Ebook Foundations of Modern Physics 1St Edition Weinberg Online PDF All Chapter
Ebook Foundations of Modern Physics 1St Edition Weinberg Online PDF All Chapter
Ebook Foundations of Modern Physics 1St Edition Weinberg Online PDF All Chapter
Edition Weinberg
Visit to download the full and correct content document:
https://ebookmeta.com/product/foundations-of-modern-physics-1st-edition-weinberg/
More products digital (pdf, epub, mobi) instant
download maybe you interests ...
https://ebookmeta.com/product/the-foundations-of-spacetime-
physics-1st-edition-antonio-vassallo/
https://ebookmeta.com/product/the-first-three-minutes-a-modern-
view-of-the-origin-of-the-universe-2022-edition-steven-weinberg/
https://ebookmeta.com/product/elements-of-modern-physics-1st-
edition-s-h-patil/
https://ebookmeta.com/product/university-physics-with-modern-
physics-third-edition-wolfgang-bauer/
JEE Advanced Physics-Modern Physics 3rd Edition Rahul
Sardana
https://ebookmeta.com/product/jee-advanced-physics-modern-
physics-3rd-edition-rahul-sardana/
https://ebookmeta.com/product/elements-of-early-modern-physics-j-
l-heilbron/
https://ebookmeta.com/product/foundations-of-modern-global-
seismology-2nd-edition-charles-j-ammon/
https://ebookmeta.com/product/modern-physics-kenneth-s-krane/
https://ebookmeta.com/product/political-grammars-the-unconscious-
foundations-of-modern-democracy-1st-edition-davide-tarizzo/
Foundations of Modern Physics
PREFACE
4 RELATIVITY
4.1 Early Relativity
Motion of the Earth □ Relativity of motion □ Speed of light □
Michelson–Morley experiment □ Lorentz–Fitzgerald contraction
4.2 Einsteinian Relativity
Postulate of invariance of electrodynamics □ Lorentz
transformations □ Space inversion, time reversal □ The Galilean
limit □ Maximum speed □ Boosts in general directions □ Special
and general relativity
4.3 Clocks, Rulers, Light Waves
Clocks and time dilation □ Rulers and length contraction □
Transformation of frequency and wave number
4.4 Mass, Energy, Momentum, Force
Einstein’s thought experiment □ Formulas for energy and momentum □ E = mc2 □ Force
in relativistic dynamics
5 QUANTUM MECHANICS
5.1 De Broglie Waves
Free-particle wave functions □ Group velocity □ Application to
hydrogen □ Davisson–Germer experiment □ Electron microscopes
□ Appendix: Derivation of the Bragg formula
5.2 The Schrödinger Equation
Wave equation for particle in potential □ Boundary conditions □
Spherical symmetry □ Radial and angular wave functions □
Angular multiplicity □ Spherical harmonics □ Hydrogenic energy
levels □ Degeneracy
5.3 General Principles of Quantum Mechanics
States and wave functions □ Observables and operators □
Hamiltonian □ Adjoints □ Expectation values □ Probabilities □
Continuum limit □ Momentum space □ Commutation relations □
Uncertainty principle □ Time dependent wave functions □
Conservation laws □ Heisenberg and Schrödinger pictures
5.4 Spin and Orbital Angular Momentum
Doubling of sodium D-line □ The idea of spin □ General action of
rotations on wave functions □ Total angular momentum operator □
Commutation relations □ Spin and orbital angular momentum □
Multiplets □ Adding angular momenta □ Atomic fine structure and
space inversion □ Hyperfine structure □ Appendix: Clebsch–
Gordan Coefficients
5.5 Bosons and Fermions
Identical particles □ Symmetric and antisymmetric wave functions
□ Bosons and fermions in statistical mechanics □ Hartree
approximation □ Slater determinant □ Pauli exclusion principle □
Periodic table of elements □ Diatomic molecules: para and ortho □
Astrophysical cooling
5.6 Scattering
Scattering wave function □ Representations of the delta function □
Calculation of the Green’s function □ Scattering amplitude □
Probabilistic interpretation □ Cross section □ Born approximation □
Scattering by shielded Coulomb potential □ Appendix: General
transition rates
5.7 Canonical Formalism
Hamiltonian formalism □ Canonical commutation relations □
Lagrangian formalism □ Action principle □ Connection of
formalisms □ Noether’s theorem: symmetries and conservation
laws □ Space translation and momentum
5.8 Charged Particles in Electromagnetic Fields
Vector and scalar potential □ Charged particle Hamiltonian □
Equations of motion □ Gauge transformations □ Magnetic
interactions □ Spin coupling
5.9 Perturbation Theory
Perturbative expansion □ First-order perturbation theory □ Dealing
with degeneracy □ The Zeeman effect □ Second-order
perturbation theory
5.10 Beyond Wave Mechanics
State vectors □ Linear operators □ First postulate: values of
observables □ Second postulate: expectation values □ Probabilities
□ Continuum limit □ Wave functions as vector components
6 NUCLEAR PHYSICS
6.1 Protons and Neutrons
Discovery of the proton □ Integer atomic weights □ Nuclei as
protons and electrons? □ Trouble with diatomic nitrogen □
Discovery of the neutron □ Nuclear radius and binding energy □
Liquid drop model □ Stable valley and decay modes
6.2 Isotopic Spin Symmetry
Neutron–proton and proton–proton forces □ Isotopic spin rotations
□ Isotopic spin multiplets □ Quark model □ Pions □ Appendix: The
three–three resonance
6.3 Shell Structure
Harmonic oscillator approximation □ Raising and lowering
operators □ Degenerate multiplets □ Magic numbers □ Spin–orbit
coupling
6.4 Alpha Decay
Coulomb barrier □ Barrier suppression factors □ Semi-classical
estimate of alpha decay rate □ Level splitting □ Geiger–Nuttall law
□ Radium alpha decay □ Appendix: Quantum theory of barrier
penetration rates
6.5 Beta Decay
Electron energy distribution □ Neutrinos proposed □ Fermi theory
□ Gamow–Teller modification □ Selection rules □ Strength of weak
interactions □ Neutrinos discovered □ Violation of left–right and
matter–antimatter symmetries □ Neutrino helicities □ Varieties of
neutrino
ASSORTED PROBLEMS
BIBLIOGRAPHY
AUTHOR INDEX
SUBJECT INDEX
Preface
This book grew out of the notes for a course I gave for
undergraduate physics students at the University of Texas. In this
book I think I go farther forward than is usual in undergraduate
courses, giving readers a taste of nuclear physics and quantum field
theory. I also go farther back than is usual, starting with the struggle
in the nineteenth century to establish the existence and properties of
atoms, including the development of thermodynamics that both
aided in this struggle and offered an alternative program.
I fear that some readers may want to skim through this early part
and hurry on to what they regard as the good stuff, quantum
mechanics and relativity. That would be a pity. In my experience
physics students who aim at a career in atomic or nuclear or
elementary particle physics often manage to get through their formal
education without ever becoming familiar with entropy, or
equipartition, or viscosity, or diffusion. That was true in my own
case. This book, or a course based on it, may provide some students
with their last chance to learn about these and other matters needed
to understand the macroscopic world.
Readers may find this book unusual also in its strong emphasis on
history. I make a point of saying a little about the welter of
theoretical guesswork and ill-understood experiments out of which
modern physics emerged in the twentieth century. This, it seems to
me, is a help in understanding what otherwise may seem an
arbitrary set of postulates for relativity and quantum mechanics. It is
also a matter of personal taste. Research in physics seems to me to
lose some of its excitement if we do not see it as part of a great
historical progression. Some valuable historical works are listed in a
bibliography, along with collections of original articles that I have
found most helpful.
But this is not a work of history. Historians aim at uncovering how
the scientists of the past thought about their own problems – for
instance, how Einstein in 1905 thought about the measurement of
space and time separations in developing the special theory of
relativity. For this aim of historical writing it is necessary to go deeply
into personal accounts, institutional development, and false starts,
and to put aside our knowledge of subsequent progress. I try to be
accurate in describing the state of physics in past times, but the aim
of this book in discussing the problems of the past is different: it is
to make clear how physicists think about these things today.
This book is intended chiefly for physics students who are well into
their time as undergraduates, and for working scientists who want a
brief introduction to some area of modern physics. I have therefore
not hesitated to use calculus and matrix algebra, though not in
advanced versions. As required by the subject matter, the
mathematical level here slopes upwards through the book. Where
possible I have chosen concrete rather than abstract formulations of
physical theories. For instance, in Chapter 5, on quantum mechanics,
I mostly represent physical states as wave functions, only coming at
the end of the chapter to their representation as vectors in Hilbert
space. In some sections detailed material that can be skipped
without losing the thread of the theory is put into appendices. Two
of these appendices present what in my unbiased opinion are
improved derivations of important results: the appendix to Section
2.6 gives a revised version of Einstein’s derivation of his formula for
the diffusion constant in Brownian motion, and the appendix to
Section 6.4 presents a revision of Fermi’s calculation of the rate of
alpha decay.
In my experience, with some judicious pruning, the material of the
book up to about the middle of Chapter 5 can be covered in a one-
term undergraduate course. But I think that to go over the whole
book would take a full two-term academic year.
This book treats such a broad range of topics that it is impossible
to go very far into any of them. Certainly its treatment of quantum
mechanics, statistical mechanics, transport theory, nuclear physics,
and quantum field theory is no substitute for graduate-level courses
on these topics, any one of which would occupy at least a whole
year. This book presents what I think, in an ideal world, the
ambitious physics student would already know when he or she
enters graduate school. At least, it is what I wish that I had known
when I entered graduate school.
In any case, I hope that the student or reader may be sufficiently
interested in what I do discuss that they will want to go into these
topics in greater detail in more specialized books or courses, and
that they will find in this book a good preparation for such further
studies.
I am grateful to many students and colleagues for pointing out
errors in the lecture notes on which this book is based and for the
expert and friendly assistance I have received from Simon Capelin
and Vince Higgs, the editors at Cambridge University Press who
guided the publication of this book.
STEVEN WEINBERG
1
Temperature Scales
A word must be said about the phrase “at constant temperature.”
Boyle lived before the establishment of our modern Fahrenheit and
Celsius scales, whose forerunners go back respectively to 1724 and
1742. But, although in Boyle’s time no meaningful numerical value
could be given to the temperature of any given body, it was
nevertheless possible to speak with precision of two bodies being at
the same temperature: they are at the same temperature if when
put in contact neither body is felt to grow appreciably hotter or
colder. Boyle’s glass tube could be kept at constant temperature by
immersing it in a large bath, say of water from melting ice. Later the
Fahrenheit temperature scale was established by defining the
temperature of melting ice as 32 ◦ F and the temperature of boiling
water at mean atmospheric pressure as 212 ◦ F, and defining a 1 ◦ F
increase of temperature by etching 212 − 32 equal divisions
between 32 and 212 on the glass tube of a mercury thermometer.
Likewise, in the Celsius scale, the temperatures of melting ice and
boiling water are 0 ◦ C and 100 ◦ C, and 1 ◦ C is the temperature
difference required to increase the volume of mercury in a
thermometer by 1% of the volume change in heating from melting
ice to boiling water. As we will see in the next chapter, there is a
more sophisticated universal definition of temperature, to which
scales based on mercury thermometers provide only a good
approximation.
After the temperature scale was established it became possible to
carry out a quantitative study of the relation between volume and
temperature, with pressure and mass kept fixed by enclosing the air
in a vessel with flexible walls, which expand or contract to keep the
pressure inside equal to the air pressure outside. This relation was
announced in an 1802 lecture by Joseph Louis Gay-Lussac (1775–
1850), who attributed it to unpublished work in the 1780s by
Jacques Charles (1746–1823). The relation, subsequently known as
Charles’ Law, is that at constant pressure and mass the volume of
gas is proportional to T − T0, where T is the temperature measured
for instance with a mercury thermometer and T0 is a constant whose
numerical value naturally depends on the units used for
temperature: T0 = −459.67 ◦ F = −273.15 ◦ C. Thus T0 is absolute
zero, the minimum possible temperature, at which the gas volume
vanishes. Using Celsius units for temperature differences, the
absolute temperature T ≡ T − T0 is known today as the temperature
in degrees Kelvin, denoted K.
Theoretical Explanations
In Proposition 23 of his great book, the Principia, Isaac Newton
(1643–1727) made an attempt to account for Boyle’s law by
considering air to consist of particles repelling each other at a
distance. Using little more than dimensional analysis, he showed that
the pressure p of a fixed mass of air is inversely proportional to the
volume V if the repulsive force between particles separated by a
distance r falls off as 1/r. But as he pointed out, if the repulsive force
goes as 1/r2, then p ∝ V −4/3. He did not claim to offer any reason
why the repulsive force should go as 1/r and, as we shall see, it is
not forces that go as 1/r but rather forces of very short range that
act only in collisions that mostly account for the properties of gases.
It was the Swiss mathematical physicist Daniel Bernoulli (1700–
1782) who made the first attempt to understand the properties of
gases theoretically, on the assumption that a gas consists of many
tiny particles moving freely except in very brief collisions. In 1738, in
the chapter, “On the Properties and Motions of Elastic Fluids,
Especially Air” of his book Hydrodynamics, he argued that in a gas
(then called an “elastic fluid”) with n particles per unit volume
moving with a velocity υ that is the same (because of collisions) in
all directions, the pressure is proportional to n and to υ2, because
the number of particles that hit any given area of the wall in a given
time is proportional to the number in any given volume, to the rate
at which they hit the wall, which is proportional to υ, and to the
force that each particle exerts on the wall, which is also proportional
to υ. For a fixed mass of gas n is inversely proportional to the
volume V, so pV is proportional to υ2. If (as Bernoulli thought) υ2
depends only on the temperature, this explains Boyle’s law. If υ2 is
proportional to the absolute temperature, it also gives Charles’ law.
Bernoulli did not give much in the way of mathematical details,
and did not try to say to what else the pressure might be
proportional besides nυ2, a matter crucial for the history of
chemistry. These details were provided by Rudolf Clausius (1822–
1888) in 1857, in an article entitled “The Nature of the Motion which
We Call Heat.” Below is a more-or-less faithful description of
Clausius’ derivation, in a somewhat different notation.
Suppose a particle hits the wall of a vessel and remains in contact
with it for a small time t, during which it exerts a force with
component F along the inward normal to the wall. Its momentum in
the direction of the inward normal to the wall will decrease by an
amount Ft, so if the component of the velocity of the particle before
it strikes the wall is υ⊥ > 0, and it bounces back elastically with
normal velocity component −υ⊥, the change in the inward normal
component of momentum is −2mυ⊥, where m is the particle mass,
so
Now, suppose that this goes on with many particles hitting the wall
over a time interval T ≫ t, all particles with the same velocity vector
v. The number N of particles that will hit an area A of the wall in
this time is the number of particles in a cylinder with base A and
height υ⊥T, or
(1.1.1)
This is for the unphysical case in which every particle has the same
value of υ⊥, positive in the sense that the particles are assumed to
be going toward the wall. In the real world, different particles will be
moving with different speeds in different directions, and Eq. (1.1.1)
should be replaced with
(1.1.2)
the brackets indicating an average over all gas particles, with the
factor 1/2 inserted in the first expression because only 50% of these
particles will be going toward any given wall area.
To express in terms of the root mean square velocity,
Clausius assumed without proof that “on the average each direction
[of the particle velocities] is equally represented.” In this case, the
average square of each component of velocity equals and the
average of the squared velocity vector is then
(1.1.3)
(1.1.4)
(1.1.5)
where k is a constant, in the sense of being independent of p, n, and
T. But the choice of notation does not tell us whether k varies from
one type of gas to another or whether it depends on the molecular
mass m. Clausius could not answer this question, and did not offer
any theoretical justification for Boyle’s law or Charles’ law. Clausius
deserves to be called the founder of thermodynamics, discussed in
Sections 2.2 and 2.3, but these are not questions that can be
answered by thermodynamics alone. As we will see in the following
section, experiments in the chemistry of gases indicated that k is the
same for all gases, a universal constant now known as Boltzmann’s
constant, but the theoretical explanation for this and for Boyle’s law
and Charles’ law had to wait for the development of kinetic theory
and statistical mechanics, the subject of Section 2.4.
As indicated by the title of his article, “The Nature of the Motion
which We Call Heat,” Clausius was concerned to show that, at least
in gases, the phenomenon of heat is explained by the motion of the
particles of which gases are composed. He defended this view by
using his theory to calculate the specific heat of gases, a topic to be
considered in the next chapter.
1.2 Chemistry
Elements
The idea that all matter is composed of a limited number of
elements goes back to the earliest speculations about the nature of
matter. At first, in the century before Socrates, it was supposed that
there is just one element: water (Thales) or air (Anaximenes) or fire
(Heraclitus) or earth (perhaps Xenophanes). The idea of four
elements was proposed around 450 BC by Empedocles of Acragas
(modern Agrigento). In On Nature he identified the elements as “fire
and water and earth and the endless height of air.” Classical Chinese
sources list five elements: water, fire, earth, wood, and metal.
Like the theory of atoms, these early proposals of elements did
not come accompanied with any evidence that these really are
elements, or any suggestion how such evidence might be gained.
Plato in Timaeus even doubled down and stated that the difference
between one element and another arises from the shapes of the
atoms of which the elements are composed: earth atoms are tiny
cubes, while the atoms of fire, air, and water are other regular
polyhedra – solids bounded respectively by 4, 8, or 20 identical
regular polygons, with every edge and every vertex of each solid the
same as every other edge or vertex of that solid.
By the end of the middle ages this list of elements had come to
seem implausible. It is difficult to identify any particular sample of
dirt as the element earth, and fire seems more like a process than a
substance. Alchemists narrowed the list of elements to just three:
mercury, sulfur, and salt.
Modern chemistry began around the end of the eighteenth
century, with careful experiments by Joseph Priestley (1733–1804),
Henry Cavendish (1743– 1810), Antoine Lavoisier (1743–1794), and
others. By 1787 Lavoisier had worked out a list of 55 elements. In
place of air there were several gases: hydrogen, oxygen, and
nitrogen; air was identified as a mixture of nitrogen and oxygen.
There were other non-metals on the list of elements: sulfur, carbon,
and phosphorus, and a number of common metals: iron, copper, tin,
lead, silver, gold, mercury. Lavoisier also listed as elements some
chemicals that we now know are tightly bound compounds: lime,
soda, and potash. And the list also included heat and light, which of
course are not substances at all.
(1.2.1)
(1.2.2)
(1.2.3)
Avogadro’s Number
Incidentally, a mole of any element or compound of molecular
weight μ is defined as μ grams, so in Eq. (1.2.3) the ratio M/μ
expressed in grams equals the number of moles of gas. Since N =
M/m1 μ, one mole contains a number of molecules equal to 1/m1
with m1 given in grams. This is known as Avogadro’s number. But of
course Avogadro did not know Avogadro’s number. It is now known
to be 6.02214 × 1023 molecules per mole, corresponding to unit
molecular weight m1 = 1.66054 × 10−24 grams. The measurement
of Avogadro’s number was widely recognized in the late nineteenth
century as one of the great challenges facing physics.
1.3 Electrolysis
Early Electricity
Electricity was known in the ancient world, as what we now call
static electricity. Amber rubbed with fur was seen to attract or repel
small bits of light material. Plato in Timaeus mentions “marvels
concerning the attraction of amber.” (This is where the word
electricity comes from; the Greek word for amber is “elektron.”)
Electricity began to be studied scientifically in the eighteenth
century. Two kinds of electricity were distinguished: resinous
electricity is left on an amber rod when rubbed with fur, while
vitreous electricity is left on a glass rod when rubbed with silk. Unlike
charges were found to attract each other, while like charges repel
each other. Benjamin Franklin (1706–1790) gave our modern terms
positive and negative to vitreous and resinous electricity,
respectively.
In 1785 Charles-Augustin de Coulomb (1736–1806) reported that
the force F between two bodies carrying charges q1 and q2
separated by a distance r is
(1.3.1)
Early Magnetism
Magnetism too was known in the ancient world, as what we now call
permanent magnetism. The Greeks knew of naturally occurring
lodestones that could attract or repel small bits of iron. Plato’s
Timaeus refers to lodestones as “Heraclean stones.” (Our word
magnet comes from the city Magnesia in Asia Minor, near where
lodestones were commonly found.)
Very early the Chinese also discovered the lodestone and used it
as a magnetic compass (a “south-seeking stone”) for purposes of
geomancy and navigation. Each lodestone has a south-seeking pole
at one end, attracted to a point near the South Pole of the Earth,
and a north-seeking pole at the other end, attracted to a point near
the Earth’s North Pole. Magnetism was first studied scientifically by
William Gilbert (1544–1603), court physician to Elizabeth I. It was
observed that the south-seeking poles of different lodestones repel
each other, and likewise for the north-seeking poles, while the south-
seeking pole of one lodestone attracts the north-seeking pole of
another lodestone. Gilbert concluded that one pole of a lodestone is
pulled toward the north and the other toward the south because the
Earth itself is a magnet, with what in a lodestone would be its south-
seeking and north-seeking poles respectively near the Earth’s North
Pole and South Pole.
Electromagnetism
It began to be possible to explore the relations between electricity
and magnetism quantitatively with the invention in 1809 of electric
batteries by Count Alessandro Volta (1745–1827). These were stacks
of disks of two different metals separated by cardboard disks soaked
in salt water. Such batteries drive steady currents of electricity
through wires attached to the ends of the stacks, with positive and
negative terminals identified respectively as the ends of the stacks
from which and towards which electric current flows.
In July 1820 Hans Christian Oersted (1777–1851) in Copenhagen
noticed that turning on an electric current deflected a nearby
compass needle, and concluded that electric currents exert force on
magnets. Conversely, he found also that magnets exert force on
wires carrying electric currents.
These discoveries were carried further in Paris a few months later
by Andrè-Marie Ampère (1775–1836), who found that wires carrying
electric current exert force on each other. For two parallel wires of
length L carrying electric currents (charge per second) I1 and I2, and
separated by a distance r ≪ L, the force is
(1.3.2)
where km is another universal constant. The force is repulsive if the
currents are in the same direction; attractive if in opposite directions.
One ampere is defined so that F = 10−7 × L/r newtons if I1 = I2 = 1
ampere. (That is, km ≡ 10−7 N/ampere2.) The electromagnetic unit
of electric charge, the coulomb, is defined as the electric charge
carried in one second by a current of one ampere. A modern
ammeter measures electric currents by observing the magnetic force
produced by current flowing through a wire loop.
The connection between electricity and magnetism was
strengthened in 1831 by Michael Faraday (1791–1867), at the Royal
Institution in London. He discovered that changing magnetic fields
generate electric forces that can drive currents in conducting wires.
This is the principle underlying the generation of electric currents
today. Electricity began soon after to have important practical
applications, with the invention in 1831 of the electric telegraph by
the American painter Samuel F. B. Morse (1791–1872).
Finally, in the 1870s, the great Scottish physicist James Clerk
Maxwell (1831–1879) showed that the consistency of the equation
for the generation of magnetic fields by electric currents required
that magnetic fields are also generated by changing electric fields. In
particular, while oscillating magnetic fields produce oscillating electric
fields, also oscillating electric fields produce oscillating magnetic
fields, so a self-sustaining oscillation in both electric and magnetic
fields can propagate in apparently empty space. Maxwell calculated
the speed of its propagation and found it to equal 2
Discovery of Electrolysis
Electrolysis was discovered in 1800 by the chemist William Nicholson
(1753–1815) and the surgeon Anthony Carlisle (1768–1840). They
found that bubbles of hydrogen and oxygen would be produced
where wires attached respectively to the negative and positive
terminals of a Volta-style battery were inserted in water. Sir
Humphrey Davy (1778–1829), Faraday’s boss at the Royal
Institution, carried out extensive experiments on the electrolysis of
molten salts, finding for instance that, in the electrolysis of molten
table salt, sodium, a previously unknown metal, was produced at the
wire attached to the negative terminal of the battery and a greenish
gas, chlorine, was produced at the wire attached to the other,
positive, terminal. Davy’s electrolysis experiments added several
metals aside from sodium to Lavoisier’s list of elements, including
aluminum, potassium, calcium, and magnesium.
A theory of electrolysis was worked out by Faraday. In modern
terms, a small fraction (1.8 × 10−9 at room temperature) of water
molecules are normally dissociated into positive hydrogen ions (H+),
which are attracted to the wire attached to the negative terminal of
a battery, and negative hydroxyl ions (OH−), which are attracted to
the wire attached to the positive terminal. At the wire attached to
the negative terminal, two H+ ions combine with two units of
negative charge from the battery to form a neutral H2 molecule. At
the wire attached to the positive terminal, four OH− ions give one O2
molecule plus two H2O molecules plus four units of negative charge,
which flow through the battery to the negative terminal.3
Likewise, a small fraction of molten table salt (NaCl) molecules are
normally dissociated into Na+ ions and Cl− ions. At the wire attached
to the negative terminal of a battery, one Na+ ion plus one unit of
negative charge gives one atom of metallic sodium (Na); at the wire
attached to the positive terminal, two Cl− ions give one chlorine (Cl2)
molecule and two units of negative charge, which flow through the
battery to the negative terminal.
In Faraday’s theory, it takes one unit of electric charge to convert
a singly charged ion such as H+ or Cl− to a neutral atom or
molecule, so since molecules of molecular weight μ have mass μm1,
it takes M/m1 μ units of electric charge to convert a mass M of singly
charged ions to a mass M of neutral atoms or molecules of molecular
weight μ. Experiment showed that it takes about 96 500 coulombs
(e.g., one ampere for about 96 500 seconds) to convert μ grams
(that is, one mole) of singly charged ions to neutral atoms or
molecules. (This is called a faraday; the modern value is 96 486.3
coulombs/mole.) Hence Faraday knew that e/m1 ≃ 96 500
coulombs/gram, where e is the unit of electric charge, which was
called an “electrine” in 1874 by the Irish physicist George Johnstone
Stoney (1826–1911). Having measured the faraday, if physicists
knew the value of e then they would know m1, but they didn’t have
this information until later. Also, no one then knew that e is the
charge of an actual particle.
so
Heat as Energy
In 1798 Benjamin Thompson (1753–1814), an American expatriate
in England, offered evidence against the idea that heat is a fluid.
(Thompson is also known as Count Rumford, a title he was given
when he later served as military adviser in Austria.) It was well
known that boring a cannon produces heat, which might be
supposed to be due to the liberation of caloric from the iron, but
Rumford observed that if the heat is carried away by immersing the
cannon in running water while it is being bored there is no limit to
the heat that can be produced.
The first measurement of the energy in heat was provided in the
mid-1840s by James Prescott Joule (1818–1889). In his apparatus a
falling weight turned paddles in a tank of water, heating the water.
The gravitational force on a mass m kilograms is m times the
acceleration of gravity, 9.8 meters/sec2 or 9.8 newtons per kilogram.
Work is force times distance, so dropping one kilogram a distance of
one meter gave it an energy equal to 9.8 newton meters, now also
known as 9.8 joules. Joule found that the paddles driven by this
dropping weight would raise the temperature of 100 grams of water
by 0.023 ◦ C, so the paddles produced heat equal to 0.023 × 100
calories, the calorie being defined as the heat required to raise the
temperature of one gram of water by one degree Celsius. Hence
Joule could conclude that 9.8 joules is equivalent to 2.3 calories, so
one calorie is equivalent to 9.8/2.3 = 4.3 joules. The modern value is
4.184 joules.
In 1847 the Prussian physician and physicist Hermann von
Helmholtz (1821–1894) put forward the idea of the universal
conservation of energy, whether in the form of kinetic or potential or
chemical energy or heat. But what sort of energy is heat? For some
nineteenth century physicists the question was irrelevant. They
developed the science of heat known as thermodynamics, which did
not depend on any detailed model of heat energy. But there was one
context in which the nature of heat energy seemed evident. In his
great 1857 paper, The Nature of the Motion which We Call Heat,
Clausius found that at least part of the heat energy of gases is the
kinetic energy of their molecules.
Kinetic Energy
The concept of kinetic energy was long familiar. If a steady force F is
exerted on a particle of mass m, it produces an acceleration F/m, so
after a time t the velocity of the body is υ = Ft/m. The distance
traveled in this time is t times the average velocity υ/2, and the work
done on the particle is the force times this distance:
(2.1.2)
Multiply this with mAmB and add the square of Eq. (2.1.1), so that
the scalar products cancel. Dividing by 2(mA + mB), the result is
another conservation law,
(2.1.3)
Equations (2.1.1) and (2.1.3) have been derived here only for the
case in which the particles are in contact only for a brief time
interval during which the force acting between the bodies is
constant, but this is not an essential requirement for we can break
up any time interval into a large number of brief intervals in each of
which the change in the force is negligible, Then, since mA vA + mB
vB and do not change in each interval, they do
not change at all, as long as the bodies exert forces on each other
only when they are in contact.
In 1669 Christiaan Huygens (1629–1695) reported in Journal des
Sçavans that he had confirmed the conservation of the total of
mv2/2, probably by observing collisions of pendulum bobs, for which
initial and final velocities could be precisely determined. Newton in
the Principia called the conserved quantity mv the quantity of
motion, while Huygens gave the name vis viva (“living force”) to the
conserved quantity mv2/2. These two quantities have since become
known as momentum and kinetic energy.
On the other hand, it was essential in deriving the conservation of
kinetic energy that we assumed that particles interact only when in
contact. This is generally a good approximation in gases, but it is not
valid in the presence of long range forces, such as electromagnetic
or gravitational forces. In such cases kinetic energy is not conserved
– it is only the sum of kinetic energy plus some sort of potential
energy that does not change.
Specific Heat
The total kinetic energy of N molecules of gas of mass m and mean
square velocity 〈v2〉 is Nm〈v2〉/2. Clausius had found the relation
(1.1.4) between mean square velocity and absolute temperature,
according to which m〈v2〉/2 = 3kT/2, where k is some constant (later
identified as a universal constant of nature), so the total kinetic
energy is 3NkT/2. A mass M of gas of molecular weight μ contains N
= M/μm1 molecules, so the total kinetic energy is 3MRT/2μ, where R
= k/m1 is the gas constant (1.2.4). Clausius concluded that to raise
the temperature of a mass M of gas of molecular weight μ by an
amount dT at constant volume, so that the gas does no work on its
container, requires an energy dE = 3MRdT/2μ. The ratio dE/MdT is
known as the specific heat, so Clausius found that the specific heat
of a gas at constant volume is
(2.1.4)
(2.1.5)
(2.1.6)
(2.1.7)
(2.1.8)
(2.1.10)
Adiabatic Changes
It often happens that work is done adiabatically, that is, without the
transfer of heat. In this case the conservation of energy tells us that
the work done by an expanding fluid must be balanced by a
decrease in its internal energy EV :
(2.1.11)
For an ideal gas, the internal energy per unit volume E is given by
Eq. (2.1.10), so this tells us that
"Te tapoitte hänet sen — sen vuoksi?" kysyi Josephine. "Ben tunsi
teidät, hän tiesi, että ette aikonut viedä minua Laskariin.
"Näet siis, että kun en kerran voi ase kädessä saada hänestä
voittoa ja kun vihaan häntä siinä määrin kuin teen, että minun täytyy
käyttää muita keinoja musertaakseni hänet ja saadakseni hänet
harmista vääntelehtimään. Sen vuoksi, että Brannon pitää sinusta —
sen vuoksi vien sinut siis Panya Cacheen. Ja toisena syynä siihen
on se, että minä pidän itse sinusta.
Viimeinen paljastus, tullen heti sen jälkeen kuin hän oli päässyt
selvyyteen Lattimerin todellisesta luonteesta, vei kokonaan hänen
voimansa. Hän tunsi Lattimerin lujat käsivarret ympärillään ja taisteli
häntä vastaan, vaikka tiesi sen toivottomaksi. Hän kuuli hänen
puhuttelevan häntä töykeästi ja tunsi, että häntä kannettiin ja sitten
käyden epätoivon vimmalla satulan nuppiin hän tunsi nousevansa
ylöspäin.
Yhdesneljättä luku.
Musta kiiti alussa tuulena tasangon yli. Sen vauhti hiljeni vähitellen
kun se kierteli muutaman matalan kukkulan juuria, joille alkuvuorista
irtautuneita kallionlohkareita oli vierinyt sen tielle. Se kiipesi
äkkijyrkkää rinnettä taitavasti kuin kissa ja juoksi sitten vuoren
harjannetta pitkän matkaa, syöksyi saccaton-heinää kasvavalle
lakeudelle huimaavaa vauhtia, riensi taasen jyrkkää vuorenkuvetta
ylös ja poikki toisen ruohokentän saman korkean vuorenhuipun
juurella, jolta Josephine edellisenä yönä oli katsellut suureen
laaksoon.
Brannon tunsi siis tien varsin hyvin. Siitäkin huolimatta olisi hänellä
ollut sangen helppo seurata Lattimeria ja Josephineä, sillä heidän
jälkensä näkyivät selvästi matkan pehmeillä osilla.
Kahdesneljättä luku.
Tottumattomana pitkiin ratsastusmatkoihin ja lopen väsyneenä
siitä jännityksestä, minkä matkan vaarat olivat aiheuttaneet, oli
Josephinen pakko jatkaa kulkuaan. Ja ajatellessaan toivotonta
tulevaisuuttaan, itki hän lohduttomasti hevosen laskeutuessa vuoren
syrjännettä heinäiselle tasamaalle.
Ja nyt, kun hän johti hevostaan rinnettä alas, kuoli hänen toivonsa
kokonaan ja hän puhkesi itkuun. Lattimer ratsasti hänen luokseen ja
pani toisen kätensä hänen olkapäälleen.
Se oli Denver. Hän oli ilmestynyt kuin näky. Mutta Josephine tiesi,
että hän olisi saattanut ratsastaa heitä kohti tuntikausia heidän
tietämättään mitään hänestä. Sillä se, joka tahtoi kulkea salassa,
saattoi menestyksellä käyttää hyväkseen niitä kukkuloita, laaksoja,
kallioita ja syvänteitä, joita seutu oli täynnään. Ja hän tiesi myös, että
Denver ei olisi näyttäytynyt jos hän vain oli huomannut heidät,
ennenkuin hän oli varma heidän henkilöllisyydestään. Se kuului
miehen luonteenominaisuuksiin.
"Brannon!"
Kolmasneljättä luku.
Ei ollut muuta kuin yksi keino ja sitä hänen täytyi käyttää. Brannon
oli ollut hänen ystävänsä koko ajan, koettaen turhaan näyttää
hänelle kuinka mielettömät hänen puuhansa olivat ja hän oli
käyttänyt oveluutta, jotta Josephine ei olisi loukkaantunut. Josephine
ei senvuoksi saattanut uhrata Brannonia edes siten pelastuakseen
itse uhkaavasta häpeästä. Koettaen vaientaa sisässään riehuvaa
myrskyä ja epätoivoisesti pyrkien tyyneksi, jotta ei Brannon epäilisi
häntä, onnistui hänen hymyillä hänelle Mustan pysähtyessä hänen
eteensä ja Brannonin hypähtäessä satulasta.
"Oli sitä mieltä, että te teitte väärin!" kertasi Brannon. "Mitä sillä
tarkoitatte?"
Mutta hän näkikin nyt, että hän oli erehtynyt ja että Brannon ei ollut
niinkään altis jättämään häntä. Josephine tunsi, että Brannonin täytyi
epäillä hänen valehtelevan tai että hänellä oli jotakin sanomista
Lattimerille Whitmanin ampumisen johdosta — vaikka, jos kerran
Whitman eli, niin —.
Jos Brannon sen vuoksi aikoi tavata Lattimeriä, niin täytyi hänen
estää se.
"Tuletteko?"