You are on page 1of 82

1

Abstract
In these notes I present an overview of electrodynamics and quantum mechanics
which (together with statistical mechanics) are the foundation of much of today’s
technology: electronics, chemistry, communication, optics, etc.
CONTENTS

1 Introduction: the Unity of Science 4


2 Quantum Mechanics 5
2.1 The puzzles of matter and radiation 6
2.1.1 Planck’s Black-body radiation 8
2.1.2 The photo-electric effect 12
2.1.3 Bohr’s atom 14
2.1.4 Adsorption, stimulated emission and the laser 16
2.2 Quantum Mechanical formalism 19
2.3 Simple QM systems 21
2.3.1 The chiral amonia molecule 21
2.3.2 The amonia molecule in a constant electric field 24
2.3.3 The amonia maser and atomic clocks 26
2.3.4 The energy spectrum of aromatic molecules 28
2.3.5 Conduction bands in solids 29
2.4 Momentum and space operators 31
2.4.1 Heisenberg uncertainty principle 34
2.5 Schroedinger’s equation 35
2.5.1 Diffraction of free particles 35
2.5.2 Quantum interference observed with C60 37
2.5.3 QM tunneling and the Scanning Tunneling Micro-
scope 38
2.6 The correspondance principle 41
2.6.1 Gauge invariance and the Aharonov-Bohm effect 42
2.7 Dirac’s equation: antiparticles and spin 44
2.7.1 Angular momentum and spin 49
2.8 The Hydrogen atom and electronic orbitals 52
2.8.1 Spin-orbit coupling 54
2.8.2 Many electron systems 55
2.8.3 The periodic table 56
2.9 The chemical bond 59
2.9.1 Hückel’s molecular orbital theory 63
2.9.2 Molecular vibrational spectrum 65
2.9.3 Molecular rotational spectrum 67
2.10 Time independent perturbation theory 68
2.10.1 The polarizability of atoms in an electric field 71
2.10.2 Atom in a constant magnetic field: the Zeeman
effect 73
2.10.3 Degenerate eigenstates 76
2.11 Time dependent perturbation theory 77

3
1
INTRODUCTION: THE UNITY OF SCIENCE

4
2
QUANTUM MECHANICS

Quantum mechanics (QM) is a theory of matter and its interactions with force fields
(here we will only care about electromagnetic fields). While classical mechanics and
electromagnetism are intuitive (one has a direct experience of gravitation, light, elec-
tricity, magnetism, etc.) quantum mechanics is not. The description of matter that arises
from the QM formalism is totally at odds with our daily experience: particles can pass
through walls, can be at two different places and in different states at the same time,
can behave as waves and interfere with each other. Worse, QM is a non-deterministic
description of reality: it only predicts the probability of observing events. This aspect
deeply disturbed Einstein who could not accept that QM was the correct final descrip-
tion of reality (as he famously quipped: ”God does not play dice”). He and many others
came up with alternative descriptions of QM introducing hidden variables (unknow-
able to the observer) to account for its non-deterministic aspects. But in 1964 John
Bell showed that if hidden variables existed some measurements would satisfy certain
inequalities. The experiments performed by Alain Aspect and his collaborators in the
1970’s showed that the Bell inequalities were violated as predicted by QM, but not by
the hidden variable theories thereby falsifying them.
Yet, for all its technical prowess Aspect’s experiment was only addressing a philo-
sophical issue concerning the foundations and interpretation of QM. The theory itself
had been amply vindicated earlier by its enormous predictive power: QM explains the
stability of atoms, their spectra, the origin of the star’s energy and of the elements and
their properties, the nature of the chemical bond, the origin of magnetism, conductivity,
superconductivity and superfluidity, the behaviour of semi-conductors and lasers, etc.,
etc.. All of today’s micro-electronic industry is derived from applications of QM (tran-
sistors, diodes, integrated circuits, etc.), the development of the chemical industry is a
result of the QM understanding of the chemical bond and the nuclear industry would of
course been impossible without an understanding of the nucleus and the nuclear forces
that QM provided.
So, for all its weirdness Quantum Mechanics is the most successfull explanation of
the World ever proposed by Mankind. It beats Platonicism, the Uppanishads, Kabbalah,
Scholasticism, etc., yet is non-intuitive and cannot be understood except by following
its mathematical formalism to its logical conclusions. ”The great book of Nature is
written in the language of mathematics”, Galileo’s quip is truer for QM more than for
any other scientific theory. More recently one of the founder’s of QM, Eugene Wigner
wrote in an article entitled ”the unreasonable effectiveness of mathematics in the natural
sciences”, that ”the miracle of the appropriateness of the language of mathematics for
the formulation of the laws of physics is a wonderful gift which we neither understand
nor deserve”.

5
6 QUANTUM MECHANICS

It is with this mind set that I would like you to approach the study of QM. Like an
apprentice sorcerer learning the tricks of his master without fully understanding them,
yet always at awe confronting their power. As we have done with electromagnetism,
we will approach QM by following as far as we can the historical narrative. We will
see why the radiation of a Black Body was such a puzzle that it prompted Max Plank to
introduce the idea that energy was quantized; why the stability of atoms and their spectra
prompted Bohr, Sommereld and others to suggest that the energy levels of atoms were
also quantized; how the idea that particles could also have wavelike behaviour was first
suggested by de Broglie and brought to fruition by Schroedinger, Heisenberg and Dirac.
And how from then on, QM revolutionized the understanding of matter, the chemical
bond, magnetism, conduction, etc.

2.1 The puzzles of matter and radiation


At the end of the 19th century, scientists disposed of a very succesfull theoretical frame-
work that could explain many of the problems known at that time and which was tech-
nologically revolutionary. Newtonian mechanics was amazingly successful in predict-
ing the motion of celestial bodies. Its most striking success was the prediction by Le
Verrier in 1846 of the existence of the planet Neptune. Analysing some anomalies in
the motion of Uranus, he predicted Neptune’s precise location in the sky, a prediction
which was immediately confirmed by German astronomers. In 1861, Maxwell unified
electricity, magnetism and optics opening the area of electrical appliances and wireless
communication: Edison invented the light bulb in 1879 and founded ”General Electric”
in 1892 while Marconi established the ”Marconi Wireless Telegraph Company” in 1897.
Finally, thermodynamics was sustaining the advance of the industrial revolution as ther-
mal engines were driving industry and railways. In spite of terrible social inequalities
(as described by C.Dickens, E.Zola and others) this was a time of peace, prosperity and
optimism, illustrated by the nascent Impressionist movement.
Yet, many fondamental scientific questions remained unsolved and paradoxical. The
chemical properties of the various elements were not understood. The periodicity of
these properties as a function of the mass of the elements as determined in 1869 by
Mendeleev in his famous Periodic Table of the Elements was a mystery. Nonetheless
on the basis of his ad-hoc classification Mendeleev predicted the existence of two new
elements, Gallium and Germanium, which were duly discovered in 1875 and 1886 and
are essential in today’s semiconductor industry! The existence of atoms (indivisible
particles of matter characteristic of each element) postulated by Dalton to explain the
properties of molecules was not generally accepted. Because of the successful applica-
tions of continuum mechanics (in the design of bridges, buildings (e.g. the Eiffel tower),
etc.) and fluid dynamics (in explaining the tides, water waves, etc.), matter was gener-
ally believed to be some sort of continuum akin to a gel not a swarm of particles. It was
Einstein who in 1905 finally managed to convince the scientific world of the existence
of atoms and molecules by showing that the erratic motion exhibited by dust particles
on the surface of water (first observed by the botanist R.Brown in 1827) was due to the
shocks of the water molecules. The continuum pre-conception also sustained the inter-
pretation of electromagnetic waves. Since all known waves at the times were observed
THE PUZZLES OF MATTER AND RADIATION 7

F. 2.1. The emission spectrum in the visible range for a few elements. Notice the fine
spectral lines and the different spectral characteristics for the different elements.
This was one of the puzzles that QM solved.

to propagate in a continuum medium (such as water, air, etc.) at a velocity v = κ/ρ


p

(where κ is the compressibility and ρ the density of the medium), the electromagnetic
waves predicted by Maxwell and discovered by Hertz were assumed to propagate in
some continuum medium: the ether which properties determined their velocity. How-
ever all attempts to detect the ”wind” of ether resulting from the motion of the Earth in
that medium proved negative. This prompted Einstein to formulate his theory of rela-
tivity which postulated the constancy of the speed of light and got rid of any notion of
ether, see Appendix.
Then there were questions related to the emission and absorption spectra of elements
that exhibited discrete lines rather than an undifferentiated continuum of absorption or
emission, as was the case for sound and water waves and as we have seen for scattered
and refracted light. Not only did the elements exhibit specific adsorption lines but those
differed from element to element. These observations did not fit with the then prevailing
conception of matter as a continuum.
Finally there was the problem of the radiation from a Black Body, a material (such as
to a good approximation graphite) which adsorbs radiation uniformly at all frequencies
and which can therefore also emit radiation uniformly at all frequencies. Notice how-
ever that many bodies (e.g. the elements just mentioned) are not black-bodies as they
adsorb/emit only at certain frequencies. At a given temperature, the radiation inside a
black body cavity is at thermal equilibrium with the walls of the cavity that absorb and
re-emit it. When computing the electromagnetic radiation energy emitted by a black-
body at a given temperature, one found its energy to diverge because the number of
modes at high frequencies diverged. This was not only absurd but also in contradiction
8 QUANTUM MECHANICS

with the experiments which studied the energy distribution inside the cavity by measur-
ing the energy leaking out of the cavity (for example through a small hole) as a function
of frequency.
To see how this comes about, consider a square cavity (an oven) of size a at tempera-
ture T . As we have seen, a body at a given temperature emits electromagnetic radiation,
see Fig.??. Imagine that the walls of the cavity are made of small oscillators emitting ra-
diation at frequency ω (like the oscillators we considered when studying the frequency
dependence of the refraction index). Stationary waves of the form sin ~k · ~r cos ωt will
be present in the cavity if its walls are reflecting (though for the energy to equilibrate
between oscillators the walls cannot be 100% reflective). To satisfy the boundary con-
ditions on the walls we shall require that: k x = πl/a, ky = πm/a, kz = πn/a (n, m, l ≥ 0).
Hence we have: k2 = (l2 + m2 + n2 )(π/a)2 ≡ (πρ/a)2 . The number of modes dNlnm with
such wavelength is:

4πρ2 ka adk k dk
dNlnm = 2 dρ= π( )2 = 8πa3 ( )2
8 π π 2π 2π
The factor 2 results from the two possible polarizations of the fields, while the factor
(4πρ2 /8)dρ counts the number of modes in a shell in the positive octant (n, m, l ≥ 0).
According to the equipartition theorem of statistical mechanics (see below) the average
energy of each oscillatory mode is: < E >= kB T . Using the relations: k = ω/c ≡ 2πν/c
(ν like f is the frequency), the energy density of the emitted radiation du = dNlnm <
E > /a3 becomes:
8πkB T 2
du = ν dν (2.1)
c3
hence the total energy, the integral of the energy density over all frequencies, diverges as
ν3 . This divergence became known as the Jeans’ (or ultra-violet) catastrophy. While the
data agreed with that formula at low frequencies, it differed at high frequencies (small
wavelengths).

2.1.1 Planck’s Black-body radiation


Rather than questioning the equipartition theorem which was verified in other contexts
or the possibility of atoms to emit light of arbitrarily high frequencies, Planck suggested
in 1900 that light was emitted by the cavity walls in very small discrete quantities,
quantas of energy: e = hν, where h, the Planck constant is:

h = 6.626 10−27 erg sec = 4.135 10−15 eV sec

, so that light of energy En is made up of n quantas: En = nhν. In that case the average
energy emitted at frequency ν is the sum over all possible energies En , weighted by their
Boltzmann probability (see below the Chapter on Statistical Mechanics):

e−En /kB T
P(En ) = P −E /k T
ne
n B

So that the average energy is:


THE PUZZLES OF MATTER AND RADIATION 9

X hν
< E >= En P(En ) =
n
ehν/kB T − 1

When hν  kB T , one recovers the previous result: < E >= kB T , however at large
emission frequencies the average energy decays as < E >∼ hν exp(−hν/kB T ). Planck
therefore suggested to modify the previous result, Eq.2.1 to yield, see Fig.2.2:

8πhν3 1
ρ(ν) ≡ du/dν = (2.2)
c3 ehν/kB T − 1
Where ρ(ν) is known as the spectral density of radiation. Identifying the smallest quanta
of energy with a light particle (a photon of energy hν), Eq.2.2 states that the density of
photons in a Black-body is:

dN p ρ(ν) 8πν2 1
= = 3 hν/k T (2.3)
dν hν c e B −1
Notice that the total energy density in the cavity is now finite:

8π5 k4B 4
Z ∞
8π(kB T )4 ∞ x3
Z
Utot = ρ(ν)dν = 3 3
dx x = T (2.4)
0 hc 0 e − 1 15h3 c3
R∞
Where we used the equality: 0 dx x3 /(e x − 1) = π4 /15. The total radiated power per
unit area trough a small hole in the cavity becomes,

cUtot 1
Z Z
cUtot
Irad = k̂ · n̂dΩ = cos θd(cos θ) = σS B T 4 (2.5)
4π 2 0

which is known as Stefan’s law and where the Stefan-Boltzmann constant:


2π5 k4B
σS B = = 5.67 10−5 erg sec−1 cm−2 ◦ K−4 = 5.67 10−8 W m−2 ◦ K−4
15h3 c2
Therefore by measuring the total intensity of the radiation leaking out from a cavity
(for example an oven) one can measure the temperature of that cavity. One can test
the validity of Planck’s law (actually how close to a black-body the cavity really is)
by measuring the dependence of the intensity on the radiation wavelength. From the
wavelength λmax at which the intensity is maximal an other estimate of the temperature
can be deduced: kB T ' hc/5λmax . For example at 300K (which corresponds to a thermal
energy kB T ' 25 meV), the maximum of emission is at λmax ∼ 10µm. The thermal cam-
eras that visualize humans and warm animals (see Fig.??) must therefore be sensitive to
far-infrared light.

The temperature of the Sun and the Earth


The Sun is to a very good approximation a black body, see Fig.2.2. The radiations
emitted by the fusion reactions occuring at its core (at temperature of 13 106 K) are
10 QUANTUM MECHANICS

F. 2.2. The emission spectra of the sun and the universe. The sun emission spectra is
pretty well fit by the spectrum of a lack body at 5770K, however notice the existence
of some specific adsorption bands in the visible and UV spectrum. The universe on
the other hand presents a spectrum that is perfectly matched by a black-body at
2.726K.
at thermal equilibrium with the reacting nuclear particles and diffuse out to the Sun’s
surface which is much cooler. By fitting the spectrum of the sunlight to Planck’s for-
mula one can determine the Sun’s surface temperature: T S = 5770K. The total power
generated at the Sun’s core and emitted at its surface is:

PS = 4πR2S σS B T S4 = 3.85 1026 W

where RS = 6.96 108 m is the Sun’s radius. Since the radius of the Sun’s core is es-
timated to be ∼ RS /5 the volume of the core is: Vcore = 1.13 1025 m3 and the average
power per unit volume generated in the Sun’s core is: PS /Vcore = 34 W/m3 . This is less
than the power generated by our body to keep warm!!
There are a few ways to verify that. Let us assume an average daily calory intake of '
3000kcal ' 1.2 107 J, which comes to a power consumption of ' 150W. Approximating
a man as a cylinder of height L = 2m and radius r = 0.2m, the power consumption per
unit volume is ' 600 W/m3 of which about half goes to metabolic activity. Alternatively
one can use Stefan’s law to estimate the losses between a body at 37◦ C (T b = 310◦ K)
and an environment at 27◦ C (T e = 300◦ K) (this is a crude estimate since other effects
such as perspiration regulate our temperature): ∆I = σS B (T b4 − T e4 ) ' 64W/m2 which
yields a power per unit volume ' 640W/m3 . From these consistent estimates we de-
duce that our power consumption per unit volume is much larger than the Sun’s!! What
makes the Sun so bright and hot is its huge mass, not its rather inefficient thermonuclear
reactions.

Let us now estimate the temperature of the Earth T E resulting from its adsorption of
THE PUZZLES OF MATTER AND RADIATION 11

the Sun’s radiation and its own radiation at T E . The sunlight impinging on the Earth at
a distance from the Sun RS E = 1.496 1011 m has an intensity:

IE = PS /4πR2S E = (RS /RS E )2 σS B T S4 = 1.37 kW/m2

Of that radiation a fraction (known as the Earth’s albedo) α ∼30% is reflected, mostly
by the clouds, snow and ice-caps. The Sun radiation power arriving at the surface of the
Earth is thus about 1 kW/m2 . It is an important number to remember when designing
solar energy plants: its sets the maximal power per unit area available from the Sun.
Notice that by measuring the radiation arriving on Earth and the angle sustained by
the Sun: θS = (RS /RS E ) one can also get an estimate of the Sun’s temperature: T S =
(IE /σS B θS2 )1/4 . The energy absorbed by the Earth heats it and is reradiated (to a good
approximation) like a black body at temperature T E . We can compute the temperature
of the Earth by a simple energy balance. At steady-state the energy radiated is equal to
the energy absorbed:
σS B T E4 4πR2E = (1 − α)IE πR2E
From which we get T E = ((1 − α)IE /4σS B )1/4 = 255K = -18C. The Earth is actually
slightly warmer because of the green house effect that reflects part of the emitted energy
back to Earth.

The Universe as a perfect black-body


While it is difficult to design a perfect black-body, since as we shall see below
bound electrons adsorb at their resonance frequency (as is the case for the Sun’s spec-
trum for example), the Universe as a whole turned out to be the best known example
of a black-body, see Fig.2.2. The Universe is bathed in a uniform radiation field of
very low frequency whose spectral distribution is perfectly matched by a black-body
at 2.726K. This phenomena was predicted by George Gamow in 1948 and observed
serendipitously by Arno Penzias and Robert Wilson in 1964 when measuring the noise
of a microwave antenna they had built. It was higher than they had expected as they
were actually detecting the 3K radiation of the Universe. This background radiation is
the most striking evidence for the existence of the Big-Bang. According to this scenario,
the Universe began as a big explosion of matter and radiation. At the beginning light
and matter interacted continuously and were in thermal equilibrium (as they are in the
Sun’s core). But then as the Universe expanded it cooled. When it reached a temper-
ature of ∼ 3000K Hydrogen atoms started to form that could not absorb non-resonant
light: radiation decoupled from matter. At that point the radiation spectrum was that of a
black body at the temperature of decoupling. It is the relics of that original radiation that
we are observing today as an isotropic cosmic background radiation. Let us see why it
exhibits a black-body spectrum at a temperature of 2.7K.
Since once hydrogen atoms formed radiation largely stopped to interact with matter,
the number of photons at frequency ν (see Eq.2.3): a3 dN p remained constant. But as the
universe continued to expand to a size a0 > a so did the radiation wavelength (recall
that in a box k = 2π/λ is a multiple of π/a), i.e. the frequency of the radiation decreased
12 QUANTUM MECHANICS

by the expansion factor αe = a0 /a. So that the energy density du0 of the background
photons at frequency ν0 obeys now:
8πν2 1
(a0 )3 du0 = hν0 a3 dN p = hν0 a3 3 hν/k T −1

c e B

02
0 3 8πν 1
= (a0 )3 α−3
e hν αe dν0 (2.6)
c3 ehν0 /kB (T/αe ) − 1
which is the energy density of a black-body at a temperature T 0 , smaller than the tem-
perature at decoupling T by the expansion factor αe : T 0 = (T/αe ):
8πhν0 3 1
du0 = 3 hν 0 /k T 0 dν0
c e B −1
Because the Universe expanded by a factor αe ∼ 1100 since the decoupling time, one
obtains a current temperature for the background radiation of T 0 = 2.72◦ K. The precise
agreement on the value of that temperature is not very important as is the observation
that the cosmic background radiation is the best Black-Body ever observed. It is also
highly isotropic in the rest frame of the Universe. As our galaxy the Milky Way moves
at about 600 km/sec with respect to the Cosmic background, the Doppler effect red-
shifts the radiation in one direction and blue-shifts it in the opposite one. This effect can
be subtracted from the measured distribution of radiation intensities. One also needs to
subtract the contribution from the stars in the galaxy (which fortunately emit at much
higher frequencies, in the visible mostly). The measured variations in the temperature of
the Universe at different angular positions are then smaller than 10−5 K, yet these small
fluctuations served as the nucleation points for the galaxies and can account for their
observed distribution, see Fig.2.3. As E.Wigner wrote it is a ”miracle ... that we neither
undersand nor deserve” that a theory devised to explain (approximatively) the radia-
tion of hot bodies has turned out to provide such an amazingly precise and powerfull
description of the Universe!
2.1.2 The photo-electric effect
Besides the emission spectrum of atoms and the black body radiation, an other ex-
periment stood in apparent contradiction with Maxwell’s electromagnetic theory: the
photo-electric effect which observed that electron were emitted from a conducting ma-
terial with an energy that depended on the color (the frequency) of the radiation not on
its intensity. This was at odds with Maxwell’s electromagnetic theory that asserted that
the energy of radiation was related to its intensity (see Eq.??) not its frequency! Ein-
stein knew the solution for Black-body radiation for which Planck had to assume that
the radiation emitters in the walls’ cavity could only emit light in small quantas. In 1905
Einstein went further and assumed that all light actually comes in small bunches, pho-
tons, which energy is proportional to their frequency: E = hν. When such a photon is
absorbed by an electron its energy is used to tear the electron from the binding potential
Φ of the material and move it at velocity v:
1 2
hν = mv + Φ (2.7)
2
THE PUZZLES OF MATTER AND RADIATION 13

F. 2.3. The temperature of the cosmic microwave background measured across the
sky by the COsmic Background Explorer (COBE) satellite. The top image is the
raw data which is red/blue shifted due to the movement of our galaxy through the
universe at ∼ 600km/s. Correcting for this Doppler shift yields the middle image
which is still ”polluted” at the equator by the light emitted from the stars in our
galaxy. Substracting that measurable emission yields the bottom image where the
temperature fluctuations of the microwave background across the Universe are as
small as 10µ◦ K. These small fluctuations nonetheless served as nucleation points
for the formation of the stars and galaxies shortly after the decoupling time.

Hence an electron can only be observed if light of high enough frequency is used to
remove it from the material. The kinetic energy of the electron increases then linearly
with the illumination frequency. The current emitted is however proportional to the
number of adsorbed photons, i.e. to the light intensity. At the time Einstein proposal
was revolutionary since it assumed that energy came in discrete packets that could not be
infinitely divided and it appeared to contradict Maxwell’s equations. It took 16 years and
confirming experiments to establish the validity of his model, for which he got the Nobel
prize in 1921 (and not for his more profound and revolutionary theories of relativity
and gravitation). Incidently from Einstein’s relation between energy and momentum:
E 2 = (mc2 )2 + (pc)2 (see Appendix) one deduces that if the energy of the photon (of
mass m = 0) is quantized so must its momentum be: p = E/c = h/λ ≡ ~k (~ ≡ h/2π).
14 QUANTUM MECHANICS

Einstein’s understanding of the photo-electric effect has had enormous technologi-


cal impact. All digital cameras are based upon it. These CCD (Charge Coupled Device)
cameras consist of an array of small capacitors (a few micron in size) each defining
a pixel (= picture element). When light (with frequency in the infrared or higher) im-
pinges on a given pixel it kicks off an electron from one side of the capacitor to the other
and charges it with an amount which is proportional to the light intensity. The charges
in a given row of capacitors are then read out by transfering them from one capacitor to
the next along the line like in a ”bucket brigade”. Thus is the image read and stored. By
covering the array of pixels with a mask-array that filters different colors (Red, Green or
Blue) the device can be transformed into a color camera where adjacent pixels respond
to different colors.
Similarly all of today’s solar cells are based on the photo-electric effect using light
to generate a current by transfering electrons in a semiconductor material from the so-
called valence band (and leaving a positively charge ”hole” behind) into the conduction
band (on which more below).

2.1.3 Bohr’s atom


Following on the footsteps of Planck and Einstein who proposed that energy and mo-
mentum were quantized: E = n~ω and p = n~k, Niels Bohr suggested in 1913 that
the angular momentum of electrons in an atom was similarly quantized: L ≡ mvr = n~
thereby explaining their paradoxical stability (see above). Indeed given the balance of
electrostatic and centrifugal forces: mv2 /r = e2 /r2 and the assumed quantization of an-
gular momentum one deduces that in the hydrogen atom the orbits of the electron are
quantized with a radius: r = n2 ~2 /me2 ≡ n2 r0 (r0 ≡ ~2 /me2 = 0.53Å is known as the
Bohr radius of Hydrogen), velocity v = e2 /n~ and energy:
1 2 e2 me4 e2 1 13.6eV
En = mv − =− 2 2 =− 2
=− (2.8)
2 r 2~ n 2r0 n n2
Thus the energy to ionise a hydrogen atom, i.e. kick off its electron from its ground state
at n = 1 is 13.6 eV. Because of energy quantization an electron orbiting the nucleus will
not radiate continuously, but emit (or absorb) radiation in quantas of energy:
1 1
hνnm = Em − En = 13.6eV( 2
− 2) (2.9)
n m
Hence the emission or adsorption spectra of atoms consists of discrete lines correspond-
ing to electronic transitions between states with different quantum numbers. For the hy-
drogen atom only the lowest energy transitions to n = 2 (the so-called Balmer series) lie
in the visible range with wavelengths in the red: 656.3nm (m = 3); in the blue 486.1nm
(m = 4) and in the UV range: 434.1nm (m = 5) and 410.2nm (m = 6), see Table 2.1 and
Fig.2.1. The explanation of the stability of atoms and the emission lines of hydrogen
was a major success of the Bohr model for which he was awarded the Nobel prize in
1922.
THE PUZZLES OF MATTER AND RADIATION 15

m= 2 3 4 5 6 7 8
Lyman series (n=1) 121.6 102.6 97.2 94.9 93.7 93.0 92.6
Balmer series (n=2) - 656.3 486.1 434.1 410.2 397.0 388.9
Pashen series (n=3) - - 1870 1280 1090 1000 954

Table 2.1 The major emission lines in the hydrogen atom. The wavelength (in nm) is
tabulated for various values of the initial (m) and final state (n)

Bohr’s approach to the hydrogen atom was generalized in 1915 by Arnold Som-
merfeld who proposed that for any bound particle (atom, harmonic oscillator, etc.) a
quantity known in classical mechanics as the action was quantized:
I
~p · d~q = nh (2.10)

where ~p, ~q are the momentum and coordinate of the particle. The Wilson-Sommerfeld
quantization rule, Eq.2.10, reduces to Bohr’s quantization of the angular momentum in
the case of the hydrogen atom since the integral pdq = 2πmvr. But the same rule also
H

explains why the oscillators of frequency ω assumed by Planck to exist in the walls of
a back-body would emit√ radiation in quantas of energy ~ω. For a harmonic oscillator
with frequency ω = k/m, the energy is:
p2 kq2 p2 mω2 q2
Eosc = + = +
2m 2 2m 2
from which one derives:
I Z qm q Z qm q
pdq = 2 2mEosc − (mωq)2 = 2mω q2m − q2 dq = πmωq2m
−qm −qm

with q2m ≡ 2Eosc /mω2 . Hence Eq.2.10 implies: Eosc = n~ω, which is the assumption
made by Planck.
The Wilson-Sommerfeld quantization rule can also be used to find the energy level
of a quantum rotator rotating at frequency ω about one of its major axes with moment
of inertia I . In that case the angular moment Iω = l~ (l = 0, 1, 2, ...) and the angular
energy is El = Iω2 /2 = ~2 l2 /2I.
In 1924, Louis de Broglie (in his Ph.D thesis!) generalizing Sommerfeld’s idea sug-
gested that all matter are described by waves. So, just as a photon possesses a momen-
tum p = ~k, thus an electron is characterized by wavevector: k = p/~. The Wilson-
Sommerfeld rule is thus equivalent to the requirement that the electron wave in a bound
system interfers constructively to form a standing wave. As we shall see in the following
de Broglie’s analogy opened the way for the formal development of quantum mechan-
ics from its analogy with optics: classical mechanics becoming to quantum mechanics
what geometrical optics is to electromagnetic waves.
16 QUANTUM MECHANICS

F. 2.4. (A) The absorption of radiation by an atom in its ground state. (B) the stimu-
lated emission of a photon in presence of radiation by an atom in its excited state:
notice that this process is the time reversal of adsorption. (C) The spontaneous emis-
sion of a photon (in absence of radiation) by an atom in its excited state.

2.1.4 Adsorption, stimulated emission and the laser


Based on the Bohr-Sommerfeld model, Einstein proposed in 1917 a simple theory of
light-matter interaction which could account for Planck’s formula and would be (40
years later) the basis for the invention of the laser. First he pointed out that since micro-
scopic processes are reversible the adsorption of a photon is indistinguishible from the
process of stimulated emission, see Fig.2.5. In other words in presence of an external
electro-magnetic field the transition rate B21 to state 2 from 1 should be the same as the
transition rate B12 to state 1 from 2. The transition being due to the interaction between
radiation and matter, the overall rate T i j (i, j = 1, 2) will depend on the spectral density
of radiation at the transition energy ρ(ν = ∆E/h) and on the density of states 1 and
2: n1 = N1 /V and n2 = N2 /V (where Ni is the number of atoms in state i and V the
volume): T i j = Bi j ρN j /V. Einstein also recognized that in abscence of interaction with
the external field, an atom in an excited state 2 could spontaneously return to the ground
state 1 by emission of a photon of energy hν. That process depends on the lifetime τ s
of the excited state and occurs at a rate A12 = 1/τ s . To summarize in steady state the
transition rates to and from each state should balance:
B21 ρN1 /V = B12 ρN2 /V + A12 N2 /V
From which since B12 = B21 we derive:
A12 /B12
ρ(ν) =
N1 /N2 − 1
Since at thermal equilibrium the probability of being in the excited state is smaller than
in the ground state by the Boltzmann factor N2 /N1 = exp(−∆E/kB T ) one obtains:
A12 1 8πhν3 1
ρ(ν) = =
hν/k
B12 e B − 1T c3 ehν/kB T − 1
THE PUZZLES OF MATTER AND RADIATION 17

F. 2.5. Principle of operation of a laser. An amplifying medium is pumped by an


external energy source (e.g. a flash lamp) to generate a higher density of excited
states than of ground states (population inversion). The medium is placed in a cavity
with reflecting mirrors, one of which lets a small fraction of the light out. The light
reflected in the cavity is amplified by the stimulated emission of the excited states.
When the threshold for lasing is achieved the losses in the cavity are balanced by
the gain from the amplifying medium.

with the identification: A12 /B12 = 8πhν3 /c3 or:

c3 λ3
B12 = =
8πhν τ s 8πhτ s
3

Einstein model could account for Planck’s Black-body radiation if atoms capable
of absorbing radiation at all frequencies are uniformly present. It also made possible
decades later the invention of the laser, acronym for Light Amplification by Stimulated
Emission of Radiation. A laser consists of a light amplifying medium inside a highly
reflective optical cavity, which usually consists of two mirrors arranged such that light
bounces back and forth, each time passing through the gain medium, see Fig.2.5. Typi-
cally one of the two mirrors is partially transparent to let part of the beam exit the cavity.
To achieve light amplification the excited state in the medium of a laser cavity emitting
at frequency ν0 = ∆E/h has to be more populated than the lower energy state. Since at
thermodynamic equilibrium low energy states are always more populated than higher
energy ones, energy must be injected in the medium to ”pump” (i.e. excite) atoms from
the ground state into the light emitting state.
Let us therefore consider light of intensity I(z, ν) = ρ(z, ν)c (0 < z < l) propagating
in a cavity of length l and cross section S . Due to stimulated emission of power dPemit
in a volume dV = S dz the increase in the light intensity is:
18 QUANTUM MECHANICS

dI
= dPemit /dV = hν0 (n2 − n1 )B12 ρ(ν)δ(ν − ν0 )
dz
c2
= (n2 − n1 ) I(z, ν)δ(ν − ν0 ) (2.11)
8πν02 τ s

For various reasons that we shall discuss later, the transition frequency between two
R the δ-function in the preceeding equation
states is never infintely sharp and one replaces
by a function g(ν) which approximates it: g(ν)dν = 1. Quite often the response g(ν)
of a resonant oscillator is appropriate (see Appendix on Fourier transforms):
γ/π
g(ν) =
(ν − ν0 )2 + γ2
One then obtains:
dI c2
= (n2 − n1 ) g(ν)I(z, ν) ≡ γ(ν)I(z, ν)
dz 8πν02 τ s

As argued above, the light intensity grows exponentially when the population of atoms
in the medium is inverted, i.e. when n2 > n1 . As light propagates in the cavity it suffer
losses (due to adsorption for example) of magnitude α cm−1 . Since light has to come
out of the cavity there are also losses due to the fact that only a portion R of the intensity
is reflected back into the cavity. For a laser to operate the losses must equal the gain:
R exp[(γ(ν) − α)l] = 1, which imply that the population inversion at threshold has to
satisfy:
8πτ s 1
(n2 − n1 )t = 2 (α − ln R)
λ g(ν) l
Hence the longer the wavelength the smaller the required population inversion for las-
ing. That is one of the reasons that masers (lasers in the microwave range) were the
first to be invented while X-ray lasers, even though of great utility, have been difficult
to develop. During steady-state laser operation the balance of losses and gain imply that
the population inversion remains at threshold. The more the ground state is pumped, the
more the excited state is induced to emit by the increased light density in the cavity, thus
keeping the difference between the density of the two states fixed at its threshold.
QUANTUM MECHANICAL FORMALISM 19

2.2 Quantum Mechanical formalism


The early 20th century investigations by Planck, Einstein, Bohr, Sommerfeld, de Broglie
, etc. revealed a picture of matter that was different from the infinitely divisible contin-
uum of energy and momentum that prevailed untill then. Many of the properties of
matter could be explained by assuming that these quantities were discrete rather than
continuous. Thus could Planck explain the radiation spectrum of black-bodies, Einstein
the photo-electric effect and Bohr the stability of atoms and the emission spectrum of
hydrogen (though not of other elements). These early efforts suggested that matter and
radiation shared similar properties: light came as photons, particles of zero mass but
possessing definite energy and momentum. Similarly electrons had wave-like proper-
ties and could interfere with themselves, as in the orbitals of Bohr’s atom. What was
missing was a conceptual framework that would unite these observations and models
with classical mechanics. The breakthrough came with the works of Werner Heisenberg
and Max Born in 1925 and Erwin Schroedinger in 1926. The later in particular wrote an
equation for the probability of finding a particle at a given position that was inspired by
the analogy pointed out by de Broglie between waves and particles. As we shall see later
the eigenvalues of the famed Schroedinger equation yield the energy levels of a bound
system, much as one determines the resonant modes of electromagnetic radiation in a
cavity (in both cases one solves a Helmholtz equation, see Appendix).
Heisenberg proposed a matrix formulation of Quantum Mechanics that was later
shown by Schroedinger to be equivalent to his own formulation. Heisenberg’s approach
however inspired the mathematically rigorous and clear formulation of Quantum Me-
chanics presented in 1930 by Paul A.M.Dirac in his landmark book (”the Principles
of Quantum Mechanics”, Clarendon press, Oxford). In the following we shall follow
Dirac’s lead.

According to Dirac, a physical system is characterized by its state |Ψ > (also called
wave-function by Schroedinger). These states are complex unit vectors in a so-called
Hilbert space, defined so that their scalar product < Ψ|Ψ >= 1. When an observable
O is measured, the system is perturbed and ends up in an eigenstate |n > of O with
eigenvalue On : O|n >= On |n > (in linear algebra the eigenstates are the vectors that
diagonalize the matrix O). As in linear algebra one can expressing the vector-state |Ψ >
in terms of the eigenvectors |n > of O:
X
|Ψ >= αn |n > (2.12)
n

Since |Ψ > is a complex vector its conjugate is: < Ψ| = n α∗n < n|. The physical
P
interpretation of the amplitudes αn is at the core of Dirac’s formulation: the probability
of observing a system in state |n > after the measurement of O is: |αn |2 . So that the
probability of measuring a value On when the system is in state |Ψ > is:

P(On ) = | < n|Ψ > |2 = |αn |2 (2.13)


20 QUANTUM MECHANICS

Notice that |Ψ > being a unit vector: n |αn |2 = 1 as it should for |αn |2 to be interpreted
P
as a probability. This is the physical interpretation of the wave-function and the intrinsic
indeterminism of QM that so annoyed Einstein. The average value of O in state |Ψ > is:
X
< Ψ|O|Ψ > = (< n|α∗ n )O(αm |m >)
n,m
X
= Om α∗ n αm < n|m > (2.14)
n,m
X X
= |αn |2 On = P(On )On =< O > (2.15)
n n

where we assumed the eigenstates to be orthonormal < n|m >= δnm . Quantum me-
chanics in Dirac’s formulation is thus reduced to linear algebra: the states are complex
vectors and the observables complex matrices with real eigenvalues, i.e. Hermitian ma-
trices satisfying Amn = A∗nm . If the eigenstates of one operator (observable) A are also
the eigenstates of an other operator B then A and B commute:

AB|n >= Abn |n >= an bn |n >= bn an |n >= bn A|n >= BA|n >

If the operators commute they can be both measured simultaneously: they are diagonal-
ized by (i.e they share) the same eigenstates. If on the other hand the eigenstates of A
and B are not identical, their simultaneous measurement is not possible. Let {|n >} be
the eigenvectors of A, then
X X
BA|n >= Ban |n >= an B|n >= an |m >< m|B|n >= an Bmn |m >
m m

On the other hand:


X X X
AB|n >= A|m >< m|B|n >= am Bmn |m >, an Bmn |m >= BA|n >
m m m

hence the operators do not commute [A, B] ≡ AB − BA , 0. We shall see later that
the non-commutability of operators (which is quite common with matrix operations)
is at the core of Heisenberg uncertainty principle which states that the position and
momentum of a particle cannot be simultaneously determined.
The energy being an important observable in physics, the energy operator (or Hamil-
tonian, H) plays an important role in Quantum Mechanics. Its eigenvalues are the pos-
sible measured energies of the system (which can be discrete or continuous) and its
eigenmodes are like the resonant modes of an oscillator or the specific orbits of the
electron in Bohr’s atom:
H|n >= n |n >
From Planck and Einstein, we know that there is a relation between energy and fre-
quency: n = ~ωn , so that an eigenstate with given frequency ωn evolves in time as
exp(−iωn t) = exp(−in t/~). Notice that if the initial state of the system is one of the
SIMPLE QM SYSTEMS 21

eigenstates |n >, it remains there with probability P = | exp(−iωn t)|2 = 1. If however the
initial state is |Ψ(0) >= n αn |n > then:
P

X
|Ψ(t) >= αn e−in t/~t |n >
n

From which one derives Schroedinger’s equation:

∂ 1 X 1
|Ψ(t) >= αn n e−in t/~t |n >= H|Ψ(t) >
∂t i~ n i~

or in the more common notation:



i~ |Ψ(t) >= H|Ψ(t) > (2.16)
∂t
This is essentially all of Quantum Mechanics: a definition of physical states as vec-
tors in a complex space, measurements as matrix operations on these vector states, the
outcome of the measurements as eigenvalues of those matrices and a description of the
time evolution of the physical states by Schroedinger’s equation. The rest is application
of this linear algebra formalism!

2.3 Simple QM systems


2.3.1 The chiral amonia molecule
Our first application of the QM formalism described above will be the amonia molecule
and the amonia maser (the microwave equivalent of the laser discussed earlier). We will
consider here the chiral amonia molecule NHDT (D and T stand for the isotopes of Hy-
drogen: Deuterium and Tritium), rather than the achiral NH3 considered by Feynman (in
Vol3 of his ”Lectures on Physics”). This choice allows us not to care about rotational
motions and it exemplifies the queer nature of QM better than the achiral molecule.
NHDT is a tetrahedron with the four atoms sitting at the four apexes. The molecule
possesses distincts enantiomers, i.e. distinct states |1 > and |2 >, which are mirror im-
ages of each other depending on whether the Nitrogen atom is on the right or the left
of the HDT plane, see fig.2.6. These states are not eigenstates of the Hamiltonian as
the Nitrogen in state |1 > can end up in state |2 > by passing through the HDT plane,
like a left-handed glove can be transformed into a right-handed one by turning it inside
out. Even though this energetically costly transition is classically impossible there is in
QM always a small probability for such a process to happen (this tunneling through an
energetically forbidden zone (a wall, see below) is one of the oddities of QM).
Due to the symmetry of states |1 > and |2 >, the Hamiltonian of this two-state
system can thus be written as: !
E0 −A
H0 =
−A E0
22 QUANTUM MECHANICS

F. 2.6. The chiral amonia molecule, NHDT consist of a nitrogen atom bound to the
different isotopes of hydrogen(H): deuterium(D) and tritium(T). This molecule exist
with different chirality: left-handed or right-handed which are mirror images of each
other. Since nitrogen is slightly more electrophilic (negatively charged) than the
hydrogen isotopes the molecule possesses a small electric dipole moment µ.
0
Its eigenvalues (eigen-energies) are: E I,II = E0 ± A and its eigenvectors are:
! !
1 1
|I >= √1 |II >= √1
2 −1 2 1

You can check that H0 |I >= E I0 |I > and H0 |II >= E II


0
|II >. In the eigenvector basis the
Hamiltonian is diagonal:
E0 + A 0
!
H0 =
D
0 E0 − A
It is a general result from linear algebra that the matrix of eigenvectors:
!
1 1 1
Λ= √
2 −1 1

diagonalizes the original matrix: H0D = ΛT H0 Λ.


Notice that the energy eigenstates for this chiral amonia molecule consist of a co-
herent superposition of a left- and a right-handed state with probability 1/2! This is a
classically absurd situation akin to Schroedinger’s famous cat paradox, see Fig.2.7. In
this gedanken experiment he proposed to couple a cat enclosed in a box to a two-state
QM system: the cat is dead if the system is in state |1 > and alive if in state |2 >. Ac-
cording to QM, before one looks into the box the cat exists as a superposition of the two
states: dead and alive; just like our chiral amonia molecule is described as a superpo-
sition of left and right-handed states. However once a measurement is made the cat is
either dead or alive; just as the chiral amonia molecule is -when observed- either left-
or right-handed.
SIMPLE QM SYSTEMS 23

F. 2.7. The gedanken (thought) experiment that Schroedinger proposed to test the va-
lidity of QM. In a closed box isolated from the external world there is a cat and a
radioactive source of say β−particles (electrons). If a particle is emitted and detected
by a Geiger counter a poison vial is broken that kills the cat. Otherwise the cat is
alive. As the state of the particle is a superposition of bound and emitted particle,
one must consider the cat to be in a superposition of a live and dead animal. For Ein-
stein that thought experiment demonstrated that QM was incomplete since it gave
rise to absurd assumptions. Yet, when isolated from the external world (a tall order
for macroscopic systems) all experiments so far are consistent with this ”absurd”
superposition of states.

So did Einstein ask: how was the cat ”really” before we looked into the box? He
claimed that the superposition is non-sensical and the cat is either dead or alive. He then
argued that some unknown factors (hidden variables) in the description of the micro-
scopic two-state sytem that determines the cat’s observed state results in the QM proba-
bilistic and verified predictions. However as mentioned earlier, the experimental viola-
tion of Bells’ inequalities suggest that Einstein was wrong and that such ”Schroedinger
cats” exist as a superposition of dead and alive states, even if we have no clue what that
means! We will discuss again that point more quantitatively below.
A crucial ingredient enters into the QM picture and it is the measurement process.
For a system to exhibit the interference effects resulting from QM state superpositions
it must not interact with the environment. These interactions are like independent mea-
surements of the system and they destroy its coherence (e.g. the two-state superposition)
by implicating (entangling them with) many external uncontrolled states. Now it is very
easy for a macroscopic system like a cat to interact with the world outside the box (for
example through the radiation it emits or adsorbs). Hence quantum experiments with
large objects are notoriously difficult to perform. So far the largest molecule for which
quantum interference effects have been demonstrated is the buckyball: C60 (see below).
Since the chiral states are given by:
24 QUANTUM MECHANICS


|1 > = ( |I > + |II >)/ 2

|2 > = (−|I > +|II >)/ 2 (2.17)
if the molecule has left-handed chirality to begin with: |Ψ(0) >= |1 >, it will evolve as:
0 0
e−iE I t/~ |I > + e−iE II t/~ |II >
|Ψ(t) >= √
2
The probability of finding the system with a left-handed chirality (state |1 >) after a
time t is: 0 0
|e−iE I t/~ + e−iE II t/~ |2
P1 (t) = | < 1|Ψ(t) > |2 = = cos2 (At/~)
4
and the probability of finding the system with right-handed chirality is P2 (t) = sin2 (At/~).
When a given molecule is measured its chirality is well defined (either left or right), but
measurements over many molecules yield the oscillating probability distribution just
computed. Since cos2 (At/~) = (1 + cos 2At/~)/2 the oscillation frequency is related to
the difference between the energy levels of the eigenstates ω0 = 2A/~. For NH3 that
frequency ν0 = ω0 /2π = 24 Ghz is in the microwave range. For NHDT it is slighly
lower due to the higher mass of Deuterium and Tritium.

The essence of Bell’s inequalities may be grasped from this simple example. Imag-
ine that a NHDT molecule prepared in a definite chiral state (|1 > or |2 >) is observed
a time δt later so that the probability of finding it in the same state is 99%. If it is
measured again a time δt later it will be observed in the same state as previously with
probability 99%. One may now ask: what is the probability of observing the system
in its initial state if we look at it a time not δt but 2δt later? If as Einstein believed
the system is at δt in a definite state that is only once in a hundred times different
from the initial state, then at 2δt the state of the system would be at worst twice in
a hundred times different from the initial state, namely the probability of observing
the system in its initial state would be at worst 98%. However the QM prediction is:
P(2δt) = cos2 (2Aδt/~) = 1 − 4(Aδt/~)2 = 0.96 (since we assumed that P(δt) = 0.99,
i.e. Aδt/~ = 0.1). The QM mechanical prediction violates the lower bound (the Bell
inequality) set by ”realistic” theories which assume that the system is in a definite state
which we have simply no way of determining, not a ”meaningless” superposition of
left and right-handed molecules. As mentioned earlier the experimental results (mea-
sured not on chiral amonia molecules but some other two-state system) vindicate the
QM prediction and rule out the ”realistic” theories.

2.3.2 The amonia molecule in a constant electric field


Since nitrogen is more electrophilic than hydrogen, it tends to be slightly more nega-
tively charged than the hydrogen isotopes and the molecule ends up with a permanent
SIMPLE QM SYSTEMS 25

electric dipole moment ~µ, as shown in Fig.2.6. In presence of an electric field E~ the en-
~ (Eq.??). If the electric field is along the x-axis in Fig.2.6,
ergy of a dipole is W = −~µ · E,
then we expect state |1 > to have higher energy than state |2 >. The Hamiltonian of the
molecule in an external electric field is thus:
E + W −A
!
H= 0
−A E0 − W
Notice that in the eigenbasis representation (where the unperturbed Hamiltonian H0 is
diagonal), the perturbed Hamiltonian can be written as:

E0 + A 0
! !
0 W
H 0 = ΛT HΛ = H0D + δH = + (2.18)
0 E0 − A W 0
√ √
The eigenvalues of H (and H 0 ) are: E I,II = E0 ± A2 + W 2 . Defining tan θ ≡ ( A2 + W 2 −
W)/A the eigenvectors of H are:

cos θ sin θ
! !
|I >= |II >=
− sin θ cos θ

When W = 0 one recovers the previous result. When W  A: |I >' |1 > and |II >'
|2 >, in which case the enantiomers are also eigenstates. In practice however W =
µE  A and the energies vary as: E I,II = E I,II0
± µ2 E 2 /2A. State |I > (with its dipole
essentially opposite to the electric field) has higher energy than state |II >. Notice that
we can write the energies E I,II in the following form (we will see later that this is a
general result when the diagonal Hamiltonian is perturbed by an amount δH)
δH12 δH21
E I = E I0 + (2.19)
E I0 − E II
0

δH21 δH12
E II = E II
0
+ 0 (2.20)
E II − E I0
These results suggest a way to separate the eigenstates by passing a beam of amonia
molecules through a strong electric field gradient. This field gradient generates a force
on the molecules:
~ I,II = ∓ µ ∇E
2
F~ I,II = −∇E ~ 2
2A
which separates them: state |I > is deflected to regions of small electric fields, while
state |II > is deflected to regions of high electric fields. Thus can a sub-population
inversion be generated where high energy amonia molecules are separated from lower
energy ones. This high energy sub-population can then be used to amplify microwave
radiaton by stimulated emission. The resulting device, known as a maser, was the first
implementation of a stimulated radiation amplification device and served as the first
atomic clock.
26 QUANTUM MECHANICS

F. 2.8. Principle of operation of a maser. An amonia beam (here the chiral molecule
NHDT) is sent trough a slit into a beam splitter that consists of a strong inhomoge-
nous electric field. At high field one can separate the different enantiomers which
are eigenstates of the energy. The high energy eigenstate (|I >) is sent into a cavity
where it goes into the low energy state |II > by emitting stimulated radiation at the
resonant frequency ω0 = (E I − E II )/~

2.3.3 The amonia maser and atomic clocks


In an amonia maser, Fig.2.8, the high energy state |I > selected as described, enters
a resonant cavity (tuned to the transition frequency ω0 ). A population inversion of the
amonia molecules is thus generated in the cavity and amplification of radiation can be
expected. The emitted radiation stimulated by the presence in the cavity of an external
source generates a highly coherent microwave beam.
To analyse the operation of a maser, let us assume that amonia molecules enter a
cavity in which they experience a time varying electric field: E~ = Ee−iωt x̂. This field
couples with the dipole moment of the molecules to modulate their energy by W(t) =
−µEe−iωt . The perturbed Hamitonian in the eigenbasis (|I > and |II >) is then, see
Eq.2.18: ! !
EI 0 0 W(t)
H = H0 + δH(t) =
0 D
+ (2.21)
0 E II W ∗ (t) 0
Looking for a general solution: Ψ(t) >= C I (t)|I > +C II (t)|II >, Eq.2.16 yields:
i~∂t C I = E I C I + W(t)C II
i~∂t C II = W ∗ (t)C I + E II C II (2.22)
Looking for a solution C I,II = αI,II (t) exp(−iE I,II t/~) yields the following equation for
αI,II :
i~∂t αI = W(t)e−i(E II −EI )t/~ αII = −µEe−i(ω−ω0 )t αII
i~∂t αII = W ∗ (t)e−i(E I −E II )t/~ αI = −µEei(ω−ω0 )t αI (2.23)
At the resonance: ω = ω0 , one obtains:
SIMPLE QM SYSTEMS 27

∂2t αI,II + Ω2 αI,II = 0 with Ω = µE/~

which solution is αI (t) = cos Ω(t − t0 ) and αII = sin Ω(t − t0 ). The probability of being in
state |I > (α2I ) or II > (α2II ) oscillates with frequency 2Ω: the molecule go periodically
from stimulated emission in state |I > to adsorption in state II >. These oscillations
are different from the oscillations between the enantiomers |1 > and |2 > observed at
frequency ω0 in absence of electric field. This oscillation in the probability of observing
a specific enantiomer is not associated to emission of any radiation, it results from the
fact that the enantiomers are not eigenstates of the Hamiltonian: in absence of a time
varying electric field, if the molecule is in eigenstates |I > or |II > it remains there.
If the frequency of the electric field is slightly off resonance ω − ω0 ' 0 and if the
molecule remains in the cavity for a short time t  1/Ω we may assume that αI ' 1
and integrate the equation for αII to yield:

ei(ω−ω0 )t − 1
αII (t) = iΩ
i(ω − ω0 )
The probability of transition from state |I > to state |II > is then:

sin2 (ω − ω0 )t/2
PI→II (t) = |αII |2 = (Ωt)2
[(ω − ω0 )t/2]2
The function sinc x ≡ sin x/x decays rapidly for values of |x| > π. Hence the transition
rate is significant only for frequencies which are very close to resonance: |ω − ω0 | <
2π/t. If the molecule remains in the cavity for 1 sec, the relative possible detuning :
|ν/ν0 − 1| = 1/ν0 t ∼ 4 10−11 . Only frequencies within that very narrow range can be
amplified by stimulated emission. This allows for very high Q resonators, Eq.??, i.e.
very precise frequency generators and clocks. Because of the sharpness of the function
sinc2 (ω − ω0 )t/2 one often rewrites the previous equation using the limit


lim sinc2 (ω − ω0 )t/2 = δ(ω − ω0 )
t→0 t
to obtain the transition rate from state |I > to |II >:

dPI→II (t) 2π|δH12 |2


BI,II = = 2πΩ2 δ(ω − ω0 ) = δ(E I − E II − ~ω) (2.24)
dt ~
Notice that if we had assumed the amonia beam to enter the cavity in state |II >, i.e.
with αII = 1, then we could have similarly integrated the equation for αI to yield the
transition probability from state |II > to |I >, which comes to be:

2π|δH21 |2
BII,I = δ(E I − E II − ~ω) = BI,II
~
which was the intuitive assumption Einstein made: the rate of stimulated emission BI,II
is equal to the rate of adsorption BII,I . Since the energies of the excited states are never
28 QUANTUM MECHANICS

perfectly sharp, due to the Heisenberg uncertainty principle that we shall see below and
due to thermal motion that leads to Doppler broadening of the emission lines (see sec-
tion??), the δ-function is replaced by the densityRof states with the appropriate energy:
ρ(E) (with E = E II + ~ω) and the normalisation ρ(E)dE = 1).
The amonia maser was the first atomic (or molecular) clock. The present genera-
tion of atomic clocks use a microwave transition in Cesium (Cs) as the reference fre-
quency for the clock. The atoms are cooled to very low temperatures (µ◦ K) to reduce
the Doppler broadening of their emission lines and keep them for as long as possible
in the cavity, usually a Penning trap (see ????). As a result the record for the frequency
precision of atomic clock is: |ν/ν0 − 1| ∼ 10−16 .

2.3.4 The energy spectrum of aromatic molecules


A simple generalization of the two-state system that we considered previously is the
n-state system consisting for example of a circular chain of n identical atoms such as
benzene C6 H6 (n = 6) around which electrons can hop. If an electron has energy E0
when associated with a particular atom |n > and can hop only between nearest neighbors
the Hamiltonian for this system in the basis of the atom’s position |n > is:

 E0 −A 0 . . . −A 
 
 −A E −A . . . 0 
0
H =  . .. 
 
(2.25)
 .. . 
−A . . . 0 −A E0
 

The eigenstates of that system obey: H|Ψ >= E|Ψ >, with |Ψ >= n Cn |n > From
P
which we derive the equations:
E − E0
C1 + C2 + Cn = 0
A
E − E0
C1 + C2 + C3 = 0
A
..
.
E − E0
C1 + Cn−1 + Cn = 0
A
Looking for a solution: Clm = exp i(2πlm/n) where l = 1, 2, ...., n we get:
El − E0
= −(ei2πl/n + e−i2πl/n )
A
Thus the eigen-energies of the system are:
El = E0 − 2A cos 2πl/n (2.26)
For benzene (n = 6) we have: E6 = E0 − 2A; E1,5 = E0 − A; E3 = E0 + 2A and E2,4 =
E0 + A. Energies E1,5,6 which are smaller that E0 are associated to so-called bonding
SIMPLE QM SYSTEMS 29

F. 2.9. Band-gap theory of material. (a) If the gap between the valence and conduction
band is large (a few electron-Volts) the material is an insulator. (b) Variation of the
band-gap energy with wave-vector k. Top curve: energy of electrons Ee ; bottom
curve: energy of holes Eh (increases towards the bottom). (c) A semi-conductor
which due to thermal excitation or illumination has some electrons in the conduction
band and some holes in the valence band. (d) a conductor is a material for which
the valence and conduction band overlap or equivalently for which an energy band
is not filled.
orbitals or wave-functions, whereas energies E2,3,4 > E0 are associated to anti-bonding
orbitals, on which we shall have more to say later. Notice that the eigen-state associated

to the maximally bonding orbital (l = 6) is fully symmetric: |Ψ6 >= (1, 1, 1, 1, 1, 1)/ 6,
while that associated to the maximally anti-bonding orbital (l = 3)√is antisymmetric to
the permutation of nearest neighbors: |Ψ3 >= (1, −1, 1, −1, 1, −1)/ 6.

2.3.5 Conduction bands in solids


An interesting generalization of the previous analysis is the case of a long chain of
n atoms a distance a apart. In that case we have: Clm = exp[i(2πl/na)ma]. Since the
position of atom m is x = ma we may write Ck (x) = exp ikx with k = 2πl/na. Like the
oscillation modes on a string of length na, the electron’s eigenstates are 1D transverse
waves of wavelength λ = 2π/k = na/l. The energy of such a mode is:

E(k) = E0 − 2A cos ka (2.27)

The energy of the electron is bounded: E0 − 2A < E(k) < E0 + 2A. If each atom
contributes two electrons, these 2n electrons will occuy all the n-energy states (we shall
see later that each state can accomodate 2 electrons) and electron hoping in this energy-
band will be impossible. If however the on-site energy E0 can possess discrete values
(El as it does indeed, for example in Bohr’s model), the coupling of the n−atoms will
generate energy bands around each energy value El into which electron may hop. This
forms the basis of the band-theory of conduction in materials, see Fig.2.9: a material
will conduct if there are empty states into which electrons can hop. If the low energy
30 QUANTUM MECHANICS

F. 2.10. A p-type semiconductor consists of a material (usually Silicon, Si) doped
with an element (such as Aluminium) which having less electrons in its outer shell
than Si tend to trap electrons from the lattice, leaving a electron vacancy instead:
a hole. This depletes the valence bands of electrons allowing conduction through
motion of holes. A n-type semiconductor consists of a material doped with an elec-
tron donor, an element (such as Phosphate) which has more electrons in its outer
shell than Si. This injects electrons into the conduction band allowing the material
to conduct electricity. A pn-junction is formed when p-type and n-type semiconduc-
tors are brought in contact. Such a junction can be used as a Light Emitting Diode
(LED) when electron from the n-side of the junction recombine with holes from the
p-side. The wavelength of emitted light is determined by the energy-gap between
the conduction and valence and bands.
(valence) band is filled with electrons and the next (conduction) band is empty but
many electron-volts (eV) above it, then the electrons have no states in which to go and
the material is an insulator. If the conduction band overlaps with the valence band then
the electrons have empty states to hop to and current can flow in the material. The
interesting and technologically important case is the situation where the band-gap Eg
between the conduction and valence bands is small: Eg . 1eV (the gap for Silicon
(Si) is: Eg = 1.1eV; for Germanium (Ge): Eg = 0.72eV). In that case electrons can be
transfered from the valence to the conduction band by thermal agitation or via the photo-
electric effect resulting in a material that can behave either as a metal or an insulator
depending on the external conditions (temperature, voltage, illumination wavelength
and intensity, etc.).
In particular the introduction of atomic impurities (doping) that donate electrons to
(or accept electrons from) a semiconductor lattice creates a situation where few elec-
trons occupy an almost empty conduction band (or few holes (electron vacancies) oc-
cupy an almost full valence band). As we shall see later the electrons in an atom oc-
cupy certain orbitals or shells around the nucleus. Atoms (such as Phosphate or Ar-
senic) that have more electrons in their outer shell than the bulk semiconductor (the
so-called majority carrier, usually Silicon) will usually donate an electron to the lat-
tice, whereas atoms that have less electrons in their outer shell (such as Boron or Alu-
minium) will accept an electron from the lattice. The doping creates so called n- and
MOMENTUM AND SPACE OPERATORS 31

p-type semiconductors that have electrons in their conduction-band (n-type) or holes in


their valence-band (p-type), see Fig.2.10. The energy of the electrons moving near the
bottom Emin = E0,c − 2Ac of the conduction band (i.e. when ka << 1 in Eq.2.27) is:
E(k) = Emin + ~2 k2 /2mc
where the effective mass of the electron in the conduction band is defined as: mc =
~2 /∂2k E(k) = ~2 /2a2 Ac which can be quite different from the mass of a free electron.
Notice that if, as proposed by de Broglie, we identify the momentum of the electron
as p = ~k , then Eq.2.3.5 represents the kinetic energy of an electron with mass mc at
the bottom of the conduction band. Similarly the energy of the hole near the top of the
valence band is:
Eh (k) = Emax − ~2 k2 /2mv
The movement of holes in the valence band is similar to that of air bubbles in water: as
the moving water displaces the bubble up, the electrons moving in the opposite direction
to the hole minimize their energy by displacing it to the top of the valence band. As a
result holes have minimal energy at the top of the valence band: their energy increases
as k increases.

Doped semiconductors are the basic ingredients of all the semi-conductor industry
(transistors, diodes, integrated circuits, etc.). For example the coupling of p-type and
n-type semiconductors, generate a pn-junction which acts like a diode (current flows in
only one direction). It can also be used as a powerful and efficient light source. Electrons
in the n-type part of the junction can recombine with holes in the p-type (i.e. transit to
the valence band) with emission of light (just as in the atomic or molecular transitions
discussed earlier in the context of stimulated emission and the laser). The advantage
of these Light Emitting Diodes (LED) is that by appropriate tuning of the energy gap
(appropriate choice of the semi-conducting material) one can tune the wavelength at
which the LED will emit light. For example in the red and infrared part of the spectrum
Galium-Arsenide (GaAs) is the material of choice for LEDs. The intensity of light is
controlled by the current flowing through the junction. These LED are used in all kind
of electronic displays, in new high efficiency spot-lamps and traffic lights, in the remote
control of various electronic device, in the laser diode of DVD players, etc.

2.4 Momentum and space operators


In the example above we considered the case of a QM system that could occupy only
discrete states |n >. It is easy to generalize that to free particles that can be found at
continuous positions |x >. In that case the general wave-function in position (or real)
space can be formally written as:
X
|Ψ >= Ψ(x)|x >
x
32 QUANTUM MECHANICS

Where Ψ(x) is the probability amplitude of finding the particle at position |x >, so that:

P(x) = Ψ∗ (x)Ψ(x) (2.28)

Of course the probability of finding the particle somewhere is one, so that the wave-
function Ψ(x) is normalized:
Z
dxΨ∗ (x)Ψ(x) = 1 (2.29)

The mean position of the particle is:


Z Z
< x >= dxP(x)x = dxΨ∗ (x)xΨ(x) (2.30)

For example, the wavefunction can be:


1
e−(x−x0 ) /4σ
2 2
Ψ(x) =
(2πσ2 )1/4
for which the probability distribution:
1
e−(x−x0 ) /2σ
2 2
P(x) = √
2πσ
corresponds to a gaussian distribution with mean < x >= x0 and standard deviation
< (x− < x >)2 > = σ. In Fourier-space (see Appendix) we can write:
p

Z Z
1 1
Ψ(x) = √ dkΨ(k)eikx = √ d pΨ(p)eipx/~ (2.31)
2π 2π~
where we used de Broglie’s relations: p = ~k. Conversely:
Z
1
Ψ(p) = √ dxΨ(x)e−ipx/~ (2.32)
2π~
Ψ(p) is the wavefunction in the momentum (or Fourier) space: |Ψ >= p Ψ(p)|p >. For
P
the example chosen above the momentum wavefunction is:

e−ipx0 /~
Z ∞
dxe−(x−x0 ) /4σ e−ip(x−x0 )/~
2 2
Ψ(p) = √
2π~(2πσ ) 2 1/4 −∞
−ipx0 /~ −p2 σ2 /~2 Z ∞
e e
dxe−(x−x0 +2ipσ/~) /4σ
2 2
= √
2π~(2πσ2 )1/4 −∞

2σ/~ −ipx0 /~ −p2 σ2 /~2
= e e
(2π)1/4
Being the Fourier transform of Ψ(x), Ψ(p) satisfies the normalization condition:
MOMENTUM AND SPACE OPERATORS 33

Ψ∗ (p0 )Ψ(p)
Z Z Z
0
1= dxΨ∗ (x)Ψ(x) = d pd p0 dxei(p−p )x/~
2π~
Z Z
= d pΨ (p)Ψ(p) =

d pP(p) (2.33)

where we used the identity (see Appendix on Fourier transforms):


Z
dxeipx/~ = 2π~δ(p) (2.34)

Eq.2.33 is known in the theory of Fourier transforms as Parseval’s theorem. For a Gaus-
sian wavefunction Ψ(x) (see abovepexample) P(p) corresponds to a Gaussian centered
on p = 0 with standard deviation < (p− < p >)2 > = ~/2σ. The mean value of the
momentum satisfies:
Z Z
< p >= d pΨ∗ (p)pΨ(p) = d p0 d pΨ∗ (p0 )δ(p0 − p)pΨ(p)

Using the identity Eq.2.34 and Eq.2.31 allow us to express the momentum operator in
real space as:
Z
1 0
<p>= dxd p0 d pΨ∗ (p0 )ei(p−p )x/~ pΨ(p)
2π~
~ ∂
Z Z Z
1 0
= dx d p0 Ψ∗ (p0 )e−ip x/~ d p Ψ(p)eipx/~
2π~ i ∂x
~ ∂
Z Z
= ∗
dxΨ (x) Ψ(x) ≡ dxΨ∗ (x) p̂Ψ(x)
i ∂x
where we identify the momentum operator in real space as:
~ ∂
p̂ = ≡ −i~∂ x (2.35)
i ∂x
Similarly by writing the mean position < x > in momentum space we come to identify
the position operator in momentum space as:

x̂ = i~ ≡ i~∂ p (2.36)
∂p
Notice that momentum and space-operators do not commute. In real space:
Z Z
~
< x|[x, p]|x >= dxΨ∗ (x)(x p̂ − p̂x)Ψ(x) = dxΨ∗ (x)[x∂ x Ψ − ∂ x (xΨ)] = i~
i
We would have obtainedR the same result by computing the commutator in momentum
space: < p|[x, p]|p >= d pΨ∗ (p)( x̂p − p x̂)Ψ(p) = i~. Since two observable cannot
share the same eigenstates if they do not commute, the position and momentum of a
particle cannot be simultaneously determined with absolute precision. The same result
holds also for the commutator of time and energy:
Z
< x|[H, t]|x >=< x|Ht − tH|x >= dxΨ∗ (x)(i~∂t t − ti~∂t )Ψ(x) = i~
34 QUANTUM MECHANICS

2.4.1 Heisenberg uncertainty principle


From the result that momentum and space do not commute we can derive the Heisenberg
uncertainty principle. Consider the action on state |ψ > of the operator xo +iλpo where λ
is a number and the operators xo and po are the deviation of the position and momentum
operators from their mean: xo = x− < x > and po = p− < p >:

|φ >= (xo + iλpo )|ψ >

Since [xo , po ] = [x, p] = i~ the positiveness of the probability implies that:

0 ≤ < φ|φ > = < ψ|(xo − iλpo )(xo + iλpo )|ψ >
= < ψ|xo2 + iλ[xo , po ] + λ2 p2o |ψ >
= < ψ|xo2 |ψ > −~λ+ < ψ|p2o |ψ > λ2 = P2 (λ)

For the quadratic polynomial P2 (λ) to be non-negative for any real λ, its determinant
has to satisfy:
~2 − 4 < ψ|xo2 |ψ >< ψ|p2o )|ψ >≤ 0

Or in terms of the position and momentum variables:

~2
< ∆x2 >< ∆p2 > ≥ (2.37)
4

This is Heisenberg’s principle:


√ it sets a limit on the precision withpwhich one can mea-
sure both the position δx = < ∆x2 > and the momentum δp = < ∆p2 > of a physi-
cal system. The smallest uncertainty (the equality in Eq.2.37) is obtained for a Gaussian
probability distribution as can be verified from the example worked out above. Since
the Hamiltonian and time operators do not commute similar uncertainty relation can be
obtained for the energy and time uncertainties:

~2
< ∆E 2 >< ∆t2 > ≥ (2.38)
4

Notice that Heisenberg principle is a direct mathematical consequence of the QM


description of physical systems by a complex wave-function Ψ(x) (Eq.2.13) and of de
Broglie’s relation between wavelength and momentum. Heisenberg uncertainty princi-
ple is a tautology: a consequence of the definition of Fourier transforms (see Appendix).
In the context of communication it has been known for a long time: a very short time
signal is spread over a very large frequency spectrum. In the context of optics we have
also encountered it in the diffraction pattern from a hole which is larger the smaller the
hole is.
SCHROEDINGER’S EQUATION 35

2.5 Schroedinger’s equation


We have determined the representation of the position operator in momentum space and
of the momentum operator in real space. Note that in its eigenspace the momentum
space p̂ is diagonal, i.e. it is a number. We can now write the representation of the
Hamiltonian in any of these Hilbert spaces. It is often easier to work in real space, in
which case the Hamiltonian which is the sum of kinetic and potential energy is written
as:
p̂2 ~2 ~2
H= + V(~x) = − (∂2x + ∂2y + ∂2z ) + V(~x) = − ∇2 + V(~x) (2.39)
2m 2m 2m
and Schroedinger’s equation, Eq.2.16 can be recast as:
∂ ~2
i~ Ψ(~x, t) = − ∇2 Ψ(~x, t) + V(~x)Ψ(~x, t) (2.40)
∂t 2m
Multiplying Eq.2.40 by Ψ∗ , its complex conjugate by Ψ and subtracting the two yields
a conservation law for the probability distribution P(x) = |Ψ(x)|2 :
~2 ∗ ~ 2 ~ 2 Ψ∗ )
i~∂t P = − (Ψ ∇ Ψ − Ψ∇
2m
~2 ~ ~ − Ψ∇Ψ
~ ∗)
= ∇ · (Ψ∗ ∇Ψ
2m
which can be recast in the usual form (see Eq.?? for the charge distribution in EM):

~ · J~ = 0 ~ ~ Ψ∗ v̂Ψ + c.c.
∂t P + ∇ with J~ = Ψ∗ ∇Ψ + c.c. = (2.41)
2im 2
where c.c stands for the complex conjugate and v̂ = ~∇/im ~ is the velocity operator.
Eq.2.41 is a very important self-consistency check of Quantum Mechanics, since for
P(x) to be interpreted as a probability distribution it must satisfy a conservation law.
This equation expresses the intuitive expectation that the change in the probability of
finding a particle at a given position is equal to the particle flux gradient.

2.5.1 Diffraction of free particles


If the potential is null V(x) = 0 then the eigen-solutions of Schroedinger’s equation,
Eq.2.40:
dΨ ~2
i~ = − ∇2 Ψ (2.42)
dt 2m
are plane waves:
~
Ψ(~x, t) = eik·~x−iEk t/~
with ~p = ~~k and Ek = ~2 k2 /2m. These plane-waves are eigenmodes of the momentum
operator:
~ x, t) = ~p Ψ(~x, t)
p̂Ψ(~x, t) = −i~∇Ψ(~
36 QUANTUM MECHANICS

F. 2.11. The double-slit or Young’s experiment. (a) a wave passing through two slits in
a screen generates two wave-sources which interference creates on a far-away screen
a pattern of interference consisting of alternating minima and maxima of intensity.
(b) a particle in state |O > impinging on a double slit generates two states |I > and
|II > corresponding to its passage through slit 1 or 2. The phase of these states
evolves as exp ikl. If their coherence is maintained they can interfere on a screen a
large distance z from the slits, generating an oscillating pattern related to their phase
difference: φint = k(l2 − l1 ) = kd sin θ. (c) Observation of the interference pattern of
electron passing through a double slit and impinging on a camera. Each electron is
observed as a particle (white dot) with a well defined position on the camera. The
QM interference pattern is only visible when a sufficiently large number of particles
has been observed (A.Tonomura, Proc.Natl.Acad.Sci. 102, 14952 (2005).

One of the most striking confirmations of the QM mechanics picture is the obser-
vation of a diffraction pattern like the one seen with electro-magnetic radiation when
a free particle is passed through one or a few slits (see ??? and Fig.2.11(a)). Let the
particle be in an eigenstate

|0 >= Ψ(~x, t) = eikz−iEk t/~

as it impinges on a screen that is absoring except for two apertures of size a a distance
d apart. In the far-field, i.e. at distances z  d2 /λ the wave amplitude is given by
Huygens’s principle, Eq.??:

eik(z+(x +y )/2z
2 2 Z
dx0 dy0 e−i(kx x +ky y )
0 0
Ψdi f f (x, y, z) = (2.43)
iλz
Where k x = kx/z and ky = ky/z. Thus the probability of detecting a particle on a screen
a distance z from the slits is:

4a2
|Ψdi f f |2 = sinc2 k x a cos2 k x d/2 (2.44)
(λz)2
where sinc x ≡ sin x/x. If the distance between the diffraction slits is much larger than
their width (d  a), the probability oscillates with a period: δx = λz/d. While each
SCHROEDINGER’S EQUATION 37

particle is observed as such, i.e. localized at a given position, the overall probability
distribution of the particles varies in space just as the diffraction pattern of EM radiation
does.
There is an other way to derive that result, see Fig.2.11(b). The wave-function of the
particle on the screen can be written as: |Ψ >= |I > +|II > where state |I >∼ eikl1 |O >
corresponds to the particle passing through the first slit and state |II >∼ eikl2 |O > to the
particle passing through the second slit. If the slits are assumed to be thin, i.e. a  d,
then

|Ψdi f f |2 = < Ψ|Ψ >∼ |1 + eik(l2 −l1 ) |2 < O|O >= 4 cos2 kd sin θ/2
= 4 cos2 k x d/2 ≡ 4 cos2 (φint /2)

since l2 − l1 ' d sin θ ' xd/z. Notice that the interference pattern is observed only if the
phase-coherence between states |I > and |II > is maintained. For example if one wanted
to see through which slit the particle has passed, one would have to use radiation with
wavelength λo . d. Such radiation would impart on the detected particle a momentum
uncertainty δpo ∼ h/λo . In the far-field d2 /λz  1 the observation imparts on the
observed state a phase uncertainty:

δφo = δpo z/~ ∼ 2πz/λo > 2πx/d  2πxd/λz ∼ φint

which is enough to destroy the interference pattern. This is the essence of Heisenberg’s
principle: if the particle interacts with radiation of small enough wavelength as required
to determine its position the perturbation to its momentum will be so large as to destroy
the coherence of the states. Trying to determine through which slit the particle has
passed, namely to localize its position to better than the distance between the slits d,
imparts on the particle a momentum uncertainty large enough to destroy the interference
pattern, replacing it with an image of the two slits.

2.5.2 Quantum interference observed with C60


In 1999, the group of Anton Zellinger observed QM interference effects in the largest
molecule investigated so far: the fullerene C60 , a molecule that can be seen with a
Scanning Tunneling Microscope (STM, see below), Fig.2.12(a). A beam of molecules
at ∼ 900◦ C with a velocity distributed around 220 m/sec was shot at slits of size
a ∼ 40nm and inter-distance d ∼ 100nm. At this velocity the wavelength of the parti-
cles is λ = 0.025Å! They were observed on a detector positioned a distance z ' 1.25m
from the slits (i.e. in the far-field λz  d2 ). Since the size of the slits a is only slightly
smaller than d the intensity is strongly damped (see Eq.2.44) by the factor sinc2 k x a when
x > λz/2a. Hence only a few oscillations in the probability of detection can be observed
in Fig.2.12(c). In particular the first peaks (fringes) at x = 0 and x = ±λz/d ∼ 30µm
can be seen. The contrast is not as high as expected since the velocity of the molecules
(hence their momentum and the distance between fringes) is thermally spread. It is in-
teresting however that the coherence of the molecule passing through the slits can be
38 QUANTUM MECHANICS

F. 2.12. The double-slit or Young’s experiment performed with the C60 molecule (Na-
ture 401, 680 (1999)). (a) Scanning tunneling microscope (STM) images of a C60
cluster and the reconstructed theoretical image of the cluster. (b) Schematic drawing
of the double slit experiment with C60 and the expected pattern, which is a cosine
wave modulated by the one-slit diffraction pattern (sinc2 k x a). (c) Upper plot: results
of the diffraction of the molecule from the slits (the continuous line is a best fit to
the theoretical prediction). Notice that the small side peaks are observed at a dis-
tance δx = λz/d ∼ 30µm from the central peak as expected for the diffraction from
two slits (see text). Bottom plot: the results when no slits are present and hence no
diffraction is expected (control experiment).

maintained with molecules at such high temperatures. The reason for that coherence is
that the particle is emitting black-body radiation at a wavelength of a few µm which is
much larger than the distance between the slits d ∼ 0.1µm. Hence the uncertainty in the
molecule’s momentum imparted by its black-body radiation is not enough to destroy
the coherence of the wave-function passing through contiguous slits.

2.5.3 QM tunneling and the Scanning Tunneling Microscope


A consequence of Heisenberg’s principle and a queer feature of Quantum mechanics is
the possibility for particles to pass through energetically forbidden regions. Imagine a
particle moving with momentum p along the x-axis and encountering a wall (an energy
barrier of height V > p2 /2m = E and size x = d, see Fig.2.13(a)). In classical mechanics
the particle is reflected from such a wall. However in QM, the particle has a finite
probability of passing through the wall and emerging on its other side! On the left side
of the wall the particle’s wavefunction is the sum of an incident and a reflected wave:
Ψl (x) = Aeikx + Be−ikx
with k2 = 2mE/~2 , while on its right side the transmitted wavefunction is:
Ψr (x) = Ceikx
SCHROEDINGER’S EQUATION 39

In the forbidden region (i.e. within the wall: 0 < x < d) Schroedinger’s equation,
Eq.2.40
~2
− ∇2 Ψ + VΨ = EΨ
2m
has solutions of the form:
Ψo (x) = De−κx + Eeκx

with κ = 2m(V − E)/~. At the interface with the wall (at x = 0 and x = d) the
continuity of the wavefunction and its first derivative require that:

A+B = D+E
ik(A − B) = κ(−D + E)
Ceikd = De−κd + Eeκd
ikCeikd = κ(−De−κd + Eeκd )

The transmission probability is therefore:

C 4k2 κ2
T = | |2 =
A 4k2 κ2 + (k2 + κ2 ) sinh2 κd
4E(V − E) 16E −d/λt
= → e when : E  V and κd  1
4E(V − E) + V sinh κd
2 2 V
Where λt = 1/2κ is the typical tunneling length. For an energy barrier V = 1eV: λt '
1Å which is the typical size of an atom. The transmission probability thus decreases
exponentially with the size d of the gap which can be but a few Å wide.

This sensitivity of QM tunneling to atomic size dimension has led to the develop-
ment in 1981 of the Scanning Tunneling Microscope (STM) by Gerd Binning and Hein-
rich Rohrer. They used it to image surfaces with atomic resolution and were awarded
for it the Nobel prize in Physics in 1986. In a STM a conducting tip is scanned above
a conducting surface, see Fig.2.13(b). The electrons tunnelling between the tip and the
surface (across an insulating gap, usually vacuum) generate a current which amplitude
is exponentially sensitive to the size of the gap and the voltage across it. A piezo-electric
tube displaces the tip above the surface while maintaining a constant current, namely a
fixed distance between the tip and the surface via an appropriate feedback loop. Mea-
suring the amplitude of the feedback signal while scanning the tip generates an image of
the surface with atomic resolution, Fig.2.13(c). This invention revolutionized the study
of matter at the nanoscale. It also led to the development of many other scanning probe
microscopes, such as the Atomic Force Microscope (AFM) which uses the deflection
of a small cantilever, like the needle in a gramophone, to map a surface with atomic
resolution.
40 QUANTUM MECHANICS

F. 2.13. (a) An electron wave of energy E is incident on a barrier of energy V. Some
of the wave is reflected and some tunnels through the barrier to be transmitted.
(b) The principle of the Scanning Tunneling Microscope (STM). A conducting tip
mounted on a piezo-electric drive is scanned across a surface. The current tunneling
between the surface and the tip is kept constant via an appropriate feedback loop that
maintains a constant distance between tip and surface. The feedback signal is used
to generate an image of the surface with atomic resolution. (c) Image of the surface
of a gold crystal taken in vacuum at 77◦ K. To minimize their surface energy the gold
atoms deviate from the bulk crystal structure and arrange in columns several atoms
wide with regularly-spaced pits between them.

The STM also made possible the study of QM problems which for decades were
assumed to be thought experiments, such as the problem of finding the eigen-energies
of a particle confined in a 2D box. With the help of an STM such a situation could be
constructed and studied: consider for example the case of a free particle (an electron
on the surface of a conductor (Cu) with effective mass m∗ = 0.38me ) confined in a two
dimensional ring of radius a = 71.3Å formed by 48 atoms of iron (Fe) deposited and
manipulated with an STM to form a corral, Fig.2.14. The Schroedinger equation for
such a particle is:
~2
− ∗ ∇2 Ψ = EΨ
2m
for r < a and Ψ = 0 for r > a. This is a typical Helmholtz equation (see Ap-
pendix) which solutions are: |n, l >≡ Ψn,l (r, φ) = Jl (κn,l r)eilφ where the eigen-energies
En,l = ~2 κn,l
2
/2m∗ are determined by the boundary condition Jl (κn,l a) = 0. Fig.2.14
presents the results from the group of Don Eigler at IBM who observed with an STM
the wavefunction of an electron trapped in such a corral. The measured tunneling cur-
rent is proportional to the probability of finding the electron under the scannig tip, which
THE CORRESPONDANCE PRINCIPLE 41

F. 2.14. Electrons on the surface of copper (Cu) with effective mass m = 0.38me
are confined in a corral of radius a = 71.3Å formed by 48 atoms of iron (Fe), see
Crommie et al., Science 262, 218 (1993). The probability of finding an electron at a
given place within the corral is proportional to the intensity of the current tunneling
into the scanning conducting tip of an STM. The observations are remarquably fit
by a linear superposition of just three eigen-functions, solutions of Schroedinger’s
equation for this problem.

is remarquably fit by a linear combination of just three eigenstates: |5, 0 >, |4, 2 > and
|2, 7 > with similar energies (since: κ5,0 a = 14.931, κ4,2 a = 14.796 and κ2,7 a = 14.821).
In this corral the surface electrons with the highest energy are thus occupying the top of
the conduction band from which they hop into the tip of the STM.

2.6 The correspondance principle


One of the main self-consistency checks of the interpretation of QM is that its equations
yield the classical results when the system is large and its wavefunction is localized so
that one can replace the operators with their mean values. The correspondance principle
therefore states that the time evolution of the mean of an observable should satisfy the
equation of classical mechanics.
If Ô is an operator that does not depend explicitely on time, the evolution of its mean
value is given by:
d d d d
< Ô > = < Ψ|Ô|Ψ >= ( < Ψ|)Ô|Ψ > + < Ψ|Ô |Ψ >
dt dt dt dt
1 1
= − < Ψ|Ĥ Ô|Ψ > + < Ψ|ÔĤ|Ψ >
i~ i~
where we used Eq.2.16 with Ĥ as the Hamiltonian operator. Hence the time evolution
of the mean value of an operator obeys:
d
i~ < Ô >=< Ψ|[Ô, Ĥ]|Ψ >=< [Ô, Ĥ] > (2.45)
dt
Consider for example the problem of a particle moving in a potential V(x). Its Hamilto-
nian is Ĥ = p̂2 /2m + V(x). The mean velocity of the particle is
42 QUANTUM MECHANICS

d < [ x̂, Ĥ] > < [ x̂, p̂2 ] >


< v >= < x̂ >= = =< p > /m
dt i~ 2i~m

which is indeed the classical result ~p = m~v. Similarly the evolution of the velocity
operator yields:

d < [ p̂, Ĥ] > < [ p̂, V(x)] > ∂V


m < v̂ >= = =−< >=< F >
dt i~ i~ ∂x
Which is Newton’s law: mdv/dt = −∂ x V = F. For a particle in a magnetic field the
Hamiltonian is:
( p̂ − (q/c)Â)2
Ĥ = (2.46)
2m
~ is constant along the z−axis, the vector potential is: A =
If the magnetic field B
(−yB, xB, 0)/2. The mean velocity v x is:

d < [ x̂, Ĥ] > < [ x̂, ( p̂2 − (q/c)(Â p̂ + p̂Â))] >
< x̂ > = =
dt i~ 2i~m
< [ x̂, p̂2x ] > −2(q/c)Â x [ x̂, p̂ x ] >
=
2i~m
< p x − (q/c)A x >
=
m

~ Notice that the canonical momentum


Which is the classical result: m~v = ~p − (q/c)A.
~p is different from the kinetic momentum ~π = m~v. We will leave it as an exercise for the
reader to derive the Lorentz force Law: d~π/dt = (q/c)~v × B, from the time evolution of
the velocity operator: v̂ = ( p̂ − (q/c)Â)/m.

2.6.1 Gauge invariance and the Aharonov-Bohm effect


The Hamiltonian of a charged particle in a magnetic field, Eq.2.46, appears to violate
the principle of gauge invariance. This principle implies that Nature is unaffected by
~ that leave the real electromagnetic fields (E,
a change in the potential fields (Φ, A) ~ B)
~
unchanged. A ~ is defined up to the addition of the gradient of an arbitrary function f :
~→A
A ~+∇ ~ f which does not affect the real field B
~=∇~ × A,
~ see Eq.??. But such a gauge
transformation does modify the Hamiltonian:

~ f )2
( p̂ − (q/c)Â − (q/c)∇
Ĥ 0 =
2m
However this transformation can be compensated by a similar gauge transformation on
the wavefunction:
Ψ0 (x) = eiq f /~c Ψ(x)

In that case:
THE CORRESPONDANCE PRINCIPLE 43

F. 2.15. The double-slit or Young’s experiment performed in presence of a magnetic


field confined in a region isolated from the interfering electrons, see Tonomura et al.,
Phys.Rev.Lett. 56, 792 (1986). An electron passes through two slits and interferes on
a far-field screen. On its way it is incident on a microscopic toroidal ring of ferro-
magnetic material (with azymuthal magnetization) enclosed by a superconducting
Niobium (Nb) shield to prevent any flux leakage. Notice the phase shift δφ = π in
the interference pattern of the electrons that have passed in the center of the ring as
compared with those that have passed outside (bright stripes appear instead of dark
ones). This phase shift is due to the flux ∆Φ = hc/2e (= 2 10−7 Gauss cm2 = 2 10−15
Tesla m2 ) enclosed by the paths surrounding the ring (the value of this flux quantum
is itself a result of the Aharonov-Bohm effect on the superconducting current with
charge 2e flowing and interfering constructively in the Niobium shield ).

1 ~ f )2 eiq f /~c Ψ(x)


Ĥ 0 Ψ0 (x) = ( p̂ − (q/c)Â − (q/c)∇
2m
1 ~ f )eiq f /~c ( p̂ − (q/c)Â)Ψ(x)
= ( p̂ − (q/c)Â − (q/c)∇
2m
1
= ( p̂ − (q/c)Â)2 Ψ(x) = ĤΨ(x)
2m

In general the extra phase q f /~c has no effect on the probability of detecting a parti-
cle at a given position, since it depends on the absolute value of Ψ(x). There is however
a situation first pointed out by Aharonov and Bohm in 1959 where this gauge transfor-
mation leads to detectable effects. This happens when the probability of detecting an
electron in presence of a magnetic field results from the interference between different
paths the electron can follow, as in the diffraction experiment shown in Fig.2.11 and
Fig.2.15. Following our earlier derivation, the wave-function of the particle on a screen
far away from the diffraction slits is:
44 QUANTUM MECHANICS

~ r ~ r
|Ψ > = |I > +|II >= eikl1 +i(q/~c) |O > +eikl2 +(q/~c)
R R
1
A·d~ 2
A·d~
|O >
~ r
H
= (1 + e ik(l2 −l1 )+i(q/~c) A·d~
)|O >
= (1 + e ik(l2 −l1 )+iδφ
)|O >

where the phase difference δφ is related to the magnetic flux Φ enclosed by the interfer-
ing paths:
I I I
~ ~ ~ ~
δφ = (q/~c) A · d~r = (q/~c) ∇ × A · dS = (q/~c) B ~ · dS~ = qΦ/~c

Hence the enclosed magnetic flux can change the pattern of interference of the elec-
tron wavefunction, even if the electron does not (and cannot) pass through the region
containing the magnetic field and no force is exerted on it! As shown in Fig.2.15, this
effect has been clearly evidenced using superconducting materials to prevent leakage
of the magnetic field. Notice that the interference pattern is periodic in Φ with a pe-
riod ∆Φ = hc/q a property which is used to measure the magnetic field in extremely
sensitive Superconducting Quantum Interference Devices (SQUID).

2.7 Dirac’s equation: antiparticles and spin


We have seen the power of the QM formalism in predicting effects that are not only
counter-intuitive but at odds with classical mechanics: molecules that can exist as su-
perpositions of enantiomers, particles that behave like waves and diffract and electrons
that sense the magnetic field in regions they cannot accede. The most striking demon-
stration of the power and validity of QM has been the prediction of spin (a property
of electrons which has no classical equivalence) and the existence of anti-particles that
came out of the successful attempt by P.A.M. Dirac to unify QM with relativity. Since
spin is crucial in understanding the properties of atoms and molecules, we shall see how
it arises from Dirac’s equation.
The Hamiltonian in Shroedinger’s equation describes a non-relativistic particle for
which the energy Ep= p2 /2m  mc2 . For a relativistic particle, Einstein showed (see
Apendix) that E = p2 c2 + m2 c4 . However writing the relativistic analogue of Eq.2.40:

∂Ψ √ 2 2 2
i~ = −~ c ∇ + m2 c4 Ψ
∂t
led to insurmountable problems related to the interpretation of the square-root of the
momentum operator. Similarly writing H 2 = p2 c2 + m2 c4 to derive the so-called Klein-
Gordon equation:
∂2 Ψ
= [∇2 − (mc/~)2 ]Ψ (2.47)
∂(ct)2
yields an equation which is second order in time (Schroedinger’s equation is first order)
and does not allow for the interpretation of |Ψ|2 as a probability density. Dirac searched
for a relativistic formulation that like Eq.2.16 would be first order in time, yet would
treat time and space in a relativistic invariant way. He noticed that he could obtain
DIRAC’S EQUATION: ANTIPARTICLES AND SPIN 45

such an equation if he rewrote the relativistic invariant Klein-Gordon equation for a


wavefunction |Ψ(x, t) > that is a vector rather than a scalar:

∂2 X X
[∇2 − ]|Ψ(x, t) >= ( Ak ∂ xk + i(β/c)∂t )( Ak ∂ xk + i(β/c)∂t )|Ψ(x, t) >
∂(ct)2

Where Ak , β are hermitian matrices that satisfy the following relations: A2k = β2 = 1 and
Ak A j + A j Ak = Ak β + βAk = 0 (for k , j). To satisfy these 10 equations the matrices
have to be at least 4 × 4 matrices and the eigenvectors are 4-vectors which satisfy:
X
( Ak ∂ xk + i(β/c)∂t )|Ψ >= mc/~|Ψ >

Further application of the operator ( Ak ∂ xk + i(β/c)∂t ) yields the Klein-Gordon equa-


P
tion, Eq.2.47 which is compatible with relativity. Multiplying the preceding equation on
the left by the matrix ~cβ yields the Dirac equation:
X
i~∂t |Ψ(x, t) >= (c αk pk + mc2 β)|Ψ(x, t) > (2.48)
k

which is formally equivalent to Eq.2.16 and where: αk = −iβAk and pk = −i~∂ xk . As


mentioned by Dirac (see ”Principles of Quantum Mechanics”, p.253ff) the matrices αk , β
describe some new internal degree of freedom of the electron which commutes with the
space and time operators. They can be written as:

0 σk
! !
I 0
αk = β= (2.49)
σk 0 0 −I

where I is the 2 × 2 unit matrix and σk are the 2 × 2 Pauli matrices:


! ! !
01 0 −i 10
σx = σy = σz = (2.50)
10 i 0 0 −1

In Dirac’s equation as in Schroedinger’s equation the wavefunction has a natural


interpretation as a probability distribution. Multiplying Eq.2.48 by < Ψ(x, t)| and its
complex conjugate
X
−i~(< Ψ(x, t)|∂t ) = [< Ψ(x, t)|(c αk pk + mc2 β)]
k

by |Ψ(x, t) > and subtracting the two yields:


X X
i~∂t < Ψ|Ψ >= c pk < Ψ|αk |Ψ >= −i~c ∂ xk < Ψ|αk |Ψ >
k k

which expresses a conservation law for the probability density P(x, t) =< Ψ(x, t)|Ψ(x, t) >
with current density Jk = c < Ψ|αk |Ψ >:
∂P ~ ~
+∇·J =0
∂t
46 QUANTUM MECHANICS

Let us now investigate Eq.2.48. Let us first consider a free electron with zero mo-
mentum. The equation is then:

i~∂t |Ψ >= mc2 β|Ψ >

which solutions are |Ψ0 >1,2 = e−i(mc /~)t (φ̃0 , 0, 0) and |Ψ0 >3,4 = ei(mc /~)t (0, 0, χ̃0 ) where
2 2

φ̃0 and χ̃0 are 2D vectors. The first solution describes an electron with positive energy
Ee = mc2 and the second a hole with energy Eh = −mc2 on which we will have more to
say shortly.

In the presence of an electro-magnetic field Dirac’s equation is generalized just as


Schroedinger’s equation was generalized: by replacing the momentum pk by πk = pk −
(q/c)Ak and adding the potential energy qΦ
X
i~∂t |Ψ(x, t) >= [c αk (pk − (q/c)Ak ) + qΦ + mc2 β]|Ψ(x, t) > (2.51)
k

Let us look for a solution in terms of the two-component vector: |Ψ >= (φ̃, χ̃). Dirac’s
equation can then be recast as:

φ̃ χ̃ φ̃ φ̃
! ! ! !
= c k σk πk + qΦ + mc2
P
i~∂t (2.52)
χ̃ φ̃ χ̃ −χ̃

To obtain the low energy limit of Dirac’s equation we look for a solution which is a
perturbation of the previous solution for positive energy Ee = mc2 at zero momentum:

φ̃ −i(mc2 /~)t φ(x, t)


! !
|Ψ(x, t) >= =e (2.53)
χ̃ χ(x, t)

Where φ and χ are slowly time-varying fields (i.e. slower than ω0 = mc2 /~), solutions
of the coupled equations:

φ χ φ
! ! ! !
0
= c k σk πk + qΦ − 2mc2
P
i~∂t (2.54)
χ φ χ χ

In the low energy limit (when mc2 is larger than any other energy) the field χ is slaved
to φ (i.e. all terms in χ are negligible with respect to the 2mc2 χ term) and thus:
σk πk φ
P
χ= k
2mc
which yields the following equation for φ:

( σk πk )2
P
i~∂t φ = [ k + qΦ]φ
2m
Using the following property of Pauli matrices:
DIRAC’S EQUATION: ANTIPARTICLES AND SPIN 47

~ σ · B)
σ · A)(~
(~ ~ =A
~·B
~ + i~ ~×B
σ·A ~

one obtains:
σ · ~π)2 = π2 + i~
(~ σ · ~π × ~π
= π + i~
2
σ · (−i~∇ ~ − (q/c)A) ~ − (q/c)A)
~ × (−i~∇ ~ (2.55)
= π − (~q/c)~
2 ~ ×A
σ·∇ ~ = π2 − (~q/c)~ ~
σ·B
From which one obtains the low energy version of Dirac’s equation:
( p̂ − (q/c)Â)2 ~ + qΦ] φ
i~∂t φ = [ σ·B
− (~q/2mc)~ (2.56)
2m
which ressembles Schroedinger’s equation except for the additional term on the right
hand side which is the energy of a dipole m ~ = (~q/2mc)~ ~ Hence
σ in a magnetic field B.
Dirac’s equation implies that the electron possesses an intrinsic magnetic moment, a
fact that was known from the spectrum of atoms in a magnetic field (see below). Exem-
plifying yet again the ”unreasonable effectiveness of Mathematics” it is amazing that
this intrinsic properties of the electron is a mathematical consequence of the unification
of QM and relativity!
Moreover Dirac’s was able to relate this magnetic moment to the intrinsic angular
momentum of the particle. Consider the time evolution in abscence of magnetic field of
the angular momentum ~L = ~r × ~p. As we have seen the time evolution of an observable
obeys:
X
i~∂t < L x > = < [L x , H] >=< [ypz − zpy , (c αk pk + mc2 β)]
k
= c < αy [y, py ]pz − αz [z, pz ]py >= i~c < αy pz − αz py >, 0
Thus in Dirac’s equation < L x > is not a constant of motion, in contrast with the
Schroedinger’s equation for a free particle where it is. Now consider the evolution of σ̂ x
(the 4 × 4 matrix which diagonal elements are the Pauli σ x matrices):
X
i~∂t < σ̂ x > = < [σ̂ x , H] >=< [σ̂ x , (c αk pk + mc2 β)]
k
= c < [σ̂ x , αy ]py + [σ̂ x , αz ]pz >= 2ic < αz py − αy pz >
Where we used the following property of Pauli’s matrices: [σ x , σy ] = 2iσz , [σ x , σz ] =
−2iσy . Therefore L x + ~σ̂ x /2 commutes with the Hamiltonian, i.e. it is a constant of
the motion. Dirac interpreted this result to mean that the particle has a spin angular
momentum S~ = ~~ σ/2 which must be added to the orbital angular momentum ~L to get
the total angular momentum, J~ = ~L + S~ which is a constant of the motion.
According to Eq.2.56, the magnetic dipole moment of the electron is related to its
spin by:
m σ = (qg/2mc)S~
~ = (~q/2mc)~ (2.57)
Where g = 2 is known as the gyromagnetic ratio. This is in contrast with the mag-
~ = (q/2mc)~L (see Eq.??) for which
netic moment arising from the angular momentum m
48 QUANTUM MECHANICS

g = 1. The prediction of the gyromagnetic ratio of the electron was an other success
of Dirac’s equation. In fact the value of g has become the most stringent test of quan-
tum electro-dynamics (QED), the theory that generalized Dirac’s approach and unified
Electromagnetism and Quantum Mechanics. It is the most precisely known constant of
nature: its value has been tested to 12 decimal places g = 2.0023193043617(!!), making
QM and QED the most precisley tested theories ever proposed.

Let us now return to the hole solution with energy Eh = −mc2 . We have already seen
such solutions when studying the band theory of solids. In fact the original interpretation
of Dirac was very similar. He assumed that all negative energy levels were occupied by a
”sea” of electrons. Out of this sea, electrons could be excited to an energy level Ee = mc2
leaving behind a hole (or positron) of energy Eh = mc2 . This process, electron-positron
creation (and its reverse, annihilation) is routinely observed in particle accelerators once
an energy E > 2mc2 ∼ 1MeV is achieved. The dynamic behaviour of the positron is
similar to that of an electron with opposite charge. To derive the positron wave-equation
at low energies, we shall follow our previous study of the electron in a similar regime.
We look for a low energy solution:

φ̃ i(mc2 /~)t φ
! !
|Ψ >= =e (2.58)
χ̃ χ

Which leads to the following equation for the positron-wavefunction at low energies:

σk πk )2
P
(
i~∂t χ = [− k
+ qΦ]χ
2m
Since the Pauli matrices are hermitian, the equation for the complex conjugate of χ is:

σk (i~∂ xk − (q/c)Ak )2
P
( k
i~∂t χ = [

− qΦ]χ∗ (2.59)
2m
( k σk (−i~∂ xk + (q/c)Ak )2
P
=[ − qΦ]χ∗ (2.60)
2m
This is the low energy limit of Dirac’s equation for a particle of mass m and charge −q.
As seen above in the band-gap theory of conduction in solids, the hole solution describes
the behaviour of a particle of the same mass as the electron and of opposite charge. This
prediction of Dirac was vindicated by the discovery of the positron three years later by
Carl Anderson, for which they both shared the 1933 Nobel prize in Physics.
To summarize, the unification in 1929 of relativity and QM by Dirac led to many
unexpected predictions of which we have seen three:
1) the existence of anti-particles (e.g. the positron) of mass identical to the particle
and with opposite charge. These are created when an energy 2mc2 is available to create
for example an electron-positron pair (raise the electron from the top of the ”sea” at
Eh = −mc2 to the electron rest energy Ee = mc2 ).
DIRAC’S EQUATION: ANTIPARTICLES AND SPIN 49

2) The association of an intrinsic angular momentum variable, the spin, S = ~σ/2


for both the electron and its anti-particle.
3) The existence of a magnetic dipole moment linked to the spin of the particle in
similitude to the classical relation between magnetic moment and angular momentum
but larger by a factor g = 2.

2.7.1 Angular momentum and spin


The Hamiltonian for a free particle is:

~2 2 ~2 1 ∂2 1 ∂ ∂ 1 ∂2
H=− ∇ =− { r+ 2 sin θ + } (2.61)
2m 2m r ∂r 2 r sin θ ∂θ ∂θ r2 sin θ ∂φ2
2

~2 ∂2 2 ∂ ~2 2
=− { 2 + }− L (2.62)
2m ∂r r ∂r 2mr2
The operator : L̂2 = −~2 L2 is known as the angular momentum operator, since its
R L̂ /2mr is the classical angular momentum kinetic
2 2
contribution to the Hamiltonian, i.e.
energy: E L = L2 /2I (whereI = d3 rρ(r)r2 is the moment of inertia). As shown in
the Appendix on the Helmholtz equation, the eigenfunctions of L2 are the spherical
harmonics Ylm [θ, φ) with eigen-value −l(l + 1), hence:
L̂2 |l, m > = ~2 l(l + 1)|l, m >
s
(2l + 1)(l − m)! m
with :|l, m > = Ylm [θ, φ) = Pl (cos θ)eimφ (2.63)
4π(l + m)!

The linear angular momentum operator L̂ = r̂ × p̂ can be written as:


∂ ∂ ∂ ∂
!
L̂ x = −i~ y − z = i~ sin φ + cot θ cos φ (2.64)
∂z ∂y ∂θ ∂φ
∂ ∂ ∂ ∂
!
L̂y = −i~ z − x = i~ cos φ − cot θ sin φ (2.65)
∂x ∂z ∂θ ∂φ
∂ ∂ ∂
!
L̂z = −i~ x − y = −i~ (2.66)
∂y ∂x ∂φ

The eigenfunctions of L̂2 are also eigenfunctions of L̂z with eigenvalue ~m:
L̂z |l, m >= ~m|l, m >
Notice that L̂z does not commute with L̂ x or L̂y . In fact these operators satisfy the fol-
lowing commutation relations:
[L̂ x , L̂y ] = i~L̂z
[L̂ x , L̂z ] = −i~L̂y
[L̂y , L̂z ] = i~L̂ x
or : [L̂i , L̂ j ] = i~i jk L̂k (2.67)
50 QUANTUM MECHANICS

where i, j, k stands for x, y or z and i jk = 1 for an even permutation of the indices and
−1 for an odd permutation. As we shall see below, it is usefull to define the raising and
lowering operators:

∂ ∂
!
L̂± = L̂ x ± L̂y = ~e ±iφ
± + i cot θ
∂θ ∂φ

which satisfy the following commutation rules:

[L̂+ , L̂z ] = −~L̂+


[L̂− , L̂z ] = ~L̂−
[L̂+ , L̂− ] = 2~L̂z

Notice that L̂2 = L̂2x + L̂y2 + L̂z2 = L̂z2 + L̂+ L̂− − ~Lz = L̂z2 + L̂− L̂+ + ~Lz commutes with L̂z
but also with L̂ x , L̂y , L̂± .

[L̂2 , L̂ x ] = [L̂y2 , L̂ x ] + [L̂z2 , L̂ x ] = L̂y [L̂y , L̂ x ] + [L̂y , L̂ x ]L̂y + L̂z [L̂z , L̂ x ] + [L̂z , L̂ x ]L̂z
= −i~L̂y L̂z − i~L̂z L̂y + i~L̂z L̂y + i~L̂y L̂z = 0

These operators therefore do not mix eigenstates with different values of l but since they
do not commute with L̂z they do mix eigenstates with different values of m. Using their
commutation rules one can show that L̂± |l, m >∼ |l, m ± 1 >. Consider for example the
commutation relations for L̂+ :

−~L̂+ |l, m > = [L̂+ , L̂z ]|l, m >= (L̂+ L̂z − L̂z L̂+ )|l, m >
= ~mL̂+ |l, m > −L̂z L̂+ |l, m >
namely : L̂z (L̂+ |l, m >) = ~(m + 1)(L̂+ |l, m >)

Hence:
L̂± |l, m >= a±lm |l, m ± 1 >= ~ (l + 1) − m(m ± 1)|l, m ± 1 >
p
(2.68)
Where the coefficient a+lm is determined by the normalization condition:

|a+lm |2 = |a+lm |2 < l, m + 1|l, m + 1 >=< l, m|L̂+∗ L̂+ |l, m >=< l, m|L̂− L̂+ |l, m >
= < l, m|L̂2 − L̂z2 − ~Lz |l, m >= ~2 (l(l + 1) − m(m + 1))

Notice that the action of L+ on a state |l, m > is to generate the higher m-state
|l, m + 1 > whereas the action of L− on |l, m > generates the lower m-state l, m − 1 >,
which justifies their name as raising and lowering operators.

Since the spin is also an angular momentum operator the same commutation rela-
tions exist between Ŝ 2 , Ŝ x , Ŝ y , Ŝ z and Ŝ ± . For spin one-half particles, the eigenstates
DIRAC’S EQUATION: ANTIPARTICLES AND SPIN 51

|s, sz > of Ŝ 2 = ~2 σ2 /4 and Ŝ z = ~σz /2 with eigenvalues s(s + 1)~2 = 3~2 /4 and
~sz = (±1/2)~ are:
! !
1 0
|1/2, 1/2 >= |1/2, −1/2 >=
0 1

We have seen that one of the results of Dirac’s equation is that spin and angular mo-
mentum operators L̂z , Ŝ z do not commute with the Hamiltonian and are therefore not
good (i.e. conserved) quantum numbers. Rather the total angular momentum operator:
Jˆz = L̂z + Ŝ z with eigenvalue k = m + sz is conserved, together with Jˆ2 , L̂2 and Ŝ 2 . Thus
the eigenstate of the total angular momentum is: | j, l, s, k >. Since m = k ∓ 1/2 and the
eigenstates of the angular momentum are Ylm we can write:
! !
1 0
| j, l, s, k >= c1 |l, k−1/2 > |s, 1/2 > +c2 |l, k+1/2 > |s, −1/2 >= c1 Yl,k−1/2 + c2 Yl,k+1/2
0 1

For this state to be an eigenstate of Jˆ2 = L̂2 + Ŝ 2 + 2L̂ · Ŝ it has to satisfy:

3 L̂σ
Jˆ2 | j, l, s, k >= ~2 [l(l + 1) + + ]| j, l, s, k >= ~2 γ| j, l, s, k >
4 ~
Using the relation:
!
Lz L−
L̂σ = L̂ x σ x + L̂y σy + L̂z σz ==
L+ −Lz

and the previously derived equation for the raisng and lowering operators, Eq.2.68, one
obtains the following eigenvalue equation for γ:

l(l + 1) + 1/4 + k − γ l(l + 1) − (k − 1/2)(k + 1/2)


= 0


l(l + 1) − (k − 1/2)(k + 1/2) l(l + 1) + 1/4 − k − γ

which yields the quadratic equation:

[(l + 1/2)2 − γ]2 − (l + 1/2)2 = 0

which solution is γ = j( j+1) with j = l±1/2. Hence the eigenvalues of the total angular
momentum operator Jˆ2 are ~2 j( j + 1) with eigenfuctions:

√ l + k Yl,k−1/2
!
1
| j, l, s, k > = |l + 1/2, l, 1/2, k >= √
2l + 1 l − k + 1 Yl,k+1/2

k + 1 Yl,k−1/2
!
1 l−
| j, l, s, k > = |l − 1/2, l, 1/2, k >= √ √
2l + 1 − l + k Yl,k+1/2
52 QUANTUM MECHANICS

2.8 The Hydrogen atom and electronic orbitals


We have seen above how the formalism of QM could be applied to simple cases (two
state systems, free particle, particle in a box, etc.). However its first and major success
was in solving for the energy level of the hydrogen atom and in explaining the properties
of the elements as classified in Mendeleev’s periodic table. Schroedinger’s equation for
a hydrogen-like atom is:

~2 2 Ze2
− ∇ Ψ(~x, t) − Ψ(~x, t) = En Ψ(~x, t) (2.69)
2me r
where the −Ze2 /r term stands for the Coulomb interaction between the orbiting electron
(of mass me ) and a nucleus containing Z protons (in Hydrogen Z = 1). In spherical
coordinate the differential operator can be written as:

∂2 2 ∂ 1 1 ∂ ∂ 1 ∂2
∇2 = + + [ (sin θ ) + ]
∂r2 r ∂r r2 sin θ ∂θ ∂θ sin2 θ ∂φ2
As for the Helmholtz problem (see Appendix) we look for a eigen-function:

Ψnlm (r, θ, φ) = Rnl (r)Ylm (θ, φ)

where the spherical harmonics Ylm (θ, φ) are eigen-functions of the angular operator (see
above):

1 ∂ ∂ 1 ∂2
L2 Ylm (θ, φ) = [ (sin θ ) + ]Ylm (θ, φ) = −l(l + 1)Ylm (θ, φ)
sin θ ∂θ ∂θ sin2 θ ∂φ2
The equation for the radial function Rnl (r) then becomes:

~ 2 ∂2 2 ∂ ~2 l(l + 1) Ze2
[− ( 2+ )+ − ]Rnl = En Rnl
2me ∂r r ∂r 2me r2 r

The first term on the left is the radial part of the kinetic energy, the second term is the
centrifugal potential (the angular part of the kinetic energy) and the third term is the
Coulomb repulsion. Writing Rnl (r) = unl (r)/r yields the following equation for u(r):

d2 unl 2me Ze2 l(l + 1) 2me En


2
+ ( 2
− 2
)unl = − 2 unl
dr ~ r r ~

Defining a = 2Zme e2 /~2 ≡ 2/r0 and λ2n = −2me En /~2 one can recast the previous
equation as:
a l(l + 1)
u00nl + ( − )unl = λ2n unl
r r2
In the limit r → ∞, the terms in the bracket on the left hand side are negligible and
the solution in that limit is unl (r) ∼ exp(−λn r). Let us therefore look for a solution:
unl (r) = µnl (r)exp(−λn r). The equation that µnl (r) obeys is:
THE HYDROGEN ATOM AND ELECTRONIC ORBITALS 53

a l(l + 1)
µ00nl − 2λn µ0nl + ( − )µnl = 0
r r2

Notice that close to r = 0 the solution behaves as µnl (r) ∼ rl+1 . Let us therefore look
for a polynomial solution: µnl (r) = nj=1 c j r j . The coefficient of the term of order j − 1
P
reads:
j( j + 1)c j+1 − 2λn jc j + ac j − l(l + 1)c j+1 = 0

which yields:
2λn j − a
c j+1 = cj
j( j + 1) − l(l + 1)
To be defined for all values of j the series {c j } must start at j = l + 1 (the small r limit
of µnl (r)) and to be finite it must end at a value l < j = n for which: a = 2nλn , i.e. for a
value of the eigen-energy, En :

~2 λ2n ~2 1
En = − =−
2me 2me r02 n2

which is the result obtained by Bohr, see Eq.2.8. Notice that while the radial eigen-
functions depend on the two quantum numbers n, l, the eigen-energies depend only on
n. This is a peculiarity of central potentials (such as the Coulomb potential between
proton and electron). In non-hydrogen atoms that degeneracy is raised due to the extra
Coulomb repulsion between the electrons. It is customary to label the various orbitals
(the hydrogen atom’s eigensolutions) by their shell number n = 1, 2, 3, .... and their
main angular momentum number l, such that l = 0 is called the s-orbital; l = 1 the
p−orbital; l = 2 the d−orbital; l = 3 the f −orbital; l = 4 the g−orbital ... etc. (in
alphabetical order skipping j). With this nomenclature we can proceed to write down
the radial eigen-functions Rnl (r) = µnl (r) exp(−r/nr0 )/r for the first orbitals:

1s orbital : R10 (r) = 2r0−3/2 e−r/r0


r −r/2r0
2s orbital : R20 (r) = (2r0 )−3/2 (2 − )e
r0
1 r
2p orbital : R21 (r) = √ (2r0 )−3/2 e−r/2r0
3 r0

In abscence of other terms in the Hamiltonian, the orbitals of the electron in the
Hydrogen atom are degenerate, i.e. for different values of the spin and angular quantum
numbers (l, m) the energy of the electron is the same. Since there are 2 possible spin
states for every eigenfunction Ψnlm and 2l + 1 states with different values of m (and
same value of l: −l ≤ m ≤ l) and since 0 ≤ l ≤ n − 1, the degeneracy of a state
with shell number n is: 2n2 . As we shall see next this degeneracy is partially raised by
various relativistic effects (for example by the interaction between the electron spin and
the magnetic field the electron generates when orbiting the nucleus, so-called spin-orbit
coupling), by the action of external fields or by the presence of other electrons.
54 QUANTUM MECHANICS

2.8.1 Spin-orbit coupling


We have seen that Schroedinger’s equation, Eq.2.56 is only an approximation to the
more correct relativistically invariant Dirac’s equation, Eq.2.51. Although Dirac’s equa-
tion for the Hydrogen atom can be solved exactly (see Appendix), it is instructive to
investigate the various terms that were neglected when deriving Eq.2.56.
One of these terms is the so called spin-orbit coupling, which is due to the interaction
between the electron spin and the magnetic field experienced by the orbiting electron.
~ = −r̂∂r Φ =
In the rest frame of the nucleus there exists only an electric field E~ = −∇Φ
e~r/r . However in the frame of the electron orbiting at velocity ~v = ~p/me , the nucleus is
3

moving at velocity −~v and is generating both electric and magnetic fields. If the electron
frame of reference was inertial (i.e. if the electron was moving in a straight line) then
the magnetic field experienced by the electron would be

~ = − ~p × ~r e = e ~L
~ = −~v × E/c
B
me c r3 me cr3

This magnetic field which is parallel to the angular momentum ~L = ~r × ~p, interacts with
the electron spin S~ = ~~
σ/2 to contribute an energy, see Eq.2.56:

δH so =
~e ~ = e S~ · B
~ ·B
σ ~
2me c me c
Since the electron frame is not inertial the final result differs by a factor 2 from this
estimate:
e2 ~ ~
δH so = S · L ≡ g so S~ · ~L
2m2e c2 r3
As was argued by Dirac (see above), Lz and S z are not eigenvalues of the relativistic
Hamiltonian, rather the projection of the total angular momentum Jz = Lz + S z is.
Rewriting 2S~ · ~L = J 2 − L2 − S 2 one can write the Spin-orbit contribution as:
e2
δH so = (J 2 − L2 − S 2 )
4m2e c2 r3
Its contribution is best estimated in the basis which commutes with Dirac’s Hamiltonian,
i.e. in terms of eigenstates | j, l, s, k > of J 2 , L2 , S 2 and Jz :
δE so =< j, l, s, k|δH so | j, l, s, k >
To compute how that spin-orbit coupling modifies the energy spectrum of the Hydrogen
atom computed above, first notice that it couples between eigenstates with the same
value of l but different values of j, i.e. j = l ± 1/2. Hence the term: J 2 − L2 − S 2
with eigenvalue: j( j + 1) − l(l + 1) − 3/4 can have two values: ~2 l (if j = l + 1/2) and
−~2 (l + 1) (if j = l − 1/2). The term in 1/r3 in δH so can be computed directly from the
radial solutions Rnl of the Hydrogen atom with the result:
e2 e2
gnl ≡< n, l| |n, l >=
4m2e c2 r3 2m2e c2 a30 n3 l(2l + 1)(l + 1)
THE HYDROGEN ATOM AND ELECTRONIC ORBITALS 55

Thus because of the relativistic coupling between the electron’s angular momentum and
its spin the energy levels of the Hydrogen atom with same values of n and l are split in
two:
e2 ~2
(
l
δE ±so =
2m2e c2 r3 n3 l(2l + 1)(l + 1) −l − 1
0

Notice that that the spin-orbit interaction is a very small (yet measurable) perturbation
to the electronic energies of the hydrogen atom:

|En | e2
δE so ∼ ∼ 7.4 10−4 eV  |En |
me c2 r0

Where we used the fact that e2 /2r0 = |E1 | = 13.6 eV and me c2 = 0.5MeV.

2.8.2 Many electron systems


Untill now we have mainly dealt with the problem of finding the energy level of a QM
system consisting of a single electron (hoping between atoms, confined in a box or orbit-
ing the nucleus). When considering a QM system consisting of two (or more) electrons
there arises besides the question of their interaction (which as we shall see below can
be treated perturbatively) and the issue of their identity. In classical mechanics while
two particles (e.g. billiard balls) may look similar they are never identical: they can be
identified from their different trajectories. In QM not only do all electrons ”look” sim-
ilar (for example all hydrogen atoms have the same spectrum) but they cannot be told
apart: they are identical. As we shall see, this identity of QM particles has profound
consequences on the properties of matter.

Consider the Hamiltonian of a two particle system Ĥ(1, 2). Because of the iden-
tity of the particles the Hamiltonian is unchanged upon a permutation of the particles:
Ĥ(1, 2) = Ĥ(2, 1). Hence the permutation operator P̂ commutes with the Hamiltonian:

P̂Ĥ(1, 2)|1, 2 >= Ĥ(2, 1)|2, 1 >= Ĥ(1, 2)P̂|1, 2 >

and is a constant of motion. Two successive permutations brings us back to the initial
state: P̂2 = 1 which implies that the eigenvalues of the permutation operator are ±1:
|2, 1 >= P|1, 2 >= ±|1, 2 >. These eigenvalues are constants of motion, namely the
wavefunction of a many particles system is either even or odd under permutation of its
(identical) particles. Like spin this is a QM property of the particle. Particles known
as fermions (electrons, protons, neutrons, etc.) are odd under permutations: |2, 1 >=
−|1, 2 >, while particles which are even under permutation |2, 1 >= |1, 2 > are known
as bosons (e.g. photons). In 1940 Wolfgang Pauli, requiring QM to be invariant under a
change of reference frames (relativistic invariance), related this permutation property to
the spin of the particle: particles with half integer spin are fermions while particles with
integer spin are bosons.
56 QUANTUM MECHANICS

If the particles can occupy states |ψ > and |φ >, then for two bosons the joint wave-
function |1, 2 > (symmetric under exchange of the particles) is:
|ψ(1) > |φ(2) > +|ψ(2) > |φ(1) >
|1, 2 >= √ = |2, 1 >
2
For two fermions the joint wavefunction is:
|ψ(1) > |φ(2) > −|ψ(2) > |φ(1) >
|1, 2 >= √
2
which is an eigenfunction of the permutation operator P with eigenvalue -1: |2, 1 >=
P|1, 2 >= −|1, 2 >. This wavefunction implies that in contrast to bosons, two fermions
cannot share the same eigenfunctions: |ψ >, |φ >. If their spatial wavefunction is iden-
tical then their spin-state must differ and if their spin-state is identical then their spatial
wavefunction must differ. This extremely important property of fermions is known as
Pauli’s exclusion principle. It implies that the lowest energy level of an electronic sys-
tem can be occupied by at most two electrons (with opposite spins). Contrast that with
the situation for bosons where the lowest energy state can be occupied by an arbitrary
number of particles (for example photons of same energy as in Plank’s black body radi-
ation).

2.8.3 The periodic table


Pauli’s exclusion principle and the solution of Schroedinger’s equation for the hydro-
gen atom allows for a classifications of atoms according to the degree to which their
orbitals are successively filled by the orbiting electrons: the so-called Aufbau principle.
As noticed above due to electron-electron interactions the near angular degeneracy (ne-
glecting spin-orbit coupling) existing in the hydrogen atom is lifted in multi-electron
systems. For the same value of the shell number n, orbitals with smaller values of l
(more symmetric angular wavefunctions) have lower energy, see Fig.2.16. These ener-
gies cannot be calculated analytically, but numerical schemes have been developped to
estimate them with great accuracy, see perturbation method below.
Hydrogen (H) has one electron in the lowest orbital: 1s. Its spin S = 1/2 imply
that there are 2s + 1 = 2 possible states in a s−orbital. In terms of Dirac’s eigenstates
| j, l, s, k > with total angular momentum j = l + 1/2 = 1/2 that state can also be written
as 2S +1 L J =2 S 1/2 , a notation known as spectroscopic notation.
Helium (He) has two electrons occupying the lowest orbital (written as (1s)2 ) with
anti-parallel spin, i.e with total spin S = 0. The spectroscopic notation of that state is
2S +1
L J =1 S 0 .
Lithium (Li) has three electrons, two occupying with opposite spin the lowest or-
bital (1s)2 and one electron occupying the next orbital: 2s, hence the electron occupancy
is : (1s)2 (2s). Since the chemical properties of atoms are determined by their more outer
electrons (so called valence electrons which can engage in bonding interactions), the
THE HYDROGEN ATOM AND ELECTRONIC ORBITALS 57

F. 2.16. The orbital energies of many-electron atoms. The energy levels are ordered
along the abscissa according to the main shell number n, which is also indicative
of the size of the atom (the spread of the radial wavefunction increases roughly as
nr0 ). Notice that levels with different angular quantum number l (different colors)
have different energies, though they are still degenerate with respect to the azymutal
quantum number m: the number of circles indicate the level of degeneracy 2l + 1.
Notice that in abscence of an external field there is no preferred axis that would
break that azymutal symmetry. To minimize the energy electrons fill these levels
from the bottom up while satisfying Pauli’s principle: no more than two electrons
(with opposite spin) per energy level. The first level to be filled is 1s, then 2s, 2p,
etc... Notice that the more outer orbitals 4s having lower energy is filled before 3d;
5s before 4d and 6s before 4 f and 5d.

property of Lithium with one valance electron in the same s−state as the electron in
Hydrogen imply similar chemical properties, as indeed all the alkalines (the elements in
the first column of Mendeleev’s table, see Fig.2.17: Sodium (Na), Potassium (K), Ru-
bidium (Rb), Cesium (Cs), Francium (Fr)) which all have a single electron in an s−state.
In spectroscopic notation the state of the electron in the unfilled shell is: 2S +1 L J =2 S 1/2 ,
which is the same as Hydrogen. Since (as we shall see below) it is the electronic state in
the outer shell that determines the chemical characteristics of the element the chemical
properties of Lithium and hydrogen are similar.
Beryllium (Be) has four electrons, two occupying with opposite spin the lowest
orbital (1s)2 and two electrons occupying the next orbitals (2s) with opposite spin. Its
electron occupancy is thus: (1s)2 (2s)2 which is similar to Helium. Beryllium is however
more reactive than Helium since its 2p orbital has an energy only slightly higher than
58 QUANTUM MECHANICS

F. 2.17. The periodic-table of the elements. The periodic table devised by Mendeleev
and ordered according to increasing atomic number (number of electrons) and sim-
ilar chemical properties. In comparison the table is redrawn as deduced from the
filling up of the orbitals from the bottom up, taking into account Pauli’s exclusion
principle and ordering the columns according to the number of valence (outer shell)
electrons. The agreement between the two is one of the greatest success of QM.

2s which result in the electron easily hoping to the 2p orbital. Its outer shell electron oc-
cupancy is identical to that of Magnesium (Mg), Calcium (Ca), Strontium (Sr), Barium
(Ba) and Radium (Ra) which have similar chemical properties.
Boron (B) has 5 electrons, two in orbital 1s, two in orbital 2s and one in orbital 2p:
(1s)2 (2s)2 (2p). Its outer shell occupancy and chemical properties are similar to that of
Aluminium (Al), Gallium (Ga), Indium (In) and Thalium (Tl). These elements are often
used as p-type dopants (electron acceptors) in semi-conductors. The value of the total
angular momentum is j = l − s = 1/2 (due to spin-orbit coupling (see above) the state
THE CHEMICAL BOND 59

j = l + s = 3/2 has higher energy). Its outer electronic state is thus 2 P1/2 .
Carbon (C) has 6 electrons, two in orbital 1s, two in orbital 2s and two in orbital 2p.
Since orbital 2p is triply degenerate, the question arises as to where the two electrons
in 2p settle. Because of Pauli’s principle, electrons with identical wavefunction (i.e.
orbitals) have opposite spin, while electron in different orbitals may have similar spin.
Since electrons in different orbitals overlap and repel each other less, that situation is
energetically favorable giving rise to Hund’s rule: ”the state of highest spin has the
lowest energy”. Hence in Carbon the electrons in the 2p orbital have similar (parallel)
spin and occupy orbitals with different azymutal number m. The outer shell properties
of Carbon and its chemical properties are similar to Silicon (Si) and Germanium (Ge),
i.e. semiconductors. Tin (Sn) and Lead (Pb) have smaller band gaps (see above) and are
in fact conductors at room temperatures.
Nitrogen (N) has 7 electrons, two in orbital 1s, two in orbital 2s and three in orbital
2p. These outer electrons have parallel spin because of Hund’s rule and occupy the
three 2p orbitals with different azymutal number m. The outer shell occupancy and
chemical properties of Nitrogen are similar to that of Phosphorus (P), Arsenic (As),
Antimony (Sb) and Bismuth (Bi). These elements are often used as n-type dopants
(electron donors) in Si and Ge semi-conductors.
Oxygen (O) has 8 electrons, two in 1s, two in 2s and four in 2p. Two of those are
paired (they share the same orbital with opposite spin) and two have parallel spins and
share orbitals with different m. Its outer shell electron occupancy is identical to that of
Sulphur (S), Selenium (Se), Tellurium (Te) and Polonium (Po).
Fluorine (F) has nine electrons, the last five of which occupy the three 2p orbitals,
four are paired (they share the same orbital with opposite spin) and one occupy the
third orbital. The outer shell occupancy of Fluorine is similar to that of other Halogens:
Chlorine (Cl), Bromine (Br), Iodine (I) and Astatine (At).
Neon (Ne) has ten electrons which occupy all the 1s, 2s and 2p orbitals. All electrons
are paired and therefore not disponible for bonding. Neon like Helium and the other
Noble gases: Argon( Ar), Krypton (Kr), Xenon (Xe), Radon (Rn) have a similar outer
electron shell and share a very low chemical reactivity.
An interesting situation occurs when orbitals 4s and 5s fill up (for elements Ca and
Sr). The next ten elements have their electrons fill the more inner 3d or 4d shells. These
elements are known as transition metals. Similarly when orbitals 6s fill up (for Ba), the
next shell to fill up is the inner 4f shells which defines a set of 14 rare-earth elements
(Lanthanides) with similar chemical properties. Once this shell is filled the next one, 5d
is also an inner shell and the next ten elements share similar properties with transition
metals.

2.9 The chemical bond


The role that the Hydrogen atom has played in the understanding of the periodic table
of the elements has been played by the Hydrogen molecule H+2 in understanding the
chemical bond. Setting the protons at positions ±~a = ±aẑ, the problem is to solve the
Schroedinger equation for a single electron:
60 QUANTUM MECHANICS

~2 2 e2 e2 e2
HΨ(~x, t) = − ∇ Ψ(~x, t) − ( + − )Ψ(~x, t) = En Ψ(~x, t) (2.70)
2me |~x + ~a| |~x − ~a| 2a

By rewriting the differential operators in prolate spheroidal coordinates:

x = a sinh µ sin ν cos φ


y = a sinh µ sin ν sin φ
z = a cosh µ cos ν

this equation can be solved by separation of variables (i.e. looking for a solution Ψ(µ, ν, φ) =
M(µ)N(ν)Φ(φ)) as was done for the hydrogen atom (in spherical coordinates). Noticing
that : |~x ± ~a| = a(cosh µ ± cos ν, we leave it as an exercise to the reader to show that
Eq.2.70 then yields the following equations:

∂2 Φ
+ m2 Φ = 0
∂φ2
∂2 M ∂M m2
+ coth µ + (2α cosh µ − + En0 sinh2 µ − λ)M = 0
∂µ2 ∂µ sinh2 µ
∂2 N ∂N m2
+ cot ν + (− 2 + En0 sin2 ν + λ)N = 0 (2.71)
∂ν 2 ∂ν sin ν

Where α ≡ 2a/r0 (with r0 = ~2 /me e2 , the Bohr radius of Hydrogen), En0 ≡ (2me a2 /~2 )(En −
e2 /2a) and m = 0, ±1, ±2, . . . and λ are quantum numbers (playing the role that m, l
played in the hydrogen atom). The energy levels En have been solved in 1927 (only one
year after Schroedinger solved the Hydrogen atom) by O.Burrau who then looked for
the distance 2a between the atoms that would minimize the ground state energy E0 .

Since most problems in physics are not exactly soluble and in particular the com-
putation of energy spectra of complex molecules, the existence of an exact solution for
H+2 was very helpfull in the development and testing of approximate methods to com-
pute eigen-energies and eigen-functions (orbitals) for more complex QM systems. One
of these approximation is known as the variational method (for details see L.P.Pauling
and E.B.Wilson, ”Introduction to quantum mechanics”, Chap.7). The idea is pretty sim-
ple. One assumes that in some limit a given problem can be solved exactly, i.e. a set
of eigen-functions can be computed for the Hamiltonian in that limit. One then uses a
linear superposition of these solutions as a trial function and one estimates the energy of
that state for the complete Hamiltonian. This energy is then minimized with respect to
the parameters of the superposition (for example the linear coefficients). Since one does
not expect the trial function to yield the exact solution of the full problem, the estimated
minimal energy Emin is an upper bound to the true energy of the system: Emin > E0 . In
general the more complete the set of functions used as a trial, the better the approxima-
tion. However the main problem with the variational approach is that there is no way of
THE CHEMICAL BOND 61

F. 2.18. The ground state energy level of the Hydrogen molecule ion H+2 as a function
of the distance R between the two protons. The exact solution E0 (R) is shown in
dashed and compared to the results obtained from the variational method for a sym-
metric and an anti-symmetric electronic wavefunction. Only the symmetric solution
describes a bound state (energy E s smaller than the hydrogen ground state energy
E1 ).

estimating how good an approximation it is. For that one needs to resort to perturbation
expansions, which we will study later.
In the case of the hydrogen molecule ion H+2 its Hamiltonian can be written as:

H = HH (r1 ) − e2 /r2 + e2 /2a = HH (r2 ) − e2 /r1 + e2 /2a

where HH is the Hamiltonian of the Hydrogen atom. When the two protons are far apart
(a  r0 ) the solutions for the electron wavefunction are simply those computed for a
hydrogen atom centered on either proton, i.e. |Ψnlm (r1 ) > or |Ψnlm (r2 ) > with ~r1 = ~r − ~a;
~r2 = ~r + ~a. When the atoms are brought in proximity, we shall use for simplicity a trial
wavefunction |ΨT > which is a superposition of hydrogen 1s ground q states (with energy
E1 = −e2 /2r0 = −13.6eV) : |Ψ(r) >= R10 (r)Y00 (θ, φ) = exp(−r/r0 )/ πr03 :

|ΨT >= c1 |Ψ(r1 ) > +c2 |Ψ(r2 ) >

Where the coefficients c1 , c2 are parameters to be determined such as to minimize the


energy E of that state. Inserting the trial wavefunction into the full Schroedinger equa-
tion yields:
62 QUANTUM MECHANICS

< ΨT |H|ΨT >= E < ΨT |ΨT >


which defining H11 =< Ψ(r1 )|H|Ψ(r1 ) >; H12 =< Ψ(r1 )|H|Ψ(r2 ) >= H21 ; H22 =<
Ψ(r2 )|H|Ψ(r2 ) > and the overlap function: ∆ ≡< Ψ(r2 )|Ψ(r1 ) > is equivalent to:

c21 (H11 − E) + 2c1 c2 (H12 − ∆E) + c22 (H22 − E) = 0 (2.72)

We now have to determine the coefficients c1 , c2 that minimize E, i.e. for which ∂E/∂c1 =
∂E/∂c2 = 0. Taking the derivative of the previous equation with respect to c1 and c2
yields at the minimum of the energy a set of coupled homogenous equations:

(H11 − E)c1 + (H12 − ∆E)c2 = 0


(H12 − ∆E)c1 + (H22 − E)c2 = 0

Which has a solution when the determinant:



H11 − E H12 − ∆E
H − ∆E H − E = 0
12 22

Since H11 = H22 and H12 = H21 we obtain:

H11 + H12
Es =
1+∆
H11 − H12
Ea =
1−∆
The first solution corresponds to a symmetric trial wavefunction: c1 = c2 :

1
|Ψ s >= √ (|Ψ(r1 ) > +|Ψ(r2 ) >)
2 + 2∆
The second solution corresponds to an anti-symmetric wavefunction: c2 = −c1 :

1
|Ψa >= √ (|Ψ(r1 ) > −|Ψ(r2 ) >)
2 − 2∆
The various integrals can then be evaluated numerically or in this case even exactly to
yield:
r0
H11 = E1 (1 − )+J
a
r0
H12 = ∆E1 (1 − ) + K
a
(2.73)

where J known as the Coulomb integral, K the resonance integral and ∆ the overlap
integral are:
THE CHEMICAL BOND 63

Z  r
0 r0 
J= d3 rΨ(r1 )(−e2 /r2 )Ψ(r1 ) = −2E1 − + e−4a/r0 (1 +
2a 2a
Z
2a
K= d3 rΨ(r1 )(−e2 /r2 )Ψ(r2 ) = 2E1 e−2a/r0 (1 + )
r0
Z 2
2a 4a
∆= d3 rΨ(r1 )Ψ(r2 ) = (1 + + 2 )e−2a/r0
r0 3r0
The energies of the symmetric and antisymmetric wavefunctions can then be recast as:
r0 J+K
E s = E1 (1 − )+ (2.74)
a 1+∆
r0 J−K
Ea = E1 (1 − ) + (2.75)
a 1−∆
Only the symmetric solution has a minimal energy E s = −15.36eV lower than E1 =
−13.6eV at an internuclear distance 2a ∼ 1.3Å. The anti-symmetric solution has an
energy Ea always larger than E1 and thus cannot represent a bound state of the H+2
ion. Comparing these values to the exact solution (E0 = −16.4eV at 2a = 1.06Å), see
Fig.2.18, suggests that in this case the simple variational approach yields a reasonable
approximation to the ground state of the system.

Just as we constructed the orbitals in atoms, by filling up the (perturbed) Hydro-


gen orbitals from the bottom up taking into consideration Pauli’s exclusion principle we
may arrange the two electrons in the H2 molecule by filling up the molecular orbitals
just determined taking into account Pauli’s principle. The problem in that case is that
the electron-electron repulsion (e2 /|~r1 − ~r2 |) may not be negligible (as in the atomic
construction where it is somewhat neutralized by the higher positive charge of the nu-
cleus). However because the electron charges are spread over a region of size ∼ r0 , their
repulsive interaction turns out to alter only slightly the ground state energy which is
still stable for the symmetric wavefunction and unstable for the antisymmetric one. The
two electrons in H2 thus share the same ground state orbital with opposite spin (they
are paired). This is a general result: a chemical bond between two elements is formed
by electron pairs whose spatial wavefunction is symmetric with respect to particle ex-
change and of opposite spin (to satisfy Pauli’s exclusion principle).
2.9.1 Hückel’s molecular orbital theory
The variational approach used to estimate the ground state of H+2 has been used by
Eric Hückel in 1930 to estimate the eigen-energies and orbitals of hydro-carbon chains
and rings of n carbons. As we shall now show this Hückel Molecular Orbital (HMO)
theory justifies a-posteriori our initial approach to multi-state systems (the amonia and
aromatic molecules).
Hückel’s idea was to use a trial molecular wavefunction |Ψ MO > which is a super-
position of atomic wavefunctions |φ > (p-orbitals) centered on the individual Carbon
atoms:
64 QUANTUM MECHANICS

F. 2.19. The band gap ∆E of a few alkenes with n = 7, 9, 11 measured by the adsorp-
tion of photons of energy hν = ∆E. Notice that the measured energies are propori-
onal to sin π/(n + 1) as predicted from Hückel’ molecular orbital theory.
X
|Ψ MO >= ci |φi >
i

Just as we have done for H+2the idea is to determine the coefficients {ci } that minimize
the energy of the full Hamiltonian:

< Ψ MO |H|Ψ MO >= E < Ψ MO |Ψ MO >

Neglecting the overlap between neighboring atomic orbitals, i.e. assuming: ∆ =< φi |φ j >'
0 one gets: X
ci c j Hi j − Ec2i δi j = 0
ij

Where Hi j =< φi |H|φ j >. Hii and Hi j are respectively the Coulomb and resonance inte-
grals we saw above. If only the overlap between nearest neighbours j = i ± 1 contribute
significantly to the resonance integral, minimization of the energy with respect to the
vector of coefficients {ci } yields:

 E0 −A 0 . . . 0
 
  c 
 c1 
 
 −A E −A . . . 0   1 
0   .   . 
 .. ..   ..  = E  .. 
  
(2.76)
 . .     
 cn cn
0 . . . 0 −A

E0
THE CHEMICAL BOND 65

Where we defined: E0 ≡ Hii and A ≡ Hi,i+1 = Hi−1,i . Notice the similarity with the
model studied earlier for aromatic molecules, Eq.2.25. The eigen-energies can then be
computed as follows:

2 cos θ 1 0 . . .

0
1 2 cos θ 1 . . . 0

∆n = . = 0
..
. . . 0 1 2 cos θ

0

with 2 cos θ ≡ (E − E0 )/A. Computing the determinant yields a recursion relation: ∆n =


2 cos θ∆n−1 − ∆n−2 with ∆0 = 1 and ∆1 = 2 cos θ. Its solution is ∆n = sin(n + 1)θ/ sin θ
and the eigenvalues are consequently determined by: ∆n = 0, i.e. θ = mπ/(n + 1)
(m = 1, 2, ..., n):
mπ lπ
E = E0 + 2A cos = E0 − 2A cos (2.77)
n+1 n+1
with l = n + 1 − m, compare with Eq.2.26. The energy gap betwen the highest occupied
molecular orbital(HOMO, the top of the valence band) with lhomo = n/2 (or (n + 1)/2 for
odd n) and the lowest unoccupied molecular orbital (LUMO, bottom of the conduction
band) with llumo = lhomo + 1 is: ∆E = −4A sin π/(n + 1). Hückel’s computation of the
energy gap of alkene chains has been measured for chains of various lengths and com-
pares nicely with this prediction, see Fig.2.19. The filling up of the valence orbitals is
similar to the filling up of atomic orbitals, each orbit accomodating an electron pair with
opposite spin. The small gap in long alkene chains such as poly-acetylene and progress
in their doping (with electron donors or acceptors) have led to the development of con-
ducting polymers for which Alan J. Heeger, Alan MacDiarmid and Hideki Shirakawa
were awarded in 2000 the Nobel prize.

2.9.2 Molecular vibrational spectrum


Wa have seen that the nuclei of diatomic molecules such as H2 are stabilized by the ex-
change of paired electrons at the bottom of an effective potential V(r), see Fig.2.18. The
Hamiltonian describing a system of two bound atoms of mass m1 , m2 at an equilibrium
distance d = 2a is thus:

~2 ∂2 ~ 2 ∂2
H = −( + ) + V(|~r1 − ~r2 |)
2m1 ∂r12 2m2 ∂r22

Defining the center of mass (CM) coordinate : R ~ CM = (m1~r1 + m2~r2 )/(m1 + m2 )| and
~r12 = ~r1 − ~r2 allows to rewrite the previous Hamiltonian as: H = HCM + Hosc , where
HCM is the Hamiltonian of the center of mass (free particle of mass m1 + m2 ) and Hosc
is the Hamiltonian describing the two bound atoms in the CM coordinate system. Near
equilibrium the binding potential V(r12 ) can be expanded in a Taylor series:

1 ∂2 V ~ 2 + ...
V(r12 ) = V(d) − | (~r − d)
2 r12 =d 12
2 ∂r12
66 QUANTUM MECHANICS

F. 2.20. (a) The first eigen-functions of the harmonic oscillator. Notice the inversion
symmetry of the even function and the antisymmetry of the odd ones. (b) The com-
puted and measured eigenenergies of the vibrational spectrum of C6 H++ 4 (Vanoshi et
al., Theor. Chem. Account 120,45 (2008)). Notice the equal spacing of the spectral
lines and their value (∼ 0.1eV) typical of vibrational energies in molecules.

Defining k ≡ ∂2 V/∂r12
2
|r12 =d and looking at the deviations from equilibrium: ~r = ~r12 − d~
yields the Hamiltonian of an harmonic oscillator:

~ 2 ∂2 1
Hosc = − + kr2 (2.78)
2m ∂r2 2
with m = m1 m2 /(m1 + m2 ). The determination of the eigen-energies and eigen-functions
of a harmonic oscillator turns out to be one of those rare examples of an exactly soluble
problem in QM (the hydrogen atom is an other).
As done earlier, we shall look for a separable solution Ψ(x, y, z) = X(x)Y(y)Z(z) of
Schroedinger’s equation:
Hosc Ψ(x, y, z) = EΨ(x, y, z)
Since r2 = x2 + y2 + z2 we obtain:

∂2
X(x) + (λ x − α2 x2 )X(x) = 0 (2.79)
∂x2

with α = mω/~, ω = k/m and similar equations for Y(y) and Z(z). The eigenenergy is
given by:
~2 (λ x + λy + λz )
E=
2m
In the limit x → ∞, Eq.2.79, simplifies to X 00 = α2 x2 X, for which the asymptotic
bounded solution is X∞ ∼ exp(−αx2 /2). As we have done for the solution of the radial
equation of the Hydrogen atom, let us then look for a solution:

X(x) = f (x) exp(−αx2 /2)

where f (x) is a finite polynomial. Inserting that Ansatz into Eq.2.79 yields:
THE CHEMICAL BOND 67

f 00 − 2αx f 0 + (λ x − α) f = 0
Looking for a polynomial solution: f (x) = n an xn yields the recursion:
P

n(n + 1)an+2 + ((λ x − (2n + 1)α)an = 0


which ends at a value n = 0, 1, 2, .... such that: λ x = (2n + 1)α. Thus for a one dimen-
sional oscillator the energy levels are:
~2 λ x
En1D = = ( n + 1/2 )~ω (2.80)
2m
which (except for the 1/2) is the result obtained using the Bohr-Sommerfeld rule,
Eq.2.10. The energy levels in the harmonic oscillator are equally spaced by ~ω. For
molecules this spacing is typically of order 0.1eV (see Fig.2.20),
√ i.e. in the infrared.
The polynomials f (x) are known as Hermite polynomials of ξ = αx:
n 
2 d
Hn (ξ) = (−1)n eξ n e−ξ
2


so that the eigenfunctions read:
2
/2
√ α/π1/4
Xn (x) = An e−αx Hn ( αx) with : An = √ (2.81)
2n n!
The first Hermite polynomials describing the lowest eigenstates are:
H0 (ξ) = 1 (2.82)
H1 (ξ) = 2ξ (2.83)
H2 (ξ) = 4ξ2 − 2 (2.84)
H3 (ξ) = 8ξ3 − 12ξ (2.85)
The first eigenfunctions of the 1D-harmonic oscillator are shown in Fig.2.20. Notice
that that the even functions in n are symmetric with respect to inversion x → −x,
whereas the odd functions are antisymmetric. For identical di-atoms the spin of the
nuclei in the ground state (n = 0) must thus be opposite (Pauli’s principle). The 3D-
harmonic oscillator can be decomposed into three components with resulting energy
levels: En1 ,n2 ,n3 = (n1 + n2 + n3 + 3/2)~ω.
2.9.3 Molecular rotational spectrum
A diatomic molecule may not only vibrate but also rotate in space. In the CM coordinate
system the rotation Hamiltonian is:
~2 J 2 ~2 1 ∂ ∂ 1 ∂2
Hrot = =− [ (sin θ ) + ] (2.86)
2I 2md2 sin θ ∂θ ∂θ sin2 θ ∂φ2
where I = mr2 is the moment of inertia. Notice the similarity with the angular part of
the Hydrogen atom Hamiltonian. The eigen-energies of the rotation Hamiltonian are
therefore:
68 QUANTUM MECHANICS

~2 j( j + 1)
Ej = (2.87)
2I
The eigenfunctions are the spherical harmonics (see Appendix): Y jm (θ, φ) and as in the
Hydrogen atom, each eigenstate is 2 j+1 -fold degenerate (since m = − j, − j+1, ..., j). If
the molecule both vibrates and rotates its energy levels are only approximately a sum of
rotational and vibrational levels (in fact the system does not separates nicely into pure
vibrational and pure rotational modes).
For more complex molecules with different moments of inertia, the analysis be-
comes more complicated. For a symmetric top, i.e. if two of the moments of inertia are
equal (I1 = I2 , I3 , the Hamiltonian can still be written as:

~2 J12 ~2 J22 ~2 J32 ~2 J 2 ~2 1 1


H= + + = + ( − )J32
2I1 2I2 2I3 2I1 2 I3 I1

where the body-fixed angular momentum operators J1 , J2 , J3 , J 2 are defined in the


rotating frame of the top (and not with respect to a fixed frame as is the case in the
Hydrogen atom for the angular momentum operator J). ~ In particular the commutation
relations of Ji are anomalous:

[J1 , J2 ] = −iJ3

Notice the different sign as compared to the commutation relations of J~ ([J1 , J2 ] = iJ3 ).
These different angular momentum operators commute: [Ji , J j ] = 0 and satisfy J 2 =
J 2 . Hence the eigenstates of the symmetric top Hamiltonian are characterized by three
quantum numbers that define the eigen-values of J 2 (or J 2 ), J3 and J3 . For a given
eigenstate j, k of J 2 and J3 the eigen-energies are:

~2 ~2 1 1
E j,k,m = j( j + 1) + ( − )k2
2I1 2 I3 I1

When k > 0, the energy levels are 2(2 j + 1)-fold degenerate (the factor 2 j + 1 because
of the degeneracy in m (m = − j, ...., j) and the factor 2 because of the degeneracy
k → −k). The typical energies associated with molecular rotational levels corresponds
to microwave radiation energies (meV).

2.10 Time independent perturbation theory


Most of the problems in physics and in QM in particular are not exactly soluble. There-
fore the development of methods that yield approximate solutions is of particular im-
portance. We have seen earlier one such method (the variational approach) and have
pointed out its main drawback: the abscence of a systematic way of estimating the error
made. In the following we will study a systematic method to get an approximate solu-
tion and to estimate the error made within this approximation. The idea is to express the
TIME INDEPENDENT PERTURBATION THEORY 69

Hamiltonian H as the sum of a known exactly soluble one H0 plus a perturbation λH1 ,
where λ is a small parameter (λ  1):

H = H0 + λH1

One seeks the eigenstates |Ψn > of the full Schroedinger’s equation:

H|Ψn >= En |Ψn >

knowing the eigenstates |φn > of the soluble problem:

H0 |φn >= n0 |φn >

If these eigenstates span the Hilbert space of H (the space of its solutions) we may look
for an eigenstate of H as: X
|Ψn >= anm |φm >
m
Inserting that Ansatz into the full Schroedinger equation yields:
X X X X
anm H|φm >= anm (H0 + λH1 )|φm >= anm (m0 + λH1 )|φm >= anm En |φm >
m m m m

multiplying this equation on the left by < φl | yields:


X X
anm (En − m0 ) < φl |φm >= λ anm < φl |H1 |φm >
m m

Since < φl |φm >= δlm we obtain:


X
(En − l0 )anl = λ anm < φl |H1 |φm > (2.88)
m

When λ = 0: En = n0 and |Ψn >= |φn >, i.e. anl = δnl . When λ  1 we will therefore
look for a solution as an expansion in λ:

anl = δnl + λa(1)


nl + λ anl + . . .
2 (2)

En = n0 + λEn(1) + λ2 En(2) + . . .

Inserting that Ansatz into Eq.2.88 yields:


X
(n0 −l0 +λEn(1) +λ2 En(2) +. . . )(δnl +λa(1)
nl +λ anl +. . . ) = λ
2 (2)
< φl |H1 |φm > (δnm +λa(1)
nm +λ anm +. . . )
2 (2)

Equating terms order by order in λ yields:

λ: nl =< φl |H1 |φn >


En(1) δnl + (n0 − l0 )a(1) (2.89)
X
λ2 : En(2) δnl + En(1) a(1)
nl + ( n
0
− l
0 (2)
)anl = < φl |H1 |φm > a(1)
nm
m
..
.
70 QUANTUM MECHANICS

At order λ we obtain:

when n=l: En(1) =< φn |H1 |φn >


< φl |H1 |φn >
when n,l: nl =
a(1) (2.90)
n0 − l0

Notice that a(1)


nn is determined by the normalization condition of
X
|Ψn >= (1 + λa(1)
nn )|φn > +λ nm |φm >
a(1)
m,n

< Ψn |Ψn >= 1 at order λ, which implies: nn + ann


a(1) (1)∗
= 0. Hence a(1)
nn is purely imaginary:
nn = iα , so that to first order in λ:
a(1) (1)

X
|Ψn >= eλα |φn > +λ
(1)
nm |φm >
a(1)
m,n

The phase-shift by λα(1) is of no physical consequences (at this order it can be adsorbed
into a phase shift of |Ψn >) and can be chosen as zero. To second order in λ, Eq.2.90
yields:
X
when n = l : En(2) + En(1) a(1)
nn = < φn |H1 |φm > a(1)
nm
m
X
when n,l: (n0 − l0 )a(2)
nl = < φl |H1 |φm > a(1) (1) (1)
nm − E n anl
m

Using the first-order results,i.e. Eq.2.90, yields:


X
when n = l : En(2) = < φn |H1 |φm > a(1)
nm
m,n
X < φn |H1 |φm >< φm |H1 |φn > X | < φn |H1 |φm > |2
= = (2.91)
m,n n0 − m0 m,n n0 − m0
1 X < φl |H1 |φm >< φm |H1 |φn > < φn |H1 |φn >< φl |H1 |φn >
when n , l : a(2) = 0 −
nl
n − l0 m,n n0 − m0 n0 − l0

The coefficient a(2)


nn is determined as before from the normalization condition up to sec-
ond order in λ of:
X
|Ψn >= (1 + λ2 a(2)
nn )|φn > + nm + λ anm |φm >
(λa(1) 2 (2)

m,n

which yields:
1 X | < φm |H1 |φn > |2
nn = −
a(2)
2 m,n (n0 − m0 )2
One may compute higher order terms in this expansion, but usually only the first
non-zero contribution is computed. The advantage of this perturbation expansion over
TIME INDEPENDENT PERTURBATION THEORY 71

the variational approach seen earlier is that the error in a perturbation expansion up to
order n is of O(λn+1 ).

Example: Energy levels in non-Hydrogen atoms


We have seen that in Hydrogen, the energy levels have an accidental degeneracy in
the angular momentum quantum number l. While the wavefunctions Rnl depend on both
quantum numbers, the energy levels themselves depends only on n. To see how this de-
generacy is lifted in non-Hydrogen atoms consider the alkalines which have their inner
shells filled and one electron on a new shell. The electrostatic potential experienced by
the outer electron is:
e2 (Z − 1)e2 e2
V(r) = − − (1 − p(r)) ≡ − − δV(r)
r r r
The term δV(r) represents the Coulomb interaction of the outer electron with the Z − 1
partially screened charges of the nucleus (p(r) represent the effective negative charge
within a sphere of radius r due to the inner Z − 1 electrons: p(0) = 0 and limr→∞ p(r) =
1). The shift in the first order energy levels is given by Eq.2.90:
∆Enl
(1)
= − < Ψnlm |δV|Ψnlm >
Since δV is maximal at r = 0 (the nuclear charge is less screened the closer one gets to
the nucleus) , the energy shift is larger the larger the probability of finding the electron
near the nucleus at r = 0. When solving the Schroedinger equation for the Hydrogen
atom, we found that near r = 0: Rnl ∼ (r/rn )l with rn = nr0 . Assume for simplicity that
δV = δV0 exp(−r/r s )/r. Then:
δV0 ∞
Z ∞ Z
∆Enl = −
(1)
drr Rnl (r)δV(r) = −
2 2
dxx2l+1 e−rn x/rs
0 rn 0
δV0
=− (2l + 1)!(r s /rn )2l+2
rn
Since the inner electrons are spread on a distance r s < rn the negative energy shift is
maximal for l = 0 and decreases as l increases. This justifies the lift in the accidental
degeneracy of the Hydrogen energy levels in multi-electron atoms which we used when
filling-up their electronic orbitals (see Aufbau principle above and Fig.2.17).
2.10.1 The polarizability of atoms in an electric field
~ we shall estimate its
To study the polarizability of an atom in a constant electric field E,
change in energy when subjected to this field. We shall deduce its electric susceptibility
χ (see Chapter on electrostatics) from the electrostatic energy E p of an induced dipole
~
~ = χE:
P
E p = −P ~ = −χE2 /2
~ · E/2 (2.92)
Without loss of generality we will assume that the direction of E~ defines the z-axis.
Thus to the unperturbed Hamiltonian H0 of the atom one needs to add the perturbation
due to the electric field:
72 QUANTUM MECHANICS

H = H0 + ezE
To first order the energy shift, Eq.2.90:
∆Enl
(1)
=< Ψnlm |eEr cos θ|Ψnlm >= 0
Indeed, using the mathematical identities:
l+1−m m l+m m
cos θ Pm
l (cos θ) = Pl+1 (cos θ) + P (cos θ) (2.93)
2l + 1 2l + 1 l−1
s s
(l + 1 + m)(l + 1 − m) (l + m)(l − m)
cos θ Ylm (θ, φ) = Yl+1,m (θ, φ) + Yl−1,m (θ, φ)
(2l + 3)(2l + 1) (2l + 1)(2l − 1)
≡ Cl+1,m Yl+1,m (θ, φ) + Cl−1,m Yl−1,m (θ, φ)
we may compute the first order energy shift (with |Ψnlm >= Rnl (r)Ylm (θ, φ)) as follows:
Z
∆Enl
(1)
= ∗
r2 drdΩRnl (r)Ylm (θ, φ)eEr cos θRnl (r)Ylm (θ, φ)
Z
= eE r3 drdΩR2nl (r)Ylm∗
(θ, φ)[Cl+1,m Yl+1,m (θ, φ) + Cl−1,m Yl−1,m (θ, φ)] = 0

where we used the orthogonality of the spherical harmonics (see Appendix). The second
order correction in the energy, Eq.2.91 is then:
X | < Ψnlm |eEr cos θ|Ψn0 l0 m > |2
(2)
Enl = (2.94)
n0 ,l0 ,n,l
nl0 − n00 ,l0
Notice that this sum is well defined only for non-Hydrogen atoms where the denom-
inator is non-zero (as argued above). In Hydrogen the degeneracy of the energy with
respect to the angular quantum number l necessitates a perturbation analysis specific for
degenerate eigenstates which we will present below. In view of Eq.2.93, we see that in
the sum the only terms that contribute have l0 = l ± 1. In fact since the lifting of the
l−degeneracy is a small effect with respect to the variation of the energy with the major
quantum number n, the main contribution in the sum appearing in Eq.2.94 will come
from levels n0 = n.
X | < Ψnlm |r cos θ|Ψnl0 m > |2
(2)
Enl = e2 E2
l0 =l±1
nl0 − nl0 0
Since the shift in the angular l-degeneracy is larger the smaller l is: nl0 − n,l−1
0
> n,l+1
0

nl0 > 0 and the energy shift Enl is in general negative, see Fig.2.21. This raising of the
(2)

degeneracy in l with an applied electric field is known as the Stark effect.


Assuming the outer electrons to be in their ground states so that only levels with
l0 = l + 1 contribute in the prevuious equation, we may now compare the change in
energy of this atom in an electric field with the energy of an induced dipole, Eq.2.92
and identify the electric susceptibility χ of a non-Hydrogen atom as:
| < Ψnlm |r cos θ|Ψn,l+1,m > |2
χ = 2e2
n,l+1
0
− nl0
TIME INDEPENDENT PERTURBATION THEORY 73

F. 2.21. The shift in the spectral line of an alkaline, Rubidium excited to a high n
level (35s state). Notice the negative parabolic shift as a function of electric field,
from V.Bendkowsky et al., Nature, 458,1005 (2009).

F. 2.22. (a-b) Strong magnetic field lines pierce the surface of the Sun at its spots.
These are regions on the Sun’s surface that are cooler by about 1000◦ K than their
surrounding due to the damping of the underlying convective motion by the mag-
netic field. (c) This field of a few thousand Gauss causes the splitting of the spectral
emission lines measured along the vertical black line passing through a Sun’s spot
shown in (b).

2.10.2 Atom in a constant magnetic field: the Zeeman effect


Just as we have studied the lifting of the angular l−degeneracy resulting from the ap-
plication of an electric field, we shall now investigate the response of a Hydrogen-like
atom to a constant magnetic field. We have seen that the low energy approximation of
74 QUANTUM MECHANICS

Dirac’s equation yields the Schroedinger equation with a magnetic correction due to the
electron spin S~ with magnetic moment m~ = −(~e/2me c)~σ = −(eg/2me c)S~ , see Eq.2.57:

~ 2 Ze2
 
 (~p + (e/c)A)
 eg ~ ~ 
+ Ψ = EΨ
 
− §·B

2me r 2me c


 

~ ·A
In the Coulomb gauge: ∇ ~ = 0 (i.e. A
~ = (B~ × ~r)/2) this equation becomes:

~2 ∇2 Ze2 e2 ~ 2
( )
e ~ eg ~ ~
− − + A · ~p + A + § · B Ψ = EΨ
2me r me c 2me c2 2me c
In order to know what type of perturbation we need to consider let us estimate the
order of magnitude of the various terms on the left. The first two terms (the kinetic
and Coulomb energies) are of order of the ionization energy of the atom (typically 1-
10eV). The third term is of order ~eB/me c ' 10−8 B eV/Gauss, which for a strong
magnetic field B ∼ 104 Gauss is of O(10−4 eV). The fourth term on the left is of order
(e2 r02 /2me c2 )B2 ∼ 10−9 eV and is thus negligible with respect to the third one. The
Hamiltonian can thus be written as:
e ~ egB
H = H0 + ( B × ~r) · ~p + Sz
2mc 2me c
~ × ~r) · ~p = B
Since: ( B ~ · (~r × ~p) = B
~ · ~L, one obtains:

eB
H = H0 + (Lz + gS z )
2me c
Since H0 commutes with Lz (which eigenvalues are ~m; m = −l, ..., l) and with S z (with
eigenvalues ±~/2), the energy shift is:
e~B
∆E = (m ± 1) = ~ωL (m ± 1) (2.95)
2mc
where we set: g = 2 (see Eq.2.57). The frequency ωL = eB/2mc is known as the Larmor
frequency: it is the frequency of precession of a magnetic dipole about the direction of
the magnetic field. Notice that the magnetic field lifts the degeneracy in the azymuthal
quantum number m. It should also be noted that the present description of the Zeeman
effect is only an approximation valid for strong magnetic fields. Indeed as we have seen
when studying Dirac’s equation Lz and S z do not commute with the full Hamiltonian.
The reason that they do commute with H0 is that this Hamiltonian which we derived
from Dirac’s equation is only the lowest term in a low energy expansion of Dirac’s
relativistic Hamiltonian. In a strong magnetic field however, the magnetic perturbation
term is larger than the next (relativistic) terms in Dirac’s expansion which can thus
be neglected (for example the spin-orbit coupling discussed previously). The Zeeman
effect has been used to estimate the strong magnetic field coming out of the Sun’s spots
through the spliting of the Zeeman lines of Hydrogen atoms at the Sun’s surface, see
Fig.2.22
TIME INDEPENDENT PERTURBATION THEORY 75

F. 2.23. A Magnetic Resonance Imaging machine consists of an apparatus in which a


strong permanent magnet and various coils allow for the generation of strong mag-
netic field gradients along a pre-determined direction. This field gradient results in a
splitting of the proton energy levels with a gap that varies with space. By measuring
the adsorption of radio-waves emitted by a transmitter at given (Larmor) frequency
(corresponding to a given gap) one can deduce the concentration of Hydrogen atoms
at the corresponding location. By measuring the lifetime of the excited state, one
can further distinguish betwen different tissues. The result is a very detailed image
of various cuts through the body, here the head.

Nuclear magnetic resonance imaging (MRI).


One of the main applications of the Zeeman effect is in nuclear Magnetic Resonance
Imaging which functioning principle is based on the use of a magnetic field to align the
nuclear spin of atoms (mainly Hydrogen), in other words to split their nuclear energy
levels according to Eq.2.95: ∆E = ~ωLp σz (where ωLp is the Larmor frequency of the pro-
ton which for the typical fields of MRI, i.e. 1 Tesla=104 Gauss is in the radio-frequency
range: 42Mhz). The low energy state (spin aligned with the field) will be more occupied
than the high energy one. As seen earlier in the study of the Amonia molecule, radiation
at the resonant frequency (ω = ωLp ) will be absorbed by the protons. As a result they
will be excited into their high energy state (spin anti-parallel to the magnetic field) from
which they return to the ground state by spontaneous emission. In MRI one applies a
strong magnetic field gradient on a patient. As a result the protons (e.g. the nuclei of
hydrogen in H2 O)) display a Larmor frequency that varies with distance along the gra-
dient. By measuring the adsorption of radio-waves at a specific frequency, one can thus
deduce the concentration of protons at the corresponding location. The Fourier trans-
form of the received radio-signal (i.e. the intensity of the signal at varying frequencies)
therefore provides a map of proton density at varying locations along the field gradient.
Moreover by measuring the lifetime of the excited state (which depends on the chemical
environment of the Hydrogen nucleus, e.g. whether it is bound to Oxygen or Carbon)
one can distinguish between different tissues (e.g. richer in fat or water). The result is a
high resolution image of various cuts through the patient’s body, see Fig.2.23.
76 QUANTUM MECHANICS

2.10.3 Degenerate eigenstates


When eigenstates have the same energy the perturbation expansion we described above
diverges at second order since the denominator in Eq.2.91 is zero for degenerate states
(i.e. when n0 = m0 ). The way to deal with degenerate eigenstates is to neglect all tran-
sitions to non-degenerate eigenstates and to solve the equations exactly taking into ac-
count only the transitions between states of the same energy. This will usually raise the
degeneracy with the new eigenstates expressed as a superposition of the original un-
perturbed eigenstates. One may then proceed using non-degenerate perturbation theory
with these new eigenstates as zeroth order approximation.
As above we assume that the Hamiltonian is
H = H0 + λH1
And we want to solve for it taking into account only the N degenerate eigenstates |φn >
(n = 1, . . . , N) of H0 with energy 0 :
H0 |φn >=  0 |φn >
One seeks the new eigenstates |ψn > of H as a superposition of these degenerate eigen-
states: X
|ψn >= anm |φm >
m
which yields: H|ψn >= En |ψn > or:
X
anm (( 0 − En ) + λH1 )|φm >= 0
m

Defining Vnm ≡ λ < φn |H1 |φm > yields the following equation:

 + V11 − En ...
0
V12 V1N
V21  0 + V22 − En . . . V2N
. . . . = 0 (2.96)
.. .. .. ..


... VN,N−1  0 + VNN − En

VN1
The solution of this N th order equation for En will usually yield N eigenvalues and N
orthogonal eigenstates ψn > which can then be used instead of the |φn > in a perturbation
expansion together with the other non-degenerate eigenstates of H0 .
Example: The Stark effect in Hydrogen.
When previously studying the response of atoms to a constant electric field E~ we
mentioned that the case of Hydrogen was particular since its energy levels are degen-
erate with respect to the quantum numbers l (and m) and thus the second order non-
degenerate perturbation expansion breaks down. We shall now see how the degeneracy
of the electronic energy levels in hydrogen are affected by the presence of an electric
field. The Hamiltonian is:
H = H0 + ezE
Consider the case of the n = 2 level of hydrogen for which the states: l = 0; m = 0
and l = 1; m = 0, ±1 all have the same energy: E2 = E1 /4 (with E1 = −13.6eV).
TIME DEPENDENT PERTURBATION THEORY 77

Following our previous perturbation analysis for degenerate energy levels, we need
to diagonalize the Hamiltonian matrix: < 2, l0 , m0 |H|2, l, m >. While the unperturbed
Hamiltonian H0 is diagonal with energy E2 , using Eq.2.93 we deduce that the dipole
coupling < 2, l0 , m0 |ezE|2, l, m > is non-zero only if m0 = m and l0 = l ± 1, i.e. only when
coupling states |2, 0, 0 > and |2, 1, 0 >. In that case:
Z Z
< 2, 1, 0|ezE|2, 0, 0 > = eE r drR21 (r)R20 (r) dΩY10 (θ, φ) cos θY00 (θ, φ)
3

= −3er0 E ≡ −∆E2

The eigen-energies E are thus determined by the equation:



E2 − E −∆E2 0 0
−∆E2 E2 − E 0 0
= 0
0 0 E 2 − E 0
0 0 0 E2 − E

I
which solutions are: E so = E2 + ∆E2 = E2 + 3er0 E and E so
I
= E2 − ∆E2 with eigenstates:
|2, 0, 0 > −|2, 1, 0 >
|2, I > = √
2
|2, II >= |2, 0, 0 > +|2, 1, 0 >

2
III,IV
and E so = E2 with eigenstates |2, 1, 1 > and |2, 1, −1 >. Notice that for the hydrogen
atom the Stark effect, i.e. the shift in the energy level, is linear with electric field and
not quadratic as it is for the other atoms for which the electronic energy levels are non-
degenerate, see Eq.2.94.

2.11 Time dependent perturbation theory


Untill now, we have dealt mostly with time independent problems (the one exception
was our investigation of the amonia maser), namely systems for which the Hamiltonian
did not depend explicitly on time and for which the goal was to determine the eigen-
states and eigen-energies. Very often though to investigate these states, one drives the
system with a time dependent external perturbation, such as an electro-magnetic field,
and studies the induced transitions between the various eigenstates. In such a situation
the Hamiltonian of the system can be written as: H(t) = H0 + λH1 (t), where H0 is the
time independent unperturbed Hamiltonian for which the eigenstates and eigen-energies
are known:
H0 |φn >= n0 |φn >
To solve the time dependent Schroedinger equation:

i~∂t |Ψ(t) >= H|Ψ(t) >= H0 |Ψ(t) > +λH1 (t)|Ψ(t) > (2.97)

with initial condition: |Ψ(0) >= |φi >, we shall look for a solution as an expansion in
terms of the eigenstates of H0 :
78 QUANTUM MECHANICS

X
|Ψ(t) >= an (t)|φn >
n

Inserting that Ansatz into Eq.2.97 yields:


X dan X
i~ |φn >= [n0 + λH1 (t)]an (t)|φn >
n
dt n

Multiplying this equation on the left by < φm | and recalling that the eigenstates are
orthonormal: < φm |φn >= δmn yields:
dam (t) X
i~ = m0 am (t) + λ < φm |H1 (t)|φn > an (t)
dt n

0
Define: am (t) = µm (t)e−im t/~ to obtain the following equation for µm (t):
dµm (t) X 0 0
i~ =λ < φm |H1 (t)|φn > ei(m −n )t/~
dt n

Integrating yields:

λ X t
Z
µm (t) − δmi = dt < φm |H1 (t)|φn > eiωmn t µn
i~ n 0

with ωmn = (m0 − n0 )/~ and the initial condition µm (0) = δmi . For short enough times,
the system remains essentially in its initial state |φi > and for states m , i we may write:

λ t
Z
µm (t) ' dt < φm |H1 (t)|φi > eiωmi t
i~ 0
Now consider a periodic perturbation, such as an electro-magnetic wave, so that: λH1 (t) =
(Ve−iωt + V ∗ eiωt )/2 = |V| cos(ωt + θ):
< φm |V|φi > i[ωmi −ω]t < φm |V ∗ |φi > i[ωmi +ω]t
Z t
1
µm (t) ' dt e + e
i~ 0 2 2
= −[Vmi F(ω) + Vmi∗
F(−ω)]

with Vmi =< φm |V|φi > /2 and :


t
ei[ωmi −ω]t − 1
Z
F(ω) = dt ei[ωmi −ω]t =
0 m0 − i0 − ~ω

The probability that the system is in state |φm > at time t is:

Pm (t) = |am (t)|2 = |µm (t)|2 = |Vmi F(ω) + Vmi



F(−ω)|2

This transition probability will be large when either F(ω) or F(−ω) are large, which
occurs when m0 − i0 = ~ω or m0 − i0 = −~ω. The former case corresponds to absorption
TIME DEPENDENT PERTURBATION THEORY 79

of energy from the perturbing field (m0 > i0 ), the latter to emission of radiation from
the initial state (i0 > m0 ).

In case of adsorption the probability of finding the system in state |φm > at times t is
thus:
2π|Vmi |2
Pm (t) = |Vmi |2 |F(ω)|2 = |Vmi |2 (t/~)2 sinc2 [(ωmi − ω)t/2] → tδ(ωmi − ω) (2.98)
~2
where we used the mathematical identity: lima→0 sinc2 ax = (π/a)δ(x), which is a valid
approximation as long as (ωmi − ω)t < 1. Hence the transition rate from state |φi > into
state |φm >: T mi = dPm /dt is:
2π|Vmi |2 2π|Vmi |2 0
T mi = 2
δ(ωmi − ω) = δ(m − i0 − ~ω)
~ ~
Since the frequency of the initial beam is not infinitely sharp and if there exists a dis-
tribution of final state energies (which is usually the case due to Doppler broadening of
the final state energy, etc.) then the δ-function in the above
R equation has to be replace by
the energy density ρ(m ) of the final state (such that 1 = dδ( − i0 − ~ω) = dρ()):
R

2π|Vmi |2 0
T mi = ρ(m ) (2.99)
~
This equation is known as Fermi’s Golden rule. It relates the transition rate from the
ground state |φi > to the final state |φm > to this state’s energy density and to the coupling
between initial and final states. This is the result we obtained when studying the amonia
maser, see Eq.2.24.
In case of emission, the probability of finding the system in a lower energy state
|φm > after a time t is computed exactly as before:
∗ 2
2π|Vmi | 2π|Vmi |2 0
T mi = δ(m0 − i0 + ~ω) = ρ(i )
~ ~
which is the same result as that for adsorption (notice that here it is the energy density of
the initial state that appears in Fermi’s Golden rule, as the ground state being infinitely
lived is infinitely sharp). This is also in agreement with Einstein treatment of radiation,
see above.

Let us know investigate in more details the factors affecting the coupling element:
Vmi . Let the perturbation be a plane electromagnetic wave:
~=A
A ~0 ei(kx−ωt) + A
~∗0 e−i(kx−ωt)

propagating along the x-axis and let us choose the z-axis as parallel to A ~0 (which is
always possible unless some external field exist which defines the z-axis). The electric
80 QUANTUM MECHANICS

field is : E~ = −(1/c)∂t A
~ and B ~ × A.
~=∇ ~ The average intensity (Poynting vector) of this
electromagnetic wave is, see Eq.??:

I = (c/4π) < E~ × B
~ >= ω2 |A
~0 |2 /2πc

In presence of this EM-wave, the momentum ~p in the Hamiltonian is replaced by


~
~p − qA(t)/c), see Eq.2.46. The perturbation to the Hamiltonian is:
2
q ~ · ~p) + q A
~+A ~2
λH1 (t) = − (~p · A
2mc 2mc2
In the following we shall neglect the second term on the right, which is usually smaller
than the first (if however the matrix element computed with the first term vanishes, the
second term must be retained). One obtains:
Z
qA0
Vmi =< φm |λH1 |φi >= − d3 xφm (~x) p̂z φi (~x)eikx
mc
Since the electromagnetic wavelength λ is usually much larger than atomic scales we
may assume that kx  1 and thus approximate eikx ' 1, hence:
Z
i~qA0
Vmi ' d3 xφm ∂z φi
mc
The integral can be evaluated by studying the evolution of the following transition ma-
trix element,see Eq.2.45, :

~2
Z
d i~
i~ < φm |z|φi > = < φm |[z, H]|φi >= < φm |pz |φi >= d3 xφm ∂z φi
dt m m
! !
d d d
but : i~ < φm |z|φi > = i~ < φm | z |φi > + < φm | z i~ |φi > = −(m0 − i0 ) < φm |z|φi >
dt dt dt

Since by energy conservation: ωmi = (m0 − i0 )/~ = ω:

iqA0 ωzmi
Vmi =
c
And the transition rate is then:
2π|Vmi |2 2πq2 |A0 |2 ω2
T mi = 2
δ(ωmi − ω) = 2 |zmi |2 δ(ωmi − ω)
~ ~ c2
2πq 2
=( ) cI(ω) |zmi |2 δ(ωmi − ω) (2.100)
~c
The transition rate depends on the dipole moment between the initial and final states:
< φ f |qz|φi >. It is therefore known as the dipole transition rate. When investigating the
polarizability of atoms in a constant electric field by time independent perturbation the-
ory, see above, we have already been led to compute the transition element z f i between
TIME DEPENDENT PERTURBATION THEORY 81

various states |n, l, m > of the hydrogen atom, see Eq.2.94. We have seen there that this
element is zero unless m = m0 and l0 = l ± 1. If an other field (e.g. a constant mag-
netic field) exist, then the polarization of the EM-wave will in general be along an other
axis (say y) and the transition element will be non-zero if m0 = m ± 1. Therefore only
transitions between states differing in their angular momentum l by ±1 and in m by 0
or ±1 are possible (at leading order), for example transitions between s and p states are
possible, but not between two s or two p states.

Example: Rotational-vibrational transitions in diatomic molecules.


Since the rotational eigenstates of a diatomic molecule are the same as the Hydrogen
atom angular wavefunctions, the same selection rules hold for dipole transitions in a
diatomic molecule as in the Hydrogen atom. With respect to the initial state, the final
states differ in their angular quantum numbers j by ±1 and m by 0 or ±1.
If vibrational states of the diatomic molecule are excited, then the transition rate
depend on the dipole moment between initial and final eigenstates, |n1 ,2 , n3 >, of the
harmonic oscillator.
Z
< n1 , n2 , n3 |z|n1 , n2 , n3 >=
0 0 0
dzXn03 (z)zXn3 (z)

Using the results of our previous investigations, √


see Eq.2.81, the eigenfunctions of the
1D-harmonic oscillator are: Xn (z) ∼ e−αx /2 Hn ( αz) (for the 3D-oscillator the eigen-
2

functions are products of 1D-eigenstates). Using the property of the Hermite polynomi-
als:
2zHn = Hn+1 + 2nHn−1
It is easy to check that the dipole moment between initial and final states will be non-
zero only if: n0 = n ± 1.
Diatomic molecules can both vibrate and rotate. Their vibrational energy levels are
given by:
En1 ,n2 ,n3 = (n1 + n2 + n3 + 3/2)~ω
They are separated by an energy gap: ∆Evib = ~ω of O(0.1eV). The diatomic rotational
energy spectrum is:
~2 j( j + 1)
Ej =
2I
with a splitting between levels: ∆Erot = ~2 (2 j + 1)/2I of O(1meV) which is much
smaller than the energy gap between successive vibrational energy level. Electromag-
netically, i.e. light-induced transitions between vibrational-rotational energy levels in
diatomic molecules occur between vibrational levels with ∆n = ±1 and ∆ j = ±1.
Hence the emission spectrum from the first excited vibrational level to the ground state
will exhibit two bands known as R- and P-bands corresponding to transitions between
rotational energy levels differing by ∆ j = ±1, see Fig.2.24.
82 QUANTUM MECHANICS

F. 2.24. (a) Transitions between vibrational energy level n = 0 and n = 1 and rota-
tional energy levels differing by ∆ j = ±1 in a diatomic molecule such as CO. These
transitions define two characteristic spectral bands known as P-band when ∆ j = −1
and R-band when ∆ j = 1.

You might also like