (cf. Fig. II.1.3
for the potential energy of intermolecular interaction).
V 1.3.2 Structure and Spectra of Molecules 519

5°. The bond length is defined as the equilibrium inter-


alomic distance r0 in a moleculer(Par. 4°). The dissociation
energy (bond energy) D (see Fig. V4.3.2) of a molecule is the
energy numerically equal to the work that has to be done to
break the chemical bonds injthe molecule. This energy is

Fig. VI.3.1 Fig. VI.3.2

needed to break up the molecule into its constituent atoms


(or ions) and remove the latter to beyond the range of action
of interatomic forces. The dissociation energy numerically
equals the energy liberated in the formation of a molecule,
but is opposite in sign: the dissociation energy is negative,
whereas the energy liberated in the formation of a molecule
is positive.

2. Ionic Molecules
1°. An ionic (heteropolar*) molecule is one that consists
of oppositely charged ions (111.3.1.1°) of the chemical ele­
ments in the molecule (a molecule with a heteropolar bond).
The sum of the positive and negative charges of the ions in
n molecule equals zero, consequently, ionic molecules are
electrically neutral. The chemical bond ( V I.3.1.1°) is mainly
Ibe result of electrostatic attraction of oppositely charged
ions. Typical examples of ionic molecules are the alkali hal-
ides NaCl, RbBr, Csl, etc. These molecules are formed
* From the Greek heteros meaning other or different.
520 Atomic and Nuclear Physics VI.3.2

when atoms of chemical elements of the first and seventh


groups of the periodic table (VI.2.9.1°) combine.
2°. The formation of ionic molecules is determined by
the increased stability of the outer eight-electron subshell
in the atoms (VI.2.8.6°, Table VI.2.2).
Let us consider as an example the formation of the mole­
cule of NaCl. An atom of Na, like the other atoms of metals
of the first group (VI.2.9.1°), has an eleventh valence ele­
ctron that is bound more weakly to the nucleus than the
inner 10 electrons. An ionization energy of 5.1 eV is needed
to detach this electron. An atom of Cl and the other atoms
of this group have seven outer valence electrons.
The electron affinity 2?aff is defined as the amount of energy
that is liberated when an electron is attached to a metal­
loid atom. For chlorine, i?aff = 3.8 eV. The transfer of an
electron from an atom of Na to one of Cl leads to the formation
of the ions N a+ and Cl”. Each of them has a stable outer sub­
shell (VI.2.8.6°, Table VI.2.2).
3°. The electrostatic attraction of the ions N a+ and Cl"
brings them closer together. At very small distances between
the ions, the forces of attraction are replaced by forces of
repulsion that prevent further approach of the ions. The
forces of repulsion are mainly due to repulsion between the
nuclei of sodium and chlorine when the distances between
them are small. The ions N a+ and Cl" arrange themselves at
the equilibrium distance r0 from one another (VI.3.1.5°)
corresponding to equilibrium between the forces of attrac­
tion and repulsion. The stable molecule of NaCl is formed
with an ionic bond.
4°. The fact that the ionization energy of Na is greater
than the electron affinity of chlorine by 5.1 — 3.8 = 1.3 eV
signifies that the transfer of an electron from a sodium atom
to a chlorine one requires the expenditure of energy. On the
other hand, energy is liberated when a molecule is formed.
When the ions N a+and Cl" approach each other, the energy
of their electrostatic attraction is liberated. The ions are
formed and approach each other simultaneously. The mol­
ecule of NaCl is formed only after the atoms come so close
to each other that together with the formation of ions the
liberated energy will be sufficient for producing a stable
molecule.
V I.3.3 Structure and Spectra of Molecules 521

3. Molecules with a Covalent Chemical Bond


1°. Molecules formed when neutral, frequently identical
atoms, combine are called atomic molecules (molecules with
a covalent or homopolar* bond). Examples of such molecules
are diatomic molecules such as H2, 0 2, N 2, and also HF,
NO, NH3, and CH4. A covalent bond has the property of
saturation, which is expressed in a definite valency of the
atoms. A hydrogen atom can combine with only one other
hydrogen atom. A carbon atom can bind only four atoms of
hydrogen, etc.
2°. A homopolar bond has a quantum-mechanical nature.
Its explanation is based on:
(a) the wave properties of an electron (VI.1.1.3°) and the
distribution of the probability of finding an electron in an
atom (VI.2.6.3°), particularly a spherically symmetrical
electron cloud in 5-states of atoms ,
(VI.2.7.3°); and
(b) the indistinguishability in
principle of identical particles,
particularly electrons. The two
electrons in a hydrogen molecule
each [travelling about its nucle­
us—a proton—do not differ in any
way from each other.® They have
identical charges, rest masses, and
the spins of both of them equal
M2 (VI.2.8.2°).
3°. Assume ;that the electrons
in a hydrogen molecule change
places: the electron *7 formerly belonging to the nucleus
a takesj the place ofj the electron 2 belonging to the nucleus
b, and the electron 2 takes the place of the electron 7. Noth­
ing will change in the state of the hydrogen molecule be­
cause the electrons 1 and 2 are indistinguishable from each
other (Fig. VI.3.3).
4°. The exchange of electrons in a hydrogen molecule
can occur when the nuclei of the hydrogen atoms come suffi­
ciently close to each other. The “electron clouds” overlap
(see Fig. VI.3.3), and a specific quantum-mechanical ex~
* From the Greek homos meaning one and the same, alike.
5 22 Atomic and Nuclear Physics VI.3.4

change interaction occurs between the electrons. Its physical


meaning consists in that each of the electrons in a hydrogen
molecule can belong alternately to either nucleus—the elec­
trons continuously change places. This can be illustrated by
the continuous exchange of two tennis balls between two
persons close to each other. If these persons have not received
special training, then the suc­
Ep(r),eV cessful exchange of the balls is
possible only with a small dis­
tance between the partners.
5°. Quantum-mechanical cal­
culations show that if the spins
of the two electrons in a hydro­
gen molecule are antiparallel
(VI.2.8.2°), then exchange inter­
action leads to the attraction
of the two hydrogen atoms and
the formation of the stable hy­
drogen molecule. The potential
energy Ev (r) of interaction of
the two atoms has a minimum at
a distance between the atoms of
r0 = 1.6a0 = 0.83 A, where a0
is the radius of the first Bohr
orbit of a hydrogen atom
(VI.2.5.1°) (Fig. VI.3.4, curve 1). With parallel spins of the
electrons, the hydrogen atoms repel each other and no hydro­
gen molecule is formed (Fig. VI.3.4, curve 2). The equilib­
rium distance r0 and the dissociation energy D (VI.3.1.5°)
in a hydrogen molecule calculated in quantum mechanics
are in good agreement with the experimentally obtained
values of these quantities.

4*. Molecular Spectra


1°. The emission and absorption spectra of separate
molecules (V.3.4.1°) are a combination of hands formed by
closely arranged spectral lines. Molecular spectra are called
hand spectra (Par. 6°). The bands in molecular spectra are
observed in the infrared, visible, and ultraviolet range of
electromagnetic waves (V.3.7.10).
V I.3.4 Structure and Spectra of Molecules 523

2°. A spectral line in a molecular spectrum, as in atomic


spectra, appears when the energy of the molecule changes.
The total energy E of a molecule consists of five parts,
which in the first approximation are independent of one
another: EtT—the energy of translational motion of the cen­
tre of mass (1.2.3.4°) of the molecule; ETot—the energy of
rotational motion of the molecule as a whole about certain
axes; Evlh—the energy of vibrational motion of the nuclei
of the atoms in the molecule; 2?ei—the energy of motion of
the electrons in the atoms of the molecule; and i?nuc —the
energy of the nuclei of the atoms in the molecule, i.e.
E — ^ t r 4“ E t0{ 4“ ^ v ib + Ee\ 4“ ^nuc
3°. The energy E ' of a molecule determining its spectrum
consists of three parts:
= -^rot 4" ■fi’vib 4“&e\
E'
The changes in the nuclear energy Enuc do not affect
the spectra of molecules. The energy Etr changes continuous­
ly, and its changes are not associated with the optical prop­
erties of molecules.
Each of the three addends in E' changes discretely.*
The changes in the corresponding parts of the energy of a
molecule A£Vot > Ai?vlb, and AEc] also have discrete values,
and their sum is AE':
AEf = A-fiVot + A2?vlb + A2?ei
According to Bohr’s second postulate (the frequency con­
dition) (VI.2.4.3°) the frequency v of the quantum emitted
by a molecule when its energy state changes is
, __A E ' ___A E rof I A/?vlb i A E e]
h ~ h ^ h ^ h
As shown by experiments and theoretical calculations,
A£Vot Ai?vlb AZ^ei-
4°. The formation of bands in the emission and absorp­
tion spectra of molecules can be understood by using the in­
equalities at the end of Par. 3°. If an electromagnetic wave

* The substantiation of this important result obtained in quantum


mechanics is beyond the scope of this course.
524 Atomic and Nuclear Physics VI.3.4

falls on a substance, it is absorbed. When the wavelength


of the incident light is from 0.1 to 1 mm (the far infrared
region of the spectrum), the quanta of energy hv correspond
to a change in the rotational energy of a molecule A£Vot*
The absorption of a photon by a molecule transfers it from
one rotational energy level to another higher one and results
in the appearance of a spectral line of the rotational spectrum
of the molecule.* The combination of all the lines is the
complete rotational spectrum of the molecule.
5°. The absorption by a substance of electromagnetic
waves in the infrared region with wavelengths of from 1 to

v vib-rot
vuib vet-uib
vel

V ib r a tio n a l R o ta tio n a l
le v e ls le v e ls E le c t r o n V ib r a t io n a l
le v e ls le v e ls

F ig . V I . 3 .5 F ig . V I .3 .6

10 \im (microns) results in a change in Transitions


between the vibrational energy levels lead to the appearance
of a vibrational spectrum of a molecule. Upon a change in the
vibrational energy levels of a molecule, however, its rota­
tional energy states change simultaneously. The transitions
between two vibrational energy levels are attended by a
change in the rotational energy states, and a vibrational-
rotational spectrum of the molecule appears. In Fig. VI.3.5,
each transition of a molecule between two vibrational levels
producing a line with the frequency vvli, is attended by tran­
sitions between the rotational levels. The result is the for­
mation of a spectrum with the frequencies vvlb_rot. It con­
sists of groups of very close lines formed by rotational Iran-

* T h e t r a n s it io n o f a m o le c u le fr o m a n u p p e r r o t a t io n a l e n e r g y
l e v e l to a lo w e r o n e le a d s t o t h e a p p e a r a n c e o f a lin e o f a r o t a t io n a l e m is ­
sion s p e c t r u m .
Vi.3.4 Structure and Spectra of Molecules 525

sitions corresponding to one vibrational transition. All these


lines merge into a single vibrational-rotational band.
6°. The absorption of electrcflnagnetic waves of the vis­
ible ami ultraviolet range (V.3.7.10) leads to changes in
A-Eei and to transitions of a molecule between different elec­
tron energy levels, i.e. to the appearance of an electron spec­
trum of a molecule. Various possible vibrations of the nuc­
lei of the atoms in the molecule, i.e. a multitude of vibra-

F ig . V I .3 .7

tional energy levels, correspond to each electron energy lev­


el of the molecule. The transition of a molecule between
two electron levels is attended by many accompanying
transitions between the vibrational levels. An electron-
vibrational spectrum of the molecule appears with the fre­
quencies Vei-vib* It consists of groups of close lines forming
an electron-vibrational band (Fig. VI.3.6). It must also be
taken into consideration that a multitude of rotational lev­
els corresponds to each vibrational energy state (see
Fig. VI.3.5). The entire electron-vibrational spectrum in the
visible region and the one close to it is a system of several
groups of bands overlapping one another and forming a
broad band (a band spectrum of a molecule). Figure VI.3.7
contains a photograph of part of the spectrum of an iodine
molecule.
526 Atomic and Nuclear Physics V1.4.1

Chapter 4

Structure and Basic Properties


of Atomic Nuclei

1. General Characteristic

1°. The nucleus of an atom of any chemical element con­


sists of positively charged protons and of neutrons having
no electric charge. The charge of a proton in its absolute
value equals that of an electron. The proton and the neu­
tron are two states with different charges of a nuclear particle
called a nucleon. The number of protons in a nucleus—its
charge Z—corresponds to the atomic number of the relevant
chemical element in Mendeleev’s periodic table (VI.2.9.1°).
The number of neutrons in a nucleus is denoted by N. For
all nuclei, N ^ Z (except for 1H1 and 2He3—see Par. 2°).
For the light nuclei in the first half of the periodic table,
N!Z « 1. The nuclei of the atoms of the chemical elements
at the end of the periodic table are “overloaded” with neu­
trons—for them N/Z & 1.6.
2°. The mass number A of a nucleus is defined as the total
number of nucleons in the nucleus: A = Z -f N . Nuclei
are designated by the .symbols ZXA, or Xz, where X is the
symbol of an atom of a given element in the periodic table.
Nuclei having identical values of Z with different ^4’s are
called isotopes. The isotopes of nuclei of a given chemical
element have different numbers of neutrons in their nuclei.
Examples are the hydrogen isotopes J l 1, 1H2 (o^D 2—deu­
terium), and 1H3 (or XT3—tritium), the helium isotopes
2He3 and 2He4, and the uranium isotopes 92U235 and 92U238.
There are about three hundred stable and over a thousand
unstable (radioactive) isotopes of all the known chemical
elements.
3°. The mass of an atomic nucleus virtually coincides
with that of the entire atom because the mass of the electrons
VtA.i Structure and Properties of Atomic Nuclei 52/

in an atom is negligible. The mass of an electron* me is


1/1836 of the mass of a proton mv.
The masses of a neutron mn•and a proton rap in atomic
mass units (amu) (VI1.7.1°) are:
mn = 1.008 665 12(37) amu
mv = 1.007 276 470(11) amu
The numbers 37 and 11 in parentheses are the mean square
errors in the last figures.
The mass numbers of a neutron and a proton are the same
and equal unity.
The masses of atoms are also measured in atomic mass
units (VII.7.1°).
Every chemical element usually has a constant propor­
tion of its different isotopes. Chemically pure elements are a
mixture of isotopes differing from one another in their4rel-
ative atomic masses (VII.4.3°). Therefore, every chemical
element has a relative atomic mass that is the mean value of
the relative atomic masses of all its isotopes. The relative
atomic masses of chemical elements appreciably differ from
integers in many cases.
4°. A nucleus has a spin equal to the vector sum of the
spins of the particles forming it. The spin of every nucleon
has two possible values: =h^/2** (VI.2.8.2°). The spin of a
nucleus consisting of an even number of nucleons is an inte­
ger (in units of h) or zero. The spin of a nucleus consisting
of an odd number of nucleons is a half-integer (in units
of K).
5°. An atomic nucleus has no sharply expressed bound­
aries. This is associated with the fact that nucleons have wave
properties (VI. 1.1.3°). Therefore the size of a nucleus has a
conditional meaning.*** The volume of a nucleus is propor­
tional to the number of nucleons A in it. If we consider the

* A ll t h e in f o r m a t io n o n m a s s e s s e t o u t in C h a p . 4 r e la t e s to
r e s t m a s s e s ( V .4 .1 0 .3 ° ) , u n le s s o th e r w is e in d ic a t e d .
** M o re p r e c is e ly , t h e p r o j e c t io n o f th e s p in o n t o th e d ir e c t io n
o f th e in d u c t io n o f th e e x t e r n a l m a g n e t ic fie ld h a s th e tw o v a lu e s
Wj/2.
* * * T h is a ls o r e la t e s to t h e s iz e o f a n a t o m b e c a u s e th e e le c t r o n s
t r a v e llin g in a n a t o m r e l a t i v e to i t s n u c le u s a ls o h a v e w a v e p r o p e r tie s .
528 Atomic and Nuclear Physics Vl.4.2

nucleus to be a sphere of radius R, then R is calculated ac­


cording to the empirical formula
R = R0A 1/3
where R 0 = (1.3 to 1.7) X 10~16 m.
The heaviest nuclei, for example that of uranium, have
radii approaching 10~14 m in their order of magnitude.
6°. The mean density p of the nuclear substance is deter­
mined by the formula
M nuc
P (4 /3 ) jtR*

Here Mnuc is the mass of a nucleus. If mncn is the mass of


a nucleon*, then M nuc = mncnA. The mean density of the
nuclear substance is constant and does not depend on the
number A of nucleons in a nucleus: p = 1.3 X 1017kg/m3.
The colossal mean density p is far beyond the ordinary den­
sities of substances consisting of atoms of chemical elements
and their compounds.

2. Binding Energy. Mass Deficiency


1°. The binding energy of a nucleon in a nucleus is defined
as the physical quantity equal to the work that must be done
to remove the given nucleon from the nucleus without im­
parting any kinetic energy to it.
The {total) binding energy of an atomic nucleus A2?b
(negative in sign) is equal in absolute value to the work that
has to be done to break the nucleus up into its constituent
nucleons without imparting any kinetic energy to them. The
binding energy of an atomic nucleus is the difference between
the energy of the protons and neutrons in the nucleus and
their energy in the free state. It follows from the law of con­
servation of energy that when a nucleus is formed from its
constituent nucleons, energy equal to AEb must be liberat­
ed—the binding energy of the nucleus.

* F o r th e d iffe r e n c e b e tw e e n th e m a s s e s o f a n e u tr o n a n d a
p r o to n s e e V I .4 .1 .3 ° .
VIA.2 S tru ctu re and P ro p e rtie s of A to m ic N u clei 529

2°. The quantity AEJA equal to the average binding


energy per nucleon is called the unit binding energy of a
nucleus Aeb. Figure VI.4.1 show§ the unit binding energy
against the mass number A . The energy A6b has a maximum
value for the nuclei of atoms in the middle part of the period­
ic table (VI.2.9.1°), from 14Si28 to 56Ba138, i.e. when 28 <
<CA < 1 3 8 . For these nuclei, Aeb is about 8.7 MeV per

nucleon. The unit binding energy diminishes as the nuclei


become overloaded with neutrons. For nuclei at the end of
the periodic table (for example for uranium), the value of
Aeb is about 7.6 MeV per nucleon. In the region of small
mass numbers, the unit binding energy displays characteris­
tic maxima and minima (see Fig. VI.4.1). The maxima of
Aeb are observed in this region for nuclei having even num­
bers of protons and neutrons: 2He4, 6C12, 80 16. The minima
of Aeb correspond to nuclei with odd numbers of protons and
neutrons: XH2, 3Li6, 5B10. The dependence of Aeb on A ex­
plains the mechanism of liberation of nuclear energy
(VI.4.11.2°).
3°. A measure of the binding energy of an atomic nucleus
is the mass deficiency. The mass deficiency (or defect) Am
is defined as the difference between the total mass of all the
.'1 4 -0 2 1 1
§30_______________ A t o m i c a n d N u c l e a r P h y s i c s _____________V t . 4 . 2

nucleons of a nucleus in the free state and the mass of the


nucleus M nuc:
Am => Zmv -{-{A — Z)m u— Mnuc
Here Z = number of protons in the nucleus each of mass mv
A — Z = number of neutrons in the nucleus each of
mass mn.
If AEb is the binding energy of a nucleus liberated in
its formation (Par. 1°), then the mass corresponding to it
(V.4.11.10)
4ro = ^ :
characterizes the reduction in the total mass of all the nu­
cleons upon the formation of the nucleus. Hence,

AEh= [Zmp + {A — Z) mn— M miC] c2

The binding energy of a nucleus expressed through the mass


of the atom M Rt is
= [ZMh -\-{A — Z) mn— M at] c2

where M H is the mass of a hydrogen atom.


4°. Atomic nuclei, like atoms, have discrete quantized
values of the energy E.
If a nucleus has the minimum possible energy equal to
the binding energy AZ?b, then it is in the ground energy state.
If a nucleus has the energy E > £ mln = AZ?b, then it is
in an excited energy state. The case E = 0 corresponds to
disintegration of a nucleus into its constituent nucleons.
Unlike the energy levels of an atom spread over units of
electron-volts (see Fig. VI.2.3, the left-hand scale), the ener­
gy levels of a nucleus are spread over units of megaelectron-
volts (MeV) . This explains the origin and properties of gamma
rays (VI.4.7.80).
Problem. Calculate the mass deficiency and the binding
energy of the nucleus of the lithium isotope 3Li7.
z \ Solution: The mass deficiency is

Am = Zmp + Nmn— M nuc


V 1.4.3 S tru ctu re and P ro p ertie s of A to m ic N u c l e i 531

where Z is the atomic number of the element equal to the


number of protons in its nucleus, N is the number of neutrons
in the nucleus, and M nuc is the m&ss of the nucleus.
Since A/nuc = M ai — Zme, where M ai is the mass of
an atom, and me is the mass of an electron, then
Am = ZM n -f Nmn— M at = 3 x 1.007 83 + 4 x 1.008 67 —
— 7.018 22 = 0.041 86 amu (V II.7.1°)
The binding energy of the nucleus is
AE = Am c2, or AE = Am x 931.5 =
= 0.041 86 x 931.5 « 39 MeV
where 931.5 is the energy corresponding to one atomic mass
unit (VII.7.2°).

3. N uclear Forces. Drop N uclear Model


1°. The forces acting between the nucleons in a nucleus
and ensuring the existence of stable nuclei are called nuclear
forces. They are specific forces differing from gravitational
forces (1.2.8.1°) and forces of electromagnetic interaction
(111.1.3.1°). The interaction between nucleons is an example
of the strong interactions between elementary particles
(VI.5.2.5°).
2°. Nuclear forces have a number of specific properties:
(a) They are forces of attraction.
(b) Nuclear forces are short-range ones. They manifest
themselves only at very small distances between nucleons
comparable with the linear dimensions of the nucleons them­
selves. The distance r over which nuclear forces act is called
the range of nuclear forces (r & 2.2 X 10'15 m).
(c) They are independent of the charge. The nuclear forces
noting between two protons, between two neutrons, or
between a proton and a neutron are the same. Hence it fol­
lows that nuclear forces cannot have an electromagnetic
nature. The charge independence of nuclear forces is con­
firmed by the difference in the binding energy of mirror nuclei.
The nucleus B is called a mirror one with respect to the nuc­
leus A if the number of protons in B equals the number of
532 A to m ic an d N u clea r P h y sic s Vt.4.3

neutrons in A, and the number of neutrons in B equals the


number of protons in A. The simplest mirror nuclei of trit­
ium iH3 and helium 2He3 have binding energies equal to
8.49 and 7.72 MeV, respectively. Each of these two nuclei
has three nucleons. Their bond is stronger in the tritium nu­
cleus than in the helium one. In the helium nucleus, the mu­
tual repulsion of the two protons reduces the binding energy
by 8.49 — 7.72 = 0.77 MeV. Assuming that the poten­
tial energy Ev (r) of repulsion of the protons is Ev (r) =
= e2l(4jxe0) r = 0.77 MeV, we can assess the distance be­
tween the protons in the nucleus 2He3. It is 1.9 X 10"15 m
and corresponds to the range of action of nuclear forces.
(d) They have the property of saturation: each nucleon
interacts only with a limited number of the nearest nucleons,
and not with all the nucleons of the nucleus. This property
follows from the virtually linear relationship between the
binding energy AEh in a nucleus and the mass number A .
If there were no saturation and each of the A nucleons react­
ed with all the A — 1 nucleons, then the binding energy
would be proportional to the number of pairs of nucleons in
the nucleus, i.e. to the number of combinations of A parti­
cles taken two at a time: 0.5A (A — 1) = 0.5 (A2 — A).
The dependence of AEh on A would be quadratic instead of
linear. In the same way as the saturation of a covalent chem­
ical bond (VI.3.3.1°) leads to the formation of stable
groups of atoms—molecules—so does the saturation of the
nuclear forces lead to the stability of definite groups of
nucleons. The virtually complete saturation of the nuclear
forces is achieved in an alpha particle, which is a stable
formation of two protons and two neutrons (VI.4.4.2°).
(e) Nuclear forces are non-central forces, unlike Coulomb
and gravitational ones, which depend on the distance be­
tween the particles (central forces—1.2.8.1°). This manifests
itself in that apart from the distance between nucleons, the
nuclear forces also depend on the orientation of the nucleon
spins, on whether the latter are parallel or antiparallel
(VI.2.8.2°). It has been proved experimentally that a stream
of neutrons is scattered differently on molecules of ortho-
and para-hydrogen. A molecule of H2 in which the spins of
both protons in the nuclei are antiparallel is called para-
hydrogen. In a molecule of ortho-hydrogen, the spins of both
V I .4 .3 S tru ctu re and P ro p ertie s of A to m ic N u clei 533

protons are parallel. If the interaction of the nucleons did


not depend on the orientation of their spins, then the scat­
tering of the neutrons on ortho- *and para-hydrogen would
occur identically.
3°. Nuclear forces have not been studied in detail to
date. There is meanwhile no comprehensive theory of nu­
clear forces. A fruitful method of studying various properties
of an atomic nucleus is the method of nuclear models based
on the external analogy of the properties of atomic nuclei
and those of other systems that have been well studied in
physics.
4°. The simplest drop nuclear model uses the external
analogy of the following six properties of an atomic nucleus
and a positively charged drop of a liquid:
(a) The small range of action of the nuclear forces and
tlie forces of interaction between the molecules in a drop of
a liquid.
(b) The property of saturation of the forces acting be-
Iween the molecules of a liquid and the saturation of nuclear
forces.
(c) The constant density of the substance in a drop of a
Ii({uid not depending on the number of molecules in the drop.
The average density of nuclear substance is also constant and
does not depend on the number of nucleons in a nucleus.
(d) The particles in a drop of a liquid and an atomic
nucleus—molecules and nucleons, respectively—have a
definite mobility.
(e) The energy of intermolecular attraction in a drop of
a liquid corresponds to the energy of attraction of nucleons
in a nucleus due to nuclear forces. The binding energy in a
nucleus should diminish at the expense of the Coulomb repul­
sion of protons of like charge. This effect should grow with
an increase in the number of protons in a nucleus. In a
drop of a liquid, correspondingly, its stability diminishes
with an increase in its mass, i.e. with a growth in the number
of molecules in the drop.
(f) The molecules of a liquid on its surface experience
unidirectional attraction into the liquid characterized by
lIn* surface tension of the liquid (11.6.1.3°). The nucleons
on the “surface” of a nucleus experience unidirectional at-
lruction into the nucleus due to nuclear forces. This attrac-
534 A to m ic and N u clea r P h y sic s VI.4.4

tion can also be characterized by the surface tension of the


nucleus-drop. A nucleus can be characterized by the magni­
tude of its surface energy, like the energy of the surface layer
of a liquid (11.6.1.2°).
5°. An atomic nucleus is called stable (stability of an
atomic nucleus) if its composition does not change with
time. The relationship between the number of protons Zst
and the mass number A in a stable nucleus, according to
the drop model, is

s‘ i .98 + 0.015 A2/*


The value of Zst is taken equal to the integer that is nearest
to the result obtained by this formula. For light nuclei,
Zgt « A/2: in stable light nuclei the numbers of neutrons
and protons are equal (VI.4.1.1°).

4. N atural R adioactivity
1°. Natural radioactivity is defined as the spontaneous
transformation of nuclei of unstable isotopes of one chemical
element into the nuclei of iso­
topes of other chemical elements.
Natural radioactivity is attend­
ed by the emission of definite
particles: alpha and beta radia­
tion, antineutrinos (VI.4.7.6°),
and also electromagnetic (gam­
ma) radiation. Natural radioac­
tivity is observed, as a rule, in
the heavy nuclei of the elements
after lead in the periodic table
(VI.2.9.1°). There also exist
light naturally radioactive nu­
clei: the potassium isotope 19K40,
the carbon isotope 6C14, the ru­
Fig. VI.4.2 bidium isotope #r7Rb87, and
others.
2°. The composition of radioactive alpha, beta, and gam­
ma radiations is established according to their deviations
in a magnetic field perpendicular to the plane of Fig. VI.4.2.
V I .4 .5 S tru ctu re and P ro p ertie s of A to m ic N u clei 535

In this figure, 1 stands for a thick-walled lead container, and


2 for a radioactive source.
A Ipha particles are known to have a positive charge equal
in absolute value to a double charge of an electron and are
identical to the nucleus of a helium atom.
Beta radiation is a stream of fast electrons with an energy
reaching 10 MeV and a velocity close to the speed of light
in a vacuum.
Gamma radiation is hard electromagnetic radiation. It
has the greatest penetrating ability of all radioactive radia­
tions. The frequencies of gamma rays exceed those of the
hardest X-rays (V.3.6.10). The quantum properties of gamma
radiation (V.5.1.50) manifest themselves to an even greater
extent than those of X-rays.
3°. The properties of radioactive radiations established
according to their interaction with a substance are as follows:
(a) All radioactive radiations have chemical action to
a certain extent. Particularly, they cause the darkening of
photographic plates and films (V.5.6.20).
(b) Radioactive radiations result in the ionization of the
gases, and sometimes also of the solids and liquids, through
which they pass.
(c) Radioactive radiations result in the luminescence of
a number of solids and liquids (V.3.3.10).
The above properties underly the experimental methods
of detecting and investigating radioactive radiations
(VI.4.6.1°-6°).
Naturally radioactive transformations do ^nd on
the external conditions, and also do not de ’ *r
these transformations occur in a substance '
of a chemically pure element or a ch'
Hence it follows that radioactive transto
property of atomic nuclei.

5. D isplacem en t Law and F undam ental Law


of R adioactive D ecay
1°. Transformations of atomic nuclei attended by the
omission of alpha and beta radiation are called alpha and
beta decay, respectively. The term “gamma decay” is never
536 Atomic and Nuclear Physics V I.4.5

used. The disintegrating nucleus is called the parent, and


the nucleus of the product of disintegration the daughter
one.
2°. Nucleus displacement laws in radioactive decays:
in alpha decay ZXA- > Z_2YA_4 + 2He4
in beta- decay* ZXA - ^ z+1YA-f ^ e0
Here X = symbol of the chemical element corresponding
to the parent nucleus
Y = the same for the daughter nucleus
2He4 = nucleus of a helium isotope
^ e0 = designation of an electron: its charge equals —1
(in units of the elementary charge e), while its
mass number (VI.4.1.2°) is assumed to equal
zero because the mass of an electron is 1/1860
of that of a proton.
Alpha decay reduces the mass number of a nucleus by
four, and the charge of the nucleus by two elementary pos­
itive charges, i.e. displaces the chemical element two posi­
tions to the left in the periodic table (VI.2.9.1°).
In beta decay, the mass number does not change, while
the charge of a nucleus grows. The chemical element is
displaced one position to the right in the periodic table.
3°. Daughter nuclei (Par. 1°), as a rule, are radioactive
themselves. A radioactive series (radioactive family) is defined
as a sequence of radioactive transformations from a certain
parent nucleus. The radioactive isotopes of the chemical
elements in the corresponding positions of the periodic table
are members of radioactive families.
There are three naturally radioactive families, which
according to the parent nucleus are called the uranium fa­
mily (92U238), the thorium family (90Th232), and the actinium
family (89Ac23B). There is also an artificially produced radio­
active family beginning from the transuranium element
neptunium (93Np237) (VI. 4. 9.4°). A chain of alpha and beta
decays takes place in each of the radioactive families. In
each naturally radioactive series, the radioactive transfor-

* The symbol beta_ (P_) has been introduced for electron beta
decay because there is also beta+ (6+) positron beta decay (VI.4.10.3°).
V I.4.5 Structure and Properties of Atomic Nuclei 537

matrons terminate on stable nuclei of lead isotopes: the


uranium family on the nucleus of lead 82Pb206, the thorium
family on 82Pb208, and the actinium family on 82Pb207. The
neptunium family terminates on the nucleus of bismuth
83Bi209.
4°. The fundamental law of radioactive decay: the number
of decaying nuclei grows with an increase in their total num­
ber and in the time during which decay occurs. The number
AN of parent nuclei decaying during the interval from t
to t -f At is proportional to the number N of nuclei existing
by the moment t , and the time interval At:
AN = - X N At
The minus sign indicates that the number of nuclei dimin­
ishes as a result of radioactive decay. The positive propor­
tionality constant X is called the decay
constant for a given species of nuclei. AN/N
The decay constant is the relative
reduction in the number of nuclei
that decayed in a unit time:
_ — ANJN

The constant X has the dimension s"1


and characterizes the fraction of nu­
clei decaying in a unit time, i.e. de­
termines the rate of radioactive decay.
The quantity t = l/X is called the
mean lifetime of a radioactive isotope. The values of X
and t do not depend on the external conditions and are de-
lermined only by the properties of the atomic nucleus.
5°. The law of radioactive decay indicates that the radio­
active transformations of atomic nuclei are statistical process­
es (the statistical nature of radioactive transformations).
It is impossible to predict which nucleus of a radioactive
isotope will decay at a given instant. The decay of any of
Ilie nuclei is an event having an equal probability. There­
fore the law of Par. 4° mentions only the number of identical
nuclei AN that decay during the time interval At. Figure
VI.4.3 shows how the relative reduction AN IN in thejnum-
ber of radioactive nuclei depends on the interval At.
538 Atomic and Nuclear Physics VI.4.5

6°. From the fundamental law of radioactive decay


(Par. 4°), we get the law of the reduction of the number of
radioactive nuclei with time:
N = N 0e~H
Here N 0 = initial number of radioactive nuclei that existed
at the initial moment of time, i.e. at t = 0
N = number of radioactive nuclei at the moment t
e = 2.718+ = base of natural logarithms.
7°. The stability of nuclei as regards radioactive decay
is characterized, in addition to A,, by the half-life T . The
half-life is defined as the time during which half of the orig­
inal nuclei decay, or as the time after which half of the
initial number of nuclei remains undecayed: t = T if N =
= N J 2. The relationship between T , A,, and t is expressed
by the formula

7*»1L“ ®= 0 .6 9 3 t, or T= - i « f + f » 1 .4471

The half-lives of different radioactive isotopes vary with­


in very broad limits: for uranium it is 4500 million years,
for radium 1590 years, for radon 3.825 days, and for one of
the polonium isotopes it is 1.5 X 10“4 s. For some of the
artificially radioactive elements (VI.4.10.1°), T is hundred
millionths of a second. The constancy of T (or A,) for a given
species of radioactive nuclei confirms the statistical nature
of radioactive transformations (Par. 5°).
Problem 1. What per cent of the initial amount of a
radioactive chemical element decays during the time equal
to the mean lifetime of this element?
Given: t = t .
Required: 100 (N q — N )/N 0.
Solution: According to the law of radioactive decay,
N = N 0e - Xt, where N is the number of undecayed atoms by
the moment t, and A, is the decay constant. The mean life­
time t = 1/A,.
Hence, N = N 0e~WM = N 0e~\ or N/ N0 = i/e =
= 1/2.7 = 0.37. Hence,

^ = 0 .3 7 ^ 0 and N° ~ N- X 100 = ^ 0~ ^ 37 N° x 100 = 63.0%


VI.4.6 Structure and Properties of Atomic Nuclei 539

Problem 2. The activity of the carbon isotope 6C14 in


ancient wooden articles is four-fifths of the activity of this
isotope in freshly felled trees. The half-life of the carbon
isotope 6C14 is 5570 years. Find the age of the ancient
articles.
Given: a '= 0.8a0, T = 5570 years.
Required: t.
Solution: The activity of a radioactive substance is the
number of nuclei decaying in a unit time: a = | AN |/At =
= XN. According to the law of radioactive decay N =
= N 0e - Xt, hence a = XN0e~u . At the initial moment of
time, the activity a0 = XN0. Consequently, a = aQe~u ,
where X = (In 2)/T is the decay constant, and T is the half-
life.
Hence, a/a0 = e~u , or In a/a0 = — Xt = (—t In 2)/T.
This gives us
— T In (a/a0) — 5570 In (0.8) — 5570 (— 0 .225) _
In 2 ]n 2 “ 0.693 ^
« 1 8 0 0 years

6. Experimental Methods
of Studying Particles
and Radioactive Radiation
1°. All the methods of detecting and studying the prop­
erties of radioactive radiations are based on the ionizing
and photochemical actions of the emitted particles and hard
light quanta, and also on the deviation of charged particles
in magnetic fields (111.4.5.2°).
2°. Scintillation counters are based on the ability of
particles striking a fluorescent screen to cause scintillations
(VI.2.1.2°). Each flash acts on the photocathode (111.3.7.2°)
of an electron multiplier (111.3.7.4°) and knocks out elec­
trons from it. The latter, passing through n cascades of the
multiplier, produce a current pulse at the output. The pulse
is then fed to an amplifier and actuates an electromechanical
pulse counter. A curve is obtained on the cathode-ray tube
(111.3.10.2°) that shows the intensity of separate pulses
which is proportional to the energy of a separately counted
540 Atomic and Nuclear Physics V1.4.6

particle. A scintillation counter registers the number of


particles and their distribution by energies.
3°. Devices functioning in the region of a self-sustained
gas discharge (111.3.5.1°) set up by collision ionization
(111.3.3.4°) are called counters operating in gas amplification
conditions.
A Geiger counter is usually a sealed glass tube with the
cathode Ca—a thin metal cylinder—adjoining its internal
walls (Fig. VI.4.4). A thin
6 wire stretched along the
IA 1 + - ii Scaling axis of the counter is the
i -------- T c ir c u it anode A. The counter is
connected to a scaling cir­
cuit. A negative potential
is fed to the cathode Ca,
Fig. VI.4.4 and a positive one to the
anode wire A. A signal in­
dicating that a particle has entered the counter is fed to
the output of the scaling circuit from the resistor R through
the capacitor C.
A particle entering the counter owing to ionization pro­
duces electrons and positive ions in it. The electrons travel
toward the anode and get into a field with a growing inten­
sity. The velocity of the electrons grows, and they set up
an avalanche of ions. The electrons that get onto the wire
lower its potential, and a current flows through the resistor
R. A voltage pulse is produced on the resistor, i.e. a signal
that is fed to the input of the scaling circuit and registers
the entrance of a particle into the counter. Together with
the registering of a particle, the conditions of gas amplifi­
cation and avalanche growth of the number of ions are ex­
tinguished in the counter. The high potential that was pre­
viously on the anode is switched over to the resistor, the
field intensity inside the counter diminishes, and the elec­
trons, having lost their velocity, stop producing ions.
4°. The functioning of a Wilson cloud chamber is based
on the fact that the ions produced by a charged particle
passing through the chamber become centres of condensa­
tion (11.5.2.1°) of a vapour. The chamber (Fig. VI.4.5) is
a cylindrical glass vessel 7 covered on top with the glass 2.
The vessel is covered at the bottom with a layer of black
V l.4.6 S tru c tu re and P ro p e rties of A to m ic N u clei 544

velvet or cloth (on the screen 3). A saturated vapour is


formed in the chamber. Rapid lowering of the piston 4
causes adiabatic expansion of th8 vapour (11.3.3.5°) and its
sharp cooling. The vapour becomes supercooled (supersatu­
rated). A charged particle passing through such a vapour
creates a chain of ions along its path. Liquid droplets form

F i g . V I . 4 .5

on these ions as on centres of condensation, and the particle


leaves a trail behind it that can be photographed.
If a Wilson cloud chamber is placed in a strong homo­
geneous magnetic field (the Wilson-Skobeltsyn method),
then the charged particles experience the action of the Lor-
entz force (111.4.5.1°), and their paths are deflected. The
radius of curvature of the path and the known velocity of
a particle permit us to find its specific charge (111.4.6.1°).
The velocity and energy of the particle are determined from
its specific charge and radius of curvature.
5°. In a bubble chamber, the trail of a passing particle
becomes visible in a superheated liquid (11.5.3.4°) that
begins to boil upon a sharp decrease in its pressure. The
ions formed along the path of a charged particle are the
centres of vaporization. The liquid in a bubble chamber has
542__________________ A to m ic and N u c le a r P h ysic s V l.4 .7

a density that is about a thousand times greater than that


of the vapour in a cloud chamber. This makes it possible
to register particles of greater energies that are retarded in
a bubble chamber over distances that are a thousand times
smaller than those in a cloud chamber. In the latter, a fast
particle is photographed over a small portion of its path.
In a bubble chamber, the trail of the particle corresponds to
a length of the path that is a thousand times greater than
that in a cloud chamber.
6°. The method of nuclear emulsions is based on the black­
ening of a photographic layer under the action of the fast
charged particles passing through the emulsion (V.5.6.20).
Nuclear emulsions are used in the form of layers from 0.5
to 1 mm thick. In conventional photographic plates and
films, the photolayer is from 10 to 20 p,m thick. Particles
with an energy of the order of 10 MeV form a non-vanishing
trail about 0.1 mm long that can be thoroughly studied for
any amount of time. A stack of plates including a large num­
ber of them is used to study the trails of particles of very
high energies producing long trails. The stack is placed at
an angle to the trail of a particle so as to extend it.

7. Origin of Alpha, Beta,


and Gamma Rays
1°. Two stages are distinguished in the process of an
alpha decay: the formation of a particle from two protons
and two neutrons in a nucleus, and the emission of the alpha
particle by the nucleus. The separation of the four nucleons
into an independent particle is facilitated by the saturation of
the nuclear forces (VI.4.3.2°, d). The formed alpha particle
is subjected to a smaller action of the nuclear forces of attrac­
tion (VI.4.3.2°, a). At the same time, Coulomb forces of
repulsion from the protons of the nucleus act on an alpha
particle more than on separate protons.
2°. The emission by a nucleus of an alpha particle is
a specific quantum-mechanical tunnelling effect. It con­
sists in the tunnelling or penetration of an alpha particle,
which has wave properties (VI.1.1.3°), through the nuclear
V I.4.7 S tru ctu re and P ro p e rties of A to m ic N u c l e i 543

potential barrier. A notion of this barrier can be obtained


from the following very rough reasoning. The alpha particle
and the other nucleons in a nucleus can be considered as
being within a potential well (VI.1.4.3°) in the region OA.
The well has a depth of i?Pi0 (Fig. VI.4.6). This signifies
that for any particle to leave the nucleus it ought to have an
energy of at least EVt0 to overcome the attraction of the
nuclear forces. This is customarily formulated in a concise
way as follows: “a potential barrier of a certain height and
width exists at the boundary of
the nucleus”. In Fig. VI.4.6, this EL
1
barrier is depicted in a simpli­
fied way as a rectangular barrier
AB of height 2?p<0 and “width” —

L. The alpha particle in the nu­


cleus has an energy E smaller A B
than the height of the potential
K L ^
barrier (see Fig. VI.4.6). The
particle, however, having wave F ig. V I .4 .6
properties, can penetrate or “tun­
nel” through the potential bar­
rier as shown by the arrow in Fig. VI.4.6. As a result,
the alpha particle will be outside the nucleus in the re­
gion where the nuclear forces ofj attraction no longer
act*. The tunnelling effect in the emission of an alpha
particle by a nucleus explains all the laws of alpha decay
in complete agreement with experimental data.**
3°. The beta decay of naturally radioactive nuclei can­
not be explained by the simple flying of electrons out of the
nucleus because there are no stable electrons in a nucleus.
In creating the modern ideas on the origin of beta particles,
Ilie main part was played by data on the energies of electrons
emitted by beta radioactive sources. Experiments showed
that beta particles have an exceedingly broad range of ener­
gies up to a certain maximum value i^p.max (the continuity

* I t is proved in q u an tu m m ech a n ics th a t the p r o b a b ility of the


h in n e llin g effect for th e sep arate n u cleon s in a n u cleu s th a t are not
com bined in to an alp h a p a r ticle is v a n ish in g ly sm a ll.
** In form ation on th ese la w s is beyond th e scope of the present
course.
544 A to m ic and N u c le a r P h ysic s V t.4 .7

of the energy spectrum of electrons in beta decay). Figure


VI.4.7 shows a curve of the distribution by energies of N
electrons emitted by nuclei. The continuous curve termi­
nates at the boundary i^p.max-
4°. Atomic nuclei are in definite energy states
(VI.4.2.4°). The loss by a nucleus of the energy associated
with the emission of a beta particle signifies the transition
of the nucleus from an energy state with a greater energy to
another state with a smaller energy. It is impossible to make
this agree with the fact
that an electron emitted
from an atomic nucleus can
have any energy value from
B oundary zero to -fi^max*
o ffi-sp e c tru m 5°. The difficulty in ex­
E, MeV plaining the energy of beta
particles is enhanced by
0 0 . 2 0.4 0 . 6 0 . 8 1 1.2
the difficulty with the val­
F ig. V I .4.7
ues of the spins of nuclei.
In beta decay, the mass
number of the nucleus does
not change. Consequently, the total spin of all the nu­
cleons in the nucleus (VI.4.1.4°) should not change. But
an electron having a spin of dzft/2 “carries away” its spin in
beta decay. The spin of the nucleus ought to change—instead
of an integral one (in units of K) it ought to be a half-integral
one (in units of h), or vice versa. This does not happen.
6°. Beta decay is explained by the transformation of a
neutron 0nx in a radioactive nucleus into a proton xp* (or
jH1) with the simultaneous formation of an electron ^ e0
and another particle, namely, an antineutrino 0v£*:
0n1^ 1p1+ _.ie° + ove(see also V I.4.10.3°)
Such a transformation does not occur in stable nuclei with­
out beta-radioactivity because of the interaction of the neu­
tron with other nucleons of the nucleus. The emission of a
beta particle—an electron—from a nucleus is attended by
* It is em p h asized in the sy m b o ls of a n eutron and a proton th a t
th eir m ass num bers e q u a l u n ity , w h ile th e ir charges eq u a l 0 and -\-iy
r e sp e c tiv e ly (in u n its of th e ele m en ta ry charge e ) . An a n tin e u tr in o
h as no rest m ass (V .4.1 0 .4 °) or e le ctr ic charge.
V1.4.H S tru ctu re and P ro p e rties of A to m ic N u c le i 545

the simultaneous emission of an antineutrino. The energy of


the emitted pair of particles is distributed among them in
different ways, but so that the siwn of the energies does not
exceed the upper boundary i?p,max (Par. 3°). This explains
the possibility of beta particles having different energies.
An antineutrino has a spin equal to ±h!2 (VI.2.8.2°).There­
fore, upon the simultaneous emission of an electron and an
antineutrino from a nucleus, their spins can be oriented
oppositely, and the total spin of the nucleus in beta decay
does not change.
7°. A neutron can transform into a proton not only in
a nucleus, but also when it is in the free state (the radioactiv­
ity of a free neutron). The rest mass of a neutron exceeds the
sum of the rest masses of a proton and an electron by Am =
— 0.837 X 10~3 amu (VII.7.1°). According to the mass-
energy relation (V.4.11.1°), an energy of AE = Ame2 =
-= 782 keV corresponds to this mass. Experiments show that
a free neutron is beta-radioactive. The half-life of free neu­
trons (VI.4.5.7°) is (1.01 db 0.03) X 103s. The electrons emit­
ted by free neutrons have diverse energies, the maximum
value /?p,max (Par. 3°) being 782 keV in accordance with the
above calculation.
8. Gamma decay, as a rule, is not an independent kind
of radioactivity. Gamma radiation accompanies alpha and
beta decays. The daughter nucleus (VI.4.5.1°) created in
an alpha or beta decay is usually excited (VI.4.2.4°). The
nucleus upon its transition to a normal or less excited energy
state emits a gamma photon in the same way as an atom when
passing over from an excited state to the normal one emits
a photon in the optical (VI.2.4.3°) or X-ray (V.3.6.10) ranges.
The great hardness of gamma quanta is explained by the
high values of the energies of gamma photons. The energy
difference Ai? between the energy levels of atomic nuclei
is about 0.1 MeV, whereas in atoms AE has values not ex-
reeding scores of electron-volts.

8. Nuclear Reactions
1°. Nuclear reactions are defined as artificial transforma-
Iions of atomic nuclei due to their interactions with various
particles or with one another. In the majority of cases, two
.1 :■ 0211
540 A to m ic and N u clea r P h ysics VI.4A

nuclei and two particles participate in a nuclear reaction.


One nucleus-particle pair is called the reactant pair {reactants),
and the other the product pair {products).
2°. The following symbols are used to write a nuclear
reaction:
A + a -> B -f b or A (a, b) B
where A and B = reactant ,ind product nuclei
a and h = reactant and product particles in the
reaction.
A nuclear reaction may sometimes occur ambiguously:
in addition to the reaction A - f a - > B + b, the reaction
A + a - > C - f c, and other reactions may occur. The pos­
sible paths of a nuclear reaction are called its channels (the
channels oj a nuclear reaction). The initial stage of a nuclear
reaction (A -f- a) is called its entrance channel.
3°. A nuclear reaction is characterized by the Q of the
reaction equal to the difference between the kinetic energies
of the product and the reactant pairs in the reaction. When
Q < 0, reactions proceed with the absorption of energy and
are called endothermic or endoergic. When Q > 0, reactions
proceed with the release of energy and are called exothermic
or exoergic. The latter kind of nuclear reaction has a great
practical significance (VI.4.11.2°).
4°. In all nuclear reactions, the laws of conservation of
the total electric charge and the number of nucleons are ob­
served. In addition, the laws of conservation of energy, mo­
mentum, and angular momentum are observed *.
5°. Nuclear reactions are classified according to:
(a) the energies of the particles causing the reactions;
(b) the kind of the particles participating in the reactions;
and
(c) the mass numbers of the nuclei participating in the
reactions (VI.4.1.2°).
6°. Nuclear reactions are distinguished with low, me­
dium, and high energies of the particles. Reactions with low
energies (of the order of an electron-volt) proceed mainly

* Som e other la w s of co n se r v a tio n sp ecific for n u clear p h y sics


are ob served in n u clear r eaction s. T heir c o n sid e r a tio n is beyon d the
scope of th e p resent course.
V t.4.8 S tru ctu re and P ro p e rties of A to m ic N u clei 547

with the participation of neutrons. Reactions with medium


energies (up to several MeV) also proceed under the action
of charged particles, gamma Quanta and cosmic rays
(VI.5.3.1°). Reactions with high energies lead to the decay
of the nuclei into their constituent nucleons and to the birth
of elementary particles (mesons, hyperons, etc.) (VI.5.2.1°).
7°. Nuclear reactions, apart from neutrons, are caused
by charged particles: protons (nuclei of ordinary hydrogen),
deutons (deuterons) (nuclei of heavy hydrogen-deuterium
tU2), alpha particles (nuclei of helium 2He4), and multi­
ply charged ions of heavy chemical elements. Naturally
radioactive chemical elements (VI.4.4.1°) and cosmic rays
may be sources of charged particles. Nuclear reactions may
also proceed under the action of gamma quanta—photonu-
clear reactions (the nuclear photoelectric effect).
8°. Depending on the mass numbers (VI.4.1.2°) of the
nuclei, there are distinguished reactions with light nuclei
(A < 5 0 ), reactions with medium nuclei (50 < A < 1 0 0 ),
and reactions with heavy nuclei {A > 100). Nuclear reac­
tions are very diverse as regards the nature of the nuclear
Transformations. (Some important examples are given on
following pages.)
9°. Nuclear reactions can occur either in one stage or
in two. In the latter case in the iirst stage of a reaction, the
incident particle remains in the nucleus—the target. The
energy of the particle is transferred to many nucleons of
the nucleus instead of to one. The capture by a nucleus of
an incident particle results in the formation of an intermedi­
ate nucleus {compound nucleus). The intermediate nucleus is
in the excited state. After a certain time elapses that is
greater than the characteristic nuclear time (VI.5.2.5°),
the energy in the nucleus again concentrates on one particle
and its emission from the nucleus follows—the second stage
of the nuclear reaction.
10°. Nuclear reactions induced by alpha particles were
Ihe first nuclear reactions confirming the possibility of trans­
forming chemical elements into other ones. Reactions of
this kind with the formation of protons occur according to
Ilie scheme:

zXA + 2He4 z+iYA+3 + (P1


548 A to m ic and N u c le a r P h y s ic s Vl.4.9

where X and Y are the chemical symbols of the reactant


and product nuclei, respectively. Historically, the first
nuclear reaction was that of the transformation of the nitro­
gen isotope 7N 14 into the oxygen isotope:

7N14+ 2He4 80 17+ 1P1

9. Interaction of Neutrons
with a Substance
1°. In nuclear reactions with light nuclei under the ac­
tion of alpha particles, the neutron was discovered. This
is a very important elementary particle contained in all
atomic nuclei except for that of conventional hydrogen
(VI.4.1.1°). A neutron was first obtained in the reaction of
transforming the beryllium isotope (4Be9) into the carbon
isotope (6C12):
4Be9+ 2He4 cC12+ on1

2°. The absence of an electric charge in a neutron makes


it easier for neutrons to penetrate into atomic nuclei (as
compared to charged particles). The nature of the interac­
tion of neutrons^ith nuclei differs for fast and slow neutrons.
Neutrons are called fast (fast neutrons) if their velocity v
is so great that the corresponding de Broglie wavelength
% = h/mv (VI. 1.1.3°) of the neutrons is much smaller than
the radius R of a nucleus, i.e. h/mu < R, or v > h/mR. The
energies of fast neutrons range from 0.1 to 50 MeV. U X ^ R ,
then the neutrons are called slow (slow neutrons). The ener­
gies of slow neutrons do not exceed 100 keV. Slow neutrons
with energies up to 0.5 eV are called thermal neutrons.
3°. The interaction of neutrons with nuclei consists
either in the scattering of neutrons on the nuclei or in the
capture of neutrons by the nuclei. In substances called mod­
erators (graphite, heavy water D20 , HDO, beryllium com­
pounds), fast neutrons are scattered on the nuclei, and their
energy transforms into the energy of thermal motion of the
moderator atoms. The neutrons, as a result, become thermal
ones. Their energies at room temperatures are about 0.025 eV.
vr.4.9 S tru ctu re and P ro p e rties of A to m ic N u clei
549

If the energies of thermal neutrons coincide with the


energy of a compound nucleus (VL.4.8.90), then the resonance
absorption of neutrons by nuclei (the resonance capture of neu­
trons) occurs. The capture of the neutrons results in artifi­
cial radioactivity of the nuclei of a substance (VI.4.10.1°)
and in nuclear fission (VI.4. 11.1°). ’
4°. The reactions of uranium nuclei with neutrons re­
sulted in the creation of chemical elements with atomic num­
bers Z exceeding 92. Such chemical elements are called trans­
uranium elements. In the resonance capture of a neutron bv
the most widespread uranium isotope 92U238, the radioactive
uranium isotope 92U239 is formed. It undergoes P_-decay
with a half-life of 23 minutes and transforms into the isotope
of the transuranium element neptunium 93Np239:
92U238+ oI1l 92U239- S i * 93NP239
to min
The nucleus of the neptunium isotope 93Np239 is P_-radio-
active with a half-life of 2.3 days and transforms into the
plutonium isotope 94Pu239:
„ l\J n 2 3 » .
93W P 2 .3 d ^
94Pu239

5°. The plutonium isotope ^Fu239 owing to its effective


fission under the action of thermal neutrons (VI.4.11.4°)
plays an important part in the production of nuclear energy.
The plutonium isotope' 94Pu239 is alpha-radioactive with a
half-life of 24 000 years and transforms into the stable ura­
nium isotope 92U235:
a
94Pu239 92 U 236
2 . 4 x 1 0 4 y c a rs

6°. The nuclear reaction of the uranium isotope 92U238


with a neutron may occur along a different channel
(VI.4.8.2°) and lead to the formation of the neptunium iso­
tope 93Np237, which is the parent of a radioactive familv
(VI.4.5.30): y
92IJ238 + oni _> B2u*» + 2#n‘, 02U237 91Np237
(1.8 days ' 1
The neptunium isotope 93Np237 is alpha-radioactive with
an enormous half-life: 2.21 x 10® years.
5 50 A tom ic and Nuclear Physics V I A .10

10., A rtificial R adioactivity


1°. Artificial radioactivity is defined as the radioactivity
of isotopes (VI.4.1.2°) produced as a result of nuclear reac­
tions (VI.4.8.1°). Artificial radioactivity is associated with
violation of the condition of stability of an atomic nucleus
(VI.4.3.5°).
2°. Light nuclei? (A < 50) in which a surplus number
of neutrons in comparison with the number of protons is
produced artificially, i.e. the condition in VI.4.3.5° is
violated, are (5..-radioactive. The symbol p_ signifies that we
are speaking of the emission of electrons by such nuclei. A
typical example is the transformation of the stable sodium
isotope n Na23 under the action of neutrons into the radio­
active sodium isotope nN a24. This isotope is P_-radioactive
and transforms into the stable magnesium isotope 12Mg24:
nNa24-*- i2Mg24 + - 1e° + 0v2 +V
The process occurs with the emission of an electron (_ie°),
an antineutrino (qV?) (VI.4.7.6°), and a gamma quantum.
3°. The stability of a stable nucleus is violated when
surplus protons are introduced into it. The energy of the
nucleus grows, the condition'5 of stability of the nucleus
(VI.4.3.5°) is violated, and artificial p^-radioactivity appears.
This is the name given to artificial radioactivity associated
with the emission of a positron (+ie°) from a nucleus. A pos­
itron has the same rest mass as an electron, a spin of fi/2, and
a positive elementary charge (111.1.1.3°).
A p+-radioactive decay occurs in a nucleus when a sur­
plus proton transforms into a neutron according to the scheme
1P1 — on1+ +ie0 + 0ve
A reaction of this kind is attended by the emission of a
neutrino 0v{!—an uncharged particle with a rest mass equal to
zero. The necessity for the existence of such a particle is
dictated by the same considerations that determine the foun­
dations of the* theory of p.-natural radioactive decay
(VI.4.7.3°-6°). The p+-radioactivity is possible from the
energy standpoint, although mv <Cmu. The proton 1p1 re­
ceives the energy needed for the reaction when it interacts
with other nucleons in the nucleus.
vr.4.11 S tru c tu re a n d P ro p e r tie s of A t o m i c N u c le i 551

Examples of artificial p+-radioactivity are the trans­


formation of the aluminium isotope i 3A127 into the radio­
active phosphorus isotope 15P30:
13AI27 + 2He4+ 15P30 + 0nS 15P30 14Si30 + +1e° + 0v°
and the formation of the radioactive nitrogen isotope 7N13
from the stable boron isotope 5B10:
5B10+ 2lle4 + 7N13+ 0nS 7N13-* cC12+ +te° + 0v°

11. F ission of H eavy N uclei


1°. Heavy nuclei overloaded with neutrons are not
stable (the instability of heavy nuclei). This is confirmed by
the lower unit binding energy of heavy nuclei in comparison
with that of medium nuclei (VI.4.2.2°).
By the fission of a nucleus is meant a nuclear reaction
of splitting of a heavy nucleus (for example uranium) ex­
cited by the capture of a neutron into two approximately
equal parts called the fission products (fragments). The
nucleons of the initial compound nucleus (VI.4.8.9°) are
distributed between the fission fragments in accordance with
the laws of conservation of electric charges and mass numbers.
The release of a small number of neutrons is possible here
(Par. 5°).
2°. The fission of a heavy nucleus into two fragments
is attended by the liberation of tremendous energy. The
energy liberated per nucleon in the process of fission of a
“friable” unstable nucleus equals the difference between the
unit binding energies in the product nuclei and the reactant
one, i.e. 8.7 - 7.6 = 1.1 MeV (VI.4.2.20). A nucleus of
Ilie uranium isotope 92U238 containing 238 nucleons liberates
energy of the order of 220 MeV in fission. Upon the fission of
Ihe nuclei contained in one gramme of the uranium isotope
„oU235, the amount of energy liberated is 8 X 1010 J, or
22 000 kWh.
3°. The main part of the fission energy is liberated in
(lie form of the kinetic energy of the fission fragments. At
Ihe distance r between the fragments exceeding the range
of action of nuclear forces (VI.4.3.2°, b), the potential ener-
A tom ic and Nuclear Physics V L 4 .1 t

gy Ev of repulsion of the charged nuclear fragments is Ev =


= ZlZ2e2l4ne0r, where Z^e and Z2e are the charges of these
nuclei. At the moment of completion of fission, r = R 1 -(-
4- /?2 ^ 2i?, where R 1 and R 2 are the radii of the nuclear
fragments equal to R = 1.4 X 10“15 A x/3 (VI.4.1.5°). Con­
sidering that Zx = Z2 = 92/2 = 46 and A 1 — A 2 = 238/2 =
= 119, we have Ev « 220 MeV. The potential energy Ev
of the fragments transforms into their kinetic energy, and
they fly away in different directions with tremendous vel­
ocities.
4°. The fission of some nuclei under the action of slow
neutrons proceeds more effectively than under the action
of fast ones. Thermal neutrons (VI.4 9.2°) cause the fission
of nuclei of the plutonium isotope 94Pu230 and the uranium
isotope 92U236. The energies needed for the fission of nuclei
of the uranium isotope 92U238, and also of the nuclei of the
thorium and protactinium isotopes existing in nature, are
considerably greater and are about 1 MeV.
5°. Heavy fissionable nuclei are overloaded with neu­
trons: for them N/Z ^ 1.6 (VI.4.1.1°). This signifies that
at the moment of formation of the fission fragments they are
also overloaded with neutrons. But in stable nuclear frag­
ments N/Z » 1. Consequently, in ihc fission of nuclei there
are surplus neutrons whose number equals the difference
between the number of neutrons in the initial nucleus and
their number in the nuclear fragments (fission neutrons).
The mean number n of fission neutrons per fission event
characterizes the process of neutron multiplication in the
fission of nuclei. For example, in the fission of nuclei of
the plutonium isotope 94Pu239 and the uranium isotope
92U236 under the action of thermal neutrons, the mean num­
ber n is 3.0 and 2.5, respectively.
6°. To carry out a nucleus fission reaction, a certain
amount of energy has to be spent. It is called the nuclear
fission activation energy (the fission threshold). A nucleus in
the form of a drop (VI.4.3.4°) has the greatest stability if
the sum of the surface energy contracting the drop (VI.4.3.4°)
and the electrostatic energy of repulsion of the protons of
the spherical drop-shaped nucleus is smallest. When a
drop-shaped nucleus captures a neutron (Fig. VI.4.8a), it
V 1. 4. 11 S tru c tu re a n d P r o p e r tie s of A t o m i c N u c te i 553

deforms and takes on the shape of an ellipsoid (Fig. VI.4.86).


Jn connection with the enormous density of the nuclear sub­
stance (VI.4.1.6°), the volume of the drop-shaped nucleus
does not change, but its surface area grows, as does the mag­
nitude of the surface energy of the nucleus. At the same time,
the electrostatic energy diminishes, because with a spherical
shape of the nucleus, its protons are closest to one another,

and the energy of their repulsion is greatest. The condition


for the stability of a drop-shaped nucleus is violated even
when its deformations are small. The nucleus—the charged
drop—begins to oscillate when it captures a neutron: it
alternately stretches out and contracts. At small deforma­
tions of the drop (Fig. VI.4.8c), the forces of surface tension
(VI.4.3.4°) do not permit the drop to reach the critical val­
ue of the deformation (Fig. VI.4.8d) at which fission sets
in (Fig. VI.4.8c). The intermediate states are connected
with the formation and extension of the “neck” in the drop
(Fig. VI.4.8c, d). At excitation energies of the nucleus small­
er than the fission activation energy, the deformation of
the drop-shaped nucleus does not reach the critical value,
(lie nucleus does not divide, and returns to its basic energy
slate, emitting a gamma photon.
7°. Upon the fission of a nucleus, according^ to the drop
model, the condition Z2/A > 17 must be observed, where
Z is the atomic number of the element, and A is its mass
number. The quantity Z2/A is called the fission parameter.
The above condition is observed for all nuclei beginning
with the silver isotope /l7Ag108, for which Z2!A ^ 20. Only
lor heavy nuclei, however, does the fission activation energy
(Par. 6°) make it possible to actually accomplish nuclear
fission under the action of neutrons. At the critical value of
the fission parameter of ZP/A « 49, the existence of a nucleus
554 A tom ic and Nuclear Physics VI.4.12

is impossible—the nucleus becomes unstable and divides


spontaneously. Spontaneous fission of nuclei at Z2/A < 4 9
is possible at the expense of a tunnelling effect similar to
that existing upon alpha decay (VI.4.7.2°) and was dis­
covered for the nuclei of the uranium isotope 92U238.

12. N uclear F ission Chain R eactions.


A N uclear Reactor
1°. If every fission neutron (VI.4.11.5°) interacts with
neighbouring nuclei of a fissionable substance and initiates
a fission reaction in them, in turn, then an avalanche growth
in the number of fissions occurs. Such a fission reaction is
called a chain reaction by analogy with the chain chemical
reactions whose products can again enter into reactions with
the reactants.
2°. The condition for the initiation of a chain reaction
is the presence of multiplication of the number of neutrons
(VI.4.11.5°) upon the fission of a nucleus. The neutron mul­
tiplication factor k is the ratio of the number of neutrons
appearing in a certain stage of a reaction to the number of
such neutrons in the preceding stage. An essential condition
for the development of a chain reaction is the requirement
that k ^ 1.
3°. Not all of the fission neutrons cause the development
of a chain reaction. Part of them get into the nuclei of atoms
of non-fissionable substances present in the active section
{reactor core)—the region of space where the chain reaction
occurs. Such substances include neutron moderators
(VI.4.9.3°), and the heat-transfer agents carrying heat away
from the reactor core. A part of the neutrons simply leave
the limits of the core and cannot cause the development of
the chain reaction.
4°. Apart from what has been indicated in Par. 3°, the
development of a chain reaction depends on the mean number
n of neutrons in one fission event (VI.4.11.5°), on the di­
mensions of the reactor core, and on the processes of inter­
action of the*' neutrons with the nuclei of the fissionable
substances and those of the admixtures.
V I .4 .1 2 Structure and P ropertie s of A to m ic N uclei 555

5°. A reduction in the dimensions of the reactor core


increases the fraction of neutrons leaving the core and dim­
inishes the possibility of the development of a chain reac­
tion. The losses of neutrons are proportional to the surface
area S, while the multiplication of the neutrons is proportion­
al to the mass of the fissionable substance and, consequent­
ly, to its volume V. For a fissionable substance of a spherical
shape, S' is proportional to Z?2, V to Z?3, and S/V to 1/Z?. A
decrease in Z?, i.e. a decrease in the volume and mass of the
fissionable substance, is attended by a growth in the frac­
tion of the neutrons lost by their leaving the core. The min­
imum dimensions of the core at which k ^ 1 (Par. 2°) are
called the critical dimensions. The minimum mass of fis­
sionable substances in a core of critical dimensions is called
the critical mass.
To reduce their critical dimensions and critical mass,
fissionable substances are surrounded with neutron reflectors
(mirrors)—layers of a non-fissionable substance that does not
capture neutrons, but returns to the reactor core a greater
part of the neutrons flying out from it. The same substances
that are used as neutron moderators (VI.4.9.3°) serve as
reflectors.
A chain reaction is controlled by changing the rate v
of the chain reaction, i.e. the number of events of nuclear
fission in a substance during a unit time. Apart from the
neutron multiplication factor, the rate of a chain reaction
depends on the mean time r between two consecutive fission
('vents (the mean lifetime of one “generation” of neutrons).
6°. Controllable nuclear chain reactions are conducted
in devices called nuclear reactors (atomic piles or reactors).
The main elements of a nuclear reactor are the nuclear fuel,
the neutron moderator and reflector, a heat-transfer agent
Tor removing the heat formed in the reactor, and control
rods for controlling the rate of development of the chain
fission reaction.
7°. The uranium isotopes 92U236 and 92U238, plutonium
r,Pu239, and thorium 90Th232 are the nuclear fuel (breeding
mid fissionable substances in reactors). A natural mixture
of uranium isotopes contains 140 times more 92U238 than
,,,IJ235.
The neutron moderators and reflectors (Par. 5°) are ca-
556 Atomic and Nuclear Physics VIA.12

pable of increasing the number of slow neutrons (VI.4.9.2°),


which are most effective for the development of a chain
fission reaction.
The rapid development of a chain reaction is attended
by the liberation of a large amount of heat and overheating
of the reactor. To maintain stationary conditions of the
reactor in which the neutron multiplication factor k = 1
(VI.4.12.2°) (the critical conditions of the reactor), control
rods are inserted into the reactor core. The rods are made
of materials greatly absorbing thermal neutrons, for exam­
ple of boron or cadmium.
Water, liquid sodium, and other substances are used as
the heat-transfer agent in a reactor. Special measures are
Nuclear fuel and moderator

/ Radiation shielding
Reflector
Fig. V I.4.9

used to protect the reactor attendants from the action of the


neutron streams and gamma radiation produced in a reac­
tor (VI.4.14.1°). They include shielding means and auto­
mation of the reactor control processes. The design of a nu­
clear reactor is shown schematically in Fig. VI.4.9.
8°. The conversion of nuclear fuel takes place in nuclear
reactors using fast neutrons. The capture of neutrons by
nuclei of the uranium isolope 92U238 leads to the creation
of (he plutonium isotope fi4Pu239 (VI.4.9.4°) that can be
chemically separated from 92U238. An average of 2.5 neu­
trons are formed (VI.4.11.5°) upon the fission of one nucleus
of 92U23B, of which only one is needed to sustain the chain
VJ.4.13 Structure and Properties of Atomic Nuclei

reaction. The remaining 1.5 neutrons can be captured by


nuclei of 92U238 and produce 1.5 nuclei of 94Pu239. The con­
version ratio (regeneration factor)»in special breeder (regen­
erative) reactors exceeds unity. An alloy of uranium en­
riched with the isotope 92U235 and a heavy metal (bismuth
or lead) with a low neutron absorption is placed into the
core of a breeder reactor. Such reactors have no moderators.
The reactor is controlled by automatically shifting the
reflector or changing the mass of the fissionable substances.

13. A p p lic a tio n s o f N u c le a r E n e r g y


an d R a d io a c tiv e I s o to p e s
1°. The energy released in the nuclear chain reactions
of fission of heavy nuclei is called nuclear energy*. Nuclear
energy is utilized for peaceful purposes in electrical gen­
erating stations. The power of these stations is determined
by the power of their nuclear reactors. Reactors of a suf­
ficient power are used as power plants for prime movers on
vessels and submarines. The energy of nuclear electrical
generating stations can be used for the demineralization of
sea water. Calculations show that the cost of the deminer­
alized water produced in this way will be so low that it
will be possible to use it for irrigating arid land.
2°. The development of reactor building and the con-
si antly growing power of nuclear reactors make it possible
lo expand the production of radioactive isotopes of various
chemical elements (VI.4.10.2°, 3°). Radioactive isotopes do
not differ in their chemical properties from the stable iso­
lopes of the same chemical elements. This determines the
possibility of the practical use of radioactive isotopes.
The following properties of radioactive isotopes underly
Iheir practical use in science and engineering:
(a) the high energy of p_, (3+, and gamma radiations,
which makes it possible to detect very small amounts of
radioactive substances;

* The frequently used term “atomic energy” is not correct. Atomic


1'iicrgy is associated with the processes occurring in the electron shells
i*l a loins, molecules, and their ensembles.
558 Atomic and Nuclear Physics V l.4 .1 3

(b) the independence of the radiation of an isotope from


the conditions in which its atoms are found (the external
conditions, the nature of the chemical compound, its state
of aggregation, etc.); and
(c) the different penetrating ability of beta and gamma
radiations and the specific nature of their interaction with
a substance.
3°. Radioactive isotopes are employed as sources of
high-energy particles and radioactive tracers. These isotopes
are used as sources of high-energy particles for the dosed irra­
diation of various substances in order to initiate preplanned
changes in their structure, and also in the states of their
atoms and molecules. This underlies the methods of ra­
diation chemistry and radiation biology.
In radiation chemistry:
(a) substances are produced with the maximum resistance
to destruction caused by radioactive radiations (radiobio­
logical protection, the protection of reactor walls, heat-
transfer agents and lubricating materials functioning under
the action of radiation, etc.); and
(b) new materials are created with valuable preplanned
properties (for example, new polymers).
Radiation biology studies the changes in living organ­
isms due to the action of radioactive radiations (the hered­
itary properties of animals and plants, the change in hered­
ity, etc.). High-energy particles emitted by radioactive
isotopes are used in medicine for the diagnosis and treat­
ment of certain malignant tumours and other diseases.
4°. The use of radioactive isotopes as tracers is based
on the property of the atoms of radioactive isotopes to indi­
cate their presence by the emission of radiation (the method
of tracer atoms). The investigation of a substance having an
admixture of a radioactive tracer makes it possible to solve
very diverse problems: control of the course of production
processes, determination of the content of very small
amounts of a substance, determination of the age of geological
objects and archeological artifacts, etc.
5°. An atomic bomb is a specific reactor on fast neutrons
in which a rapid uncontrolled chain reaction takes place
with a large neutron multiplication factor (VI.4.12.2°).
The pure fissionable isotopes 92U235, 94Pu239, and 92U 233 are
V I.4 J 4 Structure and Properties of Atomic Nuclei 55(J

the nuclear explosive in an atomic bomb. A rapid explosive


chain reaction occurs on fast neutrons without moderators
at definite, relatively small, dimensions and mass of the
device. The critical mass (VI.4.12.5°) is from 10 to 20 kg
and at a density of the substance of p = 18.7 glcm3 occupies
the volume of a sphere of a radius equal to 4-6 cm. Initially,
the explosive is in a state excluding the rapid development
of a chain reaction. The substance is transferred to condi­
tions in which an uncontrolled chain reaction occurs with
the maximum speed. For this purpose, for example, the
nuclear charge of a bomb is first divided into two parts in
each of which a chain reaction is impossible. To initiate an
explosion, one of the halves of the charge is shot into the
other, and an explosive chain reaction occurs almost instan­
taneously when they combine.
6 °. An explosive nuclear reaction results in the release
of tremendous energy, and a temperature of about 108 kel-
vins is reached. An enormous increase in pressure occurs,
and a powerful shock wave is formed. The colossal number
of fission fragments containing radioactive isotopes, in­
cluding long-lived ones, is a great danger to living organisms.

14*. B io lo g ic a l A c t io n o f R a d io a c tiv e
R a d ia tio n s *
1°. Radioactive radiations have a strong action on a
substance, especially on live tissues. The harmful action of
radioactive radiations on an organism is associated with
the formation of free chemical radicals and with mutations
in cells. The latter may affect reproduction, and also result
in radiation sickness and in the formation of malignant tu­
mours.
The biological action of radioactive radiations is assessed
by means of special quantities.
The radiation dose D is defined as the ratio of the radia­
tion energy to the mass of the irradiated substance. The
unit of the dose is the gray (Gy) equal to one joule per kilo-

* The units of the quantities characterizing the action of radia-


lions are given in this section in connection with its special nature#
5C0 Atomic and Nuclear Physics Vl.4.14

gram (J/kg)—the radiation dose at which one joule of ioniz­


ing radiation energy is transmitted to a mass of the irra­
diated substance of one kilogram.
A non-system unit of the dose is the rad (radiation
absorbed dose):
1 rad = 10"2 J/kg
2°. The power, or rate, of the radiation dose is the dose
related to a unit time:

Its unit is gray per second (Gy/s) or watt per kilogram (W/kg).
3°. The exposure dose D e is an energy characteristic of
radiation describing the latter according to the effect of
ionization of dry atmospheric air. Its unit is the coulomb
per kilogram (C/kg)—the exposure dose to photon, X-ray
or gamma radiation at which the sum of the electric charges
of the ions of one sign produced by the electrons released
in irradiated air with a mass of one kilogram upon the com­
plete utilization of the ionizing ability equals one coulomb.
4°. A non-system unit of the exposure dose is the roent-
gen (r):
1 r = 2.58 x 10~4 C/kg
A roentgen corresponds to the exposure dose at which a total
charge of ions of one sign equal to one absolute electrostatic
unit of charge is produced in one cubic centimetre of dry
air at standard atmospheric pressure.
5°. The power, or rate, of the exposure dose P e = D j t
is measured in amperes per kilogram (A/kg). It is the rate
of exposure to photon radiation at which the exposure dose
grows by 1 C/kg per second.
Non-system units of the rate of the exposure dose are:
1 r/s = 2.58 x 10“4 A/kg, \ r/min = 4.30 X 10“6 A/kg
1 r/h = 7.17 x 10“8 A/kg
6 °. The equivalent radiation dose is assessed according
to the biological action of the radiation. The unit is the gray
(J/kg).
V I .4 .1 5 Structure and Properties of Atomic Nuclei 561

A non-system unit of the equivalent radiation dose is the


roentgen-equivalent-man (rem) unit. This is the name given
lo the absorbed radiation energy ^biologically equivalent
to one roentgen:
1 rem = 10“2 J/kg
The biological action, apart from the equivalent dose,
depends on the energy of separate particles. Thus, identical
doses of gamma radiation, X-rays, and neutrons are not dan­
gerous to the same degree.
7°. A safe dose for the human organism is considered
one that is about 250 times greater than that produced by
Ilie cosmic background (VI.5.3,1°) and radioactive radia­
tions from the interior of the Earth. A single radiation expo­
sure exceeding 500 roentgens is considered to be dangerous
for a person. In passive treatment (without transplantation
of marrow, etc.), such an exposure gives a fatality rate of
:»() per cent.

15*. F u s io n R e a c tio n s
1°. Fusion (or thermonuclear) reactions are exoergic nuclear
reactions (VI.4.8.3°) of synthesis of light nuclei into heavier
(»nes. Fusion reactions proceed effectively at ultrahigh
lernperatures of the order of 107 to 109 K. A very high energy
is released in fusion reactions. It exceeds the energy liberat­
ed in the fusion of heavy nuclei (VI.4.11.2°). For example,
in the reaction of fusion of nuclei of deuterium XD 2 and tri­
tium XT3 (VI.4.1.2°) into a helium nucleus 2He4:
1D2 + 1T3 - > 2He4 + 0n 1
mii energy of about 3.5 MeV per nucleon is released. In
lission reactions, the energy per nucleon is about 1 MeV.
In the synthesis of a helium nucleus from four protons:
4 1p1- * 2He4 + 2+1e°
where +1e° is the symbol of a positron (VI.4.10.3°), a still
higher energy is released. It equals 6.7 MeV per particle.
The “profitability” of fusion reactions from the energy view­
point is explained by the fact that the unit binding energy
.1(1 01! 11
562 Atomic and Nuclear Physics VL4.15

of a helium nucleus considerably exceeds the unit binding


energy of the hydrogen isotope nuclei (VI.4.2.2°, Fig. VI.4.1).
2°. For the fusion of light nuclei, the potential barrier
(VI.4.7.2°) due to the Coulomb repulsion of the protons in
the nuclei with like positive charges has to be surmounted.
For the fusion of deuterium nuclei 1D21 they have to be
brought near one another to a distance of r « 3 X 10“15 m.
For this purpose, work must be done equal to the electrostat­
ic potential energy of repulsion Ev = e2/4ne0r « 0 . 1 MeV.
Deuteron nuclei can surmount such a barrier if upon col­
liding their mean kinetic energy 3 kT /2 (11.2.4.4°) equals
0 . 1 MeV. This is possible when T = 2 X 109 K. In practice
for a number of reasons (which our reader must take for
granted because they are beyond the scope of our course),
the temperature needed for the proceeding of fusion reactions
lowers by two orders and is 107 K.
3°. A temperature of the order of 107 K is characteristic
of the central region of the Sun. Spectral analysis of the
Sun’s radiation has shown that the substance of the Sun,
as of many other stars, contains up to 80 % of hydrogen and
about 20% of helium. The quantity of carbon, nitrogen,
and oxygen does not exceed one per cent of the mass of stars.
In view of the enormous mass of the Sun (about 2 X 1027 kg),
the amount of these gases is sufficiently great.
4°. Fusion reactions proceed on the Sun and the stars
and are the source of energy making up for their radiation.
Every second, our Sun emits an energy of 3.8 X 1026 J,
which corresponds to a reduction in its mass by 4.3 million
tonnes (V. 4.11.1°). The specific release of energy by the Sun,
1. e. the energy released per unit of mass of the Sun a second,
is 1.9 X 10"4 J/s-kg. It is very small and is only one per
cent of the specific liberation of energy in a living organism
in the process of metabolism. The radiated power of our
Sun has not virtually changed during the several thousands
of millions of years from the birth of the solar system.
5°. Fusion reactions on the Sun proceed according to
the carbon-nitrogen cycle in which the fusion of hydrogen
nuclei into a nucleus of helium is facilitated in the presence
of nuclei of the carbon isotope 6C12 playing the part of
catalysts (a catalyst is a substance that accelerates chemi­
cal reactions). At the beginning of the cycle, a fast proton
VI.4.15 Structure and Properties of Atomic Nuclei 563
penetrates into a nucleus of the carbon isotope 6C12 and
forms an unstable radioactive nucleus of the nitrogen isotope
7N 13 (VI.4.10.3°) with the emission#of a gamma quantum:
eC12 + 1P1 7N 18 + y
The transformation (VI.4.10.3°) ip 1 on1 4 - +1e° + 0vJ! takes
place in the nucleus of 7N 13 with a half-life of 14 minutes,
and a nucleus of the carbon isotope 6C13 is formed:
7N 13- > 6C13 + +1eo + 0v°e
In about 2.7 million years, the nucleus of 6C13 captures a
proton and forms a nucleus of the stable nitrogen isotope 7N14:
6C13 + 1P1— 7N 14 + y
After, on the average, 32 million years, the nucleus of 7N 14
captures a proton and transforms into a nucleus of the oxygen
isotope 80 16:
7N 144- iP ^ s O ^ + v
The unstable nucleus of 80 16 with a half-life of three minutes
emits a positron and a neutrino and transforms into a nuc­
leus of the nitrogen isotope 7N16:
80 16 7N 15 -f- +1e° + ove
'The cycle is completed by the nucleus of 7N1B absorbing
a proton and decaying into a nucleus of the carbon isotope
,,C12 and an alpha particle. This occurs about every
100 000 years:
7N« + 1p i ^ 8C*2+ 2He4
A new cycle again begins with the absorption by 6C12
of a proton, which occurs on the average every 13 million
years. The separate reactions of the cycle are divided in
lime by intervals that are immeasurably great according
to the Earth’s scales of time. The cycle, however, is a closed
one and occurs continuously. Therefore, the different reac-
lions of the cycle occur on the Sun simultaneously, having
Mailed at different moments of time.
6 °. As a result of one cycle, four protons fuse into a he­
lium nucleus with the appearance of two positrons and gamma
radiation. We must add to this the radiation appearing
,iu*
564 Atomic and Nuclear Physics V I .4 .1 5

when the positrons combine with the electrons of the plasma


(VI.5.5.5°). In the formation of one gramme-atom of helium,
700 000 kWh of energy is released. This amount of energy
compensates for the losses of the Sun’s energy due to radia­
tion. Calculations show that the amount of hydrogen in
the Sun will be enough to maintain fusion reactions and
the Sun’s radiation for thousands of millions of years.
7°. The carrying out of fusion reactions in the Earth’s
conditions will create enormous possibilities for producing
energy. For example, when the deuterium contained in one
litre of conventional water is used in a fusion reaction,
the energy released will be the same as that liberated in
the combustion of about 350 litres of petrol.
Conditions close to those realized in the Sun’s interior
were produced in a hydrogen bomb. A self-sustaining fusion
reaction of an explosive nature occurs in it. The explosive
substance is a mixture of deuterium XD 2 and tritium xT3.
The high temperature needed for proceeding of the reaction
is obtained by exploding a “conventional” atomic bomb
placed inside the hydrogen one.
8°. Studying of the reactions in high-temperature deu­
terium plasma (111.3.6.1°) is the theoretical foundation for
obtaining controlled fusion reactions. The principal
difficulty is the maintaining of conditions needed for a self-
sustaining fusion reaction. It is essential for such a reaction
that the rate of liberation of energy in the system where
the reaction is proceeding be not lower than the rate of with­
drawal of energy from the system. At temperatures of the
order of 108 K, fusion reactions in a deuterium plasma have
an appreciable intensity and are attended by the release
of high energy. The power released in a unit volume of plas­
ma upon the fusion of deuterium nuclei is 3 kW/m3. At a
temperature of the order of 106 K, the power is only
10- 17 W/m3.
9°. The energy losses in high-temperature plasma are
mainly associated with the escaping of heat through the
walls of the apparatus. The plasma must be thermally insu­
lated from the walls. Strong magnetic fields are used for
this purpose (the magnetic thermal insulation of plasma).
If a strong electric current is passed through a column of
plasma in the direction of its axis, then forces are set up in
VI.4.16 Structure and Properties of Atomic Nuclei 565

the magnetic field of the current that compress the plasma


into a pinch detached from the jvalls. Keeping the pinch
detached from the walls and combatting various instabili­
ties of the plasma are very intricate problems. Their solu­
tion should lead to the practical carrying out of controlled
fusion reactions.

16. Accelerators
1°. Devices for obtaining charged particles with a very
high kinetic energy are called accelerators. The following
methods of accelerating particles are distinguished: direct,
induction, and resonance. With respect to the shape of the
particle trajectories, accelerators
are classified as linear and cir- if),
cular-orbit ones. In the former,
the particle trajectories are close
to straight lines. In the latter,
the trajectories are circles or
spirals.
2°. In linear accelerators, a
particle passes once through an
iPf pPl -'
electric field of high potential
difference ((p2 — cpj) and acquires
a high kinetic energy mvV2
equal (111.1.8.5°) to <7 (qp2 — cpi),
where q is the absolute value M
of the charge of a particle.
3°. A betatron is an induc­
tion accelerator of electrons. Its
design is based on the phenome­
non of the appearance oFan eddy Fig. VI.4.10
electric field in space under the
Influence of a varying magnet­
ic field (TTI.5.3.3°). Figure VI.4.10 is a schematic view
of a betatron. An evacuated annular accelerator cham-
tier n having the shape of'a closed ring is placed between
Ilie pole shoes A and C of a strong electromagnet. The axis
»f tiie chamber coincides with the axis of symmetry 0 0 '
»f the electromagnet pole shoes. A change in the current ip
566 Atomic and Nuclear Physics VI.4.16

the electromagnet winding produces a change in the magnetic


field between the poles of the electromagnet and leads to
the setting up of an eddy electric field. The magnetic field
is symmetrical relative to the axis 0 0 '. The field lines of
the eddy electric field (111.1.3.5°) in the plane M N per­
pendicular to the axis 0 0 ' and passing through the middle
of the gap between the poles have the form of circles with
their centres at the point K . The magnitude of the electric
field intensity E at all the points of each circle is the same.
4°. If an electron is introduced into the chamber D
so that its velocity v is directed along a tangent to a circle—a
line of the electric field—then an electric force (—eE) acts
on the electron along a tangent to the field line in a direction
opposite to that of the vector E. During n revolutions around
a circle of radius r, the electron acquires a kinetic energy
equal to 2nreEn. Usually, an electron before leaving the
chamber D of a betatron covers a path with a total length
of thousands of kilometres. Special measures are taken in
a betatron to constantly keep the electron on one orbit in
the plane indicated in Par. 3°.*
Example. The uniform change in the magnetic field is
such that an electron performing one revolution around a
circle of radius r = 0.4 m acquires an energy of 20 eV.
Hence during 8.45 X 10"3 s the electron will cover a path
of 2520 km, complete 106 revolutions, and acquire an energy
of 20 MeV.
5°. Resonance circular-orbit accelerators are employed
for the acceleration of protons, deuterons, and multiply
charged ions of atoms of various chemical elements. The
particle being accelerated passes repeatedly through a
varying electric field along a closed trajectory, increasing
its energy each time. A strong transverse magnetic field
is used to control the motion of the particles and periodi­
cally return them into the accelerating electric field. A cyclo­
tron, shown schematically in Fig. VI.4.11, is the simplest
resonance accelerator. A cyclotron consists of two metal
dees M and N that are the two halves of a low thin-walled

* The conditions for the stability of the electron orbit are not
cpnsidered in the present course,
V I.4.16 Structure and Properties of Atomic Nuclei 567

cylindrical chamber divided by the narrow slot D. The dees


are confined in the closed evacuated chamber A placed be­
tween the poles of a strong electromagnet. The magnetic
field induction is directed at right angles to the plane of the
drawing. The electrodes m and n connect the dees to an elec-
trie generator that produces an alternating electric field
in the slot D.
6°. A charged particle introduced into the slot D is accel­
erated by the electric field. Assume, for example, that a
positive ion is introduced at the point C when the electric
field is maximum, and its
intensity is directed upward.
The ion will move upward in
(lie slot with uniform acceler­
ation under the action of the
electric field. There is no elec­
tric field in the metal dees M
and N. Inside the dee M , the
ion under the action of the
magnetic field will describe a
semicircle of a definite radius
(111.4.5.2°). At the moment
when the ion approaches the
slot at the point E, the direc- Fig. V I . 4 . H
lion of the electric field inten­
sity will reverse, and the ion
in the gap will be accelerated, moving downward, into t“e
<lee N. In the latter, the ion will describe a semicircle already
of a greater radius corresponding to its higher velocity. At
the moment when the ion leaves the dee N at the point tr,
t h e electric field intensity again reverses and accelerates
I Ik* ion upward. As a result of repeated acceleration of the
i n n , it will travel along an unwinding spiral, and the electric
field will impart a high kinetic energy to it.
7°. For the continuous acceleration of a particle in a
cyclotron, the motion of the particle in the magnetic field
nml the change in the electric field in the slot must be syn­
chronized. The condition for synchronization in a cyclotron
Is

To = T
568 Atomic and Nuclear Physics V1.5.1

where T$ = period of oscillations of the electric field


T = period of revolution of the particle in the mag­
netic field not depending on the velocity of
the particle and the radius of the circle
(111.4.5.3°).
The quantity T is determined by the equation
T _ 2ji m
~~ B q
where B = magnetic field induction
qlm = specific charge of a particle (111.4.6.1°).
Modern accelerators of the resonance type with the aid
of specially selected laws of the change in the magnetic
and electric fields (synchrophasotrons) make it possible
to reach a very high energy of the charged particles being
accelerated. For example, the Serpukhov accelerator in the
USSR imparts an energy of 76 GeV to protons.

Chapter 5
Elementary Particles

1. General
1°. Elementary particles are defined as particles which
at the present level of development of physics cannot be
considered as a combination of other, “simpler” particles.
An elementary particle upon interaction with other particles
or fields behaves as an integral whole. Modern physics of
elementary particles deals with the mutual transformations
of the particles, their properties manifesting themselves
when the particles interact with a substance, and the struc­
ture of the elementary particles.
2°. The question of the structure of elementary particles
is considered differently depending on the energy of these
particles. At low energies E of the particles, their structure
does not affect the results of physical phenomena caused by
them, and they are considered to be structureless (struc-
V1.5.1 Elementary Particles 569

tureless particles). The condition for such treatment is the in­


equality E < 2ra0c2, where m0c2 = is the rest energy of
a particle (V.4.11.2°). In this region of energies, elementary
particles are considered as point particles having definite
properties: a rest mass, an electric charge, spin, a lifetime,
and other properties, a number of which are not considered
in an elementary course of physics. For example, an electron
can be considered as a point particle when the intensity of
the electric field set up by it is being calculated for a point
far from it (111.1.4.1°).
Some elementary particles are deprived of certain prop­
erties. For example, neutrons have no electric charge,
neutrinos and antineutrinos have no rest mass.
3°. The notion of a point particle agrees with the requi­
rements of the theory of relativity. If an elementary particle
has finite dimensions, is an extended one, then it should not
deform because it is an integral whole. Deformation is asso­
ciated with the possibility of independent movements of
the parts of an integral whole. As applied to an elementary
particle, this should signify that external action on an integ­
ral particle should be instantaneously transmitted from
some of its parts to others. This contradicts the requirement
of the special theory of relativity on the finite speed of pro­
pagation of interactions (V.4.4.40).
4°. The notion of point elementary particles results in
serious difficulties with the energy of the particles. For exam­
ple, a stationary charged particle of charge e sets up an elec­
trostatic field of potential (p = ^/4ite0r, where r is the distance
from the particle (111.1.8.8°). A point structureless particle
sets up a field of an infinite potential at the point where it is
found ((p — oo when r -*■ 0). But in this case the potential
energy Ev of the particle equal to eq> (111.1.8.1°) also becomes
infinite, which contradicts experimental facts.
5°. The presence of a definite length of elementary par­
ticles and their having a certain structure tell in the region
of very high energies E for which the condition E 2m0c2
(Par. 2°) holds. For the structure of a nucleon, see VI.5.6.1°.
6°. In classical, prerelativistic physics, the concept
of the classical radius of an electron was introduced. If an
electron is treated as a sphere of radius rcl and it is consid­
ered that the charge e of the electron is distributed ovpr the
570 Atomic and Nuclear Physics V1.5.2

surface of the sphere, then the radius rc1 is determined from


the condition that the potential energy of the charged sphere
(111.1.8.9°) is its rest energy (V.4.11.20): e2/4jie0rcl = m0c2.
Hence, the classical radius of an electron rC] =
= e2/4n?0m0c2 = 2.8 X 10"15 m.

2. Classification of Elementary Particles


and Their Interactions
1°. There are several groups of elementary particles
differing in their properties and the nature of their inter­
actions with one another. Some of the most important prop­
erties of elementary particles are given in Table VI.5.1.
It contains only the properties of particles that can be con­
sidered at the level of an elementary course in physics.
The rest masses of the particles are given in energy units
(MeV). This is customary practice in nuclear physics, on
the basis of the mass-energy relation (V.4.11.1°). To find
out how many electron rest masses me the mass of a particle
equals, we have to multiply the mass expressed in MeV by
two because me « 0 . 5 MeV.
The classification of elementary particles is based on the
difference in their rest masses. According to the values of the
rest masses, there are distinguished leptons (light particles),
mesons (intermediate particles), and baryons (heavy parti­
cles). The latter, in turn, are divided into nucleons and hy~
perons (Table VI.5.1).
2°. According to the sign of the electric charge measured
in units of the elementary charge e (111.1.1.3°), there are
distinguished positive and negative particles carrying a
charge, and electrically neutral particles. Most of the ele­
mentary particles listed in Table VI.5.1 have a spin equal
to 1/2 (in units of h). Particles having no spin, however, also
exist (pions and kaons). As regards their lifetimes, stable
and unstable elementary particles are distinguished. The
unstable particles are characterized by their lifetime. The
photon (V.5.1.20) is placed in a special group: it is not asso­
ciated with the structure of a substance, has no rest mass,
and its spin equals h.
3°. Apart from the elementary particles listed in
Table VI.5.1, a great number of particles exists that are
V I .5 .2 Elementary Particles 571

Table VI.5.1
is--------
Electric Mass, Lifetime, Spin,
Name Symbol charge, e MeV s h

Photon Y 0 0 Stable 1

Leptons
Neutrinos
Electron neutrino, ve ve 0 0 0 Stable 1/2
antineutrino
p-Meson neutrino, 0 0 0 Stable 1/2
antineutrino
Electrons
Electron, positron e~ e+ -1 +1 0.511 Stable 1/2

Muons
pr-Meson, ^ -m e­ -1 +1 106 2.2 X 10-6 1/2
son

Mesons
Pions
a+

ji+-Meson, Ji“-me­
ai

+1 -1 140 2.6X 10-8 0


son
:x0-Meson JC° 0 135 0.8 X lO -16 0
Kaons
K+-Meson, K"-me- K+ K" +1 -1 494 1.2X10-8 0
son ^,
K°-Meson, anti- K° K° 0 0 498 K®: 0.86 X lO -1® 0
K°-meson K°: 5 .3 8 X 10-8 0
r]°-Meson T]° 0 549 2.4 X 10-19 0

Baryons
Nucleons
Proton, anti proton PP +1 -1 938.2 Stable 1/2
Neutron, anti­ nn 0 0 939.6 0.93 X 103 1/ 2
neutron
572 Atomic and Nuclear Physics V I.5.2

Table VI .5.1 (concluded)

Electric Mass, Lifetime, Spin,


Name Symbol charge, e MeV s h

Hyperons
A°-Hyperon, A° A° 0 0 1116 2 .5 X IQ"10 1/2
anti-A°-hyperon
2 +-Hyperon, 2 + 2" +1 -1 1189 0 .8 X 1 0 - 10 1/2
anti-2+-hyperon M

+
2"-Hyperon,
l
-1 +1 1197 1 .5 X 1 0 - 10 1/2
anti-2~-hyperon
2°-Hyperon, 2° 2° 0 0 1192 < lO"14 1/2
anti-2°-hyperon
S “-Hyperon, 3“ 3+ -1 +1 1321 1 .7 X 10"10 1/2
anti-S"-hyperon
S°-Hyperon, go go 0 0 1315 3 x 1 0 -1 ° 1/2
anti-S°-hyperon
Q“-Hyperon, Q- Q+ -1 +1 1672 1 .3 X 10" 10 3/2?
anti-Q“-hyperon

called resonance particles. These short-lived formations


have lifetimes characteristic of strong interactions (Par. 5°),
have'the properties listed in Par. 1°, and also definite chan­
nels of decay reactions (VI.4.8.2°). This is why the resonance
particles have been included among elementary particles.
4°. Three of the four types of interactions in nature
(1.2.2.2°) occur between elementary particles: strong, elec­
tromagnetic, and weak interactions. Each of them has a
definite relative strength and a characteristic lifetime (Ta­
ble VI.5.2).
5°. Strong interactions characterize the processes occur­
ring with baryons, pions, and kaons (see Table VI.5.1).
Table VI.5.2

fRelative Characteristic lifetime,


Type of interaction strength s

Strong 1 10“23 to 10-22


Electromagnetic 1/137 10-20 to 10"18
Weak IO-14 10-10 to 10“8
VI.5.2 Elementary Particles 573

The nuclear forces between nucleons in atomic nuclei


(VI.4.3.2°), and also the processes of the formation and decay
of mesons and hyperons (VI. 5.4.3^ 4°, 8°) in nuclear inter­
actions at high energies are due to strong interactions.
Nuclear forces are distinguished by charge independence.
The nature of nuclear forces does not depend on the pres­
ence or absence of an electric charge in particles. Therefore,
the nuclear forces between protons and neutrons, protons
and protons, and neutrons and neutrons are the same
(VI.4.3.2°, c). Processes in which strong interactions mani­
fest themselves are called fast ones and have a character­
istic lifetime—the nuclear time. The latter is the time t nuc
during which a particle of energy of the order of 1 MeV
and a velocity of 107 to 108 m/s covers a distance which in
its order of magnitude equals the diameter of a nucleus—
about 10“16 m:
__ 10-15 10"22 to 10"23 s
l nuc (107 to 1 0 8 )

Strongly interacting particles are called hadrons (large, mas­


sive particles).
6°. Electromagnetic interactions occur between electric­
ally charged particles: the Coulomb interaction of charged
particles (for example the repulsion of protons having a
like charge in nuclei), the processes of the production of
electron-positron pairs by gamma quanta (VI.5.5.4°), etc.
The intensity of electromagnetic interactions is character­
ized by a dimensionless constant—the relative strength
a = 1/137* (see Table VI.5.2). Processes attended by elec­
tromagnetic interactions are called electromagnetic processes
and occur during a time of the order of from 10“20 to 10~18 s.
7°. Weak interactions characterize processes occurring
with leptons (see Table V I.5.1). Examples are the interac­
tion of muons with nuclei, the interactions of electrons, posi­
trons, neutrinos (antineutrinos) with nuclei, and the pro-

* The constant 1/137 is the only universal dimensionless constant


that is obtained from a combination of three universal constants with
dimensions—e, ft, and c. The constant e2!hc = 1/137 is called the fine
structure constant. This name is associated with the fact that the con­
stant e2lhc determines the magnitude of the fine division (fine structure)
of the energy levels of an atom.
574 Atomic and Nuclear Physics VI.5J2

cesses of beta decays (VI.4.7.6°). In comparison with strong


interactions, the weak ones have a very low strength (see
Table VI. 5.2), and this is reflected in their name. Processes
attended by weak interactions are called slow ones, and
their duration ranges from about 10“10 to 10“8 s.
8°. Strong and electromagnetic interaction simulta­
neously exist between the nucleons in nuclei. The charge
independence of nuclear forces (Par. 5°) should result in
identical values of the rest
masses of nucleons (neutrons
-,p'>on' and protons). The addition of
electromagnetic interactions
to the strong ones results in
a. -T r different values of the rest
-5T° masses of a neutron and a pro­
QV77777777m777777777777777777777777777777, ton.
The rest mass of a nucleon
(a) (b) is the sum of the mass due
Fig. VI.5.1 to nuclear strong interactions
and of the addition to the
mass due to electromagnetic interactions. For a proton, this
addition has a negative sign, and the rest mass of a proton is
smaller than that of a neutron (see Table VI.5.1). In pions,
the rest mass of a neutral pion ji° is smaller than that of
charged pions n±. The explanation is that the electroma­
gnetic addition to the mass is positive for pions and the
same for positive and negative ones.J Therefore, the rest
masses of charged pions are identical and greater than
that of a neutral pion. Figure VI.5.1 shows conditionally
the values of the rest masses of nucleons and pions without
account taken of the electromagnetic interactions (a) and
with account taken of these interactions (6).
9°. For all the types of interactions of elementary par­
ticles, the laws of conservation of the physical quantities
characterizing the states of the particles before and after
a given interaction are observed, namely, the laws of con­
servation of energy, of momentum, of angular momentum,
and of electric charge.*
* Apart from the indicated universal laws of conservation, spe­
cial laws of conservation hold for separate kinds of interactions. Their
consideration is beyond the scope of the present course.
VI.5.3 Elementary Particles 575

3*. Cosmic Rays


1°. Cosmic rays are streams of*high-energy atomic nuclei,
mainly protons, reaching the Earth from outer space, and
also other particles produced by these nuclei in the Earth’s
atmosphere. The intensity I of cosmic rays is measured by
the density of the stream of particles—the number of particles
passing per second inside a unit solid angle (V.1.6.3°) through
a unit surface area. Cosmic rays are deflected by the Earth’s
magnetic field. Therefore the intensity of cosmic rays de­
pends on the latitude of the relevant spot on the Earth.
The maximum deflecting action of the Earth’s magnetic
field in the equatorial region is called the latitude effect.
2°. Cosmic rays outside the Earth’s atmosphere are
called primary ones. The energy E per nucleon in these rays
ranges from 1 GeV to 1020 eV. The composition and charac­
teristics of primary cosmic rays are given in Table VI.5.3.

Table V I.5.3

Intensity, Per cent in


Group of nuclei Charge Z m"2. sr-i -s"i total stream

Protons 1 1300 92.9


Helium nuclei 2 88 6.3
Light nuclei 3-5 1.9 0.13
Intermediate nuclei 6-9 5.6 0.4
Heavy nuclei > 10 2.5 0.18
Superheavy nuclei >20 0.7 0.05

At altitudes exceeding 50-60 km, a constant intensity


of the primary cosmic rays is observed. With a decreasing
altitude, a sharp increase in the intensity of the cosmic rays
takes place as a result of secondary cosmic radiation. Fig­
ure VI.5.2 shows the intensity I of cosmic rays in the units
indicated in Par. 1°.
3°. Secondary cosmic rays are formed as a result of ine­
lastic collisions (1.5.4.3°) of the primary cosmic rays with
nuclei of nitrogen and oxygen atoms in the upper layers of
the atmosphere. Below an altitude of 20 km, all cosmic rays
576 Atomic and Nuclear Physics V1.5.4

are secondary ones. The penetrability of these cosmic rays


is measured by the thickness d of the layer of lead through
which the cosmic rays penetrate. Figure VI.5.3. shows the
intensity I in relative units and also the weakening of the
intensity with an increase in the thickness d of the lead layer.

Fig. VI.5.3

Hard and soft components of secondary cosmic rays are


distinguished in accordance with their penetrability. The
hard component having a high penetrability in lead includes
heavy and fast charged particles. They lose energy mainly
for ionizing the atoms encountered along their path. The
soft component with a low penetrability includes light
charged particles—electrons and positrons, and also photons.

4*. Information
on Selected Elementary Particles
1°. Different types of elementary particles (see Tab­
le VI.5.1) are created when cosmic rays interact with the
nuclei of nitrogen and oxygen atoms in the atmosphere.
After the development of modern accelerators (VI.4.16.7°),
the birth, mutual transformations, and structure of ele­
mentary particles are being studied in laboratory conditions.
2°. The possibility of the birth of new particles when
protons interact with nuclei follows from the relationship
between mass and energy (V.4.11.1°). At a proton energy of
104 GeV that is about 104 times greater than its rest energy
mvc2 {mv is the rest mass of a proton, and c is the speed of
VI.5.4 Elementary Particles 577

light in a vacuum), the collision of a proton with a nucleus


leads to the disintegration of the nucleus, the imparting
of a great kinetic energy to the disintegration products, and
lo the birth of new particles with and without a rest mass.
3°. Muons (mu-mesons) are electrically charged particles
fi+ and p," with a rest mass approximately equal to 200 me,
where me is the rest mass of an electron. The basic character­
istics of muons are given in Table VI.5.1. Muons are not
stable particles and decay by the reactions
M
*+ +ie° + ove + ov|A
1* —>■-ie° + 0ve + 0vJt
where ^ e0 and +e° = electron and positron
0vJ! and 0vjl = electron neutrino and antineutrino
0v£ and 0vj5, = meson neutrino and antineutrino.
A meson neutrino (antineutrino) is a particle without a
rest mass and electric charge that is emitted in the processes
of the birth and decay of mesons. These particles differ from
the neutrinos (antineutrinos) that are emitted together with
electrons in beta decay processes (VI.4.7.6° and VI.4.10.3°).
This is why the two kinds of neutrinos (antineutrinos) are
distinguished by their symbols: 0v£ (0v£) and 0v[i (0vJ.). The
spin of mu-mesons is ±Ji/2, and their decay reactions comply
with the conservation laws (VI.5.2.9°). The interaction of
muons with nuclei is an example of weak interactions
(VI.5.2.7°), and muons are nuclear-inactive particles. The
interaction of electrons and positrons, and also of both
lypes of neutrinos and antineutrinos with nuclei is also
weak. All these particles relate to the same class of leptons
(see Table VI.5.1).
4°. Pions (pi-mescns) are neutral (ji°) and electrically
charged (n+ and n “) particles with a rest mass approximate­
ly equal to 300me, where me is the rest mass of an electron.
The basic characteristics of pions are given in Table VI.5.1.
I’ions, like muons, are not stable particles and decay by
I h e reactions
*+-*H ++ ovJi

ji° -> y -f 7
578 Atomic and Nuclear Physics VI.5.4

where \ i + and = positive and negative muons


0vji and 0vji = meson neutrino and antineutrino
y = gamma quantum.
Figure VI.5.4 shows schematically the sequence of
(jt-p,-e) decays. Pions are particles without a spin. This

Fig. VI.5.4

follows from their decay reactions and the requirements


of observing the laws of conservation in these reactions
(VI.5.2.9°).
5°. Pions are formed in cosmic rays (VI.5.3.1°) as a
result of fast protons and alpha particles demolishing the
nuclei of atoms of atmospheric
gases. The energy of the protons
must be about 300 MeV to pro­
duce pions.
Figure VI.5.5 shows sche­
matically how pions are pro­
duced in a laboratory. A target A
of beryllium or carbon is bom­
barded with fast protons jp1.
Pions are produced that fly out
of the target at arbitrary an­
gles. The magnetic field of the
accelerator (VI.4.16.7°) twists the pions along circles
whose radii are determined by the velocities of the pions
(111.4.5.2°). The pions flying out of the target to the front
(see Fig. VI.5.5) separate: the negative pions leave the accel­
erator chamber, and the positive ones deviate into the
chamber. For the pions that fly out of the target to the back
of the chamber, the pattern of the deviations is the opposite,
and is not shown in Fig. VI.5.5.
The masses of charged pions (see Table VI.5.1) are deter­
mined according to their energies and momenta by the meth­
ods of particle deflection in magnetic and electric fields.
The lifetime of pions (see Table VI.5.1) is determined by
VI.5.4 Elementary Particles 579

measuring the intervals between*the moment of the birth


of a pion and the moment of its decay (Par. 4°). Pions are
nuclear-active particles. Their interaction with nuclei is
an example of strong interactions (VI.5.2.5°).
Similar methods are used to study other charged particles.
Since neutral particles do not deviate in a magnetic field,
the laws of conservation (VI.5.2.9°) and reactions of the
given particle with other ones are used to study them. For
example, the reaction of a negative
pion with a proton proceeds as fol­
lows:
Jt~ + iP 1-► Ji° + on 1
The negative pion changes into a
neutral one, and the proton into a
neutron. The neutral pion ji° de­
cays into two gamma quanta (Par.
4°). The mass of the neutral pion
(see Table VI.5.1), is ^determined
according to the known masses and
energies of the proton, neutron, and negative pion, to the
energy of the gamma quanta, and also to data on the mo­
menta of the particles in these transformations.
6°. The idea of using the law of conservation of momen­
tum for studying reactions of transformation of elementary
particles can be illustrated by considering the example of
P_-decay in which an electron antineutrino 0vJ! (VI.4.7.6°)
appears. If a nucleus in p_-decay emitted only one electron,
then it would recoil in the direction opposite to that in
which the electron flies out. The momentum of the nucleus,
stationary prior to decay, would equal the momentum of
the electron in magnitude, but would be opposite in direc­
tion. If the nucleus, apart from the electron, also emits an
antineutrino, then according to the law of conservation of
momentum (1.2.6.2°) the vector sum of the three momenta—
of the electron, antineutrino, and recoil nucleus—should
equal zero as before decay (Fig. VI.5.6) (see also VI.5.5.8°).
7°. Kaons, or K-mesons, are neutral (K°) and charged
(K + and K “) particles with rest masses close to 1000/rae,
where me is the rest mass of an electron. Table VI.5.1 gives
580 A tom ic and Nuclear Physics Vi.s.5

the basic characteristics of kaons. Like other mesons, kaons


are unstable, but the channels of their decay reactions
(VI.4.8.2°) may vary, unlike the decays of pious and muons
(Paragraphs 3°, 4°). The decays of kaons result in the for­
mation of pious, muons, and neutrinos (or antineutrinos)
of both types (VI.5.4.3°). Like pions, kaons have no spin.
All these particles (pions and kaons) are combined into a
single group of particles—mesons (see Table VI.5.1).
8°. A large group of particles called hyperons has been
detected in nuclear emulsions (VI.4.6.6°). These particles
have rest masses greater than that of nucleons (see Ta­
ble VI.5.1). They approximately range from 2183rcie for a
lambda-hyperon (A0) to 3280me for a negative omega-hyperon
(Q~). Hyperons are unstable (see Table VI.5.1). The lighter
hyperons decay into nucleons and pions. The heavier ones
decay into lighter hyperons, pions and kaons. All hyperons
have a spin of h!2 except for the Q"-hyperon whose spin is
3/&/2. Nucleons (protons and neutrons) and hyperons are
relatives, and they are included in the same class of heavy
particles—baryons (see Table VI.5.1).

5. A ntiparticles
1°. Antiparticles correspond to most of the elementary
particles. Examples of particles and antiparticles are the
electron ^e0 and the positron +1e°, the muons p + and p “,
the pions jt+ and jt”, the kaons K + and K", and the electron
and meson neutrinos 0v{!, 0v|i, an(i fhe antineutrinos 0v£
and 0vji. The rest masses, spins, and lifetimes of particles
and antiparticles are the same. The electric charges of par­
ticles and antiparticles are identical in absolute value, but
opposite in sign*. Particles whose properties are completely
identical to those of their antiparticles are called truly
neutral particles. They include the photon and the neutral
mesons jt° and K°.

* Particles and anliparticles also have other characteristics


differing in sign, but their treatment is beyond the scope of the pres­
ent course.
V 1 .5 .5 E lem entary Particles 581

2°. The existence of pairs of charged particles and anti-


particles in nature expresses the principle of charge conjuga­
tion: every charged elementary particle must have an anti­
particle. In accordance with this principle, a proton has an
antiparticle—the antiproton ^p1 (see Table VI.5.1). The
principle of charge conjugation
also covers the neutral par­
ticles neutron and neutrino—
there exist the antineutron 0nx
(see Table VI.5.1) and also
the electron antineutrino 0vJ!
and meson antineutrino 0v^.
3°. The combination of a
particle and its antiparticle
is attended by the release of
energy that is not less than the
double rest energy of each of
them (V.4.11.20). The forma­
tion of a particle-antiparticle
pair requires the expenditure
of energy exceeding the dou­
ble rest energy of the particle.
This is associated with the
need to impart a momentum
and kinetic energy to the
created pair.
4°. An electron-positron
pair is formed when a hard gam­
ma quantum collides with a Fig. V I.5.7
charged particle, for example
an electron on the shell of an atom or an atomic nucleus*.
The reaction is ^ e0 -f +1e° and is possible when the
energy of the photon is at least 2mec2, i.e. hv > 2mec2 —
- 1.022] MeV, where mec2 = 0.511 MeV is the rest energy
of an electron (or positron). Figure VI.5.7 contains a photo-

* The need of having a third particle is associated with the fact


that when two particles of rest mass ///0 =/= 0 are formed from a photon,
t he total momentum of the two particles is less than that of the photon
h v / c (V.5.1.20), and a third particle is necessary to take up part of the
photon’s momentum.
582 A tom ic and N uclear Physics F/.5.5

graph of the pair formed in the medium filling a Wilson


cloud chamber (VI.4.6.4°) under the action of hard photons.
It follows from the reaction of formation of a pair under the
action of a gamma quantum that the spin of a photon must
be 0 or 1 (in units of %) because the spin of an electron and
that of a positron are h/2. Numerous experimental facts and
theoretical considerations not treated here resulted in the
conclusion that the spin of a photon equals h.
5°. The combination of an electron and a positron is
called annihilation of the pair. In the overwhelming major­
ity of cases, the reaction proceeds with the formation of
two gamma quanta:

_ie°-f-+1e°-^2Y

Before the combination of the particles e0 and +1e°, the


total momentum of the two particles in a coordinate system
associated with the centre of mass (1.2.3.4°) of the electron-
positron system was zero. After annihilation of the pair,
two photons should be formed whose momenta are directed
oppositely and whose total momentum should equal zero.*
Each of the photons carries away energy equal to hv =
= mec2 = 0.511 MeV.
6°. The minimum energy needed for the formation of a
proton-antiproton pair in a coordinate system where one
nucleon is at rest is 6mvc2 = 5.6 GeV. This energy decreases
to 4.3 GeV in the practical formation of the pair iP1--!? 1.
Considerably higher energy is needed to form a hyperon-
antihyperon pair than to form a pair of nucleons. This is
associated with the relatively great rest mass of the hyper-
ons. For example, the lightest of the antihyperons—the anti-
lambda-hyperon (A0)—is produced at energies of 5.3-5.8 GeV.
7°. A feature of antiparticles is their ability of rapid
recombination with their particles, for example positrons
with electrons, antiprotons with protons, and antineutrons
with neutrons. This is connected with the circumstance that
the substances which our surroundings consist of are built
up from electrons, protons, and neutrons. Antiparticles

* Annihilation is also possible with the formation of three pho­


tons. It obeys the laws of conservation of momentum and energy.
VI.5.6 Elementary Particles 583

produced artificially, when encountering an excess of their


particles contained in a substance, combine with them and
stop existing, causing the birth <rf new particles.
8°. The distinctions between neutral particles and their
antiparticles are established according to the different nature
of the reaction of the particles and antiparticles with a sub­
stance. Examples are the reactions of an antineutrino 0vJ!
with a proton jp1 and a neutrino 0vJ! with a neutron on1:
ove + 1P 1 —^onl + +ie° and
(*)
ov e + on1— 1P1 + - l e°
9°. The differences between electron and meson neutrinos
are established when studying the decays of charged pions
(VI.5.4.4°). If we separate the neutrinos formed in these
decays and let them be captured by neutrons, then instead
of the reaction (*) of Par. 8° the process will proceed by the
reaction
ov£ + on1-> lP1 +
where fi“ is a negative muon. This proves the difference be­
tween the electron and meson neutrinos (and antineutrinos).

6*. Concept of the Structure


of a N ucleon

1°. By the structure of a nucleon (and of any elementary


particle) is meant its extension in space and its construction.
The structure of an elementary particle cannot be considered
separately. It is associated with the structure and proper­
ties of other particles. Direct experiments involving the
investigation of the structure have been conducted only
for nucleons.* Elastic collisions (1.5.4.1°) of pions with
protons and of fast electrons with protons and neutrons are
effective ways of studying the structure of nucleons. The
* The problem of the structure of the elementary particles, and
also attempts to establish the’fundamental particles from which all
the others are built up, are the “battlefront” of high-energy physics.
In this book, we give only the most general information on the struc­
ture of a nucleon.
584 Atom ic and N uclear Physics VI.5.6

first way showed that a pion deflects insignificantly from


its initial direction upon a collision, while a proton receives
an insignificant recoil, so that the momentum Ap transferred
to the proton is not great. It follows from the uncertainty
relation (VI. 1.5.5°) that the process of the collision of a
pion with a proton occurs in a region of space with the linear
dimensions a > ft!Ap, where a characterizes the dimensions
of a nucleon.
2°. Processes of emission and absorption of particles
and antiparticles are constantly occurring in a nucleon.
A nucleon is considered as an in­
tricate combination of many par­
ticles changing in time.
The central part of a nucleon
(the bare nucleon) contains its
core with a radius of (0.2-0.4)x
X 10“15 m. In this part, a special
role, which to date is not com­
pletely clear, is played by the
heavy particles—resonance par­
ticles and nucleon-antinucleon
pairs. A pionTcloud formsrthe
outer part of a rnucleon
(Fig. VI.5.8). The notion of the
structure of a nucleon makes it possible to determine more
precisely the difference between the masses of a neutron and
a proton (VI.5.2.8°). It is associated with the energy of electro­
magnetic interaction of the nucleon core and the pion cloud.
3°. The scattering of fast electrons with an energy up
to 550 MeV on protons made it possible to study the distrib­
ution of the density of the electric charge of a proton depend­
ing on the distance r from the centre of the core. It must
be borne in mind that the charge of a proton is indivisible
and always manifests itself as a single whole. Consequently,
the distribution of the electric charge in a proton does not
signify the possibility of experimentally separating a def­
inite portion of this charge. Figure VI.5.9a shows how the
charge q contained in a spherical layer confined between
the radii r and r -f- Ar depends on r (measured in fermies).
The curve in the figure has a sharply expressed peak corres­
ponding to the distance R e = 0.77 X 10“15 m, which is
V I .5 .6 Elem entary Particles 585

called the electric radius of a proton. The area under the


curve equals the charge of a proton e. The results of similar
experiments involving the scattering of fast electrons on
neutrons are shown in Fig. VI.5.96. The scattering occurs as
if the electric radius of the neutron were equal to zero. A neu­
tron interacting with fast electrons behaves as if its pion

cloud coincided in dimensions with the core. In the inner


and outer regions of a neutron, the electric charge is nega­
tive, in the middle region it is positive. The total charge of
a neutron equal to the area under the curve is zero.
The structure of elementary particles, which is being
intensively studied at present, shows that the classical
notions of the dimensions of elementary particles (VI.5.1.6°)
should be treated as very conditional ones.
SUPPLEMENTS

1. U n it s an d D im e n s io n s
o f P h y s ic a l Q u a n titie s .
S y s t e m s o f U n it s

1°. The unit [A] of the physical quantity A is defined


as a conventionally chosen value of the physical quantity
having the same physical meaning as the quantity A .
Units that can be reproduced in the form of definite bodies,
specimens or devices are called measures. For example, a kilo­
gram weight is a measure of mass, and a metre rule is a
measure of length.
Measures made with the highest possible degree of ac­
curacy at the present level of development of measuring
equipment are called standards. For example, the standard
of mass is a definite cylinder made of a platinum-iridium
alloy, and the standard of electric resistance is a definite
electric circuit consisting of several conducting coils and
a measuring device whose readings are taken at a strictly
definite temperature.
2°. A system of units is a combination of units of phys­
ical quantities established in a definite way.
Up to 1963, several systems of units were in use for
different branches of physics [for example, the absolute
physical ^system of units cgs in mechanics, the absolute
electrostatic system of units cgse in electrostatics, and the
mks and mk(force)s systems]. Recent times have seen the
Vll.l Supplem ents 587

introduction of the International System of Units (the SI


system) as the preferable one. It is a single system for all
branches of physics. •
The basic units of a system are units established sepa­
rately for several arbitrarily selected independent physical
quantities. The basic units of the cgs system (from which
its name has been derived) are the centimetre—the unit of
length, the gramme—the unit of mass, and the second—
the unit of time (VII.3.1°). For the basic units of the SI
system, see VII.2.
Derived units are units established through the basic
units of the given system using physical laws or definitions
expressing the relationship between the physical quantities
in question and the quantities whose units have been
adopted as the basic ones.
3°. The dimension of the physical quantity B is defined
as the expression determining the relationship between the
unit of this quantity [5] and the basic units [4x1, [4 2], . . .,
[4/J of the given system. The formula of a dimension has
the form

lB] = [ Ax]n' [42] " * ...[ 4 ft]n*

where k = number of basic units in the system


n±, n2, . . ., nh = rational numbers.
4°. Homogeneous physical quantities are quantities having
identical dimensions and the same physical meaning, i.e.
differing only in their numerical value (for example, the
coordinates of the points of a body and its linear dimensions,
and the graduation of an ammeter connected to a given
portion of an electric.circuit and the current in this portion).
Similar physical quantities are quantities having the
same dimension, but differing in their physical meaning
(for example, the work of a force and the moment of the
force, and the radius of a spherical conductor and its capa­
citance in the cgse system).
Dimensionless quantities are quantities whose numerical
values do not depend on the choice of a system of units (for
example, the ratio of the length of a circle to its diameter,
the coefficient of sliding friction, and the relative permit­
tivity of a substance).
588 Supplem ents VII.S

5°. The prefixes used to designate multiples and submul­


tiples of units are given in Table V II.1.
Table V I I .1

Prefix Multiple of basic unit Symbol

Pico 10-12 P
Nano 10-9 n
Micro lO-6 P
Milli 10-3 m
Centi 10-2 c
Deci io- 1 d
Deca 10 da
Hecto 102 h
Kilo 103 k
Mega 106 M
Giga 109 G
Tera 1012 T

2. B a s ic and D e r iv e d U n it s
o f th e S I S y s t e m
The basic and derived units of the SI system, their defi­
nitions, and symbols are given in Table VII.2.

3. U n it s o f P h y s ic a l Q u a n titie s
in M e c h a n ic s
1°. The derived units of mechanical quantities in the
SI system are established through the units of length, mass,
time, and plane angle.
In addition to the SI system, the absolute physical sys­
tem of units—the cgs system—is used in mechanics. The
basic units of this system are:
the unit of length—the centimetre—one-hundredth of a
metre;
Vlt.3 Supplements 569

Table V I1.2

Quantity Unit. Symbol Definition of unit

Basic Units
Length Metre m Length equal to 1 650 763.73
wavelengths in vacuum of the
radiation corresponding to the
transition between the levels
2p10 and 5d5 of the kryp-
ton-86 atom
Mass Kilogram kg Mass of the international pro­
totype of the kilogram
Time Second s The duration of 9 192 631 770
periods of the radiation cor­
responding to the transition
between the two hyperfine lev­
els of the ground state of
the cesium-133 atom
Tempera­ Kelvin K Thermodynamic temperature
ture equal to 1/273.15 of the ther­
modynamic freezing point
of distilled water at a
pressure of 101 325 Pa
Amount of Mole mol Amount of substance of a
substance system containing as many
elementary particles (atoms,
molecules, ions, etc.) as there
are atoms in 0.012 kg of the
isotope carbon-12 (C12)
Current Ampere A Constant current which, if
maintained in two straight
parallel conductors of infinite
length, of negligible cross
section, and placed 1 metre
apart in a vacuum, would
produce between these conduc­
tors a force equal to 2 X 10-7
newton per metre of length
Luminous Candela cd Luminous intensity, in the
intensity perpendicular direction, of a
surface of 1/600 000 m2 of a
blackbody at the freezing
point of platinum under a pres­
sure of 101 325 Pa
590 Supplements Vll.4

T a b le V I 1 . 2 (co nclu d ed )

Quantity Unit Symbol Definition of unit

Supplemental Units
Plane angle Radian rad Angle between two radii of
a circle, the arc between
which equals the radius of
the circle
Solid angle Ste radian sr Solid angle with its apex at
the centre of a sphere that
cuts out an area on the surface
of the sphere equal to the
area of a square whose side
equals the radius of the sphere

the unit of mass—the gramme—one-thousandth of a


kilogram;
the unit of time—the second.
A supplemental unit for plane angles is the radian.
Table VII.3 gives the units of selected mechanical quan­
tities in the SI and cgs systems.
2°. Table VI1.4 gives selected non-system units encoun­
tered in physical and technical publications and their con­
version factors to SI units.

4. Units of Physical Quantities


in Molecular Physics and Thermodynamics
1°. The derived units of quantities in molecular physics
and thermodynamics are established in the SI system
through the units of length, mass, time, temperature, and
amount of substance.
Before the introduction of the SI system, derived units
were used in this field of physics that are related to the basic
units of the cgs system and the degree, and also non-system
units based on the unit of the quantity of heat—the calorie.
The conversion factor to be taken into account in converting
units based on the calorie to SI units is 1 cal = 4.1868 J.
V11.4 Supplements 501

2°. Two basic temperature scales are used in molecular


physics and thermodynamics: the thermodynamic temperature
scale (the previous name of the •temperature unit according
to this scale was the degree kelvin—°K, the new name is
the kelvin—K), and the international practical temperature
scale (the temperature unit is the degree Celsius—°C).
The basic point for graduation of the first scale is the
temperature of the triple point of water, taken equal to
273.16 K. The basic points for graduating the second scale
are the melting point of ice (0 °C) and the boiling point
of water (100 °C) at standard atmospheric pressure (101 325
Pa).
Since a temperature of + 0.01 °C on the practical scale
corresponds to the triple point of water, then a thermody­
namic temperature of 273.15 K corresponds to zero of the
practical scale, and temperatures can be converted by the
formulas
T (K) = 273.15 + t (°C) and t (°C) = T (K) — 273.15
where T = temperature on the thermodynamic scale
t = temperature on the practical (Celsius) scale.
3°. In connection with the decision of the Fourteenth
Conference Generale des Poids et Mesures (October, 1971)
on the introduction of the mole as the seventh basic unit
of the SI system, it became necessary to bring order into
the names and definitions of some quantities.
The amount of substance is designated by the symbol n.
The definition of the unit of the amount of substance—
the mole—is given in Table VII.2.
The relative atomic mass of an element (^4r) is a dimension­
less quantity equal to the ratio of the average atomic mass
of a natural mixture of the element’s isotopes to l/12th
of the mass of the isotope carbon-12. This quantity was pre­
viously called the atomic weight (or atomic mass). Sometimes
it was considered to be a dimensionless quantity, and some­
times it was measured in atomic mass units (amu) (VII.7.1°).
The relative molecular mass of a substance (M r) is a dimen­
sionless quantity equal to the ratio of the average molecular
mass of a natural mixture of the substance’s isotopes to
l/12th of the mass of the isotope carbon-12. This quantity
was previously called the molecular weight (or molecular
592 Supplements VII.4
Table VII.3a
VII.4 Supplements 593
594 Supplements VII.4
VITA Supplements 595
596 Supplements VII.4
Table VI1.3b
Table VI1.3b (concluded)
VII.4
Supplements
597
598 Supplements VII.4

Table VII.4

Quantity Unit Symbol Conversion to SI units

Length Micron V 1 p = 10-6 m

©
tH
O
Angstrom A

1
a
II
Mass Technical unit t.u.m . 1 t.u m. ^ 9.81 kg
of mass
Tonne — 1 tonne = 103 kg
Time Minute min 1 min = 60 s
Hour h 1 h = 3600 s
Plane angle Round angle — 2jtrad ^ 6 .2 8 3 rad
Right angle — ji/ 2 rad « 1.570 rad
Degree o 1° = jc/ 180 rad « 1.745 X
X l0 ~ 2 rad
Minute 1' = (ji/ 108) X10"2 rad »
« 2.908X 10-4 rad
Second 1" = (ji/648) x 10“3 rad «
^ 4 .8 4 8 x l 0 ”6 rad
Area Are — 1 are = 102 m 2
Hectare ha 1 ha = 104 m2
Volume Litre 1 1 1 = 10-3 m3
V elocity Kilometre per km/h 1 km /h « 2.7 7 7 X
hour XlO-1 m /s
Angle of Revolution rev 1 rev = 2xc rad
rotation
Angular Revolutions per rpm 1 rpm = 31/30 rad/s
velocity minute
Revolutions per rps 1 rps = 2n rad/s
second
Force Kilogram-force kgf 1 kgf « 9.81 N
Tonne-force tonnef 1 tonnef « 9 .8 1 X103 N
Work Watt-hour W -h 1 W -h = 3 .6 x 103 J
Power Horsepower (met­ hp 1 hp « 735.499 W
ric)
Pressure Bar bar 1 bar = 105 Pa
M illimetre of mmHg 1 mmHg « 133.322 Pa
mercury column
M illim etre of wa­ mmH20 1 mmH20 « 9.81 Pa
ter column
Technical atmo­ at 1 at » 9.81 XlO4 Pa
sphere
Physical atm o­ atm 1 atm « 101 325 Pa
sphere
VII.5 Supplements 599

mass), and sometimes it was measured in atomic mass units


(amu) (VII.7.10).
The molar mass (M ) is a dimensional physical quantity.
It equals the ratio of the mass of a substance (m) to the
amount of the substance (n):
M= —
n

The unit of molar mass in the SI system is the kilogram


per mole (kg/mol).
The numerical values of the molar mass and the relative
molecular mass of a given substance are related by the ex­
pressions
M = M r x It)"3 and M r = M x 103
4°. The units of selected quantities encountered in mole­
cular physics and thermodynamics are given in Table VII.5.
The table also gives the symbols of these quantities.

5. Units of Quantities in Electrodynamics


1°. The sixth basic unit of the SI system (in addition
to the metre, kilogram, second, kelvin, and mole) for elec­
trostatics and electrodynamics is the unit of current—the
ampere (VII.2). The proportionality constant in the equation
of Coulomb’s law for a vacuum (111.1.2.5°) is taken equal
to l/4rte0, where e0 is the electric constant. Its value is
e0 = 8.854 187 82 x l O '12 F/m
2°. Apart from the SI system, the absolute electrostatic
system of units cgse is used in electrostatics. Its basic units
are the centimetre, gramme, and second, and the relative
permittivity of a vacuum is taken equal to unity (evac = 1).
The units of the remaining quantities in this system are
called absolute electrostatic units and are designated by the
symbol “units cgse”, or by notation using the symbol of the
relevant quantity as a subscript (for example, 1 cgseq is
the absolute electrostatic unit of charge, and 1 cgsec is
the absolute electrostatic unit of capacitance).
Table V II.6 gives the units of selected quantities used
in electrostatics and electrodynamics of steady currents,
600 Supplem ents VI 1.5
Table V II .5 a
VI1.5 Supplements 601

o o o o • o 0 0
CD I cq So ^ OJ Q y l^ . ■M OJ
3-^ L.O LD
<1 |< ] -cH CD pen < ]|< 3 sf CO V f
i | Q. ■*■•
II LD
II
11 ”—1
1 I11I ^
t-H t—e mc
||
M I HH MP M
S^ M O ►—1 t—i i—• M CJ 1—1 'sJ »—1 M 1—H

pp >» >> pp IS C • co
ii a ©
S 9 h © ; ^ pP ©
© g © ©
© ^
♦3 © c c
o o 3sS
** S3 f t ^ 2 S
&tI-l >
®» © © P h © 1 g o
.2 «
“ § s
© © £ p P S3 © ©
© rt ^ J3 to .fi
© ftfl.M © S.t2
pS3C
POP©©
§ S° S3
©
©^ S3 o©
CO o
B S .S - C
g># °
'O*"" © §
pp > £
© © © co cd
,

a§§ ©£ »“• '-'© p ©pppi* © P©


crt! S3 ^ © ^ $ *
© CO •<°—» « M H
i § ° •mml S3 CO « H 0 ® g © »
CM**-« O CO © pQ © o © Ph
c f lO O © +» > ©g
CO O >»&)©
o j
>»©
g o ©
Pc W
cm m
co §r; CM .2 ’© .ts ® 3
°®aa —
| ° o ®®
c g * s
tr
©©
** S
©

O
wo
»m 43
^ «M

^
©
C d*
Pc
M O
^ .r^
.15 ©
o cd
© pP w —
« o il
l a
© o Pc
i-e 0.0 2 2
s „ « a
© •«-» K. ® 03 ©
2
sa l © pp £
p © pM
° S£|
p^ © *3 Spd
© © m $p
©«4HC
M
*»°
is : © w w© w© «
m S
o o
3 ^ «M O I S ^ pS
g ^
© CM © O
-p o —^
o* -£
^ °^d £
CM** g S
©c£+*
° pP O
O JX

a g f t s°a& S3 O
®s s o Ph ©'
s g-gg
O S S js-a a ”
h 9 cd ©
g * +9+f
* 2 2 -
111 -p
o o _
p o
&3S ss
W ? O o
a a o
•< £ w a g
n o | S
C/D «
© © s s.a
as

a F-l C
© F .M G
© Pet ^2 t-t >
S pS P PM
P c^
^a P*
©-* © cd p3 o
© 3 © © © © & ©o
U Pc
o o .S
H* > 5 p 2 5a
a © <o ^ ^

«
m r?
ca-

© ©

a
j© *«
©
© W)
PM
p
^ ©

©

<
4-»
o g

© ^
p ©
o ©
CM © © © © MC
>
© ©
o ©
M p—I
a
©

©
©
pP
pP cd
U
© w
c

cP ©
G cd © Cp >»

© §s »
_i >>
©# t9
«3 w
© §a
©
P. © Is © -M

a s
o p
© ©
P^P
CO co <■2 E CA Pc 5S & CC Pc
602 Supplements VI1.5

Table VI 1.5b
VII.7 Supplements 603

and also the relationships between these units in the cgse


and SI systems.
3°. In the SI system, many ecfhations describing electro­
magnetic phenomena contain the magnetic constant p0
equal to 4ji X 10-7 H/m.
Table VII.7 gives the units of selected quantities used
in describing electromagnetic phenomena. These units are
given only in the SI system.
4°. A non-system unit of energy is the electron-volt
(eV):
le V = 1 .6 0 2 1 8 9 2 x 10"19 J
The electron-volt is defined as the energy which an electron
acquires when it passes through a potential difference of
one volt.

6. Units of Selected Quantities


in Wave Processes and Optics
1°. The units of selected quantities used in wave process
es are given in Table VII.8.
2°. The derived SI units for photometric quantities are
established through the units of luminous intensity, length,
time, and solid angle.
Table VII.9 gives the units of selected photometric
quantities.

7. Selected Units in Atomic


and Nuclear Physics
1°. The atomic mass unit (amu) is 1/ 12th of the mass
of an atom of the carbon isotope 6C12:
1 amu = 1.660 565 5 x 10"27 kg

2°. The atomic energy unit (aeu) is the energy corres­


ponding to one atomic mass unit (V.4.11.1°):
1 aeu = 931.5016 MeV
604 Supplements VI1.7
Table VI1.6a

c Ibo <
w

a ©
© to
1a
'5
<
< a © a a
S -S * .
o0 o &-i-o
^ "0rH ^
SSSg ©. f©t < \<
Jr©PO 3a & ea ^ ^ eg Snta
1 ^«ri li ®i
ScS** t—
I |J^M I© ~5 1 *^r
o > > C M fea E
©-g 3 OfH © bo
CO f l T3
©
oO 5§©

3 p.g .2 §
© >*
G cd © &«
3*5 o h £ g S 3 * © a
A 9O ©©© 2< 2 § - £ a g'-gS a 5
«O Ta
© m§ «S fcj‘o >»© is© o i ° ©^ Q} CO

£f | o ® a«45c-d3Hty
©
0 - ^ ©
£©
1 1 .3 § QD*Q
M -
10Uo s © © © ^ ia
CDO ed cd««
Fh*°tT p©f©*£ p* ©«
©3osta © Q«© a
«cd -S
.© © © 2 *d fl
3 o
§
_Q II .£,© » § ©^ 2 ► ©*«-•
s §=
Ph bOT 3 S -2 ’S’a
o_-
^o So fl ^S .2 Op© o °
O
gja H »H 8 S g
JS>P-i o o 5 5 a © ^© o© +©» ©<©
© cd M £.21! « MClH Q) 0) •iC
—iOcPoi cd
o
A j3 ei a
a &h o >

fQ F^2
©©
a a
o traOf_
©
Ph cSl 11^o—
4 F©C
e © "O ©
P
i § o M*©
I f i a 1u u&a >© ®a
H3
o© "©
© ©

.2 ©
xJ
© ’£ ’u b©o o© ©o
tfH b
©
Fh
©© o© F
©^ M M
©
g ,2 ©
u w~ « Wp©
o ©f©
CO © W.5
VII.7 Supplements 605

N
<N
7 << • 1
< <
« i eo
i
DO 7 ©
bo bo
< bo F^J p^ bo
p* F^J
CO M rH C9
| 1 e«i*
a a a a a a

n
+ 1
eo s
o &“ o o b o
CO < vr co CO
n cr|8- ' Pa oi
1il
71 eg
II ^ V s© ^
© s
QhH-1
IN | l~~!
•t—
ii r'On ii—
— —i' S
© Z
E
Z HH
l—l OCJ
eo
HI
HI
^ )+ E

pdT3 Si ■^ * 3
O O f- ;
■p m M rl
d d O §> £}
+2 O 4S f t O
©
©.2 >
,p*1> *3 cpt ^ ®
§ °
©
2 ® S '£
Si O ^
o » *»
3 o-
2 ? .® tod © o©
ca © g ®
Si O
03 a
^ © o
a *
3 J! °-S
s b * •H © a
d
o O © M
o
> .
^ d
a- I ^
© a d
^ d boo Si© fG .2
o cd
O h OJ
Ml'S . S o * «+i d
d § SpS a © d S
I
6° a ® ® G 60 Si
•pi 03 ca a® « o •<■* § " t t ^
p ^ sj d «
cd n s
&
« ©
#■* I^ 5*2 „ s «
® _Q *
to d
o -5
^ oo 2_ „ _ § s> a
j s § »
ooc ° § | « © cd f l 5 w 03 ®
© oJ *S<=>”£ o®
I § 8 .
° Stitt'S G m "i
©
fl r ^ o oa
„ pOH .. ©
•X3 © «4H a, o ^ o ! « - ,s ® _ d
£ ^|o «+io
pda o°
G ^ o o a O 'f 5P
© U ^
cd —
i^ °-G * d
.2 S ,° ®
+j (d h d5®$s .ts
h ^ P
.15 § - h " d g
§ | .5 2 § ad ^© g o ' > > 2 5*2 2 § 0 -PO
© ©l ®©o
©
© 05 g 3
fO ©
o “ a ® S * 8.1 s g 2 ° fl
rt d 3 S --
H co © ^ o o
© © m
OfO^.2 G
o © -i pQ Q « § I 3 O5h cd
St*02 OnO © Pi
> §

eo (N
a .a a <3
6 ►—»
<

©
per
m e-

Si
pQ ©
© d
3 Si
OSi® "d © C Q3
P—I
d +*
o
©
© o O.
a
©
Si 3
Ph
a
pd
o S
u a >
©
fe
S
O ^ d3 S
°h
l-o © 4H» < Oha
© d
o
•in O
O Si * -© c
-d © Oh© ^
•'S a
03 Q « ©

oa
ad o*
o
o .a ^ © © ©
03 © £ d
©
5 a d"? bo
p ©
•^ © 2 ©M-. o
5 ° fS
d
©
8jS - o E
S3 § , 5 >» © ' d
P .
Sh h

£© • 2d -.T2 .
3 8. © d O
u ©> u ao GO O & d
d
CJ co
Table VI1.6a (concluded
606
Supplements
VII.7
VII.9 Supplements 607

3°. The unit of activity of a radioactive substance is the


number of disintegrations per second (d/s or, sometimes, s"1).
A non-system unit of radioactivity is the curie (Ci)—
the activity of a radioactive substance in which 3.7 X 1010
disintegrations occur per second:
1 Gi = 3.7 x 1010 d/s

8. Selected Physical Constants *


Molar gas constant R = 8.314 41 J/mol-K
Gravitational constant G = 6.6720 X 10~n N *m2/kg2
Planck constant h = 6.626 176 X 10~34 J -s
Charge of electron and proton e = 1.602 189 2 X 10~19C
Specific charge of proton e!mv = 9.578 758 5 X 107 C/kg
Specific charge of electron e!me = 1.758 804 7 X 1011 C/kg
Rest mass of neutron mn = 1.674 954 3 x 10~27 kg
Rest mass of proton mv = 1.672 648 5 X 10“27 kg
Rest mass of electron me = 9.109 534 X 10-31 kg
Molar volume of ideal gas in standard conditions Fm =
- 22.413 83 x 10-3 m3/mol
Avogadro constant N A = 6.022 094 3 X 1023 mol"1
Boltzmann constant k = 1.380 662 X 10~23 J/K
Rydberg constant for hydrogen i?H = 1.097 373 143 X
X 107 m"1
Faraday constant F = 9.648 456 X 104 C/mol
First Bohr radius a0 = 5.291 770 6 X 10“n m
Speed of light in vacuum c = 2.997 924 58 X 108 m/s

9. Ways of Measuring Physical Quantities


1°. Measuring a quantity consists in comparing it with
another homogeneous quantity adopted as a unit.

* The values of the constants in this book are given according to


the article “Recommended Approved Values of the Basic Physical
Constants—1973 (Report of the Working Group of CODATA on the
Basic Physical Constants, August, 1973)” in Uspekhi fizicheskikh nauky
Vol. 115, issue 4, April, 1975, pp. 623-633.
608 Supplements VII.9

Table VI1.6b
T able V II.6 b (c o n c l u d e d )
VI1.9
Supplements
609
610 Supplements VII.9
Table VII.7
iH N
1
< <
M
1
CO

< n
<
a
M
1
©
7 co
bO

bo 9 bo bo
be M 1
^5
(M (N a
a a
bC-M
.Sc
C =>
is ^ ° . 0 Co O oqo
CJH
CM o «L o 8
OTO B ^ © r ^ £
O 0) ^2© h2?2 « 11 ^ <
_< TO
■S
<
2 II ^ g ^ II W « !h « 1
A 3
cOeC - Q? 1—
S1 C1—1 ©e
< G G ^ G BG
bo3 Qq M


© $ eO b o ® : 'CO © * «4-H ©
-II . 5 co §
^ d co
2d i2
5 ©c s?°§
d *<© S o “
Sc S2 *4-1 ^
••-J§ o" =O d 6d 3.2 « = 2 cO © o ©
C d _ m
d _ o d • H , b ID
•S.2 S. o S g
* » js£
2 °*s
«+H
*d
CO W
*2 o
d n
o 3 . 9 >» o bo^S
g?3 gS
co be m CO ®
M>< © d d «
d0 . . , co f l
> g £> .2 ® § d © © ^
c
o
o
P-t O *^ 2CO2C-t ©
CO <D E>
© w
.S ^ J d 4-3 tu .-j
d d © a ° § S o
-•p-» d © £a2
!* " § •g O ft
|s§ p-t
CO
© § co °
p be co
0 d H ® ,2
©° s . *4-1 t»
>>0

-rH9*2
> I . S « a co d d ©
a 10 B 2x5^2 g
tr M * s £ 5 3 £ p co
d w d
O JO o in © OS °Jd *
*Q S3 — © © ri <-dI O .2 d o ©cd C O
H
CO © ©
d d .2
3
§* s|a
.H D o bi) D
d d - ad 0 ©
co ©
©
bo be 4-3
cO cO O ©
be A X 3 © b e co o
go o © h cO *-< 1 -J 5 pH O ^
<D
CO a as t+H d
o o a s a S.S %■* 4"» O 52*

—I > d £.ll

So *a a K a
< £
Pi .© ^2
s* P* ©
c
© P
Ph
©g © ©
© (4
>> a
■p Pi © o
d ® ^
d © 2
cs ah
©
© a< gsr s® ^^ d
2*3
o d
•-» o
• • 1
© d d
d ** be__
© © co^
o o
l | g |1 -** rt ©
: J as
g
S
B
d
J dl ig§ sp2
fl-
C 1=1
O^ O
d _
5)S al
© o
d
§?g
g § o
s 2 s
'S
U “ 3
u S 2 S tS s a a S a S S.S *o -t5 >»©
co bo d
VII.9 Supplements 611
Table VI1.8
612 Supplements VII.9
Table V II.9
VII.9 Supplements 613

The numerical value a of the physical quantity A equals


the ratio of this quantity to its unit [A I:

' If the unit of a given physical quantity changes n times


(L4]' = n [A]), then the numerical value a of this quantity
changes 1In times:
_ A _ A _ a
a ~ J A Y ~ ~ ~n\A\~ ~n
In other words, the ratio of the numerical values of
the same quantity measured in different, but homogeneous
units equals the inverse ratio of these units:
a' _ [A] _ 1
a ~ [A]' ~ n

For example, the length of a rod measured in metres is


1.53 m (a = 1.53). The length of the same rod in centimetres
(n = 10“2) is 153 cm (a' = 153), and
a' 153 1
a 1.53
100 = 10-2

If in measuring the physical quantity A we use different


units [AI and [AY, and [AY = n [A], then

\ y [A]' = a [A\ = a ) ^ -

where the fraction [AYIn expresses the unit [A] through


the unit [AY. This relationship is used not only for express­
ing quantities in multiple or submultiple units in a given
system of units, but also for conversion from one system
of units to another or to non-system units.
Example 1. Expressing the speed of 36 km/h in metres
per second, we have
36 km/h = 36 x ■3600 s = 10 m/s

In the numerator of the fraction, one kilometre is expressed


in metres, and in the denominator one hour is expressed in
seconds.
614 Supplements VII.10

Example 2. Express the pressure of 1 kgf/cm2 in pascals:

1 kgf/cm2 « = 9.81 x 104 Pa

2°. In direct measurement, a quantity is directly compared


with a measure. The pointer or scale of the measuring device
permits us to assess the value of the quantity being mea­
sured. The measurement of length with a ruler, of time with
a stopwatch, and of mass with the aid of a beam balance and
weights are examples of direct measurements.
In indirect measurement, the value of a quantity is cal­
culated according to the results of direct measurements of
other quantities with which the one being measured is asso­
ciated by a definite functional relationship. For example,
the average density of a body can be determined according
to the results of direct measurements of its mass and volume,
and the work done by a force according to direct measure­
ments of the force, the displacement, and the angle between
the force and the displacement.
Depending on the method used, some quantities can be
measured either directly or indirectly. For example, the
volume of a bolt can be measured directly if it is immersed
in a liquid poured into a measuring glass. The volume of
the bolt can be measured indirectly according to the results
of direct measurements of its length, diameter, etc.

10. Errors in M easuring P h ysica l Q uantities

1°. In measuring any physical quantity, it is impossible


in principle to determine the true value of this quantity.
Errors of measurement may be associated with technical
difficulties (the imperfection of measuring instruments,
the limited possibilities of the human eye which in many
cases takes readings of instruments, etc.), and with numer­
ous factors that are difficult or impossible to take into ac­
count (fluctuations of the ambient temperature, the motion
of air streams near a measuring instrument, vibrations of
VII.10 Supplements 615

the measuring instrument together with the laboratory


table, etc.).
The difference between the measured and the true values
of a physical quantity is called the error of measurement.
2°. Methodical (systematic) errors are due to the short­
comings of the measuring method used, to the imperfection
of the theory of the physical phenomenon which the quantity
being measured relates to, and to the lack of accuracy of
the formula used for calculations. For example, in weighing
a body on an analytical balance, the methodical error may
be associated with the fact that no account is taken of the
difference in the buoyant forces exercised by the surrounding
air on the body and on the weights.
Methodical errors can be reduced by changing and improv­
ing the method of measurement, and by introducing cor­
rections into the formula used for calculations making it
more accurate.
3°. Instrumental errors are due to imperfection of the
design and inaccurate manufacture of the measuring instru­
ments. For example, the running of a stopwatch may change
upon sharp fluctuations in the temperature, the centre of
its dial may not accurately coincide with the axis of rota­
tion of its hand, etc.
Instrumental errors are reduced by using more accurate
(but also more costly) instruments. It is impossible to com­
pletely eliminate instrumental errors.
4°. Random errors are due to many factors that cannot
be taken into account. For example, the readings of a sen­
sitive beam balance may be affected by vibrations of the
building produced by vehicles running down the street, by
dust particles settling on the balance pans during the weigh­
ing, and by the elongation of one half of the balance beam
near the experimenter’s hand.
It is impossible to completely eliminate random errors,
but they can be diminished by repeating the measurements
many times. As a result, the factors leading to an increase
and a decrease in the measuring results may partly com­
pensate one another.
Random errors are calculated on the basis of the theory
of probability and are beyond the scope of elementary
courses in physics and mathematics.
616 Supplements VII.10

5°. The arithmetic mean A m of n measurements is taken


as the result of measuring a physical quantity:
n

2 *
A — i= 1
Am~ n

The magnitude of the deviation of the result of the j-th


measurement A t from the arithmetic mean A m
AAi = | Am A it
is called the absolute error of the given measurement.
The mean absolute error A^4mof a series of n measurements
is defined as the quantity equal to

The accuracy of measuring physical quantities is com­


pared by calculating the relative error E :
__

It is usually expressed in per cent.


The final result of measuring the physical quantity A
is given in the form
A = A m ± AA
the greater of the mean absolute and instrumental errors
being taken as the absolute error A^4 (in stricter calcula­
tions the error AA is chosen from a comparison of the random
and instrumental errors). The latter equation indicates
that the true value of the measured quantity ranges from
A m — AA to A m -f AA.
6°. The so-called class of accuracy {precision) is indicated
on the scales of many measuring instruments. Its symbol
is the relevant figure in a circle. The class of accuracy deter­
mines the absolute instrumental error as a percentage of
the maximum value of the quantity that can be measured
by the given instrument. For example, an ammeter has a
VII.ll Supplements 617

scale from 0 to 5 amperes and its class of accuracy is 1.0.


The absolute error of measuring a# current with such an
ammeter is 1.0% of 5 amperes, i.e. A /lnst = ± 0 .0 5 A.
If the class of accuracy is not indicated on the scale of an
instrument, then the absolute error of the instrument is
usually taken equal to half the value of the smallest grad­
uation on its scale. For example, the absolute error in meas­
uring length with a millimetre ruler is often taken equal
to ± 0 .5 mm.
In determining the absolute error of an instrument ac­
cording to the value of its graduations, one must give atten­
tion to how the instrument is used for making measurements,
with what and how the results of measurements are record­
ed, what is the distance between adjacent marks on the
instrument scale, etc. If, for example, the distance from
the floor to a weight suspended on a thread is being measured
using a millimetre ruler without any indicators, aiming
devices and the like, then the absolute error of measurement
cannot be taken less than one millimetre. The instrumental
error is also taken equal to the value of a graduation when
the graduations are very close to one another on the scale
of the instrument, when the indicator of the instrument is
not a smoothly moving pointer, but a “jumping” one (like
the hand of a stopwatch), etc.

11. Processing the Results


of Direct Measurements

In the direct measurement of the physical quantity A,


the following operations are performed:
(a) the quantity is measured n times (A1? A 2, . . ., A n),
(b) the mean value A m of the measured quantity is found:

2 *
i= 1
Am
n
618 Supplements VII.11

(c) the absolute errors AAt of each measurement and


the mean absolute error of the entire series of measurements
are found:

2
Ai4m

(d) the result is written in the form


A — A m ± AA
and either the instrumental error or the mean absolute error
is taken as the absolute error of the result AA, depending
on which of these errors is greater;
(e) the absolute error of the result is rounded off to one
significant digit, and the mean value of the measured quan­
tity is rounded off or determined more precisely to the digital
order remaining in the absolute error after rounding off;
and
(f) the relative error of the result is computed:

Example. The diameter d of a ball was measured five


times with the aid of a micrometer whose absolute error is
Adlnst = ±0.01 mm. The results of measuring the diameter
of the ball are dx = 5.27 mm, d2 = 5.30 mm, d3 = 5.28 mm,
dk = 5.32 mm, and d5 = 5.28 mm.
The mean value of the ball diameter is
5.27 + 5.30 + 5.28 + 5.32 + 5.28
dm — 5
= 5.29 mm

The absolute errors of the measurements are Adx = 0.02 mm,


Ad2 = 0.01 mm, Ad3 = 0.01 mm, Adk = 0.03 mm, Ad5 =
= 0.01 mm, and the mean absolute error is
A, 0.02 + 0.01 + 0.01 + 0.03 + 0.01 A no
Aam= -------- 1--------—g—1------- « 0.02 mm

Since the mean absolute error is greater than the instru­


mental one, the result of measurement is
d = (5.29 ± 0.02) mm
VIL12 Supplements 619

The relative error in measuring the diameter of the ball is


£==-£t§ as 0.04 = 0.4%

12. P r o c e s s in g th e R e s u lts
o f I n d ir e c t M e a s u r e m e n ts
Let us assume that the physical quantity A and certain
other quantities A 2, . . ., A h have the following func­
tional relationship:
A — f {Ah A 2, • • •* Ak)
The quantities A x, A 21 . . ., A k may include directly meas­
ured quantities A dti, tabulated quantities A tti (whose values
in the given experiment are taken from a table instead of
being measured), and the so-called equipment data A eti
(characteristics of the experimental equipment known be­
forehand and1 not measured in the given experiment).
As a result of processing all the directly measured quan­
tities A ^ i (VII.11), for each of them the mean value C4di*)m,
the absolute error AAd,*, and the relative error E&ti will
be found.
The result for the indirectly measured quantity A in
the simplest cases is obtained according to the following
procedure:
(a) The mean value A m is calculated according to the
mean values of the quantities which the quantity A being
measured depends on:
= / (^1, mi ^2, mi • • •» Akt m)
(b) The relative error E is calculated according to the
form of the functional dependence of the quantity A on
the directly measured and tabulated quantities, and also
on the equipment data. For the simplest functional rela­
tionships, the formula for calculating the relative error E
usually has the form
E = c1El -f- c2E 2 -f- cuEk
where clt c2, . . ., ch are integral or fractional dimensionless
coefficients, and Eu E 2j . . ., Ek are the relative errors
in measuring the quantities A u A 2, . . ., Ak-
620 Supplements VII.12

Table VII.10
Dependence of
quantity A on Relative error E Absolute error AA
other quantities

+ AA2
11

A ^ + AA,
to

^1 + ^2
A^i + A A 2
A = A 1— A 2 A A x -\-A A 2
At A 2
Ai4x ttA 2
A t A A 2-\-A 2 A A l
II

A t ' A 2

a — A l .
Ai4j AA2 A.4, t Aj Ai42
A i 1 A 2 A 2 ' A\
A A i
AAX
II

A t
A = sin A 1 AA ^ cot A-^ AA± cos A x
A = cos A ! AAt tan A x AAt sin A l
1 AA x 1 AA x
A = log A x
2 .3 0 | log A t | A x 2 .30 A t

Table VII.10 gives formulas for calculating the relative E


and absolute A^4 errors in an indirectly measured quantity A
when the latter is related to the other quantities by the
simplest functional relationships most often encountered
in problems in the elementary course of physics.
The tabulated values are taken with such an accuracy
that their relative errors will be smaller than the remaining
relative errors in the preceding formula.
If the values of the equipment data are not written in
a generally adopted form (V II.lid), then it is usually con­
sidered that they are measured with an accuracy equal to
half of the last digital order indicated.
(c) The mean value of the measured quantity A m and the
relative error E are used to calculate the absolute error A^4
of the result:
kA = EAm
and it is rounded off so as to contain only one significant
digit.
V11.12 Supplements 621

(d) The final result of measurement is written in the form


A = A m + &A
The mean value A m is rounded off or defined more precisely
up to the digital order remaining in the value of the absolute
error AA after it has been rounded off.
Example 1. Determine the density p of a homogeneous
body according to the results of direct measurements of its
mass m and volume V: m = (25.4 ± 0.5) X 10"3 kg and
V = (2.94 ± 0.05) x 10-6 m3.
The mean value of the density of the body is

P » = T ^ = S I ^ ^ 8-6 5 x l0 3 k g/m3
The relative error in measuring the density Ep equals
the sum of the relative errors in measuring the mass Em
and the volume E v (see Table V II.10):
0.5x 10"3
Ep = Em-{- E v = 25.4 X 10"3
0.05 X 10-6
2.94 X 10"6
» 0.02 + 0.02 = 0.04 = 4%
The absolute error in measuring the density Ap is
Ap = pm^p = 8.65 x 103 x 0.04 « 0.3 x 103 kg/m3
Rounding off the value of pm, we write the final result
in the form
p = (8.6 ± 0.3) x 103 kg/m3
Example 2. Determine the volume V of a cylinder accord­
ing to the results of direct measurements of its diameter d
and altitude h : d = (3.46 ± 0.04) X 10"2 m, h = (4.87 ±
± 0.05) X 10”2 m.
Before calculating the mean value of the volume of the
cylinder by the formula
I
Vm= ^>dmhm

we shall assess the accuracy with which we must take the


quantity jt from the relevant table (for example, we have at
our disposal the tabulated value of n « 3.141 593).
622 Supplements VIU 2

The relative error of the result E v in the given case is


(see Table VII.10)
E v = E n + 2E d -f- E h
where
p 0.04 X 10-2 0.05 x 10-2
d 3 .4 6 x 1 0 -2
0.01 and Eh 4.87 X 10-2
0.01
i.e. E y = Efi -f- 0.03
For the inaccuracy in the value of the number jt not to
appreciably detract from the result of measurement, we must
have E n < 0 .0 1 . If, for example, we take the number n
with] an accuracy to tenths, then E n = 0.04/3.1 « 0.01,
and the relative error in the result will be E v = 0.04.
If we take the value of jt with an accuracy to hundredths,
then E n = 0.002/3.14 « 0.0006, and E v » 0.03.
Consequently, in the given case without any harm to the
accuracy of the result, we can assume that jt = 3.14. Hence,
Fm= i - X 3.14 (3.46 x 10-2)2 x 4.87 X lO*2 « 45.9 x IQ r* m3
AF = V mE v = 45.9 x lO"6 x 0.03 « 1 x 10~6 m3
V = (46±1) X 10“6 m3
Example 3. Determine the period of oscillations of a
mathematical pendulum whose length is not measured di­
rectly and is considered equal to I = 2.5 m.
In these conditions, we assume that the absolute error
of measuring the length of the pendulum is AI = ± 0.05 m,
and the relative error E x = 0.05/2.5 = 0.02.
The period of oscillations of a mathematical pendulum
is T = 2 n Y U and the relative error in its measurement is
Et = En± (E i 4- Eg)
The quantity jt in this case can be taken equal to jt =
= 3.14 because E n « 0.0006 (see Example 2), and E n <
< E i/ 2 = 0 . 01.
Now we must solve the question as to the accuracy with
which we should take the value of the acceleration of free
fall g (the tabulated value is g = 9.806 65 m/s2 for the
V1L1S Supplements 623

given locality). If we adopt the value of g = 9.8 m/s2,


then E g -■= 0.01/9.8 « 0.001, and <Eg i.e. such an
accuracy in the given case is quite acceptable, and the rela­
tive error of the result can be considered equal to E T « 0.01.
The period of oscillations of the pendulum is

and the absolute error is

AT = T E t = 3.2 x 0.01 a? 0.03 s


To write the final result, the value of the period of oscilla­
tions of the pendulum must be defined to hundredths of a
second:
T = (3.20 ± 0.03) s

13. Approximate Calculations


Without Accurate Account Taken
of Errors
1°. In processing numerous measurements, the errors of
separate results are often not computed. The error of the
approximate value of a quantity is assessed by indicating
the number of correct significant digits in this value.
Zeros to the left of a number are not considered to be
significant digits. Zeros at the middle or to the right of a
number signifying the absence in the number of digits of
the corresponding orders are significant digits. For example,
in the number 0.080 40, the first two zeros are not significant
digits, while the third and fourth zeros are significant ones.
Zeros placed at the end of an integral number instead
of unknown digits and serving only to show the digital
orders of the remaining digits are not considered to be sig­
nificant. In such cases, it is good practice not to write zeros
at the end of a number, but to replace them with the cor­
responding power of the number 10. For example, if the
number 4200 has been measured with an absolute error of
624 Supplements VII.13

± 100, then it must be written in the form 42 X 102 or


4.2 X 103. Writing in this way stresses that the given num­
ber contains only two significant digits.
2°. If the approximate value of a quantity contains
superfluous or doubtful digits, then it is rounded off: only
the true significant digits are retained, and the superfluous
ones are discarded. The following rules for rounding off
are observed:
(a) if the first digit to be discarded is greater than 4,
then the last digit being retained is increased by unity.
For example, in rounding off the number 27.3763 to hund­
redths, we should write 27.38;
(b) if the first digit to be discarded is 4 or less, then the
last digit being retained is not changed. For example, in
rounding off the number 13 847 to hundredths, we write
138 X 102; and
(c) if the part of a number being discarded consists of
the single digit 5, then the number is rounded off so that the
last digit retained will be an even one. For example, when
rounding off to tenths, 23.65 « 23.6, but 17.75 « 17.8.
3°. In performing various mathematical operations with
approximate numbers, we must observe the following rules
for calculating digits:
(a) In addition and subtraction, the same number of
decimal places is retained in the result as are present in the
number with the minimum amount of decimal places.
(b) In multiplication and division, the same number of
significant digits is retained in the result as are present in
the approximate number with the smallest significant digits.
An exception from this rule is permitted when one of
the multipliers of a product begins with unity, while the
multiplier containing the minimum number of significant
digits begins with a different figure. In these cases, one digit
more is retained in the result than there are digits in the
number with the smallest significant digits.
(c) The result of calculating the values of the functions
xn, x1/n, and the logarithm of the approximate number x
should contain the same number of significant digits as
there are in x .
When calculating intermediate results, one digit more
is retained than is recommended by rules (a) to (c) (the so-
VII.13 Supplements 625

called spare digit). The spare digit is discarded in the final


result. •
If some approximate numbers have more decimal places
(in addition or subtraction) or more significant digits (in
multiplication, division, raising to a power, extracting a
root, etc.) than others, they are preliminarily rounded off,
retaining only one superfluousj digit.
Example 1. Before adding the approximate numbers 0.374,
13.1, and 2.065, the first and third of them must be rounded
off to hundredths, and the hundredths discarded in the final
result:
13.1 + 2.06 + 0.37 » 15.5
Example 2. The result of calculating the expression
(68.04 x 7.2)/20.1 should contain only two significant
digits (according to their number in 7.2):
68.04x 7.2 68.0X7.2
20.1 20.1
24.4 24

Example 3. The result of multiplying the numbers 13.27


and 0.84 can be written with three significant digits [see
the exception to the rule (b)I:
13.27 x 0.84 ft? 13.3 x 0.84 ft? 11.2 (and not 11)
Example 4. In raising the approximate number 216 to
the third power, the result should be written only with
three significant digits:
2163 ft? 101 X 10B
Name Index

Ampere, A. M., 291 Heisenberg, W., 487


Archimedes, 114 Hooke, R., 69
Avogadro, A., 154
Kelvin, W. T., 173
Bernoulli, D., 118 Kirchhoff, G. R., 254, 431
Biberman, L. M., 481
Bohr, N., 499, 500 Lebedev, P. N., 471, 472
Boyle, R., 150 Lenz, H., 304
Carnot, S., 171 Maxwell, J., 307, 377
Charles, J., 151 Michelson, A. A., 415
Clausius, R., 173
Coulomb, C. A., 205 Newton, I., 48, 54, 57, 418, 428f
447
Davisson, C. J., 476, *>/7
De Broglie, L., 474 Oersted, H. C., 286
Einstein, A., 444, 446, 474 Pascal, B., 112
Planck, M., 433, 474
Fabrikant, V. A., 481
Faraday, M., 264, 302, 303 Rutherford, E., 494, 495, 500
Fraunhofer, J., 436
Fresnel, A. J., 418 Stoletov, A. G., 465
Sushkin, N. G., 481
Gay-Lussac, J. L., 150
Germer, L. H., 476, 477 Zhukovsky, N. E., 123
Subject Ifldex

Absorptivity, body, 430 Amplitude,


Acceleration, 26ff oscillations, pendulum, 336
amplitude, 329 resultant oscillation, 337
angular, standing wave, 374f
average, 46 stationary forced oscillations,
instantaneous, 46f 341ff
average, 26f, 31 velocity, 328
centripetal, 39 Analyser, 427
free fall, 34, 65 Analysis,
in inertial reference frame, 62 spectral, 436
instantaneous, 27 X-ray diffraction, 441
normal, 27, 39, 46 Angle,
tangential, 27 contact, 188
Accelerators, 565ff critical, total reflection, 393
linear, 565 diffraction, 422
resonance circular-orbit, 566f incidence, 392
Acceptors, 283 polar, 39
Accommodation, eye, 410 prism, 397
Acoustics, 367 reflection, 392
Action, 433 refraction, 392
Activity, radioactive substance, rotation, 47
539 solid, 405
Adiabat, 154 vision, 411
Affinity, electron, 520 Angular momentum,
Air, humidity, body, 83
absolute, 184f particle, 82f
relative, 185 Anions, 263
Amount of substance, 126, 127, Annihilation, electron-positron
591 pair, 582
Ampere, 589, 599 Anode, 263
Amplitude, Antineutrino, 544, 577
acceleration, 329 Antineutron, 581
damped oscillations, 339 Antinodes, standing wave, 374f
40 *
628 Subject Index

Antiparticles, 580ff Biology, radiation, 558


Antiproton, 581 Blackbody, 430f
Aperture, relative, 408 Body(ies),
Arc, electric, 270 absorptivity, 430
Arm, force, 80 electrified, 204
Astronomy, elongation, 198
radar, 388 emissivity, 430
radio, 388f equilibrium, 88, 115f, 147
Atom(s), 126 free, 48
Bohr model, 499ff isolated, 48
diamagnetic, 317 mass, 52f
dimensions, 127 natural frequencies of oscilla­
dynamic stability, 496 tions, 376
energy state, perfectly rigid, 18
excited, 503 rotation about fixed axis, 45ff
ground, 503 proper length, 452
normal, 503 reference, 18, 49
excited states, 503 rigid, 18
explanation of properties, 511 rotation in inertial reference
hydrogen, structure, 497 frame, 80
nuclear model, 493 thermometric, 146
orbital magnetic moment, 314f thrown vertically upward,
planetary model, 493 displacement vector, 36
Axis, motion, 35ff
optical, velocity, 36
auxiliary, lens, 401 weight, 65
mirror, 398 Boiling, 179ff
principal, Bomb,
lens, 400 atomic, 558f
spherical surface, 398 hydrogen, 564
rotation, 21 Bond(s),
chemical, 517
covalent, 279, 521
Barrier, potential, nuclear, 543 saturation property, 521
Baryons, 570, 571 homopolar, 521
Beam(s), metallic, 194
electron, 277 Broadcasting, radio, 385ff
homocentric, 395
paraxial, 398
Betatron, 565f Calorie, 162, 590
Bias, 275 Camera, 408
Subject Index 629

Candela, 406, 589 Charge(s),


Capacitance, 235ff proton, 607
isolated conductor, 236 •specific, 607
isolated sphere, 236 surface, density, 216
mutual, two conductors, surface-bound, 224
236f test, 209
parallel plate capacitor, 238 Chemiluminescence, 435
Capacitor(s), 237ff Chemistry, radiation, 558
multiplate, 238 Circuit,
parallel connection, 238 alternating current, 351
parallel plate, 237f with capacitive reactance,
plates, 237 353f
series connection, 239 impedance, 354
variable, 238 with inductive reactance, 352f
Capillaries, 190 with resistance, 352
Cathode, 263 resonance in, 356ff
Cathodoluminescence, 434 electric, 253
Cations, 263 LC, 346
Centimetre, 588 oscillator, 346
Centre(s), open, 383
gravity, 65 RLC, 346
inertia, 53 Circulation, velocity, in fluid,
mass, 53 123
closed system, 59 Clock(s),
optical, auto-oscillating, 343f
lens, 401 paradox, 455
mirror, 398 synchronization, 448
pressure, 115 Cloud, electron, 274, 504
vaporization, 180 Coefficient,
Chamber, cubic expansion, 150f
bubble, 541f linear expansion, 198
Wilson cloud, 540f quasi-elastic force, 69
Cbarge(s), secondary electron emission,
electric, 203 272f
electron, 204, 607 sliding friction, 71
specific, 607 static friction, 71
elementary, 204 temperature, resistivity, 252
f* induced, 220 thermal pressure, 151
interacting, potential energy, volume expansion, 198
228f Collision,
point, 204 central, 105
630 Subject Index

Collision, Constant,
completely inelastic, 108f fine structure, 573
perfectly elastic, 105f gas,
Compression, uniaxial, 196 molar, 154, 607
Concentration, current carriers, specific, 154
245 gravitational, 64, 607
Condensation, 178f magnetic, 603
rate, 178 Planck’s, 433, 607
Condition, frequency, 501f, 504 Rydberg, 499, 503, 607
Conductance, spring, 69
parallel connection of conduc­ Coordinate, angular, 39
tors, 256 Core, reactor, 554
unipolar, diode, 275 critical dimensions, 555
Conduction, critical mass, 555
electric, 243 Counters,
electron impurity, 282 Geiger, 540
hole impurity, 283 scintillation, 539f
impurity, 281 ff Couple, 86
intrinsic, arm, 87
activation energy, 280 forces, 86
electron, 280 moment, 86f
hole, 281 Crystallization, 201
rc-type, 280 Crystal(s), 134f
p-type, 281 atomic, 193
Conductivity, 249 elementary cell, 135
ionic, 263 ionic, 193
metal, 252 lattice, 135
Conductors, 219, 243 period, 135
first kind, 256 metallic, 194
parallel connection, 256 molecular, 194
second kind, 263 valence, 193
series connection, 255 Curie, 607
straight, 293 Current, 244
Constant, active power, 355
Avogadro, 127, 607 alternating, 351
Boltzmann, 142, 607 effective value, 355
decay, 537 instantaneous power, 355
dielectric, 206f mean power, 355
diffraction grating, 423 power factor, 356
electric, 206, 599 resonance in circuit, 356ff
Faraday, 265, 607 appearance time, 245
Subject Index 631

Current, Cycle,
average density vector, 244 reverse, 170
carriers, concentration, 245 reversible, thermal efficien­
conduction, 243 cy, 171
density in metals, 245 Cyclotron, 566ff
convection, 243
displacement, 243
eddy, 308 Decay,
retarding action, 308 alpha, 536, 542f
electric, 243 beta, 536, 543ff
direction, 244 Decibel, 611
in vacuum, 271 Deformation(s),
electron, in atom, 314 elastic, 195
in gases, 266 plastic, 195
induced, 302f relative, 195
molecular, 315 residual, 195
in parallel conductors, inter­ solid body, 194
action, 296f Degree,
photoelectric, 465 Celsius, 591
power, 262 Kelvin, 591
quasistationary, 384 Demodulation, 387
saturation, 267f, 274 Density, 54, 127
steady, 244f average, 54
calculation of multiloop net­ displacement current, 377
work, 254f energy, 242, 312f, 366
density, 251 saturated vapour, and temper­
thermionic, 272, 273 ature, 183f
Curve(s), steady current, 251
energy distribution, in black- surface charge, 216
body spectrum, 432 Detector, 386
resonance, 343, 356f Diagnostics, X-ray, 440
technical magnetization, 321 Diagram,
Cycle, 169ff stress-strain, 196
backward, 170 thermodynamic, 149
Carnot, 170ff p-7\ 151f
closed, 175 p-F, 149, 159
direct, 170 Diamagnetic(s)
forward, 170 demagnetizing, 318
irreversible, thermal efficien­ magnetizing, 318
cy, 171 relative permeability, 3161
open, 175 Diameter, collision, 133
632 Subject Index

Diascopes, 408f Dipole(s),


Dielectric(s), 221, 243 induced, quasi-elastic, 222
breakdown, 221 rigid, 223, 225
homogeneous, 221 Discharge,
isotropic, 221 arc, 270
non-polar, 222 brush, 269
polarization, 223f corona, self-maintained, 269
polar, 223, 225 gas, 266
polarization, 223, 225f self-maintained, 268ff
ionic, 226 semi-self-maintained, 267f
orientation, 226 voltage-ampere characteris­
polarized, 223 tic, 267
Diffraction, glow, 268f
conditions for, 420 spark, 269f
electrons, 477f Dispersion, light, 427ff
grating, 423 abnormal, 428
light, 420ff anomalous, 428
minima, 423 normal, 428
order, 423 Displacement, 21f, 43
neutrons, 480 Dissociation,
peaks, 422f degree, 264
order, 423 electrolytic, 263f
principal maxima, 424 Distance, 22f, 30f, 33f, 41
principal minima, 424 and displacement, 22f
spectrum, 424 in free fall, 35
Diffusion, 128f normal vision, 411
Diode, Domains, 323f
anode current, 275 Donors, 282
anode voltage, 273f Dose,
directly heated, 273f equivalent radiation, 560f
hot-cathode, 274 exposure, 560
semiconductor, 285 radiation, 559
unipolar conductance, 275 Drag, 124
volt-ampere characteristic, 273f Dynamics, 48
Diopter, 403 relativistic, 457
Dipole(s), Dynamometer, spring, 75
electric, 130, 216, 383 Dyne, 593
moment, 217
electrostatic field intensity,
216f Effect,
Hertz, 384 latitude, 575
Subject Index 633

Effect, Electron(s),
Magnus, 123 of bit, 497, 505
photoconductive, 464 orbital magnetic moment, 314
photoelectric, 464ff orbitals, 510
nuclear, 547 peripheral, 511
photoemissive, 464f shell, 510
threshold frequency, 466 spin, 315f, 508
piezoelectric, 370 subshells, 510
tunnelling, 542f systematization of states, 506
Efficiency, total energy, 502
heat engine, 176 valence, 126, 511
thermal, 171 dielectric, 222
thermodynamic, 171 velocity on orbit, 497
Elasticity, 195 Electron-volt, 603
cubic, gases, 359 Electrostatics, 203
of form, 359 Element(s),
Electrode(s), 263 periodic table, 510ff
control, 275, 276 relative atomic mass, 591
corona, 269 transuranium, 549
Electrodynamics, 208 Elongation, 198
Electroluminescence, 434 Emission,
Electrolysis, 263 photoelectric, 272
Electrolytes, 263 secondary electron, 272
Electromagnet, 296 thermionic, 272
Electron(s), 204, 571 Emissivity, 430
acceleration, centripetal, 497 Emitter(s), 272, 381, 383
affinity, 520 ultrasound, 369
charge, 204, 607 Energy,
classical radius, 569f activation,
collective, 194, 219 intrinsic conduction, 280
conduction, of metals, 219f nuclear fission, 552f
distribution in atoms, 512 bond, 519
energy levels, 484, 502f charged capacitor, 241
exchange interaction, 522 charged conductor, 240
free, 219 conversion, irreversible in con­
in metals, 245, 251 ductor, 262
gas, in metals, 194 current,
intrinsic mechanical angular intrinsic, 312
momentum, 315 proper, 312
mass, 204, 527 definition, 101
motion in potential well, 481f dissociation, 519
634 Subject Index

Energy, Energy,
electric, plasma sources, 271 transfer, 161f
electric field, 241 wave, mean volume density,
volume density, 242 366
electromagnetic field, 313 Engine, heat 174ff
volume density, 313 Epidiascopes, 408, 409
electron, 502 Episcopes, 409
fission, 551f Equation(s), see also Formula
internal, 108, 156ff Bernoulli’s, 118f
charged conductor, 240 horizontal flow tube, 118
wave, 380 continuity, steady flow of fluid,
zero, 157 118
kinetic, 102 Einstein’s photoelectric, 467
body, 102 kinetic theory of gases, funda­
gas molecule, mean, 142 mental, 141
particle, 102, 460 Mendeleev-Clapeyron, 154
system, 102 motion, particle, 19
localized particle, 490 spherical mirror, 399
magnetic field, 312 spherical wave, 367
solenoid, 312 of state, 147
volume density, 312f ideal gas, 154f
mechanical, lOlff thin-lens, 401f
conservation law, 105f, 108 wave, plane, 365
molecule, and spectrum, 523 monochromatic, 379
nuclear, 557 standing, 374
potential, 103 Equilibrium, 84
elastic interaction, 104 body in liquid, 115f
gravitational interaction, 103f dynamic, 267
interacting charges, 228f vaporization and condensa­
interaction of two molecules, tion, 179
132f indifferent, 93f
zero point, 103, 104 liquid particle, 113
proper, 459 mobile, 179
charged conductor, 240 neutral, 93
quantum, 433 rigid body, 88
rest, 459 stable, 91f
surface, 187f and potential energy, 94f
total, 156 two molecules, 133
conservation law, 108 unstable, 92f
electric, 241 Equivalent, electrochemical, 265
system, 156 Erg, 595
Subject Index 635

Error(s), Ferroelectries, 226


absolute, 616 Ferrcftnagnetic(s), 317, 320ff
mean, 616 magnetic reversal, 322
instrumental, 615 magnetization, 320
methodical, 615 saturation, 321
random, 615 relative permeability, 321
relative, 616 Field(s),
systematic, 615 electric, 208
Evaporation, 178 Coulomb, 246
rate, 178 energy, 241
Event(s), 447 extraneous, 246
simultaneity, 447 intensity, 246
relative nature, 449f homogeneous, energy, 241
Expansion, thermal, 197ff induced, 307
linear, 197ff properties, 307
volume, 198 intensity in conductor with
Experiment(s), current, 247
Millikan, 211 vortex, 307
Oersted’s, 286 electromagnetic, 208
Rutherford’s, 494f electrostatic, 208
Stoletov’s, 465 point charge, 214
Eye, 409 potential, 229ff
accommodation, 410 sphere, 215
resolving power, 411 system of point charges, 214
uniformly charged infinite
Factor, plane, 215
damping, 339, 349 gravitational, 67
neutron multiplication, 554 intensity, 67
power, 356 source, 67
refrigerating, 177 stationary, 64
safety, 197 homogeneous, 209
Fall, free, 34f magnetic, 285ff
acceleration, 34 circular loop, 293f
distance, 35 electric current, 292ff
velocity, 35 energy, 312f
Family, radioactive, 536f homogeneous, 288
Farad, 605 non-stationary, 286
Farsightedness, 411 non-uniform, 288
Feedback, solenoid, 295
negative, 344 stationary, 286
positive, 344 straight conductor, 293
636 Subject Index

Field(s), Force(s), 50ff, see also Interac­


magnetic, tion^)
uniform, 288 Ampere, 291f
varying, 286 Archimedean, 114
non-potential, 289 arm, 80
non-stationary, 208 attraction, between molecules,
potential, 229 129ff
radiation, 381 balancing, 84
stationary, 208 buoyant, 114f
varying, 208 central, 64
vortex, electric and magnetic, centripetal, 56
377 on charge in electrostatic field,
Fission, nuclear, energy, 551 ff 226
Flaw detection, coercive, 322
ultrasonic, 370 Coulomb, 205
X-ray, 440f drag, 124
Flow, driving, forced oscillations,
laminar, 116 340f
steady, 116, 117f elastic, 68f, 195
tube, 116 in solids, 131
turbulent, 116 electromotive, 247f
unsteady, 116 current source, 251
vortex, 121 effective value, 355
Fluid, see also Gas(es), Liquid(s) electromagnetic induction,
compressible, 111 303
definition, 110 external, 352
dynamics, 110 induced in moving conduc­
ideal, 117 tor, 304ff
incompressible, 111 mutual induction, 310
mechanics, 110 self-induction, 308
non-viscous, 117, 118 sinusoidal, 350
statics, 110, lllf f electrostatic,
velocity circulation, 123 attraction, 205
velocity measurement, 120f potentiality, 227
viscous, 117 repulsion, 205
Fluidity, 111 external, 51
Flux, extraneous, 246
luminous, 404f friction, 70ff
magnetic, 289f static, 70f
magnetic induction, 289f maximum, 70
F-number, 408 gravitational, 51, 63f
Subject Index 637

Force(s), Freezing, 201


gravity, 65 Frequency,
inertial, 77 carrier, 386
interaction, modulation,! 386
bodies, 62f oscillations, 332, 339
particles, 62f natural, 376
internal, 51 resonance, cyclic, 343, 356
lifting, 123 revolution, 41, 45
line of action, 50 threshold, photoemissive effect,
Lorentz, 297f 466
generalized, 299 Friction,
measurement, 72f, 75 contact, 70
non-central, 532 dry, 70
non-potential, lOOf external, 70
normal pressure, 71 internal, 70, 116ff
normal reaction, 71 in gases, 117
nuclear, 53Iff in liquids, 117
range, 531 liquid, 70
saturation property, 532 sliding, 70
potential, 64, 99f static, 70
reaction, constraints, 85 viscous, 70
repulsion, of molecules, 130f Function, work, 272, 467
restoring, 331, 333f
resultant, 51, 84
two antiparallel forces, 86 Gas(es), 134, 265, see also Fluid,
two parallel forces, 86 s Vapour
short-range, 130, 531 cubic elasticity, 359
standard, 72 degeneracy, 491ff
support reaction, 69 temperature, 491
transfer of application point, discharge, see Discharge, gas
84f electric breakdown, 268
vector components, 87f electron,
Formula, see also Equation(s) degeneracy, 492
Balmer-Rydberg, 498f in metals, 194
de Broglie’s, 475 ideal, 136f
Maxwell’s, 377f work of expansion, 158f
light pressure, 471 internal friction, 117
Thomson, 349 ionization, 266
Torricelli’s, 119 liquefaction, 184
Frame(s), reference, see Refer­ pressure, 142
ence frame(s) rarefied, 137
638 Subject Index

Generator, valve, 358f Humidity, air, 184f


Glass, magnifying, 412 Hygroscopicity, 191
Gramme, 590 Hyperons, 570, 572, 580
Graph, rest mass, 580
displacement, 22, 291 Hypersounds, 369
distance, 22, 33f Hypothesis,
velocity, 25f, 28f Ampere’s, 315
Grating, diffraction, 423 Planck’s, 433f
constant, 423 Hysteresis, magnetic, 321
spacing, 423
Gravity, zero, G6
Gray, 559, 560 Illuminance, 406
Grid, 275, 276 Illumination, 406
Gun, electron, 278 Image,
inverted, 400
negative, 473
Hadrons, 573 point, 395
Half-life, nuclei, 538 positive, 473
Harmonics, 368 real, 395
Head, resolved points, 411
kinetic, 118 upright, 400
total, 118 virtual, 395f
Heat, 160ff Impedance, alternating current
amount, 161 circuit, 354
capacity, 162ff Impulse, 55f
molar, 163 Index, refractive,
specific, 163 absolute, 391
exchange, 161 and light frequency, 428f
freezing, specific, 201 relative, 392
fusion, specific, 200 Inductance, 309
liberated in conductor, 262 mutual, 311
mechanical equivalent, 162 Induction,
pump, 177 electromagnetic, 302f
sink, 170 electrostatic, 220
source, 170 mutual, 310f
sublimation, specific, 201 Inertia, 52
vaporization, specific, 181, 201 Infrasounds, 368
and work, 161 Instruments, optical, 407ff
Henry, 610 Insulators, 243
Hertz, 593 Intensity,
Hole, positive, 280 electric field, 209
Subject Index 639

Intensity, Ionosphere, 270


electrostatic field, Ions? 263
dipole, 216f charge, 263
point charge, 213f hydrogen-like, 503
and potential, 232 recombination, 264
sphere, 215 Isoprocesses, 150
system of point charges, 214 Isotherm, 150f
uniformly charged infinite vapour, 182f
plane, 215 Isotopes, 526, 557
ionization, 266 radioactive,
luminous, half lives, 538
mean spherical, 406 mean lifetime, 537
point source, 405 uses, 557f
Interaction(s), see also Force(s)
classification, 50f
electromagnetic, 50, 208, 572f Joule, 595
gravitational, 50 Junction,
intermolecular, 129ff branched circuit, 253f
strong, 51, 572f electron-hole, 283
weak, 51, 572, 573f p-n, 283
Interference,
constructive, 416f
destructive, 416f Kaons, 571, 579f
light, 416ff rest mass, 579
pattern, 416 Kelvin,’ 589
in thin films, 418 Kenotron, 275
waves, 373 Kilogram, 589
Interferometer, Rayleigh, 419f Kinematics, 17ff
Interval, time, between events, relativistic, 457
453
Invariance,
Lorentz, 451 Lamps, daylight, 435
relativistic, 451 Lasers, 513ff
Ionization, Lattice, crystal, 135
collision, 266 period, 135
degree, 270 points, 135
gases, 266 sites, 135
impact, 266 Law(s), see also Rule(s)
surface, 268 Ampere’s, 291f
volume, 268 Avogadro’s, 154
work, 266 Boyle’s, 150
640 Subject Index

Law(s), Law(s),
Brownian motion, 128 Ohm’s
Charles’, 151 arbitrary circuit section, 249
communicating vessels, 114 circuit with e.m.f., 351f
Coulomb’s, 205 density of current in metals,
electric charge conservation, 251
204 electric circuit, 250
electrolysis, section without e.m.f., 250
combined, 265 Pascal’s, 112f
first, 264f photoemissive effect, 466
second, 265 radioactive decay, fundamental,
energy conservation, 105f, 108 537f
Faraday’s, rotational dynamics, funda­
electromagnetic induction, mental, 82, 83f
303 thermodynamics, 144f
first, 264f first, 163ff
second, 265 second, 173f
Gay-Lussac’s, 150 third, 174
Hooke’s, 69, 196 universal gravitation, 63f
illumination by point source, vector addition, 22
406 velocity addition,
Joule-Lenz, 262 in Newtonian mechanics, 62
Kirchhoff’s, for thermal radia­ relativistic, 456
tion, 431 Layer,
Lenz’s, 304 barrier electric, 284f
light reflection, 392f boundary, 121
light refraction, 393 corona, 269
mass conservation, 53 surface, 111
momentum conservation, Length,
closed system, 59f body, and velocity, 452
open system, 59f focal,
Newton’s, lens, 401
first, 48 mirror, 399
second, 54ff, 72 Lorentz contraction, 452
in relativistic dynamics, measurement, 446f
457 path, 22
third, 57 proper, body, 452
nucleus displacement, in ra­ standing wave, 375
dioactive decay, 536 Lens(es), 400ff
Ohm’s, converging, 402f
a-c circuit, 354 diverging, 402f
Subject Index 641

Lens(es), Line(s),
power, 403 sp ecia l,
thin, 400f series, 499
Leptons, 570, 571 Liquid(s), 135, see also Fluid
Lifetime, dropping, 111
mean free, 136 fluidity, 136
radioactive isotope, 537 free surface, 111
Lift, aerofoil, 123f internal friction, 117
Light, isotropy, 136
chemical action, 472f non-wetting, 188
diffraction, 420ff short-range order in, 135f
dispersion, 427ff supercooled, 201
natural, 427 superheated, 181
plane-polarized, 426 surface curvature and addition­
point source, 405 al pressure, 189
polarization, 425ff surface layer, 111
pressure, 470ff wetting, 188
quantity, 404 Loop,
quantum properties, 463f current-carrying, equilibrium
speed, position, 289
determination, 414ff hysteresis, 321 f
limiting nature, 448 iqagnetic moment, 286f
in vacuum, 607 Lumen, 612
unpolarized, 427 Luminescence, 434f
wave-particle duality of prop­ Luminophor, 435
erties, 463f, 474f Lux, 612
wave properties, 463f
Lightning, 270
Limit, Machine, perpetual motion,
elastic, 196 first kind, 165
proportionality, 196 second kind, 173
Line(s), Magnet, permanent, 322
held, Magnetics, 316ff
electric, 209f Magnetization, 320
magnetic, 289 diamagnetics, 320
Fraunhofer, 436 ferromagnetics, 320
magnetic induction, 289 paramagnetics, 320
spectral, 435 permanent, 322
in molecular spectrum, 523 residual, 322
natural width, 485 saturation, 321
reversal, 436 technical, curve, 321
642 Subject Index

Magnification, Mechanics,
angular, 412 quantum, 474ff
microscope, 413 Medium(a),
telescope, 413 active, 514
linear, amplifying, 514ff
spherical mirror, 400 continuous, llOf
thin lens, 403 elastic, 359
Masers, 513 with negative absorption of
Mass, light, 515
body, 52f optically homogeneous, 391
deficiency, 529f Melting, 200
electron, 204 Meniscus, 188
gravitational, 52 Mesons, 570, 571
inert, 52 Metals, 194
measurement, 72ff Method,
molar, 127, 599 Fizeau, 414f
neutron, 527 Michelson’s, 415f
in Newtonian mechanics, 63 nuclear emulsions, 542
proper, 457 Roemer, 414
proton, 204, 527 statistical, of investigation,
relative atomic, 591 138
relative molecular, 591 thermodynamic, of investiga­
relativistic, 457 tion, 144
rest, 457, 459 tracer atom, 558
elementary particles, 574, Wilson-Skobeltsyn, 541
577, 579, 580, 607 Metre, 589
standard, 72 Microscope, 412
Mass-spectrographs, 302 Model,
Mass-spectrometers, 302 Bohr, atom, 499ff
Materials, see also Substance(s) drop nuclear, 533f
hard-magnetic, 322 Moderators, 548
soft-magnetic, 322 Modulus,
Measurement, elasticity, 196
arithmetic mean, 616 Young’s, 196
direct, 614 Mole, 589
error, 615 Molecule(s), 126
indirect, 614 atomic, 521 f
Measures, 586 bond length, 519
Mechanics, 17 collision diameter, 133
classical, 17 dipole moment, 222
Newtonian, 17 electron spectrum, 525
Subject Index 643
Molecule(s), Motion,
electron-vibrational spectrum, at aggie with the horizontal,
525 41ff
energy and spectrum, 523 body thrown vertically up­
heteropolar, 519 ward, 35ff
ionic, 519f Brownian, 127f
path, mean free, 139f circular, uniform, 38ff
rotational spectrum, 524 curvilinear, 19, 28
spectra, 522ff decelerated, uniformly, 31, 33
time, mean free, 140 inertial, 48, 54
velocity, mechanical, 17ff
arithmetical mean, 138f non-inertial, 54
mean square, 139 non-uniform, 24, 28
vibrational-rotational spec­ oscillatory, 325
trum, 524 plane, 19
vibrational spectrum, 524 rectilinear, 19, 28
Moment, uniform, 28ff
dipole, molecule, 222 uniformly changing, 31ff
electron dipole, 217 rotational, see Rotation
induced, 222 thermal, 133f
force, 80f translational, 20f
inertia, perfectly rigid body, 21
body, 81 two-dimensional, 19
cylindrical shell, 82 uniform, 24, 28
point particle, 81 variable, 24
solid cylinder, 82 Multipliers, electron, 273
solid sphere, 82 Muons, 571, 577
thin straight rod, 82 rest mass, 577
magnetic,
atom, orbital, 314f
electron, orbital, 314 Nearsightedness, 411
loop, 286f Neutrino, 550, 571
momentum, see Angular mo­ meson, 577
mentum Neutron(s), 526
Momentum, 53f energies, 548f
angular, see Angular momentum fast, 548
localized particle, 490 fission, 552
relativistic, 457 free, radioactivity, 545
Monocrystals, 135 interaction with nuclei, 548f
Motion, multiplication, 552
accelerated, uniformly, 31, 33 resonance absorption, 549
Subject Index

Neutron(s), Nucleus(i),
rest mass, 607 heavy, instability, 551
slow, 548 and liquid drop, 533f
surplus, 552 mass, 526f
thermal, 548f mass number, 526
Newton, 593 mirror, 531f
Nodes, standing wave, 375 parent, 536
Noise, 368 spin, 527
Non-wetting, stability, 534
absence, 189 unit binding energy, 529
perfect, 189 volume, 527f
Nuclear reactions, 545ff Number(s),
chain, 554ff atomic, 494
channels, 546 Avogadro, 127
classification, 546 Loschmidt, 154
element transformation, 547f quantum, 484
endoergic, 546 magnetic, 507
endothermic, 546 magnetic spin, 508
exoergic, 546 orbital, 505f
exothermic, 546 principal, 484, 499, 503
fission, 554ff rounding off, 624
high-energy, 547 Z, 494
low-energy, 546f
medium-energy, 547 Objective, rapidity, 408
Q-value, 546 Ohm, 605
stages, 547 Optics, 390
Nucleon(s), 520, 570, 571 geometrical, 390f
binding energy, 528ff physical, 414
rest mass, 574 quantum, 462
structure, 583ff ray, 390f
Nucleus(i) wave, 414
atomic, 526 Orbit, electron, 497, 505
binding energy, 528ff Order,
charge, 526 diffraction maximum, 423
compound, 547 diffraction minimum, 423
daughter, 536 long-range, in solids, 135
energy state, short-range, in liquids, 135f
excited, 530 Ortho-hydrogen, 532f
ground, 530 Oscillation(s), 325ff
fission, 551 amplitude, 326
half-life, 538 complete, 326
Subject Index 645

Oscillation (s), Para-hydrogen, 532f


damped, 338ff Paramagnetics, 316ff
amplitude, 339 magnetizing, 319
conditional period, 339 relative permeability, 316f, 319
cyclic frequency, 339 Parameter(s),
forced, 340ff critical, 183
amplitude, 341ff fission, 553
driving force, 340f critical value, 553
electromagnetic, 350f of state, basic, 145f
free, 327 thermodynamic, 145
electromagnetic, 346ff Particle(s), 18
cyclic frequency, 349 alpha, 535
natural cyclic frequency, coordinate, uncertainty in, 486
332 elementary, 125, 568f
undamped, 327 classification, 570ff
frequency, cyclic, 326 stable, 570
harmonic, 326, 480 structure, 568f
acceleration, 328ff unstable, 570
addition, 337f equations of motion, 19
kinetic energy, 335 high-energy, sources, 558
potential energy, 335f liquid, equilibrium, 113
velocity, 328 mechanical state, 54
period, 327 momentum, uncertainty, 486f
periodic, 325 path, 19
phase, 327 point, 18, 569
resultant, relativistic, 458
amplitude, 337 resonance, 572
initial phase, 337 specific charge, 301 f
undamped, stream density, 575
cyclic frequency, 349 truly neutral, 580
period, 349 ultrarelativistic, 458
3scillator(s), Pascal, 595
Hertz, 383f Path,
linear harmonic, 480f molecule, mean free, 139f
energy levels, 483 particle, 19
zero-point energy, 483, 490f Peak(s),
magnetostriction, 369 central zero-order, 423
piezoelectric, 370 ^diffraction, 422f
quantum linear harmonic, 490 Pendulum,^
zero-point energy, 491 mathematical, 333f
)yertones, 368 acceleration} 334
646 Subject Index

Pendulum, Photorelay, 468f


mathematical, Photoresistors, 469
period, 334 Pile, atomic, 555, see also Reac­
simple, 333 tor, nuclear
spring, 325f Pions, 571, 577ff
acceleration, 329 rest mass, 574
amplitude of oscillations, 336 Plane, focal,
harmonic oscillations, 331ff lens, 401
kinetic energy, 335 mirror, 399
potential energy, 335f Plasma, 270f
period of oscillations, 332 fully ionized, 271
total energy, 335 high-temperature, 270
Perimeter, wetted, 188 weakly ionized, 270
Period, revolution, 41, 45 Point(s),
Permeability, relative, 293 boiling, 180
diamagnetics, 316f critical, 183
ferromagnetics, 321 Curie, 323
paramagnetics, 316f, 319 dew, 185
Permittivity, freezing, 201
free space, 206 lattice, 135
relative, 206 melting, 200
in varying electric field, 378 yield, 197
Phase, initial, resultant oscilla­ Polarization,
tion, 337 deformation, 224
Phenomena, dielectric, 223, 225f
capillary, 190f electron, 224
magnetostriction, 369 ionic, 226
Photocell(s), 468, 470 light, 425ff
Photoconduction, 469 orientation, 226
Photocurrent, 465f Polarizers, 427
saturation, 465f Pole, coordinate system, 39
Photoeffect, 464ff Polycrystals, 135
Photoelectrons, 465 Positron, 550
Photography, 472f Postulates,
Photoionization, 465 Bohr’s, 500f
Photoluminescence, 435 first, 500, 504
Photometry, 404ff second, 500f, 504
Photons, 462, 570 third, 501, 505
energy, 462 stationary states, 500, 504
mass, 462 Potential,
momentum, 462 conductor, and charge, 235
Subject Index 647

Potential, Process,
difference, 230f anabatic, 153, 163, 164
electrostatic field, 229ff compensating, 174
point charge, 231 cyclic, 169f
sphere, 231f electromagnetic, 573
Potentiometer, 259 equilibrium, 148ff
Power, 109f irreversible, 169
alternating current, isobaric, 150, 164
active, 355 work of expansion, 159
instantaneous, 355 isochoric, 150, 151, 164
mean, 355 isothermal, 150, 163, 164
average, 109f quasistatic, 148
exposure dose, 560 reversible, 168f
instantaneous, 110 slow, 574
mean radiated, 381f thermodynamic, 148ff
radiation dose, 560 Projectors, 408
resolving, eye, 411 diascopic, 408f
wave, mean, 367 episcopic, 408
Pressure, 111, 146 Protection, electrostatic, 220
dynamic, 118 Proton(s), 204, 526
gas, 142 charge, 204
hydrostatic, 114 electric radius, 585
in liquid, 113 rest mass, 607
measurement, 120 Pump, heat, 177
static, 118f Pumping, amplifying medium,
total, 118 515
Principle,
Archimedes’, 114f
charge conjugation, 581 Quantity(ies),
constancy of light speed, 445 dimensionless, 587
Huygens, 421 physical,
Huygens-Fresnel, 421f homogeneous, 587
independence of force action, 56 measuring, 607
minimum potential energy, 94 quantization, 483
Pauli exclusion, 509f quantized, 483
relativity, 444f, 451 similar, 587
classical, 63 Quantization, spatial, 507
Galilean, 63 Quantum,
mechanical, 63 action, 433
superposition, 371 numbers, see Number(s), quan­
electric fields, 210, 212 tum
Subject Index

Rad, 560 Reactions,


Radar, 388 nuclear, see Nuclear reactions
Radian, 590 photochemical, 472
Radiation(s), thermonuclear, 561
beta, 535 Reactor, nuclear, 555
gamma, 535, 545 breeder, 557
induced, 514 control rods, 556
infrared, 436f critical conditions, 556
isothermal, 429f fuel, 555
non-isothermal, 430 Receiver,
radioactive, 534f radio, 386f
stimulated, 514f television, 387f
thermal, 430 Recombination, 264, 267
source, 429 Reference frame(s),
ultraviolet, 437 definition, 18
X-, 437ff geocentric, 49
characteristic, 439 heliocentric, 49
diffraction, 441 inertial, 48
white, 438 Newtonian, 48
Radio broadcasting, 385ff non-inertial, 49, 77ff
Radioactivity, timing, 448
artificial, 550 Reflection,
natural, 534f angle, 392
Radius, first Bohr, 502, 607 diffuse, 393
Rate, irregular, 393
condensation, 179 regular, 393
evaporation, 178 specular, 393
exposure dose, 560 total, 393
radiation dose, 560 critical angle, 393
Ratio, voltage, 311 Reflectors, neutron, 555
Rays, Refrigerator, 176f
cathode, 277 Regions, spontaneous magnetiza­
cosmic, 575f tion, 323
light, path reversibility, 393 Relation,
Reactance, Heisenberg’s uncertainty, 487f
capacitive, 353 mass-energy, 458ff
inductive, 353 Rem, 561
Reactions, Remanence, 322
fusion, 561ff Reservoir,
carbon-nitrogen cycle, 562ff cold, 170, 175
controlled, 564 hot, 170r 17§
Subject Index 649

Resistance, Scale, temperature, see Tem-


electrical, 248f • perature scale
to motion of body, Second, 589
friction, 122 Sections, Zhukovsky, 123f
pressure, 122 Self-induction, 308ff
total, series connection of con­ Semiconductors, 243, 279f
ductors, 255 electron, 282
Resistivity, 249 hole, positive, 280
temperature coefficient, 252 impurity centres, 281
Resistor, 249 impurity hole, 283
Resonance, 343 rc-type, 282
current, 358 photoconduction, 280
voltage, 357 p-type, 283
Resonator, 383 Sensitivity, photocathode, 272
Rings, Newton’s, 418f Series,
Roentgen, 560 B aimer, 499
Roentgen-equivalent-man, 561 Lyman, 499
Rotation, 21 radioactive, 536
accelerated, 46 Set, permanent, 195
angle, 47 Shell, electron, 510
decelerated, 46 Siemens, 606
non-uniform, 46 Siren, 369
rigid body, in inertial refer­ Sites, lattice, 135
ence frame, 80 Solenoid, 294f
uniform, 45f magnetic field energy, 312
uniformly changing, 46 poles, 295
variable, 46 Solidification, 201
Rule(s), see also Law(s) Solids, 134
for calculating digits, 624f long-range order in, 135
junctions, 254 Sonar,' 370
Kirchhoff’s, Sound(s),
first, 254 audible, 368
second, 254 intensity level, 369
left-hand, 29 If loudness, 368
loops, 254 musical, 368
moments, 89 quality, 368
orbit quantization, 501 Sources,
parallelogram, 84 coherent, 372
polygon, 84 high-energy, particles, 558
right-hand, 305 incoherent, 372
right-hand screw, 287, 293 light, 405
650 Subject Index

Sources, Stress,
luminous flux, 405 normal, 196
luminous intensity, 405 shear, 196
thermal radiation, 429 Sublimation, 201
Spectrum(a), Substance(s),
band, 522 fluorescent, 495
diffraction, 424 magnetic, 316ff
dispersion, 428f strongly, 316
electromagnetic, 441 f weakly, 316
line, 439 relative molecular mass, 591
mass, 302 surface-active, 188
molecules, 522ff working, 165, 175
X-ray, continuous, 438 in refrigerator, 177
boundary, 438 Superconductivity, 253
Speed, Surface,
average, 24f equipotential, 232f
instantaneous, 25 free, 111
Spin, electron, 315f, 508 layer, 111
Spring, graduation, 74 wave, 360
Standards, 586 Surfactants, 188
State, Synchrophasotrons, 568
critical, 183 System,
equilibrium, 145, 147 auto-oscillating, 343f
inverted, 515 closed, 51, 107f
mechanical, 54 centre of mass, 59
non-equilibrium, 515 conservative, 101
overload, 67 closed, 101
quantum, stationary, 484 coordinate, 18, see also Refer­
stationary, 145 ence frame(s)
steady, 145 Cartesian, 18
thermodynamic, 145 polar, 39
weightlessness, 66 emitting, 381
Statics, 84 forces, balanced, 84
Steradian, 590 isolated, 51
Strain, 195 non-conservative, 101, 107f
Stream, fluid, 116 closed, 101, 107f
Streamline, 116, 123 open, 108
Strength, 197
ultimate, 197 Telescope(s), 413
Stress, radio, 388f
mechanical, 195 Television, 385, 387f
Subject Index 651

Temperature, 146 Time,


absolute, proper, 453
molecular-kinetic interpre­ relativistic dilation, 454
tation, 142 Tone(s), 368
zero, 147 fundamental, 368
critical, 183 pitch, 368
Temperature scale, Torque, 80
empirical, 146 Tracers, radioactive, 558
international Celsius, 147 Train, wave, 485
international practical, 591 Trajectory,
thermodynamic, 172, 591 particle, 19f
absolute, 147 and reference frame, 19
Tension, Transformation(s),
surface, 187f Galilean, 61
and temperature, 188 coordinate, 61
uniaxial, 196 Lorentz, 450f
Term, spectral, 499 Transformer, 311
Tesla, 610 voltage ratio, 311
Theorem(s), Transmitter,
Carnot’s, 171 radio, 386
kinetic energy, 102 television, 387
Theory, Triode, 275f
long-range force action, 209 anode current, 276
molecular-kinetic, 125ff cut-off voltage, 276
relativity, special, 444ff directly heated, 275
short-range force action, 209 hot-cathode, 275
Thermodynamics, 144 static grid characteristic, 276
Thermoelectrons, 272 Tube(s),
Threshold, cathode-ray, 277f
audibility, 368 deflecting plates, 278
standard, 369 electromagnetic, 278
fission, 552 electrostatic, 278
pain, 369 electron, 273
photoelectric, 466 flow, 116
Timbre, 368 Pitot, 120
Time, Pitot-Prandtl, 120
absolute nature in Newtonian X-ray, 439
mechanics, 62
mean free, molecule, 140 Ultrasounds, 368, 369f
mean relaxation, 136 Uncertainty,
nuclear, 573 in particle coordinate, 486
652 Subject Index

Uncertainty, Vector(s),
in particle momentum, 486f induction, magnetic field, 286
Unit(s), direction, 287f
absolute electrostatic, 599 magnetic induction, 286
atomic energy, 603 polygon, 51 f
atomic mass, 603 position, 18
basic, 587 radius, 18
derived, 587 wave, 462
physical quantity, 586 Velocity, 23ff
system, 586f amplitude, 328
absolute electrostatic, 599 angular,
cgs, 588, 590 average, 40
SI, 587, 588, 589f instantaneous, 40, 47
arithmetical mean, molecules,
Vacuum, 271 138f
Valves, average, 23f, 28
electron, 273ff body thrown vertically upward,
electron-beam, 277f 36
three-electrode, see Triode circular, 38
two-electrode, see Diode circulation in fluid, 123
Vaporization, 178 escape, 109
Vapour, 178, see also Gas(es) fluid, measurement, fl20f
isotherm, 182f in free fall, 35
saturated, 179 instantaneous, 24
density and temperature, 183f linear, 38, 45
pressure, 179 mean square, molecules, 139
superheated, 182 orbital, 67f
supersaturated, 183 phase, 363
unsaturated, 181 wave, 363f
pressure, 181 Viscosity, 116f
Vector(s), Vision, colour, 410
compression, 69 Volt, 605
currentjdensity, average, 244f Voltage, 248, 250
displacement, 21ff, 29, 32 drop, 248, 250
body thrown vertically up­ parallel connection of con-
ward, 36 du c I c j f , 2 5 6
displacement current density, series^connection of conduc­
377 tors, 255
electric displacement, 377 effective value, 355
elongation, 69 grid, 275
force, components, 87f ignition, 268
Subject Index 653

Volume, Wave(s),
molar, 127 pfcwer (mean), 367
ideal gas in standard condi­ radio, 385
tions, 607 classification, 385
specific, 127, 146 running, 373
sound, 360
Watt, 595, 611 velocity in gases, 363
Wave(s), 360 source, 360
acoustic, 360 spherical, 361
coherent, 372f standing, 373ff
de Broglie, 475ff amplitude, 374f
direction line, 360f antinodes, 375
elastic, 360 length, 375
velocity in liquids, 364 nodes, 375
electromagnetic, 360 surface, 360
dispersion, 378 train, 485
emission, 381ff length, 485
energy transfer, 380 transverse, 361, 363
intensity, 380f vector, 462
modulation, 385f velocity, 363
monochromatic, 379 in gases, 363
plane-polarized, 426 in liquids, 364
radiation, 381 in solids, 364
unpolarized, 426 Wavefront, 360
velocity, 377f Wavelength, 364f
volume energy density, 380 de Broglie, 475ff
energy density, mean, 366 electron, 482
frequency, 379 Weber, 610
geometrical path difference, 372 Weight, body, 65
incoherent, 372, 373 Well, one-dimensional square po­
intensity, 367 tential, 481f
interference, 373 Wetting, 188
light, 390 absence, 189
length in substance, 416 perfect, 188
longitudinal, 361ff Work, 158ff
velocity in solids, 364 elastic force, 69, 100
number, 365 electrostatic forces, 228
optical path difference, 417 elementary, 158
optical path length, 416 charge^displacement, 226f
plane, 360f force, 95f
monochromatic, 379 expansion, ideal gas, 158f
654 Subject Index

Work, Work,
force, 96ff potential forces, 103
friction, 71f, lOOf Work function, 272, 467
gravitational, 99
gravity, 64, 99f X-rays,
and heat, 161 absorption, 440
and kinetic energy, 102 properties, 440
mechanical, thermal equiva­
lent, 162
moving charge along conduc­ Yield, 186
tor, 247
negative, 158 Zero, absolute, temperature, 147
positive, 158 Zone, wave, 381
TO THE READER

Mir Publishers welcome your comments on the content,


translation, and design of this book.
We would also be pleased to receive any proposals
you care to make about our future publications.
Our address is:
USSR, 129820, Moscow 1-110, GSP
Pervy Rizhsky Pereulok, 2
Mir Publishers

Printed in the Union of Soviet Socialist Republics


THE PERIODIC TABLE
The number to the left of the chemical symbol for each element i9 its
For unstable elements, the atomic mass of the

S h e ll G roup
fille d
b y o u te r
e le c t r o n s I II II I IV V

F ir s t , H yd rogen
K - s h e ll 1 H
1 .0 0 7 9

Second, L ith iu m B e r y lliu m B oron C arbon N itr o g e n


L -s h e ll 3 Ll 4 Be 5 B 6 C 7 N
6 .9 4 i 9 .0 1 2 1 8 1 0 .8 1 12.0 1 1 1 4 .0 0 6 7

T h ir d , S o d iu m M a g n e siu m A lu m in iu m S ilic o n P h o sp h o r u s
M -s h e ll 11 Na 12 M g 13 A1 14 S i 15 P
2 2 .9 8 9 7 7 2 4 .3 0 5 2 6 .9 8 1 5 4 2 8 .0 8 5 3 0 .9 7 3 7 6

P o ta s s iu m C a lciu m S c a n d iu m T ita n iu m V a n a d iu m
19 K 20 Ca 21 Sc 22 Ti 23 V
3 9 .0 9 8 s 4 0 .0 8 4 4 .9 5 5 9 4 7 .9 0 5 0 .9 4 1 5
F o u r th ,
N - s h e ll C opper Z inc G a lliu m G erm an iu m A r s e n ic
29 Cu 30 Zn 31 Ga 32 Ge 33 A s
6 3 .5 4 e 6 5 .3 8 6 9 .7 2 7 2 .59 7 4 .9 2 1 6

R u b id iu m S tr o n tiu m Y ttr iu m Z ir c o n iu m N io b iu m
37 Rb 38 Sr 39 Y 40 Zr 41 Nb
8 5 .4 6 7 s 8 7 .6 2 8 8 .9 0 5 9 9 1 .2 2 9 2 .9 0 6 4
F if t h ,
O -sh e ll S ilv e r C a d m iu m In d iu m T in A n tim o n y
47 A g 48 Cd 49 In 50 Sn 51 Sb
1 0 7 .8 6 8 1 1 2 .4 1 1 1 4 .8 2 I I 8 .69 121. 7b

C esiu m B a r iu m H a fn iu m T a n ta lu m
55 Cs 56 B a 5 7 -7 1 * 72 H f 73 T a
1 3 2 .9 0 5 4 1 3 7 .3 3 1 7 8 .4 9 1 8 0 .9 4 7 9
S ix t h ,
P - s h e ll G old M ercury T h a lliu m L ead B is m u t h
79 A u 80 H g 81 T1 82 Pb 83 B l
1 9 6 .9 6 5 2 0 0 .5 s 2 0 4 .37 2 0 7 .2 2 0 8 .9 8 0 4

S e v e n th , F r a n c iu m R a d iu m K u r c h a to v iu m
Q -s h e ll 87 F r 88 R a 8 9 -1 0 3 ** 104 K u
(2 2 3 ) 2 2 6 .0 2 5 4 (2 6 1 )

L a n th a n u m C erium P r a se o d y m iu m N e o d y m iu m
57 La 5 8 Ge 5 9 Pr 60 Nd
* L a n th a ­ 138. 905b 1 4 0 .1 2 1 4 0 .9 0 7 7 1 4 4 .2 4
n id e
se r ie s T erb iu m D y s p r o s iu m H o lm iu m E rb iu m
65 T b 66 D y 67 H o 68 Er
1 5 8 .9 2 5 4 1 6 2 .50 1 6 4 .9 3 0 4 1 6 7 .2 e

A c tin iu m T h o r iu m P r o ta c t in iu m U r a n iu m
89 Ao 90 T h 91 Pa 92 U
** A c t in ­ (2 2 7 ) 2 3 2 .0 3 8 1 2 3 1 .0 3 5 9 2 3 8 .0 2 9
id e
se r ie s B e r k e liu m C a lifo r n iu m E in s te in iu m F erm i um
97 B k 9 8 Cf 99 E s 100 F m
(2 4 7 ) (2 5 1 ) (2 5 4 ) (2 5 7 )
OF THE ELEMENTS
atomic number Z, and the number below is its atomic mass in amu.
most stable isotope is given in parentheses.
• -----------------------------------------
N u m ber
o f e le c ­
t r o n s in
VI V II V III 0 s h e lls

H e liu m He
2 He 2
4 .0 0 2 6 0

O x y g en F lu o r in e N eon Ne
8 O 9 F 10 Ne 2. 8
1 5 .9 9 9 1 8 .9 9 8 4 0 3 2 0 .1 7

S u lp h u r C h lo rin e A rgon * Ar
16 S 17 Cl 18 A r 2. 8, 8
3 2 .0 6 3 5 .4 5 3 3 9 .9 4 s

C hrom ium M a n g a n e se Iron C o b a lt N icR el


24 Cr 25 Mn 26 F e 27 Co 28 Ni
5 1 .9 9 6 5 4 .9 3 8 0 5 5 .8 4 ? 5 8 .9 3 3 2 5 8 .7 0

S e le n iu m B r o m in e K r y p to n Kr
34 S e 3 5 Br 36 K r 2, 8
7 8 .9 0 7 9 .9 0 4 8 3 .8 0 18. 8

M oly b d en u m T e c h n e tiu m R u th e n iu m R h o d iu m P a lla d iu m


1 42 Mo 43 T e 44 R u 45 R h 46 Pd
9 5 .9 4 9 8 .9 0 6 2 101 .0 ? 1 0 2 .9 0 5 5 1 0 6 .4

T e llu r iu m Io d in e X enon Xe
52 Te 53 I 54 X e 2, 8 . 18
1 2 7 .60 1 2 6 .9 0 4 5 1 3 1 .3 0 18, 8

T u n g s te n R h e n iu m O sm iu m Ir id iu m P la tin u m
74 W 75 R e 76 Os 77 Ir 78 Pt
1 8 3 .8 5 1 8 6 .2 0 7 1 9 0 .2 1 9 2 .2 2 1 9 5 . 0e

P o lo n iu m A s t a t in e R ad on Rn
84 P o 85 A t 86 Rn 2
. 8 , 18
(2 0 9 ) (2 1 0 ) (2 2 2 ) 3 2 . 18. 8

P r o m e th iu m S a m a r iu m E u r o p iu m G a d o lin iu m
61 P m 62 Sm 6 3 Eu 64 Gd
(1 4 5 ) 1 5 0 .4 1 5 1 .9 6 1 5 7 .2 b

T h u liu m Y tte r b iu m L u t e t iu m Lu
69 Tu 70 Yb 71 Lu 2. 8 . 18
1 6 8 .9 3 4 2 173. 0a 1 7 4 .9 6 ? 32, 9, 2

N e p tu n iu m P lu to n iu m A m e r ic iu m C urium
93 Np 94 Pu 95 A m 96 Cm
2 3 7 .0 4 8 2 (2 4 4 ) (2 4 3 ) (2 4 7 )

M e n d e le v iu m N o b e liu m L a w r e n c iu m Lw
101 Md 1 02 N o 103 Lr 2 . 8 . 18
(2 5 8 ) (2 5 5 ) (2 5 6 ) 32. 32, 9, 2

You might also like