Ben Simons - Advanced Quantum Physics (2009)
Ben Simons - Advanced Quantum Physics (2009)
Ben Simons
Contents
6 Spin 53
6.1 Spinors, spin pperators, Pauli matrices . . . . . . . . . . . . . . 54
6.2 Relating the spinor to the spin direction . . . . . . . . . . . . . 55
6.3 Spin precession in a magnetic field . . . . . . . . . . . . . . . . 56
6.3.1 Paramagnetic Resonance . . . . . . . . . . . . . . . . . 56
6.4 Addition of angular momenta . . . . . . . . . . . . . . . . . . . 58
6.4.1 Addition of two spin 1/2 degrees of freedom . . . . . . . 59
6.4.2 Addition of angular momentum and spin . . . . . . . . 60
6.4.3 Addition of two angular momenta J = 1 . . . . . . . . . 61
8 Identical Particles 78
8.1 Quantum statistics . . . . . . . . . . . . . . . . . . . . . . . . . 78
8.2 Space and spin wavefunctions . . . . . . . . . . . . . . . . . . . 80
8.3 Physical consequences of particle statistics . . . . . . . . . . . . 81
8.4 Ideal quantum gases . . . . . . . . . . . . . . . . . . . . . . . . 83
8.4.1 Non-interacting Fermi gas . . . . . . . . . . . . . . . . . 84
8.4.2 Ideal Bose gas . . . . . . . . . . . . . . . . . . . . . . . 86
9 Atomic structure 89
9.1 The “real” hydrogen atom . . . . . . . . . . . . . . . . . . . . . 90
9.1.1 Relativistic correction to the kinetic energy . . . . . . . 90
9.1.2 Spin-orbit coupling . . . . . . . . . . . . . . . . . . . . . 91
9.1.3 Darwin term . . . . . . . . . . . . . . . . . . . . . . . . 93
9.1.4 Lamb shift . . . . . . . . . . . . . . . . . . . . . . . . . 93
9.1.5 Hyperfine structure . . . . . . . . . . . . . . . . . . . . . 95
9.2 Multi-electron atoms . . . . . . . . . . . . . . . . . . . . . . . . 96
9.2.1 Central field approximation . . . . . . . . . . . . . . . . 97
9.3 Coupling schemes . . . . . . . . . . . . . . . . . . . . . . . . . . 102
9.3.1 LS coupling scheme . . . . . . . . . . . . . . . . . . . . 102
9.3.2 jj coupling scheme . . . . . . . . . . . . . . . . . . . . . 105
9.4 Atomic spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
9.4.1 Single electron atoms . . . . . . . . . . . . . . . . . . . 107
9.4.2 Helium and alkali earths . . . . . . . . . . . . . . . . . . 108
9.4.3 Multi-electron atoms . . . . . . . . . . . . . . . . . . . . 109
Preface
Quantum mechanics underpins a variety of broad subject areas within physics
and the physical sciences from high energy particle physics, solid state and
atomic physics through to chemistry. As such, the subject resides at the core
of every physics programme. By building upon the conceptual foundations
introduced in the IB Quantum Physics course, the aim of Part II Advanced
Quantum Physics is to develop further conceptual insights and technical flu-
ency in the subject.
Although 24 lectures is a long course to prepare(!), it is still insufficient
to cover all of the topics that different physicists will describe as “core”. For
example, some will argue that concepts of quantum computation should al-
ready be included as an established component of the core. Simiilarly, in the
field of quantum optics, some will say that a detailed knowledge of atomic and
molecular spectroscopy is key. In the field of solid state physics, the concept
of second quantization in many-body physics is also considered central. In all
of these cases, we will be able to touch only the surface of the subject. How-
ever, the material included in this course has been chosen to cover the key
conceptual foundations that provide access to these more advanced subjects,
the majority of which will be covered in subequent optional courses in Part II
and Part III.
In the following, we list an approximate “lecture by lecture” synopsis of
the different topics treated in this course. Those topics marked by a † will be
covered if time permits.
1 Foundations of quantum physics: Overview of course structure and
organization; brief revision of historical background: from wave mechan-
ics to the Schrödinger equation.
2 Quantum mechanics in one dimension: Wave mechanics of un-
bound particles; potential step; potential barrier and quantum tunnel-
ing; bound states; rectangular well; δ-function potential well; † Kronig-
Penney model of a crystal.
3 Operator methods in quantum mechanics: Operator methods;
uncertainty principle for non-commuting operators; Ehrenfest theorem
and the time-dependence of operators; symmetry in quantum mechan-
ics; Heisenberg representation; postulates of quantum theory; quantum
harmonic oscillator.
4 Quantum mechanics in more than one dimension: Rigid diatomic
molecule; angular momentum; commutation relations; raising and low-
ering operators; representation of angular momentum states.
5 Quantum mechanics in more than one dimension: Central po-
tential; atomic hydrogen; radial wavefunction.
6 Motion of charged particle in an electromagnetic field: Classical
mechanics of a particle in a field; quantum mechanics of particle in a
field; atomic hydrogen – normal Zeeman effect; † diamagnetic hydrogen
and quantum chaos; gauge invariance and the Aharonov-Bohm effect;
free electrons in a magnetic field – Landau levels.
7-8 Quantum mechanical spin: History and the Stern-Gerlach experi-
ment; spinors, spin operators and Pauli matrices; relating the spinor to
spin direction; spin precession in a magnetic field; parametric resonance;
addition of angular momenta.
Handout
To accompany the course, a substantial handout has been prepared.1 In some
cases, the handout contains material (usually listed as Info blocks) that goes
beyond the scope of the lectures. Needless to say, the examination will be
1
I should note that, in preparing the handout, I have made use of some web-based material
– particularly the excellent lecture notes by Fowler at Virginia – and notes prepared by David
Ward in previous generations of the course. I have also included links to useful material on
the course webpage.
limited to material that is covered in lectures and not the handout. Since
this handout is substantially new, it is inevitable that there will be some
typographical errors – some of them may even be important... I would be
most grateful if you could e-mail the errors that you find to bds10@cam. I will
try to maintain a “corrected” set of notes on the web.
The overheads used in lectures can also be recovered from the course web-
site along with other relevant and useful material.
2
Problem Sets
The problem sets are a vital and integral part of the course providing the means
to reinforce key ideas as well as practice techniques. Problems indicated by a †
symbol are regarded as challenging. Throughout these notes, I have included
a number of simpler exercises which may be completed “along the way”, and
aim to reinforce some of the ideas developed in the text.
Books
As a core of every undergraduate and graduate physics programme, there is
a wealth of excellent textbooks on the subject. Choosing ones that suit is a
subjective exercise. Apart from the handout, I am not aware of a text that
addresses all of the material covered in this course: Most are of course more
dense and far-reaching, and others are simply more advanced or imbalanced
towards specialist topics. At the same time, I would not recommend relying
solely on the handout. Apart from the range of additional examples they offer,
the textbooks will often provide a more erudite and engaging discussion of the
material. Although there are too many texts on the subject to discuss every
one, I have included below some of the books that I believe to be particularly
useful.
2
http://www.tcm.phy.cam.ac.uk/~bds10/aqp.html
understanding of
1.1.1 Black-body radiation electrical circuits,
spectroscopy,
In 1860, Gustav Kirchhoff introduced the concept of a “black body”, an ob- and the emission
of black-body
ject that absorbs all electromagnetic radiation that falls upon it – none passes radiation by heated objects. He
through and none is reflected. Since no light is reflected or transmitted, the coined the term “black body”
object appears black when it is cold. However, above absolute zero, a black radiation in 1862, and two sets of
independent concepts in both circuit
body emits thermal radiation with a spectrum that depends on temperature. theory and thermal emission are
To determine the spectrum of radiated energy, it is helpful to think of a black named “Kirchhoff’s laws” after him.
body as a thermal cavity at a temperature, T . The energy radiated by the cav-
ity can be estimated by considering the resonant modes. In three-dimensions, John William Strutt, 3rd Baron
the number of modes, per unit frequency per unit volume is given by Rayleigh OM (1842-1919)
An English physi-
cist who, with
8πν 2
N (ν)dν = 3 dν , William Ramsay,
discovered the el-
c
ement argon, an
where, as usual, c is the speed of light.1 achievement for
which he earned
1 the Nobel Prize
If we take the cavity to have dimension L3 , the modes of the cavity involve wave numbers in 1904. He also
k = πn/L where n = (nx , ny , nz ) denote the vector of integers nx = 0, 1, 2, · · · ∞, etc. The discovered the
corresponding frequency of each mode is given by ν = c|k|/2π, where c is the velocity of phenomenon of Rayleigh scattering,
light. The number of modes (per unit volume) having frequencies between ν and ν + dν is explaining why the sky is blue, and
predicted the existence of surface
waves known as Rayleigh waves.
Advanced Quantum Physics
1.1. HISTORICAL FOUNDATIONS OF QUANTUM PHYSICS 2
The amount of radiation emitted in a given frequency range should be Max Karl Ernst Ludwig Planck
proportional to the number of modes in that range. Within the framework 1858-1947
German physicist
of classical statistical mechanics, each of these modes have an equal chance whose work
of being excited, and the average energy in each mode is kB T (equipartition), provided the
bridge between
where kB is the Boltzmann constant. The corresponding energy density is classical and
therefore given by the Rayleigh-Jeans law, modern physics.
Around 1900
Planck developed
8πν 2
ρ(ν, T ) = kB T . a theory explain
c3 the nature of
black-body radiation. He proposed
This result predicts that ρ(ν, T ) increases without bound at high frequencies, that energy is emitted and absorbed
in discrete packets or “quanta,”
ν — the untraviolet (UV) catastrophe. However, such behaviour stood in and that these had a definite size –
contradiction with experiment which revealed that the short-wavelength de- Planck’s constant. Planck’s finding
didn’t get much attention until the
pendence was quite benign (see, e.g., Fig. 1.1). idea was furthered by Albert Einstein
To resolve difficulties presented by the UV catastrophe, Planck hypothe- in 1905 and Niels Bohr in 1913.
sized that, for each mode ν, energy is quantized in units of hν, where h denotes Planck won the Nobel Prize in 1918,
and, together with Einstein’s theory
the Planck constant. In this case, the energy of each mode is given by2 of relativity, his quantum theory
changed the field of physics.
!∞ −nhν/kB T
n=0 nhνe hν
!ε(ν)" = ! ∞ = hν/k T ,
n=0 e−nhν/kB T e B −1
quantum theory for the specific heat of matter, which takes into account the
quantization of lattice vibrational modes, was subsequently given by Debye
and Einstein.
discovery led to the quantum revolution in physics and earned Einstein the Arthur Holly Compton 1892-
Nobel Prize in 1921. 1962:
An American
physicist, he
1.1.3 Compton Scattering shared the 1927
Nobel Prize in
In 1923, Compton investigated the scattering of high energy X-rays and γ-ray Physics with C.
T. R. Wilson for
from electrons in a carbon target. By measuring the spectrum of radiation at his discovery of
different angles relative to the incident beam, he found two scattering peaks. the Compton ef-
fect. In addition
The first peak occurred at a wavelength which matched that of the incident to his work on
beam, while the second varied with angle. Within the framework of a purely X rays he made valuable studies of
classical theory of the scattering of electromagnetic radiation from a charged cosmic rays and contributed to the
development of the atomic bomb.
interaction between electrons and high energy photons (ca. keV) results in
the electron being given part of the energy (making it recoil), and a photon
with the remaining energy being emitted in a different direction from the
original, so that the overall momentum of the system is conserved. By taking
into account both conservation of energy and momentum of the system, the
Compton scattering formula describing the shift in the wavelength as function
of scattering angle θ can be derived,4
How could the quantum hν restricting allowed radiation energies also re- Niels Henrik David Bohr 1885-
strict the allowed electron orbits? In 1913 Bohr proposed that the angular 1962
A Danish physi-
momentum of an electron in one of these orbits was quantized in units of cist who made
Planck’s constant, fundamental
contributions
to the un-
h
L = me vr = n!, != . (1.2) derstanding
2π atomic structure
and quantum
2 2 mechanics, for
As a result, one finds that RH = ( 4π$
e
0
)2 2πch3m . which he received
the Nobel Prize in Physics in 1922.
Bohr mentored and collaborated
( Exercise. Starting with the Bohr’s planetary model for atomic hydrogen, with many of the top physicists
find how the quantization condition (1.2) restricts the radius of the allowed (circular) of the century at his institute in
orbits. Determine the allowed energy levels and obtain the expression for the Rydberg Copenhagen. He was also part of the
team of physicists working on the
constant above. Manhattan Project. Bohr married
Margrethe Norlund in 1912, and one
of their sons, Aage Niels Bohr, grew
But why should only certain angular momenta be allowed for the circling up to be an important physicist who,
electron? A heuristic explanation was provided by de Broglie: just as the like his father, received the Nobel
prize, in 1975.
constituents of light waves (photons) are seen through Compton scattering
to act like particles (of definite energy and momentum), so particles such as
electrons may exhibit wave-like properties. For photons, we have seen that Louis Victor Pierre Raymond,
the relationship between wavelength and momentum is p = h/λ. de Broglie 7th duc de Broglie 1892-1987
hypothesized that the inverse was true: for particles with a momentum p, the A French physi-
cist, de Broglie
wavelength is had a mind of
a theoretician
h rather than
λ= , i.e. p = !k , (1.3) one of an
p experimenter
or engineer.
where k denotes the wavevector of the particle. Applied to the electron in de
1924
Broglie’s
doctoral
the atom, this result suggested that the allowed circular orbits are standing thesis Recherches sur la théorie des
waves, from which Bohr’s angular momentum quantization follows. The de quanta introduced his theory of
electron waves. This included the
Broglie hypothesis found quantitative support in an experiment by Davisson particle-wave property duality theory
and Germer, and independently by G. P. Thomson in 1927. Their studies of matter, based on the work of
Einstein and Planck. He won the
of electron diffraction from a crystalline array of Nickel atoms (Fig. 1.5) con- Nobel Prize in Physics in 1929 for
firmed that the diffraction angles depend on the incident energy (and therefore his discovery of the wave nature of
electrons, known as the de Broglie
momentum). hypothesis or mécanique ondulatoire.
This completes the summary of the pivotal conceptual insights that paved
the way towards the development of quantum mechanics.
Figure 1.5: In 1927, Davisson and Germer bombarded a single crystal of nickel
with a beam of electrons, and observed several beams of scattered electrons that
were almost as well defined as the incident beam. The phenomenological similarities
with X-ray diffraction were striking, and showed that a wavelength could indeed be
associated with the electrons. The first figure shows the intensity of electron scattering
against co-latitude angle for various bombarding voltages. The second figure shows
the intensity of electron scattering vs. azimuth angle - 54V, co-latitude 50. Figures
taken taken from C. Davisson and L. H. Germer, Reflection of electrons by a crystal
of nickel, Nature 119, 558 (1927).
particle as E0 ei(p·r−Et)/!. Then, one may see that the wave equation applied
to the plane wave describing particle propagation yields the familiar energy-
momentum relationship, E 2 = (cp)2 for a massless relativistic particle.
This discussion suggests how one might extend the wave equation from
the photon (with zero rest mass) to a particle with rest mass m0 . We require
a wave equation that, when it operates on a plane wave, yields the relativis-
tic energy-momentum invariant, E 2 = (cp)2 + m20 c4 . Writing the plane wave
function φ(r, t) = Aei(p·r−Et)/!, where A is a constant, we can recover the
energy-momentum invariant by adding a constant mass term to the wave op-
erator,
" # $ %
∂t2 m20 c2 (cp)2 − E 2 + m20 c4 i(p·r−Et)/!
2
∇ − 2 − 2 ei(p·r−Et)/!
=− e = 0.
c ! (!c)2
This wave equation is called the Klein-Gordon equation and correctly de-
scribes the propagation of relativistic particles of mass m0 . However, its form
is seems inappropriate for non-relativistic particles, like the electron in hydro-
gen.
Continuing along the same lines, let us assume that a non-relativistic elec-
tron in free space is also described by a plane wave of the form Ψ(x, t) =
Aei(p·r−Et)/!. We need to construct an operator which, when applied to this
wave function, just gives us the ordinary non-relativistic energy-momentum
p2
relation, E = 2m . The factor of p2 can obviously be recovered from two
derivatives with respect to r, but the only way we can get E is by having a
single differentiation with respect to time, i.e.
!2 2
i!∂t Ψ(r, t) = − ∇ Ψ(r, t) .
2m
This is Schrödinger’s equation for a free non-relativistic particle. One remark-
able feature of this equation is the factor of i which shows that the wavefunc-
tion is complex.
How, then, does the presence of a spatially varying scalar potential effect
the propagation of a de Broglie wave? This question was considered by Som-
merfeld in an attempt to generalize the rather restrictive conditions in Bohr’s Arnold Johannes Wilhelm Som-
model of the atom. Since the electron orbit was established by an inverse- merfeld 1868-1951
A German the-
square force law, just like the planets around the Sun, Sommerfeld couldn’t oretical physicist
understand why Bohr’s atom had only circular orbits as opposed to Kepler- who pioneered
developments
like elliptical orbits. (Recall that all of the observed spectral lines of hydrogen in atomic
were accounted for by energy differences between circular orbits.) and quantum
de Broglie’s analysis of the allowed circular orbits can be formulated by physics, and also
educated and
assuming that, at some instant, the spatial variation of the wavefunction on groomed a large
going around the orbit includes a phase term of the form eipq/!, where here number of students for the new era
of theoretical physics. He introduced
the parameter q measures the spatial distance around the orbit. Now, for an the fine-structure constant into
acceptable wavefunction, the total phase change on going around the orbit quantum mechanics.
must be 2πn, where n is integer. For the usual Bohr circular orbit, where
p = |p| is constant, this leads to quantization of the angular momentum
L = pr = n!.
Sommerfeld considered a general Keplerian elliptical orbit. Assuming that
the de Broglie relation p = h/λ still holds, the wavelength must vary as the par-
ticle moves around the orbit, being shortest where the particle travels fastest,
at its closest approach to the nucleus. Nevertheless, the phase change on mov-
ing a short distance ∆q should still be p∆q/!. Requiring the wavefunction to
link up smoothly on going once around the orbit gives the Bohr-Sommerfeld
quantization condition
&
p dq = nh , (1.5)
'
where denotes the line integral around a closed orbit. Thus only certain
elliptical orbits are allowed. The mathematics is non-trivial, but it turns out
that every allowed elliptical orbit has the same energy as one of the allowed
circular orbits. That is why Bohr’s theory gave the correct energy levels.
This analysis suggests that, in a varying potential, the wavelength changes in
concert with the momentum.
separated from the spatial dependence. Setting Ψ(x, t) = T (t)ψ(x), and sep-
arating the variables, the Schrödinger equation takes the form,
( 2 2 )
− !2m
∂x
ψ(x) + V (x)ψ(x) i!∂t T (t)
= = const. = E .
ψ(x) T (t)
Since we have a function of only x set equal to a function of only t, they
both must equal a constant. In the equation above, we call the constant E
(with some knowledge of the outcome). We now have an equation in t set
equal to a constant, i!∂t T (t) = ET (t), which has a simple general solution,
T (t) = Ce−iEt/!, where C is some constant. The corresponding equation
in x is then given by the stationary, or time-independent Schrödinger
equation,
!2 ∂x2
− ψ(x) + V (x)ψ(x) = Eψ(x) .
2m
!2 ∂x2
Ĥ = − + V (x)
2m
defines the Hamiltonian and the stationary wave equation can be written as
the eigenfunction equation, Ĥψ(x) = Eψ(x), i.e. ψ(x) is an eigenstate of Ĥ
with eigenvalue E.
i! ∗
j(r, t) = − [ψ (r, t)∇ψ(r, t) − ψ(r, t)∇ψ ∗ (r, t)] . (1.8)
2m
This completes are survey of the foundations and development of quantum
theory. In due course, it will be necessary to develop some more formal math-
ematical aspects of the quantum theory. However, before doing, it is useful to
acquire some intuition for the properties of the Schrödinger equation. There-
fore, in the next chapter, we will explore the quantum mechanics of bound
and unbound particles in a one-dimensional system turning to discuss more
theoretical aspects of the quantum formulation in the following chapter.
Following the rules of quantum mechanics, we have seen that the state of
a quantum particle, subject to a scalar potential V (r), is described by the
time-dependent Schrödinger equation,
where Ĥ = − !2m
2 2
∇
+ V (r) denotes the Hamiltonian. To explore its proper-
ties, we will first review some simple and, hopefully, familiar applications of
the equation to one-dimensional systems. In addressing the one-dimensional
geometry, we will divide our consideration between potentials, V (x), which
leave the particle free (i.e. unbound), and those that bind the particle to some
region of space.
Ψ(x, t) = A ei(kx−ωt) ,
i! !k p
j(x, t) = − (Ψ∗ ∂x Ψ − c.c.) = |A|2 = |A|2 .
2m m m
and E denotes the energy of the particle. As |Ψ(x, t)|2 represents a probablility
density, it must be everywhere finite. As a result, we can deduce that the
wavefunction, ψ(x), is also finite. Moreover, since E and V (x) are presumed
finite, so must be ∂x2 ψ(x). The latter condition implies that
' both ψ(x) and ∂x ψ(x) must be continuous functions of x, even if V has
a discontinuity.
Consider then the influence of a potential step (see figure) on the prop-
agation of a beam of particles. Specifically, let us assume that a beam of
particles with kinetic energy, E, moving from left to right are incident upon a
potential step of height V0 at position x = 0. If the beam has unit amplitude,
the reflected and transmitted (complex) amplitudes are set by r and t. The
corresponding wavefunction is given by
With the incident, reflected, and transmitted fluxes given by |A|2 !km< , |Ar|2 !km< ,
and |At|2 !km> respectively, one obtains
) ) ) )
) k< − k> )2 ) 2k< )2 k> 2 k> 4k< k>
R = )) ) = |r|2 , T = ) )
) k< + k> ) k< = |t| k< = (k< + k> )2 .
k< + k> )
From these results one can confirm that the total flux is, as expected, conserved
in the scattering process, i.e. R + T = 1.
' Exercise. While E −V0 remains positive, show that the beam is able to prop-
agate across the potential step (see figure). Show that the fraction of the beam that
is reflected depends on the relative height of the step while the phase depends on the
sign of V0 . In particular, show that for V0 > 0, the reflected beam remains in phase
with the incident beam, while for V0 < 0 it is reversed. Finally, when E − V0 < 0,
show that the beam is unable to propagate to the right (R = 1). Instead show that
there is an evanescent decay of the wavefunction into the barrier region with a decay
!
length set by 2π !2 /2m(V0 − E). If V0 → ∞, show that the system forms a standing
wave pattern.
2k1 k2 e−ik1 a
t= ,
2k1 k2 cos(k2 a) − i(k12 + k22 ) sin(k2 a)
translating to a transmissivity of
1
T = |t|2 = + ,2 ,
1+ 1
4
k1
k2 − k2
k1 sin2 (k2 a)
For κ2 a ) 1 (the weak tunneling limit), the transmittivity takes the form
16k12 κ22 −2κ2 a
T * e .
(k12 + κ22 )2
Finally, on a cautionary note, while the phenomenon of quantum mechan-
ical tunneling is well-established, it is difficult to access in a convincing exper- Real part of the wavefunction for
imental manner. Although a classical particle with energy E < V0 is unable E/V0 = 0.6 (top), E/V0 = 1.6
to penetrate the barrier region, in a physical setting, one is usually concerned (middle), and E/V0 = 1 + π 2 /2
with a thermal distribution of particles. In such cases, thermal activation (bottom), where mV0 a2 /!2 =
may lead to transmission over a barrier. Such processes often overwhelm any 1. In the first case, the system
contribution from true quantum mechanical tunneling. shows tunneling behaviour, while
in the third case, k2 a = π and
the system shows resonant trans-
' Info. Scanning tunneling microscopy (STM) is a powerful technique for
mission.
viewing surfaces at the atomic level. Its development in the early eighties earned
its inventors, Gerd Binnig and Heinrich Rohrer (at IBM Zürich), the Nobel Prize in
Physics in 1986. STM probes the density of states of a material using the tunneling
current. In its normal operation, a lateral resolution of 0.1 nm and a depth resolution
of 0.01 nm is typical for STM. The STM can be used not only in ultra-high vacuum,
but also in air and various other liquid or gas ambients, and at temperatures ranging
from near zero kelvin to a few hundred degrees Celsius.
The STM is based on the concept of quantum tunnelling (see Fig. 2.1). When a
STM image showing two point
conducting tip is brought in proximity to a metallic or semiconducting surface, a bias
defects adorning the copper
between the two can allow electrons to tunnel through the vacuum between them.
(111) surface. The point defects
For low voltages, this tunneling current is a function of the local density of states at (possibly impurity atoms) scat-
the Fermi level, EF , of the sample.3 Variations in current as the probe passes over ter the surface state electrons re-
the surface are translated into an image. STM can be a challenging technique, as it sulting in circular standing wave
requires extremely clean surfaces and sharp tips. patterns.
3
Although the meaning of the Fermi level will be address in more detail in chapter 8, we
mention here that it represents the energy level to which the electron states in a metal are
fully-occupied.
where the potential depth V0 is assumed positive. In this case, we will look
for bound state solutions with energies lying in the range −V0 < E < 0.
Since the Hamiltonian is invariant under parity transformation, [Ĥ, P̂ ] = 0
(where P̂ ψ(x) = ψ(−x)), the eigenstates of the Hamiltonian Ĥ must also be
From this result, we obtain that κ = maV0 /!2 , leading to the bound state
energy
ma2 V02
E=− .
2!2
Indeed, the solution is unique. An attractive δ-function potential hosts only
one bound state.
' Exercise. Explore the bound state properties of the “molecular” binding
potential V (x) = −aV0 [δ(x + d) + δ(x − d)]. Show that it consists of two bound states,
one bonding (nodeless) and one antibonding (single node). How does the energy of
the latter compare with two isolated δ-function potential wells?
Since the potential is repulsive, it is evident that all states have energy E > 0.
This potential has a new symmetry; a translation by the lattice spacing a leaves
the protential unchanged, V (x + a) = V (x). The probability density must therefore
exhibit the same translational symmetry,
which means that, under translation, the wavefunction differs by at most a phase,
ψ(x + a) = eiφ ψ(x). In the region from (n − 1)a < x < na, the general solution of
the Schrödinger equation is plane wave like and can be written in the form,
Bn + An+1 sin(ka)
Bn+1 = .
cos(ka)
Similarly, the discontinuity in the first derivative, ∂x ψn+1 |x=na −∂x ψn |na = 2maV
!2 ψn (na),
0
2maV0
An+1 = Bn cos(ka) − Bn sin(ka) + An cos(ka) .
!2 k
Similarly, replacing the expression for An+1 in that for Bn+1 , one obtains the parallel
equation,
2maV0
Bn+1 = Bn sin(ka) + Bn cos(ka) + An sin(ka) .
!2 k
With these two eqations, and the relations An+1 = eiφ An and Bn+1 = eiφ Bn , we
obtain the quantization condition,6
maV0
cos φ = cos(ka) + sin(ka) .
!2 k
√
As !k = 2mE, this result relates the allowed values of energy to the real parameter,
φ. Since cos φ can only take values between −1 and 1, there are a sequence of allowed
bands of energy with energy gaps separating these bands (see Fig. 2.2). Figure 2.2: Solid line shows the
Such behaviour is characteristic of the spectrum of periodic lattices: In the peri- variation of cos φ with ka over a
odic system, the wavefunctions – known as Bloch states – are indexed by a “quasi”- range from −1 to 1 for V0 = 2
momentum index k, and a band index n where each Bloch band is separated by an and ma2 !2 = 1. The blue line
energy gap within which there are no allowed states. In a metal, electrons (fermions) shows 0.01×E = (!k)2 /2m. The
populate the energy states starting with the lowest energy up to some energy scale shaded region represents values
of k and energy for which there
known as the Fermi energy. For a partially-filled band, low-lying excitations associ-
is no solution.
ated with the continuum of states allow electrons to be accelerated by a weak electric
field. In a band insulator, all states are filled up to an energy gap. In this case,
a small electric field is unable to excite electrons across the energy gap – hence the
system remains insulating.
6
Eliminating An and Bn from the equations, a sequence of cancellations obtains
„ «
2maV0
e2iφ − eiφ sin(ka) + 2 cos(ka) + 1 = 0 .
! k
2
Operator methods in
quantum mechanics
Alongside the ket, we can define the “bra”, #ψ|. Together, the bra and ket
define the scalar product
! ∞
#φ|ψ" ≡ dx φ∗ (x)ψ(x) ,
−∞
from which follows the identity, #φ|ψ"∗ = #ψ|φ". In this formulation, the real
space representation of the wavefunction is recovered from the inner prod-
uct ψ(x) = #x|ψ" while the momentum space wavefunction is obtained from
ψ(p) = #p|ψ". As with a three-dimensional vector space where a · b ≤ |a| |b|,
the magnitude of the scalar product is limited by the magnitude of the vectors,
"
#ψ|φ" ≤ #ψ|ψ"#φ|φ" ,
3.1 Operators
An operator  is a “mathematical object” that maps one state vector, |ψ",
into another, |φ", i.e. Â|ψ" = |φ". If
Â|ψ" = a|ψ" ,
Moreover, for any linear operator Â, the Hermitian conjugate operator
(also known as the adjoint) is defined by the relation
! !
#φ|Âψ" = dx φ∗ (Âψ) = dx ψ(† φ)∗ = #† φ|ψ" . (3.2)
From the definition, #† φ|ψ" = #φ|Âψ", we can prove some useful rela-
tions: Taking the complex conjugate, #† φ|ψ"∗ = #ψ|† φ" = #Âψ|φ", and
then finding the Hermitian conjugate of † , we have
Therefore, if we take the Hermitian conjugate twice, we get back to the same
operator. Its easy to show that (λÂ)† = λ∗ † and ( + B̂)† = † + B̂ † just
from the properties of the dot product. We can also show that (ÂB̂)† = B̂ † †
from the identity, #φ|ÂB̂ψ" = #† φ|B̂ψ" = #B̂ † † φ|ψ". Note that operators
are associative but not (in general) commutative,
i.e. #Ĥψ|ψ" = #ψ|Ĥψ" = #Ĥ † ψ|ψ", and Ĥ † = Ĥ. Operators that are their
own Hermitian conjugate are called Hermitian (or self-adjoint).
' Exercise. Prove that the momentum operator p̂ = −i!∇ is Hermitian. Fur-
ther show that the parity operator, defined by P̂ ψ(x) = ψ(−x) is also Hermitian.
' Info. Projection operators and completeness: A ‘ket’ state vector fol-
lowed by a ‘bra’ state vector is an example of an operator. The operator which
projects a vector onto the jth eigenstate is given by |j"#j|. First the bra vector dots
into the state, giving the coefficient of |j" in the state, then its multiplied by the unit
vector |j", turning it back into a vector, with the right length to be a projection. An
operator maps one vector into another vector, so this is an operator. If we sum over
a complete set of states, like the eigenstates of a Hermitian operator, we obtain the
(useful) resolution of identity
&
|i"#i| = I .
i
%
Again, in coordinate form, we can write i φ∗i (x)φi (x" ) = δ(x − x" ). Indeed, we can
%
form a projection operator into a subspace, P̂ = subspace |i"#i|.
As in a three-dimensional
% vector %
space, the expansion of the vectors |φ" and
|ψ", as |φ" = i bi |i" and |ψ" = i ci |i",%
allows the dot product to be taken
by multiplying the components, #φ|ψ" = i b∗i ci .
' Example: The basis states can be formed from any complete set of orthogonal
states. In particular, they can' be formed from' the basis states of the position or
∞ ∞
the momentum operator, i.e. −∞ dx|x"#x| = −∞ dp|p"#p| = I. If we apply these
definitions, we can then recover the familiar Fourier representation,
! ∞ ! ∞
1
ψ(x) ≡ #x|ψ" = dp #x|p" #p|ψ" = √ dp eipx/! ψ(p) ,
−∞ ( )* +
√ 2π! −∞
eipx/! / 2π!
where #x|p" denotes the plane wave state |p" expressed in the real space basis.
Û = e−iĤt/! ,
denotes
% the time-evolution operator. By inserting the resolution of identity,
1
I = i |i"#i|, where the states |i" are eigenstates of the Hamiltonian with
eigenvalue Ei , we find that
& &
|ψ(t)" = e−iĤt/! |i"#i|ψ(0)" = |i"#i|ψ(0)"e−iEi t/! .
i i
1
This equation follows from integrating the time-dependent Schrödinger equation, Ĥ|ψ! =
i!∂t |ψ!.
p̂2
' Example: Consider the harmonic oscillator Hamiltonian Ĥ = + 12 mω 2 x2 .
2m
Later in this chapter, we will see that the eigenstates, |n", have equally-spaced eigen-
values, En = !ω(n + 1/2), for n = 0, 1, 2, · · ·. Let us then consider the time-evolution
of a general wavepacket, |ψ(0)", under
% the action of the Hamiltonian. From the equa-
tion above, we find that |ψ(t)" = n |n"#n|ψ(0)"e−iEn t/! . Since the eigenvalues are
equally spaced, let us consider what happens when t = tr ≡ 2πr/ω, with r integer.
In this case, since e2πinr = 1, we have
&
|ψ(tr )" = |n"#n|ψ(0)"e−iωtr /2 = (−1)r |ψ(0)" .
n
From this result, we can see that, up to an overall phase, the wave packet is perfectly
reconstructed at these times. This recurrence or “echo” is not generic, but is a
manifestation of the equal separation of eigenvalues in the harmonic oscillator.
' Exercise. Using the symmetry of the harmonic oscillator wavefunctions under
parity show that, at times tr = (2r + 1)π/ω, #x|ψ(tr )" = e−iωtr /2 #−x|ψ(0)". Explain
the origin of this recurrence.
Û † Û = I .
i #ψ|[Û , V̂ ]|ψ"
2λ(∆B)2 + i#ψ|[Û , V̂ ]|ψ" = 0, λ=− .
2 (∆B)2
i
∆A ∆B ≥ #[Â, B̂]" .
2
i !
∆p ∆x ≥ #[p̂, x]" = .
2 2
Similarly, if we use the conjugate coordinates of time and energy, [Ê, t] = i!,
we have
!
∆E ∆t ≥ .
2
d i, -
#ψ|Â|ψ" = #ψ|Ĥ Â|ψ" − #ψ|ÂĤ|ψ" +#ψ|∂t Â|ψ" .
dt (! )* +
i
#ψ|[Ĥ, Â]|ψ"
!
This is an important and general result for the time derivative of expectation
values which becomes simple if the operator itself does not explicitly depend
on time,
d i
#ψ|Â|ψ" = #ψ|[Ĥ, Â]|ψ" .
dt !
From this result, which is known as Ehrenfest’s theorem, we see that expec-
tation values of operators that commute with the Hamiltonian are constants Paul Ehrenfest 1880-1933
An Austrian
of the motion. physicist and
mathematician,
' Exercise. Applied to the non-relativistic Schrödinger operator for a single who obtained
2
p̂ %p̂& ˙ = −#∂x V ". Dutch citizenship
particle moving in a potential, Ĥ = 2m + V (x), show that #ẋ" = m , #p̂" in 1922. He
Show that these equations are consistent with the relations, made major
. / . / contributions
d ∂H d ∂H to the field
#x" = , #p̂" = − , of statistical
dt ∂p dt ∂x mechanics and its relations with
quantum mechanics, including the
the counterpart of Hamilton’s classical equations of motion. theory of phase transition and the
Ehrenfest theorem.
 → Û † ÂÛ .
Û † ÂÛ = R[Â] .
If Ô(p̂, r̂) ≡ Ĥ, the quantum Hamiltonian, such unitary transformations are
said to be symmetries of the quantum system.
one can show that, for a constant vector a, the unitary operator
# $
i
Û (a) = exp − a · p̂ ,
!
acting in the Hilbert space of a Schrödinger particle performs a spatial trans-
lation, Û † (a)f (r)Û (a) = f (r + a), where f (r) denotes a general algebraic
function of r.
where summation on the repeated indicies, im is assumed. Then, making use of the
Baker-Hausdorff identity (exercise)
1
e B̂e− = B̂ + [Â, B̂] + [Â, [Â, B̂]] + · · · , (3.3)
2!
it follows that
1
Û † (a)f (r)Û (a) = f (r) + ai1 (∇i1 f (r)) + ai ai (∇i1 ∇i2 f (r)) + · · · = f (r + a) ,
2! 1 2
where the last identity follows from the Taylor expansion.
As we have seen, time translations are generated by the time evolution op-
erator, Û (t) = exp[− !i Ĥt]. Therefore, every observable which commutes with
the Hamiltonian is a constant of the motion (invariant under time transla-
tions),
P̂ ψ(r) = ψ(−r) .
If we require that ψ(x, 2t) = ψ ∗ (x, 0), we must have Ĥ ∗ (x) = Ĥ(x). Therefore,
Ĥ is invariant under time-reversal if and only if Ĥ is real.
' Example: For example, if the Hamiltonian commutes with the angular mo-
mentum operators, L̂i , i = x, y, z, i.e. it is invariant under three-dimensional rota-
tions, an energy level with a given orbital quantum number , is at least (2, + 1)-fold
degenerate. Such a degeneracy can be seen as the result of non-trivial actions of
the operator L̂x and L̂y on an energy (and L̂z ) eigenstate |E, ,, m" (where m is the
magnetic quantum number asssociated with L̂z ).
Ĥ = !ω(n̂ + 1/2) .
1
|n" = √ (a† )n |0" ,
n!
The operators a and a† represent ladder operators and have the effect of
lowering or raising the energy of the state.
In fact, the operator representation achieves something quite remarkable
and, as we will see, unexpectedly profound. The quantum harmonic oscillator
describes the motion of a single particle in a one-dimensional potential well.
It’s eigenvalues turn out to be equally spaced – a ladder of eigenvalues, sepa-
rated by a constant energy !ω. If we are energetic, we can of course translate
our results into a coordinate representation ψn (x) = #x|n".7 However, the
operator representation affords a second interpretation, one that lends itself
to further generalization in quantum field theory. We can instead interpret
the quantum harmonic oscillator as a simple system involving many fictitious
particles, each of energy !ω. In this representation, known as the Fock space,
the vacuum state |0" is one involving no particles, |1" involves a single particle,
|2" has two and so on. These fictitious particles are created and annihilated
by the action of the raising and lowering operators, a† and a with canoni-
cal commutation relations, [a, a† ] = 1. Later in the course, we will find that
these commutation relations are the hallmark of bosonic quantum particles
and this representation, known as the second quantization underpins the
quantum field theory of the electromagnetic field.
' Info. There is evidently a huge difference between a stationary (Fock) state
of the harmonic oscillator and its classical counterpart. For the classical system, the
equations of motion are described by Hamilton’s equations of motion,
P
Ẋ = ∂P H = , Ṗ = −∂X H = −∂x U = −mω 2 X ,
m
where we have used capital letters to distinguish them from the arguments used to de-
scribe the quantum system. In the phase space, {X(t), P (t)}, these equations describe
a clockwise rotation along an elliptic trajectory specified by the initial conditions
{X(0), P (0)}. (Normalization of momentum by mω makes the trajectory circular.)
7
Expressed in real space, the harmonic oscillator wavefunctions are in fact described by
the Hermite polynomials,
r „r « » –
1 mω mωx2
ψn (x) = #x|n! = H n x exp − ,
2n n! ! 2!
2
dn −x2
where Hn (x) = (−1)n ex dxn
e .
On the other hand, the time dependence of the Fock space state, as of any sta-
tionary state, is exponential,
If we define  = α∗ a − αa† , then F̂α = e− and F̂α† = e . If we then take B̂ = I, we
have µ = 0, and F̂α† F̂α = I. This merely means that the shift operator is unitary -
not a big surprise, because if we shift the phase point by (+α) and then by (−α), we
certainly come back to the initial position.
If we take B̂ = a, using the commutation relations,
so that µ = α, and F̂α† aF̂α = a + α. Now let us consider the operator F̂α F̂α† aF̂α .
From the unitarity condition, this must equal aF̂α , while application of Eq. (3.4)
yields F̂α a + αF̂α , i.e.
Applying this equality to the ground state |0" and using the following identities,
a|0" = 0 and F̂α |0" = |α", we finally get a very simple and elegant result:
a|α" = α|α" .
8
After R. J. Glauber who studied these states in detail in the mid-1960s, though they
were known to E. Schrödinger as early as in 1928. Another popular name, coherent states,
does not make much sense, because all the quantum states we have studied so far (including
the Fock states) may be presented as coherent superpositions.
This means that the probability of finding the system in level n is given by the
Poisson distribution, Pn = |αn |2 = #n"n e−%n& /n! where #n" = |α|2 . More importantly,
δn = #n"1/2 / #n" when #n" 0 1 – the Poisson distribution approaches the Gaussian
distribution when #n" is large.
The time-evolution of Glauber states may be described most easily in the Schrödinger
representation when the time-dependence is transferred to the wavefunction. In this
case, α(t) ≡ √2x1
(X(t) + i Pmω
(t)
), where {X(t), P (t)} is the solution to the classical
0
equations of motion, α̇(t) = −iωα(t). From the solution, α(t) = α(0)e−iωt , one may
show that the average position and momentum evolve classically while their fluctua-
tions remain stationary,
3 41/2 3 41/2
x0 ! mωx0 !mω
∆x = √ = , ∆p = √ = .
2 2mω 2 2m
In the quantum theory of measurements these expressions are known as the “standard
quantum limit”. Notice that their product ∆x ∆p = !/2 corresponds to the lower
bound of the Heisenberg’s uncertainty relation.
' Exercise. Show that, in position space, the Glauber state takes the form
# $
mω Px
#x|α" = ψα (x) = C exp − (x − X)2 + i .
2! !
P̂2 L̂2
Ĥ = + , (4.1)
2M 2I
L̂2
Ĥrot = .
2I
The eigenstates of this component of the Hamiltonian are simply the states of
the angular momentum operator. Indeed, in any quantum mechanical system
involving a radial potential, the angular momentum will be conserved, i.e.
[Ĥ, L̂] = 0 meaning that the angular component of the wavefunction can be
indexed by the states of the angular momentum operator. We therefore now
digress to discuss the quantum mechanics of angular momentum.
where, as usual, #ijk denotes the totally antisymmetric tensor — the Levi-
Civita symbol.2
$ Exercise. Show that the angular momentum operator commutes with the
p̂2
Hamiltonian of a particle moving in a central potential, Ĥ = 2m + V (r). Show that
the Hamiltonian of a free particle of mass m confined to a sphere of radius R is given
L̂2
by Ĥ = 2mR 2.
$ Info. Raising and lowering operators for angular momentum: The set
of eigenvalues a and b can be obtained by making use of a trick based on a “ladder
operator” formalism which parallels that used in the study of the quantum harmonic
oscillator in section 3.4. Specifically, let us define the raising and lowering operators,
1
In this chapter, we will index the angular momentum operators with a ‘hat’. Later, we
will become lazy and the hat may well disappear.
2
Recall that !ijk = 1 if (i, j, k) is an even permutation of (1,2,3), −1 if it is an odd
permutation, and 0 if any index is repeated.
Since each component of the angular momentum commutes with L̂2 , we can deduce
that the action of L̂± on |a, b$ cannot affect the value of a relating to the magnitude
of the angular momentum. However, they do effect the projection:
L̂z L̂± |a, b$ = L̂± L̂z |a, b$ + [L̂z , L̂± ]|a, b$ = (b ± !)L̂± |a, b$ .
Therefore, if |a, b$ is an eigenstate of L̂z with eigenvalue b, L̂± |a, b$ is either zero,
or an eigenstate of L̂z with eigenvalue b ± !, i.e. L̂± |a, b$ = C± (a, b)|a, b ± !$ where
C± (a, b) is a normalisation constant.
To fix the normalisation, we may note that the norm,
!! !!2
!! !! †
!!L̂± |a, b$!! = %a, b|L̂± L̂± |a, b$ = %a, b|L̂∓ L̂± |a, b$ ,
where we have used the identity L̂†± = L̂∓ . Then, making use of the relation L̂∓ L̂± =
L̂2x + L̂2y ± i[L̂x L̂y ] = L̂2 − L̂2z ± !L̂z , and the presumed normalisation, %a, b|a, b$ = 1,
one finds
!! !!2 " #
!! !!
!!L̂± |a, b$!! = %a, b| L̂2 − L̂2z ± !L̂z |a, b$ = a − b2 ∓ !b . (4.3)
For a given a, bmax and bmin are determined uniquely — there cannot be two states
with the same a but different b annihilated by L̂+ . It also follows immediately that
a = bmax (bmax +!) and bmin = −bmax . Furthermore, we know that if we keep operating
on |a, bmin $ with L̂+ , we generate a sequence of states with L̂z eigenvalues bmin + !,
bmin + 2!, bmin + 3!, · · ·. This series must terminate, and the only possible way for
that to happen is for bmax to be equal to bmin + n! with n integer, from which it
follows that bmax is either an integer or half an odd integer times !
At this point, we switch to the standard notation. We have established that the
eigenvalues of L̂z form a finite ladder, with spacing !. We write them as m!, and %
is used to denote the maximum value of m, so the eigenvalue of L̂2 , a = %(% + 1)!2 .
Both % and m will be integers or half odd integers, but the spacing of the ladder of
m values is always unity. Although we have been writing |a, b$ with a = %(% + 1)!2 ,
b = m! we shall henceforth follow convention and write |%, m$.
In summary, the operators L̂2 and L̂z have a common set of orthonormal
eigenstates |%, m$ with Figure 4.1: The following is a
schematic showing the angular
momentum $ scheme for % = √ 2
L̂2 |%, m$ = %(% + 1)!2 |%, m$, L̂z |%, m$ = m!|%, m$ , (4.4) with L2 = 2(2 + 1)! = 6!
and the five possible values for
the Lz projection.
where %, m are integers or half-integers. The allowed quantum numbers m
form a ladder with step spacing unity, the maximum value of m is %, and
the minimum value is −%. With these results, we may then return to the
normalization of the raising and lowering operators. In particular, making use
of Eq. (4.3), we have
$
L̂+ |%, m$ = $%(% + 1) − m(m + 1)!|l, m + 1$
. (4.5)
L̂− |%, m$ = %(% + 1) − m(m − 1)!|l, m − 1$
Beginning with the eigenstates of L̂z , the eigenvalue equation (4.4), and
making use of the expression above, we have
Since the left hand side depends only on φ, the solution is separable and takes
the form Y!m (θ, φ) = F (θ)eimφ . Note that, since m is integer, the continuity
of the wavefunction, Y!m (θ, φ + 2π) = Y!m (θ, φ), is ensured.
To determine the second component of the eigenstates, F (θ), we could
immediately turn to the eigenvalue equation involving the differential operator
for L̂2 ,
% &
1 m2 2
∂θ (sin θ∂θ ) − ∂ F (θ) = %(% + 1)F (θ) .
sin θ sin2 θ φ
However, to construct the states, it is easier to draw upon the properties of
the angular momentum raising and lowering operators (much in the same way
that the Hermite polynomials are generated by the action of ladder operators
in the harmonic oscillator problem).
Consider then the state of maximal m, |%, %$, for which L̂+ |%, %$ = 0. Mak-
ing use of the coordinate representation of the raising operator above together
with the separability of the wavefunction, this relation implies that
0 = %θ, φ|L̂+ |%, %$ = !eiφ (∂θ + i cot θ∂φ ) Y!! (θ, φ) = !ei(!+1)φ (∂θ − % cot θ) F (θ) .
From this result it follows that ∂θ F (θ) = % cot θF (θ) with the solution F (θ) =
C sin! θ, and C a constant of normalization. States with values of m lower
than % can then be obtained simply by repeated application of the angular
momentum lowering operator L̂− to the state |%, %$. This amounts to the
relation
' (
Y!m (θ, φ) = C(L̂− )!−m sin! θei!φ
' (
= C (−∂θ + i cot θ∂φ )!−m sin! θei!φ .
The eigenfunctions produced by this procedure are well known and referred
to as the spherical harmonics. In particular, one finds that the normalized
eigenstates take the form,
% &1/2
2% + 1 (% − |m|)! |m|
Y!m (θ, φ) = (−1) m+|m|
P! (cos θ)eimφ , (4.7)
4π (% + |m|)!
where
(1 − ξ 2 )m/2 dm+! 2
P!m (ξ) = (ξ − 1)! ,
2! %! dξ m+!
represent the associated Legendre polynomials. In particular, for the first
few angular momentum states, we have
Y00 = √1
)4π )
Y10 = 3
cos θ, Y11 = − 8π e sin θ
3 iφ
) 4π ) )
Y20 = 5
16π (3 cos 2 θ − 1), Y
21 = − 15 iφ
8π e sin θ cos θ, Y22 = 15 2iφ
32π e sin2 θ
Figure 4.2 shows a graphical representation of the states for the lowest spher-
ical harmonics. From the colour coding of the states, the symmetry, Y!,−m =
(−1)m Y!m∗ is manifest.
where the prefactor sin θ derives from the measure. Similarly, we have the
orthogonality condition,
+ π + 2π
∗
dθ sin θ Y!,m (θ, φ)Y!! ,m! (θ, φ) = δ!!! δmm! .
0 0
After this lengthy digression, we may now return to the problem of the
quantum mechanical rotor Hamiltonian and the rigid diatomic molecule. From
the analysis above, we have found that the eigenstates of the Hamiltonian (4.1)
are given by ψ(R, r) = √12π eiK·R Y!,m (θ, φ) with eigenvalues
!2 K2 !2
EK,! = + %(% + 1) ,
2M 2I
Figure 4.2: First four groups of spherical harmonics, Y!m (θ, φ) shown as a function
of spherical angular coordinates. Specifically, the plots show the surface generated by
|Re Y!m (θ, φ)| to fix the radial coordinate and the colours indicate the relative sign of
the real part.
$ Exercise. Using this result, determine the dependence of the heat capacity
of a gas of rigid diatomic molcules on the angular degrees of freedom. How would
this result change if the diatomic gas was constrained to just two spatial dimensions,
i.e. the axis of rotation was always perpendicular to the plane in which the molecules
can move?
p̂2
Ĥ = + V (r) .
2m
In the classical system, L2 = (r × p)2 = r2 p2 − (r · p)2 . As a result, we can
2
set p2 = Lr2 + p2r , where pr ≡ er · p denotes the radial component of the
momentum. In the quantum system, since the space and position coordinates
do not commute, we have (exercise)
we can set
% &
L̂2 2
p̂ = 2 − ! ∂r + ∂r .
2 2 2
r r
Since we already know the eigenstates of L̂2 , we can immediately take ad-
vantage of the separability of the Hamiltonian to find that ψ(r) = R(r)Y!,m (θ, φ),
where the radial part of the wavefunction is set by
% - . &
!2 2 !2
− ∂r + ∂r +
2
%(% + 1) + V (r) R(r) = ER(r) .
2m r 2mr2
Finally, we can further simplify this expression by setting R(r) = u(r)/r,
whereupon we obtain the “one-dimensional” equation
% 2 2 &
! ∂r
− + Veff (r) u(r) = u(r) , (4.9)
2m
! 2
where the effective potential, Veff (r) = 2mr 2 %(% + 1) + V (r), acquires an ad-
ditional component due to the centrifugal component of the force. Here the
equation must be solved subject to the boundary condition u(0) = 0. From
the normalization condition,
+ + ∞
1
d r |ψ(r)| =
3 2
drr2 2 |u(r)|2 = 1 ,
0 r
!2 ∂r2
− u(r) * Eu(r) ,
2m
√
having approximate solutions eκr and e−κr , where !κ = −2mE. (Here,
strictly speaking, m should denote the reduced mass of electron-proton sys-
tem.) The bound states we are looking for, of course, have exponentially
decreasing wavefunctions at large distances.
where (for reasons which will become apparent shortly) we have introduced
Ze2 κ
the dimensionless parameter ν = 4π% 0 2E
. Notice that in transforming from r
to the dimensionless variable ρ, the scaling factor depends on energy, so will
be different for different energy bound states! Consider now the behaviour of
the wavefunction near the origin. The dominant term for sufficiently small ρ
is the centrifugal component, so
%(% + 1)
∂ρ2 u(ρ) * u(ρ) ,
ρ2
for which the solutions are u(ρ) ∼ ρ−! and u(ρ) ∼ ρ!+1 . Since the wavefunction
cannot be singular, we must choose the second solution.
We have established that the wavefunction decays as e−κr = e−ρ at large
distances, and goes as ρ!+1 close to the origin. Factoring out these two asymp-
totic behaviours, let us then define w(ρ) such that u(ρ) = e−ρ ρ!+1 w(ρ). We
leave it as a tedious but straightforward exercise to show that
For large values of k, wk+1 /wk → 2/k, so wk ∼ 2k /k! and therefore w(ρ) ∼ e2ρ .
This means we have found the diverging radial wavefunction, u(ρ) ∼ eρ , which
is in fact the correct behaviour for general values of the energy.
To find the bound states, we must choose energies such that the series is not
an infinite one. As long as the series stops somewhere, the exponential decrease
will eventually take over, and yield a finite (bound state) wavefunction. Just
as for the simple harmonic oscillator, this can only happen if for some k,
wk+1 = 0. Inspecting the ratio wk+1 /wk , evidently the condition for a bound
state is that
ν = n, integer ,
in which case the series for w(ρ) terminates at k = n − % − 1. From now on,
since we know that for the functions we’re interested in ν is an integer, we
replace ν by n.
Finally, making use of the definitions of ν and κ above, we obtain the
bound state energies,
- .2
Ze2 m 1 Z2
En = − ≡ − Ry .
4π#0 2!2 n2 n2
Remarkably, this is the very same series of bound state energies found by
Bohr from his model! Of course, this had better be the case, since the series of
energies Bohr found correctly accounted for the spectral lines emitted by hot
hydrogen atoms. Notice, though, that there are some important differences
with the Bohr model: the energy here is determined entirely by n, called the
principal quantum number, but, in contrast to Bohr’s model, n is not the
angular momentum. The true ground state of the hydrogen atom, n = 1, has
zero angular momentum: since n = k + % + 1, n = 1 means both l = 0 and
k = 0. The ground state wavefunction is therefore spherically symmetric, and
the function w(ρ) = w0 is just a constant. Hence u(ρ) = ρe−ρ w0 and the actual
radial wavefunction is this divided by r, and of course suitably normalized.
To write the wavefunction
$ in terms of r, we need to find κ. Putting together
ρ = κn r, κn = −2mEn /!2 and the expression for En , we find that κn =
Ze2 m
4π%0 !2 n = na0 , where
Z
4π#0 !2
a0 = = 0.529 × 10−10 m
me2
denotes the Bohr radius – the radius of the lowest orbit in Bohr’s model.2
1 (Ze) 1
With this definition, the energy levels can then be expressed as En = − 4π%0 2a0 n
2.
Moving on to the excited states: for n = 2, we have a choice: either
the radial function w(ρ) can have one term, as before, but now the angular
momentum % = 1 (since n = k + % + 1); or w(ρ) can have two terms (so k = 1),
and % = 0. Both options give the same energy, 0.25 Ry, since n is the same,
and the energy only depends on n. In fact, there are four states at this energy,
since % = 1 has states with m = 1, m = 0 and m = 1, and % = 0 has the one
state m = 0. For n = 3, there are 9 states altogether: % = 0 gives one, % = 1
gives 3 and % = 2 gives 5 different m values. In fact, for principal quantum
number n there are n2 degenerate states (n2 being the sum of the first n odd
integers).
From now on, we label the wavefunctions with the quantum numbers,
ψn!m (r, θ, φ), so the ground state is the spherically symmetric ψ100 (r). For
this state R(r) = u(r)/r, where u(ρ) = e−ρ ρ!+1 w(ρ) = e−ρ ρw0 , with w0 a
constant and ρ = κ1 r = Zr/a0 . So, as a function of r, R10 (r) = N e−Zr/a0
with N the normalization constant:
- .3
Z 2 −Zr/a0
R10 = 2 e .
a0
For n = 2, % = 1 the function w(ρ) is still a single term, a constant, but
now u(ρ) = e−ρ ρ!+1 w(ρ) = e−ρ ρ2 w0 , and, for n = 2, ρ = κ2 r = Zr/2a0 ,
remembering the energy-dependence of κ. After normalization, we find
- .3/2 - .
1 Z Zr
R21 = √ e−Zr/2a0 .
2 6 a0 a0
The other n = 2 state has % = 0. So from n = k + % + 1, we have k = 1 and
the series for w has two terms, k = 0 and k = 1, the ratio being
wk+1 2(k + % + 1 − n)
= = −1 ,
wk (k + 1)(k + 2(% + 1))
for the relevant values: k = 0, % = 0, n = 2. So w1 = −w0 , w(ρ) = w0 (1 − ρ).
For n = 2, ρ = r/2a0 , the normalized wavefunction is given by
- .3/2 - .
1 Z 1 Zr
R20 = √ 1− e−Zr/2a0 .
2 a0 2 a0
Note that the zero angular momentum wavefunctions are non-zero and have
non-zero slope at the origin. This means that the full three-dimensional wave-
functions have a slope discontinuity there! But this is fine - the potential is
infinite at the origin. (Actually, the proton is not a point charge, so really the
kink will be smoothed out over a volume of the size of the proton - a very tiny
effect.)
In practice, the first few radial functions w(ρ) can be constructed fairly
easily using the method presented above, but it should be noted that the
differential equation for w(ρ),
where k, p are integers, and Lkp (z) is a Laguerre polynomial. The two
equations are the same if z = 2ρ, and the solution to the radial equation is
therefore,
dp −z p dp 0
L0p (z) = ez (e z ), Lkp (z) = (−1)k L (z) .
dz p dz p p+k
(These representations can be found neatly by solving Laplace’s equation using
- surprise - a Laplace transform.) The polynomials satisfy the orthonormality
relations (with the mathematicians’ normalization convention)
+ ∞
[(p + k)!]3
e−z z k Lkp Lkq dz = δpq .
0 p!
But what do the polynomials look like? The function e−z z p is zero at
the origin (apart from the trivial case p = 0) and zero at infinity, always
positive and having non-zero slope except at its maximum value, z = p. The p
derivatives bring in p separated zeroes, easily checked by sketching the curves
generated by successive differentiation. Therefore, L0p (z), a polynomial of
degree p, has p real positive zeroes, and value at the origin L0p (0) = p!, since
the only non-zero term at z = 0 is that generated by all p differential operators
acting on z p .
The associated Laguerre polynomial Lkp (z) is generated by differentiating
Lp+k (z) k times. Now L0p+k (z) has p + k real positive zeroes, differentiating
0
it gives a polynomial one degree lower, with zeroes which must be one in
each interval between the zeroes of L0p+k (z). This argument remains valid for
successive derivatives, so Lkp (z) must have p real separate zeroes.
Putting all this together, and translating back from ρ to r, the radial
solutions are given by,
- .
Zr ! 2!+1
Rn! (r) = N e−Zr/na0 Ln−!−1 (2Zr/na0 ) ,
na0
with N the normalization constant. For a given principle quantum number
n,the largest % radial wavefunction is given by Rn,n−1 ∝ rn−1 e−Zr/na0 .
$ Info. The eigenvalues of the Hamiltonian for the hydrogen exhibit an unex-
pectedly high degeneracy. The fact that En!m is independent of m is common to all
central potentials – it is just a reflection of rotational invariance of the Hamiltonian.
However, the degeneracy of different % values with the same principle quantum num-
ber is considered “accidental”, a peculiarity of the 1/r potential. To understand the
origin of the degeneracy for atomic hydrogen, it is helpful to reflect first on
the classical dynamics.
In classical mechanics, central forces also lead to conservation of angular momen-
tum leaving orbits restricted to lie in a plane. However, for 1/r type potentials, these
orbits are also closed, i.e. they do not precess. In classical mechanics, this implies
that there is an additional conserved vector, since the direction of the major axis of
the elliptical orbit is a constant of the motion. This direction is determined by the
Runge-Lenz vector,
1 e2 r
R= p×L− .
m 4π#0 r
In quantum theory, up to an operator ordering prescription, R becomes an operator,
1 e2 r
R̂ = (p̂ × L̂ − L̂ × p̂) − .
2m 4π#0 r
With this definition, one may confirm that [Ĥ, R̂] = 0 (exercise).
As a vector operator, R̂ exhibits the following commutation relations, [R̂i , L̂j ] =
i!#ijk R̂k . Similarly, [R̂i , N̂j ] = i! (−2 Ĥ)
m #ijk L̂k (exercise). Moreover, one may confirm
that
- 2 .2
e 2Ĥ 2
R̂2 = + (L̂ + !2 ) ,
4π#0 m
showing that Ĥ can be written in terms of the two constants of motion, L̂2 )
and R̂2 .
Focussing on the bound states, if we consider the Hermitian operator, K̂ = −m 2Ĥ
R̂,
which fulfil the following commutation relations, [K̂i , K̂j ] = i!#ijk L̂k , [K̂i , L̂j ] =
i!#ijk K̂k , and [L̂i , L̂j ] = i!#ijk L̂k , we find that
- 2 .2
e m
Ĥ = − .
4π#0 2(K̂ + L̂2 + !2 )
2
We can simultaneously diagonalize the operators, M̂2 , M̂z , N̂2 and N̂z ,
M̂2 |m, n, µ, ν$ = !2 m(m + 1)|m, n, µ, ν$, M̂z |m, n, µ, ν$ = !µ|m, n, µ, ν$
N̂ |m, n, µ, ν$ = ! m(m + 1)|m, n, µ, ν$,
2 2
N̂z |m, n, µ, ν$ = !ν|m, n, µ, ν$ .
where m, n = 0, 1/2, 1, 3/2, · · ·, µ = −m, −m + 1, · · · m and ν = −n, −n + 1, · · · n.
Since R̂ · L̂ = L̂ · R̂ = 0, then K̂ · L̂ = L̂ · R̂ = 0 and the only relevant states are those
for which M̂2 − N̂2 = 0, i.e. m = n. Therefore,
- 2 .2
e m
Ĥ|m, m, µ, ν$ = − |m, m, µ, ν$
4π#0 2!2 (4m(m + 1) + 1)
- 2 .2
e m
=− |m, m, µ, ν$ .
4π#0 2!2 (2m + 1)2
From this result, we can identify 2m + 1 = 1, 2, · · · as the principle quantum number.
For a given (2m + 1) value, the degeneracy of the state is (2m + 1)2 as expected.
"
the variation implies that, for any i, dt (wi ∂qi L(qi , q̇i ) + ẇi ∂q̇i L(qi , q̇i )) = 0. Then,
integrating the second term by parts, and droping the boundary term, one obtains
! # $
d
dt wi ∂qi L(qi , q̇i ) − ∂q̇i L(qi , q̇i ) = 0 .
dt
Since this equality must follow for any function wi (t), the term in parentheses in the
integrand must vanish leading to the Euler-Lagrange equation (5.1).
The curly brackets are known as Poisson brackets, and are defined for any dynamical
variables as {A, B} = ∂qi A ∂pi B − ∂pi A ∂qi B. From Hamilton’s equations, we have
shown that for any variable, f˙ = {f, H}. It is easy to check that, for the coordinates
and canonical momenta, {qi , qj } = 0 = {pi , pj }, {qi , pj } = δij . This was the
classical mathematical structure that led Dirac to link up classical and quantum
mechanics: He realized that the Poisson brackets were the classical version of the
commutators, so a classical canonical momentum must correspond to the quantum
differential operator in the corresponding coordinate.2
With these foundations revised, we now return to the problem at hand; the
infleunce of an electromagnetic field on the dynamics of the charged particle.
As the Lorentz force is velocity dependent, it can not be expressed simply
as the gradient of some potential. Nevertheless, the classical path traversed by
a charged particle is still specifed by the principle of least action. The electric
and magnetic fields can be written in terms of a scalar and a vector potential
as B = ∇×A, E = −∇φ− Ȧ. The corresponding Lagrangian takes the form:3
1
L = mv2 − qφ + qv · A.
2
1
i.e. forces that conserve mechanical energy.
2
For a detailed discussion, we refer to Paul A. M. Dirac, Lectures on Quantum Mechanics,
Belfer Graduate School of Science Monographs Series Number 2, 1964.
3
In a relativistic formulation, the interaction term
R here looks less arbitrary: the relativistic
version would have the relativistically invariant q Aµ dxµ added to the action integral, where
the four-potential Aµ = (φ, A) and dxµ = (ct, dx1 , dx2 , dx3 ). This is the simplest possible
invariant interaction between theR electromagnetic field and
R the particle’s four-velocity. Then,
in the non-relativistic limit, q Aµ dxµ just becomes q (v · A − φ)dt.
In this case, the general coordinates qi ≡ xi = (x1 , x2 , x3 ) are just the Carte-
sian coordinates specifying the position of the particle, and the q̇i are the three
components ẋi = (ẋ1 , ẋ2 , ẋ3 ) of the particle velocities. The important point is
that the canonical momentum
Reassuringly, the Hamiltonian just has the familiar form of the sum of the
kinetic and potential energy. However, to get Hamilton’s equations of motion,
the Hamiltonian has to be expressed solely in terms of the coordinates and
canonical momenta, i.e.
1
H= (p − qA(r, t))2 + qφ(r, t) .
2m
Let us now consider Hamilton’s equations of motion, ẋi = ∂pi H and
ṗi = −∂xi H. The first equation recovers the expression for the canonical
momentum while second equation yields the Lorentz force law. To under-
stand how, we must first keep in mind that dp/dt is not the acceleration: The
A-dependent term also varies in time, and in a quite complicated way, since
it is the field at a point moving with the particle. More precisely,
& '
ṗi = mẍi + q Ȧi = mẍi + q ∂t Ai + vj ∂xj Ai ,
mẍ = F = q (E + v × B) .
With these preliminary discussions of the classical system in place, we are now
in a position to turn to the quantum mechanics.
Expanding the square on the right hand side of the Hamiltonian, the
cross-term (known as the paramagnetic term) leads to the contribution
q!
− 2im (∇ · A + A · ∇) = iq!
m A · ∇, where the last equality follows from the
Coulomb gauge condition, ∇ · A = 0.5 Combined with the diamagnetic (A2 )
contribution, one obtains the Hamiltonian,
!2 2 iq! q2 2
Ĥ = −
∇ + A·∇+ A + qφ .
2m m 2m
For a constant magnetic field, the vector potential can be written as A =
−r × B/2. In this case, the paramagnetic component takes the form
iq! iq! q
A·∇= (r × ∇) · B = − L · B,
m 2m 2m
where L denotes the angular momentum operator (with the hat not shown for
brevity!). Similarly, the diamagnetic term leads to
q2 2 q2 & 2 2 ' q2B 2 2
A = r B − (r · B)2 = (x + y 2 ) ,
2m 8m 8m
where, here, we have chosen the magnetic field to lie along the z-axis.
However, there are instances when the diamagnetic contriubution can play
an important role. Leaving aside the situation that may prevail on neutron
stars, where magnetic fields as high as 108 T may exist, the diamagnetic con-
tribution can be large when the typical “orbital” scale &x2 + y 2 ' becomes
macroscopic in extent. Such a situation arises when the electrons become
unbound such as, for example, in a metal or a synchrotron. For a further
discussion, see section 5.5 below.
Retaining only the paramagnetic contribution, the Hamiltonian for a “spin-
less” electron moving in a Coulomb potential in the presence of a constant
magnetic field then takes the form,
e
Ĥ = Ĥ0 + BLz ,
2m
2
p̂ 2
where Ĥ0 = 2m e
− 4π# 0r
. Since [Ĥ0 , Lz ] = 0, the eigenstates of the unperturbed
Hamiltonian, ψl$m (r) remain eigenstates of Ĥ and the corresponding energy
levels are specified by
Ry
En$m = − + !ωL m
n2
Sir Joseph Larmor 1857-1942
where ωL = 2m eB
denotes the Larmor frequency. From this result, we expect A physicist and
that a constant magnetic field will lead to a splitting of the (2)+1)-fold degen- mathematician
who made in-
eracy of the energy levels leading to multiplets separated by a constant energy novations in the
shift of !ωL . The fact that this behaviour is not recapitulated generically by understanding
of electricity,
experiment was one of the key insights that led to the identification of electron dynamics, ther-
spin, a matter to which we will turn in chapter 6. modynamics,
and the electron theory of matter.
His most influential work was Aether
and Matter, a theoretical physics
5.4 Gauge invariance and the Aharonov-Bohm ef- book published in 1900. In 1903 he
fect was appointed Lucasian Professor of
Mathematics at Cambridge, a post
he retained until his retirement in
Our derivation above shows that the quantum mechanical Hamiltonian of a 1932.
charged particle is defined in terms of the vector potential, A. Since the latter
is defined only up to some gauge choice, this suggests that the wavefunction
is not a gauge invariant object. Indeed, it is only the observables associated
with the wavefunction which must be gauge invariant. To explore this gauge
freedom, let us consider the influence of the gauge transformation,
where Λ(r, t) denotes a scalar function. Under the gauge transformation, one
may show that the corresponding wavefunction gets transformed as
( q )
ψ " (r, t) = exp i Λ(r, t) ψ(r, t) . (5.3)
!
" Info. One physical manifestation of the gauge invariance of the wavefunction
is found in the Aharonov-Bohm effect. Consider a particle with charge q travelling
Figure 5.1: (Left) Schematic showing the geometry of an experiment to observe the
Aharonov-Bohm effect. Electrons from a coherent source can follow two paths which
encircle a region where the magnetic field is non-zero. (Right) Interference fringes
for electron beams passing near a toroidal magnet from the experiment by Tonomura
and collaborators in 1986. The electron beam passing through the center of the torus
acquires an additional phase, resulting in fringes that are shifted with respect to
those outside the torus, demonstrating the Aharonov-Bohm effect. For details see the
original paper from which this image was borrowed see Tonomura et al., Evidence
for Aharonov-Bohm effect with magnetic field completely shielded from electron wave,
Phys. Rev. Lett. 56, 792 (1986).
Here y0 = −px /qB and ω = |q|B/m coincides with the cyclotron frequency
of the classical charged particle (exercise). We now see that the conserved
canonical momentum px in the x-direction is in fact the coordinate of the centre
of a simple harmonic oscillator potential in the y-direction with frequency ω.
As a result, we can immediately infer that the eigenvalues of the Hamiltonian
are comprised of a free particle component associated with motion parallel to
the field, and a set of harmonic oscillator states,
p2z
En,pz = (n + 1/2)!ω + .
2m
The quantum numbers, n, specify states known as Landau levels.
Let us confine our attention to states corresponding to the lowest oscillator
(Landau level) state, (and, for simplicity, pz = 0), E0 = !ω/2. What is
the degeneracy of this Landau level? Consider a rectangular geometry of
area A = Lx × Ly and, for simplicity, take the boundary conditions to be
periodic. The centre of the oscillator wavefunction, y0 = −px /qB, must lie
9
The superconducting flux quantum was actually predicted prior to Aharonov and Bohm,
by Fritz London in 1948 using a phenomenological theory.
" Exercise. Consider the solution of the Schrödinger equation when working in
the symmetric gauge with A = −r × B/2. Hint: consider the velocity commutation
relations, [vx , vy ] and how these might be deployed as conjugate variables.
" Info. Integer quantum Hall effect: Until now, we have considered the
impact of just a magnetic field. Consider now the Hall effect geometry in which
we apply a crossed electric, E and magnetic field, B. Taking into account both
contributions, the total current flow is given by
# $
j×B
j = σ0 E − ,
ne
where σ0 denotes the conductivity, and n is the electron density. With the electric
field oriented along y, and the magnetic field along z, the latter equation may be
rewritten as
# $# $ # $
1 σ0 B
jx 0
ne = σ0 .
σ0 B
− ne 1 jy Ey
Inverting these equations, one finds that
−σ02 B/ne σ0
jx = Ey , jy = Ey .
1 + (σ0 B/ne)2 1 + (σ0 B/ne)2
1 23 4 1 23 4
σxy σyy
Figure 5.2: (Left) A voltage V drives a current I in the positive x direction. Normal
Ohmic resistance is V /I. A magnetic field in the positive z direction shifts positive
charge carriers in the negative y direction. This generates a Hall potential and a
Hall resistance (V H/I) in the y direction. (Right) The Hall resistance varies stepwise
with changes in magnetic field B. Step height is given by the physical constant
h/e2 (value approximately 25 kΩ) divided by an integer i. The figure shows steps for
i = 2, 3, 4, 5, 6, 8 and 10. The effect has given rise to a new international standard
for resistance. Since 1990 this has been represented by the unit 1 klitzing, defined as
the Hall resistance at the fourth step (h/4e2 ). The lower peaked curve represents the
Ohmic resistance, which disappears at each step.
These provide the classical expressions for the longitudinal and Hall conductivities,
σyy and σxy in the crossed field. Note that, for these classical expressions, σxy is
proportional to B.
How does quantum mechanics revised this picture? For the classical model –
Drude theory, the random elastic scattering of electrons impurities leads to a con-
2
me , where τ
stant drift velocity in the presence of a constant electric field, σ0 = ne τ
denotes the mean time between collisions. Now let us suppose the magnetic field is
chosen so that number of electrons exactly fills all the Landau levels up to some N ,
i.e.
eB
nLx Ly = N νmax ⇒ n = N .
h
The scattering of electrons must lead to a transfer between quantum states. However,
if all states of the same energy are filled,10 elastic (energy conserving) scattering
becomes impossible. Moreover, since the next accessible Landau level energy is a
distance !ω away, at low enough temperatures, inelastic scattering becomes frozen
out. As a result, the scattering time vanishes at special values of the field, i.e. σyy → 0
and
ne e2
σxy → =N .
B h
At critical values of the field, the Hall conductivity is quantized in units of e2 /h.
Inverting the conductivity tensor, one obtains the resistivity tensor,
# $ # $−1
ρxx ρxy σxx σxy
= ,
−ρxy ρxx −σxy σxx
where
σxx σxy
ρxx = 2 + σ2
, ρxx = − 2 + σ2
,
σxx xy σxx xy
So, when σxx = 0 and σxy = νe2 /h, ρxx = 0 and ρxy = h/νe2 . The quantum
Hall state describes dissipationless current flow in which the Hall conductance σxy is
quantized in units of e2 /h. Experimental measurements of these values provides the
best determination of fundamental ratio e2 /h, better than 1 part in 107 .
10
Note that electons are subject to Pauli’s exclusion principle restricting the occupancy of
each state to unity.
Spin
Until we have focussed on the quantum mechanics of particles which are “fea-
tureless”, carrying no internal degrees of freedom. However, a relativistic
formulation of quantum mechanics shows that particles can exhibit an intrin-
sic angular momentum component known as spin. However, the discovery
of the spin degree of freedom marginally predates the development of rela-
tivistic quantum mechanics by Dirac and was acheived in a ground-breaking
experiement by Stern and Gerlach (1922). In their experiment, they passed a
well-collimated beam of silver atoms through a region of inhomogeneous field
before allowing the particles to impact on a photographic plate (see figure).
The magnetic field was directed perpendicular to the beam, and has a strong
gradient, ∂z Bz != 0 so that a beam comprised of atoms with a magnetic mo-
ment would be bent towards the z or -z axis. As the magnetic moment will
be proportional to the total angular momentum, such an experiment can be
thought of as a measurement of its projection along z.
At the time of the experiment, there was an expectation that the magnetic
moment of the atom was generated in its entirety by the orbital angular mo-
mentum. As such, one would expect that there would be a minimum of three
possible values of the z-component of angular momentum: the lowest non-zero
orbital angular momentum is " = 1, with allowed values of the z-component
m!, m = 1, 0, −1. Curiously, Stern and Gerlach’s experiment (right) showed Gerlach’s postcard, dated 8th
that the beam of silver atoms split into two! This discovery, which caused February 1922, to Niels Bohr. It
great discussion and surprise presented a puzzle. shows a photograph of the beam
splitting, with the message, in
However, in our derivation of allowed angular momentum eigenvalues we
translation: “Attached [is] the
found that, although for any system the allowed values of m form a ladder experimental proof of directional
with spacing !, we could not rule out half-integral values of m. The lowest quantization. We congratulate
such case, " = 1/2, would in fact have just two allowed m values: m = ±1/2. [you] on the confirmation of your
However, such an " value could not translate to an orbital angular momentum theory.” (Physics Today Decem-
because the z-component of the orbital wavefunction, ψ has a factor e±iφ , ber 2003)
and therefore acquires a factor −1 on rotating through 2π! This would imply
that ψ is not single-valued, which doesn’t make sense for a Schrödinger-type
wavefunction.
Yet the experimental result was irrefutable. Therefore, this must be a
new kind of non-orbital angular momentum – spin. Conceptually, just as
the Earth has orbital angular momentum in its yearly circle around the sun,
and also spin angular momentum from its daily turning, the electron has an
analogous spin. But this analogy has obvious limitations: the Earth’s spin
is after all made up of material orbiting around the axis through the poles.
The electron spin cannot be imagined as arising from a rotating body, since
orbital angular momenta always come in integral multiples of !. Fortunately,
this lack of a simple quasi-mechanical picture underlying electron spin doesn’t
From the general formulae (4.5) for raising and lowering operators S± =
Sx ± iSy , with s = 1/2, we have S+ |1/2, −1/2# = !|1/2, 1/2#, S− |1/2, 1/2# =
!|1/2, −1/2#, or, in matrix form,
! " ! "
0 1 0 0
Sx + iSy = S+ = ! , Sx − iSy = S− = ! .
0 0 1 0
These matrices are Hermitian, traceless, and obey the relations σi2 = I, σi σj =
−σj σi , and σi σj = iσk for (i, j, k) a cyclic permutation of (1, 2, 3). These
relations can be summarised by the identity,
σi σj = Iδij + i)ijk σk .
!2 2
The total spin S2 = 4 σ = 34 !2 , i.e. s(s + 1)!2 for s = 1/2.
From this expression, we find that α/β = (nx − iny )/(1 − nz ) = e−iφ cot(θ/2)
(exercise). Then, making use of the normalisation, |α|2 + |β|2 = 1, we obtain
(up to an arbitrary phase)
! " ! "
α e−iφ/2 cos(θ/2)
= .
β eiφ/2 sin(θ/2)
Since e−iφ cot(θ/2) can be used to specify any complex number with 0 ≤ θ ≤ π,
0 ≤ φ < 2π, so for any possible spinor, there is an associated direction along
which the spin points up.
* Info. The spin rotation operator: In general, the rotation operator for
rotation through an angle θ about an axis in the direction of the unit vector n̂ is given
by eiθn̂·J/! where J denotes the angular momentum operator. For spin, J = S = 21 !σ,
and the rotation operator takes the form1 eiθn̂·J/! = ei(θ/2)(n̂·σ ) . Expanding the
exponential, and making use of the Pauli matrix identities ((n · σ)2 = I), one can
show that (exercise)
1
Warning: do not confuse θ – the rotation angle - with the spherical polar angle used to
parameterise n̂.
The angle θ between the spin and the field stays constant while the azimuthal
angle around the field increases as φ = φ0 +ω0 t, exactly as in the classical case.
|e|B
The frequency ω0 = gωc , where ωc = 2m e
denotes the cyclotron frequency. For
a magnetic field of 1 T, ωc ( 10 rads/s.
11
* Exercise. For a spin initially pointing along the x-axis, prove that )Sx (t)# =
(!/2) cos(ω0 t).
field (B0 + ω/γ) in the rotating frame. Obviously, if the frame rotates exactly
at the precession frequency, ω = ω 0 = −γB0 , spins pointing in any direction
will remain at rest in that frame – there is no effective field at all.
Suppose we now add a small rotating magnetic field with angular frequency
ω in the xy plane, so the total magnetic field,
The effective magnetic field in the frame rotating with the same frequency ω
as the small added field is then given by
Now, if we tune the angular frequency of the small rotating field so that it
exactly matches the precession frequency in the original static magnetic field,
ω = ω 0 = −γB0 , all the magnetic moment will see in the rotating frame is the
small field in the x-direction! It will therefore precess about the x-direction
at the slow angular speed γB1 . This matching of the small field rotation
frequency with the large field spin precession frequency is the “resonance”.
If the spins are lined up preferentially in the z-direction by the static field,
and the small resonant oscillating field is switched on for a time such that
γB1 t = π/2, the spins will be preferentially in the y-direction in the rotating
frame, so in the lab they will be rotating in the xy plane, and a coil will pick
up an a.c. signal from the induced e.m.f.
where the gyromagnetic ratio of the proton is about +5.6. The magnetic moment is
2.79µN (nuclear magnetons). Different nuclei will have different gyromagnetic ratios
giving more degrees of freedom with which to work. The strong static B field is chosen
to lie in the z direction and the polarization of the oscillating EM wave is chosen so
that the B field points in the x direction. The EM wave has (angular) frequency ω,
! "
gp gp Bz Bx cos(ωt)
Ĥ = µN Bz σz + Bx cos(ωt)σx = µN .
2 2 Bx cos(ωt) −Bz
condition must be satisfied well enough to get a significant transition rate. In NMR,
we observe the transitions back to the lower energy state. These emit EM radiation
at the same frequency and we can detect it after the stronger input pulse ends (or by
more complex methods).
NMR is a powerful tool in chemical analysis because the molecular field adds
to the external B field so that the resonant frequency depends on the molecule as
well as the nucleus. We can learn about molecular fields or just use NMR to see
what molecules are present in a sample. In MRI, we typically concentrate on the one
nucleus like hydrogen. We can put a gradient in Bz so that only a thin slice of the
material has ω tuned to the resonant frequency. Therefore we can excite transitions High resolution MRI scan of a
to the higher energy state in only a slice of the sample. If we vary (in the orthogonal brain!
direction!) the B field during the decay, we can recover 3d images.
For each angular momentum component, the states |j1 , m1 # and |j2 , m2 # where
mi = −ji , · · · ji , provide a basis of states of the total angular momentum
operator, Ĵ2i and the projection Jˆiz . Together, they form a complete basis
which can be used to span the states of the coupled spins,4
These product states are also eigenstates of Jˆz with eigenvalue !(m1 + m2 ),
but not of Ĵ2 .
However, for practical application, we require a basis in which the total angular
momentum operator Ĵ2 is also diagonal. That is, we must find eigenstates
|j, mj , j1 , j2 # of the four mutually commuting operators Ĵ2 , Jˆz , Ĵ21 , and Ĵ22 .
In general, the relation between the two basis can be expressed as
&
|j, mj , j1 , j2 # = |j1 , m1 , j2 , m2 #)j1 , m1 , j2 , m2 |j, mj , j1 , j2 # ,
m1 ,m2
4
Here ⊗ denotes the “direct product” and shows that the two constituent spin states
access their own independent Hilbert space.
1. As a unique entry, the basis state with maximal Jmax and mj = Jmax is
easy to deduce from the original basis states since it involves the product
of states of highest weight,
where Jmax = j1 + j2 .
2. From this state, we can use of the total spin lowering operator Jˆ− to
find all states with J = Jmax and mj = −Jmax · · · Jmax .
3. From the state with J = Jmax and mj = Jmax − 1, one can then obtain
the state with J = Jmax − 1 and mj = Jmax − 1 by orthogonality.6 Now
one can return to the second step of the programme and repeat until
J = |j1 − j2 | when all (2j1 + 1)(2j2 + 1) basis states have been obtained.
|S = 1, mS = 1, s1 = 1/2, s2 = 1/2# = |s1 = 1/2, ms1 = 1/2# ⊗ |s2 = 1/2, ms2 = 1/2# .
Now, since s1 = 1/2 and s = 1/2 is implicit, we can rewrite this equation in a
more colloquial form as
|S = 1, mS = 1# = | ↑1 # ⊗ | ↑2 # .
We now follow step 2 of the programme and subject the maximal spin state
to the total spin lowering operator, Ŝ− = Ŝ1− + Ŝ1+ . In doing so, making use
of Eq. (4.5), we find
√
Ŝ− |S = 1, mS = 1# = 2!|S = 1, mS = 0# = ! (| ↓1 # ⊗ | ↑2 # + | ↑1 # ⊗ | ↓2 #) ,
5
In fact, one may show that the general matrix element is given by
s
(j1 + j2 − j)!(j + j1 − j2 )!(j + j2 − j1 )!(2j + 1)
$j1 , m1 , j2 , m2 |j, mj , j1 , j2 % = δmj ,m1 +m2
(j + j1 + j2 + 1)!
p
X (−1)k (j1 + m1 )!(j1 − m1 )!(j2 + m2 )!(j2 − m2 )!(j + m)!(j − m)!
× .
k!(j1 + j2 − j − k)!(j1 − m1 − k)!(j2 + m2 − k)!(j − j2 + m1 + k)!(j − j1 − m2 + k)!
k
6
Alternatively, as a maximal spin state, |J = Jmax − 1, mj = Jmax − 1, j1 , j2 % can be
identified by the “killing” action of the raising operator, Jˆ+ .
i.e. |S = 1, mS = 0# = √1 (|
2
↓1 # ⊗ | ↑2 # + | ↑1 # ⊗ | ↓2 #). Similarly,
√ √
Ŝ− |S = 1, mS = 0# = 2!|S = 1, mS = −1# = 2!| ↓1 # ⊗ | ↓2 # ,
2L̂ · Ŝ
' () *
Ĵ = L̂ + Ŝ + 2L̂z Ŝz + L̂+ Ŝ− + Ŝ+ L̂− .
2 2 2
For the eigenstates of Ĵ2 , Jˆz , L̂2 and Ŝ2 we will adopt the notation |j, mj , "#
leaving the spin S = 1/2 implicit. The maximal spin state is given by7
To obtain the remaining states in the multiplet, |j = " + 1/2, mj=%+1/2 , "#, we
may simply apply the total spin lowering operator Jˆ− ,
Normalising the right-hand side of this expression, one obtains the spin state,
+ +
2" 1
|" + 1/2, " − 1/2, "# = |", " − 1# ⊗ | ↑# + |", "# ⊗ | ↓# .
2" + 1 2" + 1
7
The proof runs as follows:
and
1
Ĵ2 |$, $% ⊗ | ↑% = !2 ($($ + 1) + 1/2(1/2 + 1) + 2$ )|$, $% ⊗ | ↑%
2
= !2 ($ + 1/2)($ + 3/2)|$, $% ⊗ | ↑% .
By repeating this programme, one can develop an expression for the full
set of basis states,
+
" + mj + 1/2
|j = " + 1/2, mj , "# = |", mj − 1/2# ⊗ | ↑#
2" + 1
+
" − mj + 1/2
+ |", mj + 1/2# ⊗ | ↓# ,
2" + 1
with mj = " + 1/2, · · · , −(" + 1/2). In order to obtain the remaining states
with j = " − 1/2, we may look for states with mj = " − 1/2, · · · , −(" − 1/2)
which are orthogonal to |" + 1/2, mj , "#. Doing so, we obtain
+
" − mj + 1/2
|" − 1/2, mj , "# = − |", mj − 1/2# ⊗ | ↑#
2" + 1
+
" + mj + 1/2
+ |", mj + 1/2# ⊗ | ↓# .
2" + 1
Finally, these states can be cast in a compact form by setting
|J = 2, mJ = 2, j1 = 1, j2 = 1# = |j1 = 1, m1 = 1# ⊗ |j2 = 1, m2 = 1# ,
or, more concisely, |2, 2# = |1# ⊗ |1#, where we leave j1 and j2 implicit. Once
again, making use of Eq. (4.5) and an ecomony of notation, we find (exercise)
|2, 2# = |1# ⊗ |1#
|2, 1# = 2 (|0# ⊗ |1# + |1# ⊗ |0#)
√1
Then, from the expression for |2, 1#, we can construct the next maximal spin
state |1, 1# = √12 (|0# ⊗ |1# − |1# ⊗ |0#), from the orthogonality condition. Once
again, acting on this state with the total spin lowering operator, we obtain the
remaining members of the multiplet,
|1, 1# = 2 (|0# ⊗ |1# − |1# ⊗ |0#)
√1
Finally, finding the state orthogonal to |1, 0# and |2, 0#, we obtain the final
state,
1
|0, 0# = √ (| − 1# ⊗ |1# − |0# ⊗ |0# + |1# ⊗ | − 1#) .
3
1
Indeed, even if such a solution is formally accessible, its complexity may render it of no
practical benefit.
In the following we will address the question of how the eigenstates and
eigenenergies are modified by the imposition of a small perturbation, Ĥ (1)
(such as that imposed by an external electric or magnetic field on a charged
particle, or the deformation of some other external potential). In short, we
are interested in the solution of the Schrödinger equation,
However, note that this is not always the case. For example, as mentioned
above, an infinitesimal perturbation has the capacity to develop a bound state
not present in the unperturbed system. For now, let us proceed with the per-
turbative expansion and return later to discuss its potential range of validity.
The basic assumption that underpins the perturbation theory is that, for
Ĥ (1) small, the leading corrections are of the same order of magnitude as
Ĥ (1) itself. The perturbed eigenenergies and eigenvalues can then be ob-
tained to a greater accuracy by a successive series of corrections, each of order
"Ĥ (1) !/"Ĥ (0) ! compared with the previous. To identify terms of the same
order in "Ĥ (1) !/"Ĥ (0) !, it is convenient to extract from Ĥ (1) a dimensionless
parameter λ, characterising the relative magnitude of the perturbation against
Ĥ (0) , and then expand |n! and En as a power series in λ, i.e.
∞
!
|n! = |n(0) ! + λ|n(1) ! + λ2 |n(2) ! + · · · = λm |n(m) !,
m=0
∞
!
En = En(0) + λEn(1) + λ2 En(2) + · · · = λm En(m) .
m=0
From this equation, we must relate terms of equal order in λ. At the lowest
order, O(λ0 ), we simply recover the unperturbed equation (7.1). In practical
applications, one is usually interested in determining the first non-zero per-
turbative correction. In the following, we will explore the form of the first and
second order perturbative corrections.
and taking the inner product with the unperturbed states "n(0) |, one obtains
"n(0) |Ĥ (0) |n(1) ! + "n(0) |Ĥ (1) |n(0) ! = "n(0) |En(0) |n(1) ! + "n(0) |En(1) |n(0) ! .
(0)
Noting that "n(0) |Ĥ (0) = "n(0) |En , and exploiting the presumed normaliza-
tion "n(0) |n(0) ! = 1, one finds that the first order shift in energy is given
simply by the expectation value of the perturbation taken with respect to the
unperturbed eigenfunctions,
"m(0) |Ĥ (0) |n(1) ! + "m(0) |Ĥ (1) |n(0) ! = "m(0) |En(0) |n(1) ! + "m(0) |En(1) |n(0) ! .
(0)
Once again, with "m(0) |Ĥ (0) = "m(0) |Em and the orthogonality condition on
the wavefunctions, "m(0) |n(0) ! = 0, one obtains an expression for the first order
shift of the wavefunction expressed in the unperturbed basis,
Before turning to the second order of perturbation theory, let us first consider
a simple application of the method.
! Example: Ground state energy of the Helium atom: For the Helium
atom, two electrons are bound to a nucleus of two protons and two neutrons. If
one neglects altogether the Coulomb interaction between the electrons, in the ground
state, both electrons would occupy the ground state hydrogenic wavefunction (scaled
appropriately to accommodate the doubling of the nuclear charge) and have opposite
spin. Treating the Coulomb interaction between electrons as a perturbation, one may
then use the basis above to estimate the shift in the ground state energy with
1 e2
Ĥ (1) = .
4π$0 |r1 − r2 |
As we have seen, the hydrogenic wave functions are specified by three quantum
numbers, n, %, and m. In the ground state, the corresponding wavefunction takes the
spatially isotropic form,
" #1/2
1
"r|n = 1, % = 0, m = 0! = ψ100 (r) = e−r/a , (7.7)
πa3
2
0 !
where a = 4π"
Ze2 me = Z denotes the atomic Bohr radius for a nuclear charge Z.
a0
For the Helium atom (Z = 2), the symmetrized ground state of the unperturbed
Hamiltonian is then given by the spin singlet (S = 0) electron wavefunction,
1
|g.s.(0) ! = √ (|100, ↑! ⊗ |100, ↓! − |100, ↓! ⊗ |100, ↑!) .
2
Here we have used the direct product ⊗ to discriminate between the two electrons.
Then, applying the perturbation theory formula above (7.5), to first order in the
Coulomb interaction, the energy shift is given by
$
e2 1 e−2(r1 +r2 )/a e2 C0
En(1) = "g.s.(0) |Ĥ (1) |g.s.(0) ! = dr1 dr2 = ,
4π$0 (πa )3 2 |r1 − r2 | 4π$0 2a
% −(z1 +z2 )
where we have defined the dimensionless constant C0 = (4π) 1
2 dz1 dz1 e|z1 −z2 | . Then,
making use of the identity,
$
1 1 1
dΩ1 dΩ2 = ,
(4π) 2 |z1 − z2 | max(z1 , z2 )
where the integrations runs over the angular coordinates of the vectors z1 and z2 ,
%∞ %∞
and z1,2 = |z1,2 |, one finds that C0 = 2 0 dz1 z12 e−z1 z1 dz2 z2 e−z2 = 5/4. As a
2
result, noting that the Rydberg energy, Ry = 4π"e 1
0 2a0
, we obtain the first order
energy shift ∆E = 4 ZRy ( 34eV for Z = 2. This leads to a total ground state
5
obtain
! "m(0) |Ĥ (1) |n(0) !
En(2) = "n(0) |Ĥ (1) |n(1) ! = "n(0) |Ĥ (1) |m(0) ! (0) (0)
,
m"=n En − Em
i.e.
! |"m(0) |Ĥ (1) |n(0) !|2
En(2) = (0) (0)
. (7.8)
m"=n En − Em
where |n%m! denote the set of bound state hydrogenic wavefunctions. Although the
expression for E (2) can be computed exactly, the programme is somewhat tedious.
However, we can place a strong bound on the energy shift through the following
(0) (0) (0) (0)
argument: Since, for n > 2, |E1 − En | > |E1 − E2 |, we have
1 !
|E (2) | < (0) (0)
"100|eEz|n%m!"n%m|eEz|100! .
E2 − E1 n#=1,#,m
& &
Since n,#,m |n%m!"n%m| = I, we have n#=1,#,m |n%m!"n%m| = I−|100!"100|. Finally,
since "100|z|100! = 0, we can conclude that |E (2) | < (0) 1 (0) "100|(eEz)2 |100!. With
E2 −E1
(0) 2 (0) (0)
"100|z 2 |100! = a20 , E1 = − 4π"
e 1
0 2a0
= −Ry, and E2 = E1 /4, we have
1 8
|E (2) | < (eE)2 a20 = 4π$0 E 2 a30 .
3 2
4 e /8π$ a
0 0 3
Furthermore, since all terms in the perturbation series for E (2) are negative, the first
2
term in the series sets a lower bound, |E (2) | > |$210|eEz|100%|
(0) (0) . From this result, one
E2 −E1
can show that 0.55 × 83 4π$0 E 2 a30 < |E (2) | < 83 4π$0 E 2 a30 (exercise).3
the original circles of constant potential become ellipses, with their axes aligned
along x = ±y.
As soon as the perturbation is introduced, the eigenstates lie in the direc-
tion of the new elliptic axes. This is a large change from the original x and
y bases, which is not proportional to the small parameter α. But the origi-
nal unperturbed problem had circular symmetry, and there was no particular
reason to choose the x and y axes as we did. If we had instead chosen as
our original axes the lines x = ±y, the basis states would not have undergone
large changes on switching on the perturbation. The resolution of the problem
is now clear: Before switching on the perturbation, one must choose a set of
basis states in a degenerate subspace in which the perturbation is diagonal.
In fact, for the simple harmonic oscillator example above, the problem can
of course be solved exactly √ by rearranging the coordinates to lie along the
symmetry axes, (x ± y)/ 2. It is then clear that, despite the results of naı̈ve
first order perturbation
√ theory, there is indeed a first order shift in the energy
levels, !ω → !ω 1 ± α ≈ !ω(1 ± α/2).
When perturbing this system with an electric field oriented in the z-direction, Ĥ (1) =
qEr cos θ, a naı̈ve application of perturbation theory predicts no first-order shift in
any of these energy levels. However, at second order in E, there is a non-zero matrix
element between two degenerate levels ∆ = "200|Ĥ (1) |210!. All the other matrix
elements between these basis states in the four-dimensional degenerate subspace are
zero. So the only diagonalization necessary is within the two-dimensional degenerate
subspace spanned by |200! and |210!, i.e.
" #
0 ∆
Ĥ =
(1)
,
∆ 0
+ ,% + ,+ ,2
with ∆ = qE 1
32πa30
2− r
a0
r cos θ
a0 e−r/a0 r2 dr sin θ dθ dφ = −3qEa0 .
Diagonalizing Ĥ (1) within this sub-space, the new basis states are given by the
√
symmetric and antisymmetric combinations, (|200! ±| 210!)/ 2 with energy shifts
±∆, linear in the perturbing electric field. The states |2%, ±1! are not changed by
the presence of the field to this level of approximation, so the complete energy map
of the n = 2 states in the electric field has two states at the original energy of −Ry/4,
one state moved up from that energy by ∆, and one down by ∆. Notice that the new
√
eigenstates (|200! ±| 210!)/ 2 are not eigenstates of the parity operator -- a sketch
of their wavefunctions reveals that, in fact, they have non-vanishing electric dipole
moment µ. Indeed this is the reason for the energy shift, ±∆ = ∓2eEa0 = ∓µ · E.
bation. In the following, we will suppose that the total one-dimensional system is of
length L = N a, with periodic boundary conditions.
For the unperturbed free particle system, the eigenstates are simply plane waves
ψk (x) = "x|k! = √1L eikx indexed by the wavenumber k = 2πn/L, n integer, and
(0)
the unperturbed spectrum is given by Ek = !2 k 2 /2m. The matrix elements of the
perturbation between states of different wavevector are given by
$
1 L "
"k|V |k ! =
'
dxei(k −k)x 2V cos(2πx/a)
L 0
$ L+ ,
V " "
= ei(k −k+2π/a)x + ei(k −k−2π/a)x = V δk" −k,±2π/a .
L 0
Note that all diagonal matrix elements of the perturbation are identically zero. In
general, for wavevectors k and k ' separated by G = 2π/a, the unperturbed states are
non-degenerate. For these states one can compute the relative energy shift within
the framework of second order perturbation theory. However, for states k = −k ' =
G/2 ≡ π/a, the unperturbed free particle spectrum is degenerate. Here, and in
the neighbourhood of these k values, we must implement a degenerate perturbation
theory.
For the sinusoidal potential considered here, only states separated by G = 2π/a
are coupled by the perturbation. We may therefore consider matrix elements of the
full Hamiltonian between pairs of coupled states, |k = G/2 + q! and |k = −G/2 + q!
3 (0) 4
EG/2+q V
H= (0) .
V E−G/2+q
In particular, this result shows that, for k = ±G/2, the degeneracy of the free particle
system is lifted by the potential. In the vicinity, |q| # G, the spectrum of eigenvalues
is separated by a gap of size 2V . The appearance of the gap mirrors the behaviour
found in our study of the Kronig-Penney model of a crystal studied in section 2.2.3.
The appearance of the gap has important consequences in theory of solids. Elec-
trons are fermions and have to obey Pauli’s exclusion principle. In a metal, at low
temperatures, electrons occupy the free particle-like states up to some (Fermi) energy
which lies away from gap. Here, the accessibility of very low-energy excitations due
to the continuum of nearby states allows current to flow when a small electric field is
applied. However, when the Fermi energy lies in the gap created by the lattice po-
tential, an electric field is unable to create excitations and induce current flow. Such
systems are described as (band) insulators.
! Exercise. Using the Gaussian trial state, find the optimal value of the varia-
tional state energy, E, for an attractive δ-function potential and compare it with the
exact result.
To gain some further insight into the approach, suppose the Hamiltonian
Ĥ has a set of eigenstates, Ĥ|n! = En |n!. Since the Hamiltonian is Her-
mitian, these states span the space of possible wave functions,
& including our
variational family of Gaussians, so we can write, |ψ(α)! = n an (α)|n!. From
this expansion, we have
"ψ(α)|Ĥ|ψ(α)! !
= |an |2 En ≥ E0 ,
"ψ(α)|ψ(α)! n
for any |ψ(α)!. (We don’t need the denominator if we’ve chosen a family of
normalized wavefunctions, as we did with the Gaussians above.) Evidently,
! Example: To get some idea of of how well the variational approach works,
consider its application to the to the ground state of the hydrogen atom. Taking
into account the spherical symmetry of the ground state, we may focus on the one-
dimensional radial component of the wavefunction. Defining the trial radial wave-
function u(ρ) (presumed real), where ρ = r/a0 , the variational energy is given by
%∞ + 2 ,
0
dρ u(ρ) d
dρ2 + 2
ρ u(ρ)
E(u) = −Ry %∞ .
0
dρ u (ρ)
2
This could have been inferred from the formula for the one electron ion, where the
potential energy for the one electron is −2Z 2 Ry, one factor of Z being from the
nuclear charge, the other from the consequent shrinking of the orbit. The kinetic
energy is even easier to determine: it depends entirely on the form of the wavefunction,
and not on the actual nuclear charge. So for our trial wavefunction it has to be Z 2 Ry
per electron. Finally, making use of our calculation on page 64, we can immediately Léon Nicolas Brillouin 1889-
write down the positive contribution to the energy expectation value from the electron- 1969
electron interaction, A French physi-
cis, his father,
$ Marcel Brillouin,
e2 Z3 e−2Z(r1 +r2 )/a0 5 e2 Z 5 grandfather,
dr1 dr2 = = Z Ry .
4π$0 (πa0 )
3 2 |r1 − r2 | 4 4π$0 2a0 4 Éleuthère Mas-
cart, and great-
Collecting all of the terms, the total variational state energy is given by: grandfather,
Charles Briot,
" # were physicists
5
E = −2 4Z − Z − Z Ry .
2 as well. He made
8 contributions to quantum mechanics,
radio wave propagation in the
Minimization of this energy with respect to Z obtains the minimum at Z = 2 − 16 5
, atmosphere, solid state physics, and
information theory.
leading to an energy of 77.5 eV. This result departs from the true value by about 1 eV.
So, indeed, the presence of the other electron leads effectively to a shielding of the Hendrik Anthony “Hans”
nuclear charge by an amount of ca. (5/16)e. Kramers 1894-1952
A Dutch
physicist who
This completes our discussion of the principles of the variational approach. conducted early
and important
However, later in the course, we will find the variational methods appearing in work in quantum
several important applications. Finally, to close this section on approximation theory and
methods for stationary states, we turn now to consider a framework which electromagnetic
dispersion rela-
makes explicit the connection between the quantum and classical theory in tions, solid-state
the limit ! → 0. physics, and
statistical mechanics. He was a
long-time assistant and friend to
Niels Bohr, and collaborated with
7.4 Wentzel, Kramers and Brillouin (WKB) method him on a 1924 paper contending
that light consists of probability
waves, which became a foundation of
The WKB (or Wentzel, Kramers and Brillouin) approximation describes a quantum mechanics. He introduced
“quasi-classical” method for solving the one-dimensional time-independent the idea of renormalization, a cor-
Schrödinger equation. Note that the consideration of one-dimensional systems nerstone of modern field theory, and
determined the dispersion formulae
is less restrictive that it may sound as many symmetrical higher-dimensional that led to Werner Heisenberg’s
problems are rendered effectively one-dimensional (e.g. the radial equation for matrix mechanics. He is not as well
known as some of his contemporaries
the hydrogen atom). The WKB method is named after physicists Wentzel, (primarily because his work was not
Kramers and Brillouin, who all developed the approach independently in widely translated into English), but
his name is still invoked by physicists
1926.4 Earlier, in 1923, the mathematician Harold Jeffreys had developed as they discuss Kramers dispersion
a general method of approximating the general class of linear, second-order theory, Kramers-Heisenberg dis-
persion formulae, Kramers opacity
4 formula, Kramers degeneracy, or
L. Brillouin, (1926). “La mcanique ondulatoire de Schrödinger: une mt́hode générale
de resolution par approximations successives”, Comptes Rendus de l’Academie des Sciences Kramers-Kronig relations.
183: 2426; H. A. Kramers, (1926). “Wellenmechanik und halbzählige Quantisierung”, Z.
Phys. 39: 828840; G. Wentzel (1926). “Eine Verallgemeinerung der Quantenbedingungen
für die Zwecke der Wellenmechanik”. Z. Phys. 38: 518529.
!2 2
− ∂ ψ(x) + V (x)ψ(x) = Eψ(x) ,
2m x
will take the form A(x)e±ip(x)x/! where p(x) is the “local” value of the momen-
tum set by the classical value, p2 /2m + V (x) = E, and the amplitude, A(x), is
slowly-varying compared with the phase factor. Clearly this is a semi-classical
limit: ! has to be sufficiently small that there are many oscillations in the typ-
ical distance over which the potential varies.6
To develop this idea a little more rigorously, and to emphasize the rapid
phase variation in the semi-classical limit, we can parameterize the wavefunc-
tion as
ψ(x) = eiσ(x)/! .
5
H. Jeffreys, (1924). “On certain approximate solutions of linear differential equations of
the second order”, Proc. Lon. Math. Soc. 23: 428436.
6
To avoid any point of confusion, it is of course true that ! is a fundamental constant – not
easily adjusted! So what do we mean when we say that the semi-classical limit translates to
! → 0? The validity of the semi-classical approximation relies upon λ/L $ 1. Following the
de Broglie relation, we may write this inequality as h/pL $ 1, where p denotes the particle
momentum. Now, in this correspondence, both p and L can be considered as “classical”
scales. So, formally, we can think of think of accessing the semi-classical limit by adjusting
! so that it is small enough to fulfil this inequality. Alternatively, at fixed !, we can access
the semi-classical regime by reaching to higher and higher energy scales (larger and larger
p) so that the inequality becomes valid.
Here the complex function σ(x) encompasses both the amplitude and phase.
Then, with −!2 ∂x2 ψ(x) = −i!eiσ(x)/!∂x2 σ(x) + eiσ(x)/!(∂x σ)2 , the Schrödinger
equation may be rewritten in terms of the phase function as,
σ = σ0 + (!/i)σ1 + (!/i)2 σ2 + · · · .
At the leading (zeroth) order of the expansion, we can drop the first term
in'(7.9), leading to (∂x σ0 )2 = p2 (x). Fixing the sign of p(x) by p(x) =
+ 2m(E − V (x)), we conclude that
$
σ0 (x) = ± p(x)dx .
For free particle systems – those for which the kinetic energy is proportional
to p2 – this expression coincides with the classical action.
From the form of the Schrödinger equation (7.9), it is evident that this ap-
proximate solution is only valid if we can ignore the first term. More precisely,
we must have
5 5
5 !∂x2 σ(x) 5
5 5
5 (∂x σ(x))2 5 ≡ |∂x (!/∂x σ)| # 1 .
∂x2 σ0 ∂x p 1
∂x σ1 = − =− , σ1 (x) = − ln p(x) .
2∂x σ0 2p 2
So, to this order of approximation, the wavefunction takes the form,
C1 (i/!) R p dx C2 −(i/!) R p dx
ψ(x) = ' e +' e , (7.10)
p(x) p(x)
p2 (x)
= E − V (x) < 0 .
2m
Here p(x) is of course pure imaginary, but the same formal phase solution of the
Schrödinger equation applies provided, again, that the particle is remote from
the classical turning points where E = V (x). In this region, the wavefunction
takes the general form,
C1% R C% R
ψ(x) = ' e−(1/!) |p| dx
+ ' 2 e(1/!) |p| dx . (7.11)
|p(x)| |p(x)|
This completes our study of the wavefunction in the regions in which the
semi-classical approach can be formally justified. However, to make use of
this approximation, we have to understand how to deal with the regions close
to the classical turning points. Remember in our treatment of the Schrödinger
equation, energy quantization derived from the implementation of boundary
conditions.
E − V (x) ( F0 (x − a) ,
where F0 denotes the (constant) force. For a strictly linear potential, the
wavefunction can be determined analytically, and takes the form of an Airy
function.7 In particular, it is known that the Airy function to the right of the
classical turning point has the asymptotic form
C Rx
lim ψ(x) = ' e−(1/!) a |p| dx ,
x&a 2 |p(x)|
7
For a detailed discussion in the present context, we refer to the text by Landau and
Lifshitz.
translating to a decay into the classically forbidden region while, to the left,
it has the asymptotic oscillatory solution,
6 $ a 7 6 $ 7
C 1 π C π 1 a
lim ψ(x) = ' cos p dx − ≡' cos − p dx .
b'x<a |p(x)| ! x 4 |p(x)| 4 ! x
At the second classical turning point at x = b, the same argument gives
6 $ x 7
C% 1 π
lim ψ(x) = ' cos p dx − .
b<x'a |p(x)| ! b 4
For these two expressions to be consistent, we must have C % = ±C and
" $ x # " $ #
1 π π 1 a
p dx − − − p dx = nπ ,
! b 4 4 ! x
where, for n% even, C % = C and for n odd, C % = −C. Therefore, we have the
a
condition !1 b p dx = (n + 1/2)π, or when cast in terms of a complete periodic
cycle of the classical motion,
8
p dx = 2π!(n + 1/2) .
p2
! Example: For the quantum harmonic oscillator, H = 2m + 12 mω 2 x2 = E,
the classical momentum is given by
: " #
mω 2 x2
p(x) = 2m E − .
2
The classical turning points are set by E = mω 2 x20 /2, i.e. x0 = ±2E/mω 2 . Over a
periodic cycle, the classical action is given by The WKB wavefunction (solid)
8 $ x0 : " # and the exact wavefunction
mω 2 x2 E (dashed) for the n = 0 and
p(x)dx = 2 dx 2m E − = 2π .
−x0 2 ω n = 10 states of the quantum
harmonic oscillator.
According to the WKB method, the latter must be equated to 2π!(n + 1/2), with
the last term reflecting the two turning points. As a result, we find that the energy
levels are as expected specified by En = (n + 1/2)!ω.
In the WKB approximation, the corresponding wavefunctions are given by
" $ x #
C 1 π
ψ(x) = ' cos p(x)dx −
p(x) ! −x0 4
" $ x #
C 2π 1 π
=' cos (n + 1/2) + p(x)dx −
p(x) 4 ! 0 4
3 - " # : .4
C nπ E x x x2
=' cos + arcsin + 1− 2 ,
p(x) 2 !ω x0 x0 x0
for x > a. Note that the failure of the WKB approximation is reflected in the
appearance of discontinuities in the wavefunction at the classical turning points (see
figures). Nevertheless, the wavefunction at high energies provide a strikingly good
approximation to the exact wavefunctions.
Then, applying the continuity condition on the wavefunction and its derivative at the
classical turning points, one obtains the transmissivity,
- $ .
2 b
T (E) ( exp − |p| dx .
! a
! Info. For a particle strictly confined to one dimension, the connection formulae
can be understood within a simple picture: The wavefunction “spills over” into the
classically forbidden region, and its twisting there collects an π/4 of phase change.
So, in the lowest state, the total phase change in the classically allowed region need
only be π/2. For the radial equation, assuming that the potential is well behaved
at the origin, the wavefunction goes to zero there. A bound state will still spill over
beyond the classical turning point at r0 , say, but clearly there must be a total phase
change of 3π/4 in the allowed region for the lowest state, since there can be no spill
over to negative r. In this case, the general quantization formula will be
$
1 r0
p(r) dr = (n + 3/4)π, n = 0, 1, 2, · · · ,
! 0
with the series terminating if and when the turning point reaches infinity. In fact,
some potentials, including the Coulomb potential and the centrifugal barrier for % '= 0,
are in fact singular at r = 0. These cases require special treatment.
Identical Particles
Until now, most of our focus has been on the quantum mechanical behaviour
of individual particles, or problems which can be “factorized” into independent
single-particle systems.1 However, most physical systems of interest involve
the interaction of large numbers of particles; electrons in a solid, atoms in a gas,
etc. In classical mechanics, particles are always distinguishable in the sense
that, at least formally, their “trajectories” through phase space can be traced
and their identity disclosed. However, in quantum mechanics, the intrinsic
uncertainty in position, embodied in Heisenberg’s principle, demands a careful
and separate consideration of distinguishable and indistinguishable particles.
In the present section, we will consider how to formulate the wavefunction of
many-particle systems, and adress some of the (sometimes striking and often
counter-intuitive) implications of particle indistinguishability.
1
|χ11 % = | ↑1 ↑2 %, |χ10 % = √ (| ↑1 ↓2 % + | ↓1 ↑2 %) , |χ1,−1 % = | ↓1 ↓2 % .
2
Here the first arrow in the ket refers to the spin of particle 1, the second to
particle 2.
# Exercise. By way of revision, it is helpful to recapitulate the discussion of the Hint: begin by proving that, for
addition of spin s = 1/2 angular momenta. By setting S = S1 + S2 ,3 where S1 and S2 two spin s = 1/2 degree of free-
are two spin 1/2 degrees of freedom, construct the matrix elements of the total spin dom, S2 = S21 + S22 + 2S1 · S2 =
operator S2 for the four basis states, | ↑↑%, | ↑↓%, | ↓↑%, and | ↓↓%. From the matrix 2 × s(s + 1)!2 + 2S1z S2z + S1+ S2− +
S1− S2+ .
representation of S2 , determine the four eigenstates. Show that one corresponds to a
total spin zero state and three correspond to spin 1.
where χ(s1 , s2 ) = )s1 , s2 |χ%. For two electron degrees of freedom, the to-
tal wavefunction, Ψ, must be antisymmetric under exchange. It follows that
a pair of electrons in the spin singlet state must have a symmetric spatial
wavefunction, ψ(r1 , r2 ), whereas electrons in the spin triplet states must have
an antisymmetric spatial wavefunction. Before discussing the physical conse-
quences of this symmetry, let us mention how this scheme generalizes to more
particles.
3
Here, for simplicity, we have chosen not to include hats on the spin angular momentum
operators.
symmetric in 2 and 3. But to recover a total wave function with overall antisymmetry
it is necessary to add more terms:
# Exercise. A hydrogen atom consists of two fermions, the proton and the
electron. By considering the wavefunction of two non-interacting hydrogen atoms
under exchange, show that the atom transforms as a boson. In general, if the number
of fermions in a composite particle is odd, then it is a fermion, while if even it is
a boson. Quarks are fermions: baryons consist of three quarks and so translate to
fermions while mesons consist of two quarks and translate to bosons.
are the usual kinetic energy term, and the rotational energy of the molecule.
This is where the overall (space × spin) antisymmetric wavefunction for the
protons plays a role. Recall that the parity of a state with rotational angular
momentum $ is (−1)! . Therefore, parahydrogen, with an antisymmetric pro-
ton spin wavefunction, must have a symmetric proton spatial wavefunction,
and so can only have even values of the rotational angular momentum. Or-
thohydrogen can only have odd values. The energy of the rotational level with
angular momentum $ is E!tot = !2 $($ + 1)/I, where I denotes the moment of
inertia of the molecule. So the two kinds of hydrogen gas have different sets
of rotational energy levels, and consequently different specific heats.
# Example: As a final example, and one that will feed into our discussion of
multielectron atoms in the next chapter, let us consider the implications of particle
statistics for the excited state spectrum of Helium. After Hydrogen, Helium is
the simplest atom having two protons and two neutrons in the nucleus (Z = 2), and
two bound electrons. As a complex many-body system, we have seen already that the
Schrödinger equation is analytically intractable and must be treated perturbatively.
Previously, in chapter 7, we have used the ground state properties of Helium as a
vehicle to practice perturbation theory. In the absence of direct electron-electron
interaction, the Hamiltonian
2 # 2
" $
p̂ 1 Ze2
Ĥ0 = n
+ V (rn ) , V (r) = − ,
n=1
2m 4π&0 r
is separable and the wavefunction can be expressed through the states of the hydrogen
atom, ψn!m . In this approximation, the ground state wavefunction involves both
electrons occupying the 1s state leading to an antisymmetric spin singlet wavefunction
for the spin degrees of freedom, |Ψg.s. % = (|100% ⊗ |100%) ⊗ |χ00 %. In chapter 7, we
made use of both the perturbative series expansion and the variational method to
determine how the ground state energy is perturbed by the repulsive electron-electron
interaction,
1 e2
Ĥ1 = .
4π&0 |r1 − r2 |
Now let us consider the implications of particle statistics on the spectrum of the lowest
excited states.
From the symmetry perspective, the ground state wavefunction belongs to the
class of states with symmetric spatial wavefunctions, and antisymmetric spin (singlet)
wavefunctions. These states are known as parahelium. In the absence of electron-
electron interaction, the first excited states are degenerate and have the form,
1
|ψp % = √ (|100% ⊗ |2$m% + |2$m% ⊗ |100%) ⊗ |χ00 % .
2
The second class of states involve an antisymmetric spatial wavefunction, and sym-
metric (triplet) spin wavefunction. These states are known as orthohelium. Once
again, in the absence of electron-electron interaction, the first excited states are de-
generate and have the form,
1
|ψo % = √ (|100% ⊗ |2$m% − |2$m% ⊗ |100%) ⊗ |χ1Sz % .
2
The perturbative shift in the ground state energy has already been calculated
within the framework of first order perturbation theory. Let us now consider the shift
in the excited states. Despite the degeneracy, since the off-diagonal matrix elements
vanish, we can make use of the first order of perturbation theory to compute the shift.
In doing so, we obtain
%
p,o 1 e2 1
∆En! = d3 r1 d3 r2 |ψ100 (r1 )ψn!0 (r2 ) ± ψn!0 (r1 )ψ100 (r2 )|2 ,
2 4π&0 |r1 − r2 |
with the plus sign refers to parahelium and the minus to orthohelium. Since the
matrix element is independent of m, the m = 0 value considered here applies to all
p,o
values of m. Rearranging this equation, we thus obtain ∆En! = Jn! ± Kn! where
%
e2
|ψ100 (r1 )| |ψn!0 (r2 )|
2 2
Jn! = d 3 r1 d 3 r 2
4π&0 |r1 − r2 |
2 %
e ψ (r1 )ψn!0
∗ ∗
(r2 )ψ100 (r2 )ψn!0 (r1 )
Kn! = d3 r1 d3 r2 100 .
4π&0 |r1 − r2 |
Physically, the term Jn! represents the electrostatic interaction energy associated
with the two charge distributions |ψ100 (r1 )|2 and |ψn!0 (r2 )|2 , and it is clearly posi-
tive. By contrast, the exchange term, which derives from the antisymmetry of the
wavefunction, leads to a shift with opposite signs for ortho and para states. In fact,
one may show that, in the present case, Kn! is positive leading to a positive energy
shift for parahelium and a negative shift for orthohelium. Moreover, noting that
& ' (
3 1/2 triplet
2S1 · S2 = (S1 + S2 ) − S1 − S2 = ! S(S + 1) − 2 ×
2 2 2 2
=! 2
4 −3/2 singlet
the energy shift can be written as
& '
p,o 1 4
∆En! = Jn! − 1 + 2 S1 · S2 Kn! .
2 !
This result shows that the electron-electron interaction leads to an effective ferro-
magnetic interaction between spins – i.e. the spins want to be aligned.6
In addition to the large energy shift between the singlet and triplet states, electric Energy level diagram for ortho-
dipole decay selection rules ∆$ = ±1, ∆s = 0 (whose origin is discussed later in the and parahelium showing the first
course) cause decays from triplet to singlet states (or vice-versa) to be suppressed by order shift in energies due to
the Coulomb repulsion of elec-
a large factor (compared to decays from singlet to singlet or from triplet to triplet).
trons. Here we assume that one
This caused early researchers to think that there were two separate kinds of Helium. of the electrons stays close to the
The diagrams (right) shows the levels for parahelium (singlet) and for orthohelium ground state of the unperturbed
(triplet) along with the dominant decay modes. Hamiltonian.
Figure 8.1: (Left) Schematic showing the phase space volume associated with each
plane wave state in a Fermi gas. (Right) Schematic showing the state occupancy of a
filled Fermi sea.
2π
k= (n1 , n2 , · · · nd ), ni integer .
L
To address the quantum mechanics of the system, we start with fermions.
!2 2
EF = (6π 2 n) 3 .
2m
8
The quantization condition follows form the periodic boundary condition, φ[r+L(mx x̂+
my ŷ + mz ẑ)] = φ(r), where m = (mx , my , mz ) denotes an arbitrary vector of integers.
We can also integrate to obtain the total energy density of all the fermions,
%
Etot 1 kF
4πk 2 dk !2 k 2 !2 3
= 3 = (6π 2 n)5/3 = nEF .
L 3 L 0 (2π/L) 2m
3 20π m
2 5
of this type of supernova has been used to measure the accelerating expansion of the
universe. An estimate the neutron star radius using the formulae above leads to
Rneutron me
. . 10−3 ,
Rwhite dwarf MN
i.e. ca. 10 km. If the pressure at the center of a neutron star becomes too great, it Satyendra Nath Bose 1894-1974
collapses to become a black hole. An Indian physi-
cist who is best
known for his
work on quantum
mechanics in
8.4.2 Ideal Bose gas the early 1920s,
providing the
In a system of N spinless non-interacting bosons, the ground state of foundation for
the many-body system is described simply by a wavefunction in which all Bose-Einstein
statistics and the development of
N particles occupy the lowest energy single-particle state, i.e. in this case, the theory of the Bose-Einstein
the fully symmetrized ) wavefunction can be expressed as the product state, condensate. He is honored as the
ψB (r1 , r2 , · · · rN ) = N
i=1 φk=0 (ri ). (More generally, for a confining potential
namesake of the boson.
V (r), φk (r) denote the corresponding single-particle bound states with k the
associated quantum numbers.) However, in contrast to the Fermi system, the
transit to the ground state from non-zero temperatures has an interesting fea-
ture with intriguing experimental ramifications. To understand why, let us
address the thermodynamics of the system.
For independent bosons, the number of particles in plane wave state k with
energy &k is given by the Bose-Einstein distribution,
1
nk = ,
e(#k −µ)/kB T − 1
*
where the chemical potential, µ, is fixed by the condition N = k nk with
N the total number of particles. In a three-dimensional system, for N large,
we
* may Lapproximate
+ the sum by an integral over momentum space setting
k →
" ( 2π )3 d3 k. As a result, we find that
%
N 1 1
= n = d3 k (# −µ)/k T .
L3 (2π)3 e k B −1
!2 k2
For a free particle system, where &k = 2m , this means that
1
n= Li3/2 (µ/kB T ) , (8.1)
λ3T
* zk h2
where Lin (z) = ∞ k=1 kn denotes the polylogarithm, and λT = ( 2πmkB T )
1/2
denotes the thermal wavelength, i.e. the length scale at which the corre-
sponding energy scale becomes comparable to temperature. As the density of
particles increase, or the temperature falls, µ increases from negative values
until, at some critical value of density, nc = λ−3
T ζ(3/2), µ becomes zero (note
that Lin (0) = ζ(n)). Equivalently, inverting, this occurs at a temperature,
!2 2/3 2π
kB Tc = α n , α= .
m ζ 2/3 (3/2)
Clearly, since nk ≥ 0, the Bose distribution only makes sense for µ negative.
So what happens at this point?
Consider first what happens at zero temperature. Since the particles are
bosons, the ground state consists of every particle sitting in the lowest energy
state (k = 0). But such a singular distribution was excluded by the replace-
ment of the sum by the integral. Suppose that, at T < Tc , we have a thermo-
dynamic fraction f (T ) of particles sitting in this state.9 Then the chemical
potential may stay equal to zero and Eq. (8.1) becomes n = λ13 ζ(3/2)+f (T )n,
T
where f (T ) denotes the fraction of particles in the ground state. But, since
n = λ13 ζ(3/2), we have
Tc
# Info. Although solid state systems continue to provide the most “accessible”
arena in which to study the properties of quantum liquids and gases, in recent years, a
new platform has been realized through developments in atomic physics – in the field of
ultracold atom physics, dilute atomic vapours are maintained at temperatures close
to absolute zero, typically below some tenths of microkelvins, where their behaviour
are influenced by the effects of quantum degneracy. The method of cooling the gas
has a long history which it would be unwise to detail here. But in short, alkali atoms
can be cooled by a technique known as laser cooling. Laser beams, in addition to
carrying heat, also carry momentum. As a result, photons impart a pressure when
they collide with atoms. The acceleration of an atom due to photons can be some four
orders of magnitude larger than gravity. Consider then a geometry in which atoms
are placed inside two counterprogating laser beams. To slow down, an atom has to
absorb a photon coming towards it, and not from behind. This can be arranged by
use of the Doppler shift. By tuning the laser frequency a little bit towards the low-
frequency (“red”) side of a resonance, the laser beam opposing the atom is Doppler
shifted to a higher (more “blue”) frequency. Thus the atom is more likely to absorb
that photon. A photon coming from behind the atom is now a little bit redder, which
means the atom is less likely to absorb that photon. So in whichever direction the
atom is moving, the laser beam opposing the motion seems stronger to the atom, and Schematic of a Magneto-Optical
it slows the atom down. If you multiply this by three and have laser beams coming Trap (MOT). The invention of
from the north, south, east, west, up, and down, you get an “optical molasse”. If the MOT in 1987 at Bell Labs
you walk around in a pot full of molasses, whichever direction you go, the molasses and optical molasses was the ba-
somehow knows that is the direction to push against. It’s the same idea. sis for the 1997 Nobel Prize in
In the study of ultracold atomic gases, experimentalists are usually concerned Physics.
with addressing the properties of neutral alkali atoms. The number of atoms in a
typical experiment ranges from 104 to 107 . The atoms are conned in a trapping
potential of magnetic or optical origin, with peak densities at the centre of the trap
ranging from 1013 cm3 to 1015 cm3 . The development of quantum phenomena such as
Bose-Einstein condensation requires a phase-space density of order unity, or nλ3T ∼ 1
where λT denotes the thermal de Broglie wavelength. These densities correspond to
temperatures,
!2 n2/3
T ∼ ∼ 100nK to a few µK .
mkB
9
Here, by thermodynamic, we mean that the fraction scales in proportion to the density.
10
The Bose-Einstein condensate was first predicted by Satyendra Nath Bose and Albert
Einstein in 1924-25. Bose first sent a paper to Einstein on the quantum statistics of light
quanta (now called photons). Einstein was impressed, translated the paper himself from
English to German and submitted it for Bose to the Zeitschrift fr Physik which published it.
Einstein then extended Bose’s ideas to material particles (or matter) in two other papers.
Figure 8.2: (Left) Observation of BEC by absorption imaging. The top row shows
shadow pictures, which are rendered in a 3d plot below. The “sharp peak” is the
BEC, characterized by its slow expansion observed after 6 ms time of flight. The total
number of atoms at the phase transition is about 7 × 105 , and the temperature at the
transition point is 2 µK. (Figure from Ketterle.) (right) Figure shows the shrinking of
the atom cloud in a magnetic as the temperature is reduced by evaporative cooling.
Comparison between bosonic 7 Li (left) and fermionic 6 Li (right) shows the distinctive
signature of quantum statistics. The fermionic cloud cannot shrink below a certain
size determined by the Pauli exclusion principle. This is the same phenomenon that
prevents white dwarf and neutron stars from shrinking into black holes. At the
highest temperature, the length of the clouds was about 0.5mm. (Figure from J. R.
Anglin and W. Ketterle, Bose-Einstein condensation of atomic gases, Nature 416,
211 (2002).)
,
At these temperatures the atoms move at speeds of ∼ kB T /m ∼ 1 cm s−1 , which
should be compared with around 500 ms1 for molecules at room temperature, and
∼ 106 ms−1 for electrons in a metal at zero temperature. Achieving the regime n3T ∼ 1,
through sufficient cooling, is the principle experimental advance that gave birth to
this new field of physics.
It should be noted that such low densities of atoms are in fact a necessity. We are
dealing with systems whose equilibrium state is a solid (that is, a lump of Sodium,
Rubidium, etc.). The rst stage in the formation of a solid would be the combination
of pairs of atoms into diatomic molecules, but this process is hardly possible without
the involvement of a third atom to carry away the excess energy. The rate per atom
of such three-body processes is 1029 − 10−30 cm6 s−1 , leading to a lifetime of several
seconds to several minutes. These relatively long timescales suggest that working
with equilibrium concepts may be a useful first approximation.
Since the alkali elements have odd atomic number Z, we readily see that alkali Bosons Fermions
atoms with odd mass number are bosons, and those with even mass number are 7
Li I=3/2 6
Li I=1
fermions. Thus bosonic and fermionic alkalis have half-integer and integer nuclear 23
Na I=3/2 23
K I=4
spin respectively. Alkali atoms have a single valence electron in an ns state, so have 87
Rb I=3/2
electronic spin J = S = 1/2. The experimental star players: are shown in the table
(right). The hyperfine coupling between electronic and nuclear spin splits the ground
state manifold into two multiplets with total spin F = I ± 1/2. The Zeeman splitting
of these multiplets by a magnetic field forms the basis of magnetic trapping.
So, based on our discussion above, what happens when a Bose or Fermi gas is
confined by an harmonic trapping potential, V (r) = 12 mω 2 r2 . At high temperatures,
Bose and Fermi gases behave classically and form a thermal (Gaussian) distribution,
2 2
P (r) . e−mω r /2kB T . As the system is cooled towards the point of quantum degen-
eracy, i.e. when the typical separation between particles, n−1/3 , becomes comparable
to the thermal wavelength, λT , quantum statistics begin to impact. In the Fermi
system, Pauli exclusion leads to the development of a Fermi surface and the cloud
size becomes arrested. By contrast, the Bose system can form a BEC, with atoms
condensing into the ground state of the harmonic potential. Both features are shown
in Figure 8.2.
Atomic structure
Previously, we have seen that the quantum mechanics of atomic hydrogen, and
hydrogen-like atoms is characterized by a large degeneracy with eigenvalues
separating into multiplets of n2 -fold degeneracy, where n denotes the principle
quantum number. However, although the idealized Schrödinger Hamiltonian,
p̂2 1 Ze2
Ĥ0 = + V (r), V (r) = − , (9.1)
2m 4π"0 r
provides a useful platform from which develop our intuition, there are several
important effects which mean that the formulation is a little too naı̈ve. These
“corrections”, which derive from several sources, are important as they lead to
physical ramifications which extend beyond the realm of atomic physics. Here
we outline some of the effects which need to be taken into account even for
atomic hydrogen, before moving on to discuss the quantum physics of multi-
electron atoms. In broad terms, the effects to be considered can be grouped
into those caused by the internal properties of the nucleus, and those which
derive from relativistic corrections.
To orient our discussion, it will be helpful to summarize some of key aspects
of the solutions of the non-relativisitic Schrödinger equation, Ĥ0 ψ = Eψ on
which we will draw:
1 (p̂2 )2
Ĥ1 = − .
8 m3 c2
When compared with the non-relativistic kinetic energy, p2 /2m, one can see
that the perturbation is smaller by a factor of p2 /m2 c2 = v 2 /c2 $ (Zα)2 , i.e.
Ĥ1 is only a small perturbation for small atomic number, Z ( 1/α $ 137.
1
It may seem odd to discuss relativistic corrections in advance of the Dirac equation
and the relativistic formulation of quantum mechanics. However, such a discussion would
present a lengthy and unnecessarily complex digression which would not lead to further
illumination. We will therefore follow the normal practice of discussing relativistic corrections
as perturbations to the familiar non-relativistic theory.
energy shift can be obtained from first order perturbation theory, By making use of the form of the
radial wavefunction for the hy-
1 & 2 '2 drogen atom, one may obtain the
%Ĥ1 &n!m ≡ %n'm|Ĥ1 |n'm& = − En − 2En %V (r)&n! + %V 2 (r)&n! .
2mc2 identities,
$ %
Since the calculation of the resulting expectation values is not particularly 1 Z
=
illuminating, we refer to the literature for a detailed exposition3 and present r n! a0 n2
here only the required identities (right). From these considerations, we obtain $ %
1 Z2
the following expression for the first order energy shift, = 2 3 .
2
r n! a0 n (' + 1/2)
! " ! "
mc2 Zα 4 n 3
%Ĥ1 &n!m = − − . (9.2)
2 n ' + 1/2 4
From this term alone, we expect the degerenacy between states with different
values of total angular momentum ' to be lifted. However, as well will see,
this conclusion is a little hasty. We need to gather all terms of the same
order of perturbation theory before we can reach a definite conclusion. We
can, however, confirm that (as expected) the scale of the correction is of order
"Ĥ1 #n!m
En ∼ ( Zα
n ) . We now turn to the second important class of corrections.
2
1 1
Ĥ2 = (∂r V (r)) Ŝ · L̂ .
2m2 c2 r
2
For a hydrogen-like atom, V (r) = − 4π#
1 Ze
0 r
, and
1 Ze2 1
Ĥ2 = Ŝ · L̂ .
2m2 c2 4π"0 r3
by E = −∇V (r) = −er (∂r V ). For an electron moving at velocity v, this translates
to an effective magnetic field B = c12 v × E. The magnetic moment of the electron
q
associated with its spin is equal to µs = gs 2m S ≡ −m
e
S, and thus the interaction
energy is given by
e e e 1
−µs · B = S · (v × E) = − S · (p × er (∂r V )) = (∂r V ) L · S ,
mc 2 (mc)2 (mc)2 r
where we have used the relation p × er = p × rr = − Lr . In fact this isn’t quite correct;
there is a relativistic effect connected with the precession of axes under rotation, called
Thomas precession which multiplies the formula by a further factor of 12 .
Once again, we can estimate the effect of spin-orbit coupling by treating Ĥ2
as a perturbation. In the absence of spin-orbit interaction, one may express
the eigenstates of hydrogen-like atoms in the basis states of the mutually
commuting operators, Ĥ0 , L̂2 , L̂z , Ŝ2 , and Ŝz . However, in the presence
of spin-orbit coupling, the total Hamiltonian no longer commutes with L̂z
or Ŝz (exercise). It is therefore helpful to make use of the degeneracy of
the unperturbed Hamiltonian to switch to a new basis in which the angular
momentum components of the perturbed system are diagonal. This can be
achieved by turning to the basis of eigenstates of the operators, Ĥ0 , Ĵ2 , Jˆz ,
L̂2 , and Ŝ2 , where Ĵ = L̂ + Ŝ denotes the total angular momentum. (For a
discussion of the form of these basis states, we refer back to section 6.4.2.)
Making use of the relation, Ĵ2 = L̂2 + Ŝ2 + 2L̂ · Ŝ, in this basis, it follows
that,
1
L̂ · Ŝ = (Ĵ2 − L̂2 − Ŝ2 ) .
2
Combining the spin and angular momentum, the total angular momentum
takes values j = ' ± 1/2. The corresponding basis states |j = ' ± 1/2, mj , '&
(with s = 1/2 implicit) therefore diagonalize the operator,
! "
!2 '
Ŝ · L̂|j = ' ± 1/2, mj , '& = |' ± 1/2, mj , '& ,
2 −' − 1
where the brackets index j = ' + 1/2 (top) and j = ' − 1/2 (bottom). As for
the radial dependence of the perturbation, once again, the off-diagonal matrix
elements vanish circumventing the need to invoke degenerate perturbation
theory. As a result, at first order in perturbation theory, one obtains
! " $ %
1 !2 ' Ze2 1
%H2 &n,j=!±1/2,mj ,! = .
2m c 2
2 2 −' − 1 4π"0 r3 n!
Then making use of the identity (right),4 one obtains For ' > 0,
! " ( ) $ % ! "3
1 2 Zα 4 n 1
1 mcαZ 1
%Ĥ2 &n,j=!±1/2,mj ,! = mc j
1 . = .
4 n j + 1/2 − j+1 r3 n! !n '(' + 12 )(' + 1)
Note that, for ' = 0, there is no orbital angular momentum with which to
couple! Then, if we rewrite the expression for %Ĥ1 & (9.2) in the new basis,
! " ( 1 )
1 2 Zα 4
%Ĥ1 &n,j=!±1/2,mj ,! = − mc n j
1 ,
2 n j+1
where Qnuclear (r) = Zeδ (3) (r) denotes the nuclear charge density. Since the
perturbation acts only at the origin, it effects only states with ' = 0. As a
result, one finds that
! "4
Ze2 !2 1 Zα
%Ĥ3 &njmj ! = 4π|ψ!n (0)|2 = mc2 nδ!,0 .
4π"0 8(mc)2 2 n
n = 3, ' can take values of 0, 1 or 2. Here, the pairs of states 3S1/2 and 3P1/2 ,
and 3P3/2 and 2D3/2 each remain degenerate while the state 3D5/2 is unique.
These predicted shifts are summarized in Figure 9.1.
This completes our discussion of the relativistic corrections which develop Willis Eugene Lamb, 1913-2008
from the treatment of the Dirac theory for the hydrogen atom. However, A physicist who
won the Nobel
this does not complete our discription of the “real” hydrogen atom. Indeed, Prize in Physics
there are further corrections which derive from quantum electrodynamics and in 1955 “for his
discoveries con-
nuclear effects which we now turn to address. cerning the fine
structure of the
hydrogen spectrum”. Lamb and
9.1.4 Lamb shift Polykarp Kusch were able to pre-
cisely determine certain electromag-
According to the perturbation theory above, the relativistic corrections which netic properties of the electron.
follow from the Dirac theory for hydrogen leave the 2S1/2 and 2P1/2 states
degenerate. However, in 1947, a careful experimental study by Willis Lamb
and Robert Retherford discovered that this was not in fact the case:5 2P1/2 Hans Albrecht Bethe 1906-2005
A German-
state is slightly lower in energy than the 2S1/2 state resulting in a small shift American
of the corresponding spectral line – the Lamb shift. It might seem that such a physicist, and
tiny effect would be deemed insignificant, but in this case, the observed shift Nobel laureate
in physics “for
(which was explained by Hans Bethe in the same year) provided considerable his work on the
insight into quantum electrodynamics. theory of stellar
nucleosynthesis.”
In quantum electrodynamics, a quantized radiation field has a zero-point A versatile theo-
energy equivalent to the mean-square electric field so that even in a vacuum retical physicist,
Bethe also made important contribu-
there are fluctuations. These fluctuations cause an electron to execute an tions to quantum electrodynamics,
oscillatory motion and its charge is therefore smeared. If the electron is bound nuclear physics, solid-state physics
by a non-uniform electric field (as in hydrogen), it experiences a different and particle astrophysics. During
World War II, he was head of the
potential from that appropriate to its mean position. Hence the atomic levels Theoretical Division at the secret
are shifted. In hydrogen-like atoms, the smearing occurs over a length scale, Los Alamos laboratory developing
the first atomic bombs. There he
! "2 played a key role in calculating the
2α ! 1 critical mass of the weapons, and
%(δr)2 & $ ln , did theoretical work on the implosion
π mc αZ method used in both the Trinity test
and the “Fat Man” weapon dropped
some five orders of magnitude smaller than the Bohr radius. This causes the on Nagasaki.
electron spin g-factor to be slightly different from 2,
! "
α α2
gs = 2 1 + − 0.328 2 + · · · .
2π π
There is also a slight weakening of the force on the electron when it is very
close to the nucleus, causing the 2S1/2 electron (which has penetrated all the
way to the nucleus) to be slightly higher in energy than the 2P1/2 electron.
Taking into account these corrections, one obtains a positive energy shift
! "4 ! "
Z 8 1
∆ELamb $ nα Ry ×
2
α ln δ!,0 ,
n 3π αZ
Once again, to evaluate the expectation values on the spin degrees of free-
dom, it is convenient to define the total spin F = I + S. We then have
1 1 1
S · I = 2 (F2 − S2 − I2 ) = (F (F + 1) − 3/4 − I(I + 1))
! *2! 2
2
1 I F = I + 1/2
=
2 −I − 1 F = I − 1/2
where rij ≡ |ri − rj |. The first term represents the “single-particle” contri-
bution to the Hamiltonian arising from interaction of each electron with the
nucleus, while the last term represents the mutual Coulomb interaction be-
tween the constituent electrons. It is this latter term that makes the generic
problem “many-body” in character and therefore very complicated. Yet, as we
have already seen in the perturbative analysis of the excited states of atomic
Helium, this term can have important physical consequences both on the over-
all energy of the problem and on the associated spin structure of the states.
The central field approximation is based upon the observation that
the electron interaction term contains a large central (spherically symmetric)
component arising from the “core electrons”. From the following relation,
!
+
|Ylm (θ, φ)|2 = const.
m=−!
Here the one-electron potential, Ui (r), which is assumed central (see below),
incorporates the “average” effect of the other electrons. Before discussing how
to choose the potentials Ui (r), let us note that Ĥ0 is separable into a sum of
terms for each electron, so that the total wavefunction can be factorized into
components for each electron. The basic idea is first to solve the Schrödinger
equation using Ĥ0 , and then to treat Ĥ1 as a small perturbation.
On general grounds, since the Hamiltonian Ĥ0 continues to commute with
the angular momentum operator, [Ĥ0 , L̂] = 0, we can see that the eigenfunc-
tions of Ĥ0 will be characterized by quantum numbers (n, ', m! , ms ). However,
since the effective potential is no longer Coulomb-like, the ' values for a given Douglas Rayner Hartree FRS
n need not be degenerate. Of course, the difficult part of this procedure is 1897-1958
An English
to estimate Ui (r); the potential energy experienced by each electron depends mathematician
on the wavefunction of all the other electrons, which is only known after the and physicist
most famous for
Schrödinger equation has been solved. This suggests that an iterative approach the development
to solving the problem will be required. of numerical
analysis and its
To understand how the potentials Ui (r) can be estimated – the self- application to
consistent field method – it is instructive to consider a variational ap- atomic physics.
proach due originally to Hartree. If electrons are considered independent, the He entered
St John’s College Cambridge in
wavefunction can be factorized into the product state, 1915 but World War I interrupted
his studies and he joined a team
Ψ({ri }) = ψi1 (r1 )ψi2 (r2 ) · · · ψiN (rN ) , studying anti-aircraft gunnery. He
returned to Cambridge after the war
where the quantum numbers, ik ≡ (n'm! ms )k , indicate the individual state and graduated in 1921 but, perhaps
because of his interrupted studies,
occupancies. Note that this product state is not a properly antisymmetrized he only obtained a second class
Slater determinant – the exclusion principle is taken into account only in so degree in Natural Sciences. In 1921,
a visit by Niels Bohr to Cambridge
far as the energy of the ground state is taken to be the lowest that is consistent inspired him to apply his knowledge
with the assignment of different quantum numbers, n'm! ms to each electron. of numerical analysis to the solution
of differential equations for the
Nevertheless, using this wavefunction as a trial state, the variational energy is calculation of atomic wavefunctions.
then given by
+. ! 2 2 "
∗ ! ∇ 1 Ze2
E = %Ψ|Ĥ|Ψ& = 3
d r ψi − − ψi
2m 4π"0 r
i
. .
1 + e2
+ 3
d r d3 r! ψi∗ (r)ψi∗ (r! ) ψj (r! )ψi (r) .
4π"0 |r − r! |
i<j
.
1 + e2
Ui (r) = d3 r! |ψj (r! )|2 .
4π"0 |r − r! |
j&=i
Equation (9.4) has a simple interpretation: The first two terms relate to the
nuclear potential experienced by the individual electrons, while the third term
represents the electrostatic potential due to the other electrons. However, to
simplify the procedure, it is useful to engineer the radial symmetry of the
potential by replacing Ui (r) by its spherical average,
.
dΩ
Ui (r) -→ Ui (r) = Ui (r) .
4π
8
Note that, in applying the variation, the wavefunction ψi∗ can be considered independent
of ψi – you might like to think why.
If we compare this expression with the variational state energy, we find that
+ .
1 + e2
E= "i − d3 r! d3 r |ψj (r! )|2 |ψi (r)|2 . (9.5)
4π"0 |r − r! |
i i<j
Note that each of√the N ! terms in Ψ is a product of wavefunctions for each individual Vladimir Aleksandrovich Fock
electron. The 1/ N ! factor ensures the wavefunction is normalized. A determinant 1898-1974
A Soviet physi-
changes sign if any two columns are exchanged, corresponding to ri ↔ rj (say); this cist, who did
ensures that the wavefunction is antisymmetric under exchange of electrons i and j. foundational
Likewise, a determinant is zero if any two rows are identical; hence all the ψk s must work on quan-
be different and the Pauli exclusion principle is satisfied.9 In this approximation, a tum mechanics
and quantum
variational analysis leads to the Hartree-Fock equations (exercise), electrodynamics.
, - His primary
!2 2 Ze2
εi ψi (r) = − ∇i − ψi (r) scientific con-
2m 4π"ri tribution lies
+. 1 e2 0 1 in the development of quantum
+ d 3 rj ψ (r
∗ #
) ψ j (r#
)ψ i (r) − ψ j (r)ψ i (r#
)δ m ,m . physics, although he also contributed
4π"0 |r − r# | j s i s j
significantly to the fields of me-
j"=i
chanics, theoretical optics, theory of
The first term in the last set of brackets translates to the ordinary Hartree contribution gravitation, physics of continuous
above and describes the influence of the charge density of the other electrons, while medium. In 1926 he derived the
Klein-Gordon equation. He gave
the second term describes the non-local exchange contribution, a manifestation of his name to Fock space, the Fock
particle statistics. representation and Fock state, and
developed the HartreeFock method
in 1930. Fock made significant
The outcome of such calculations is that the eigenfunctions are, as for contributions to general relativity
theory, specifically for the many
hydrogen, characterized by quantum numbers n, ', m! , with ' < n, but that the body problems.
states with different ' for a given n are not degenerate, with the lower values
of ' lying lower. This is because, for the higher ' values, the electrons tend
to lie further from the nucleus on average, and are therefore more effectively
screened. The states corresponding to a particular value of n are generally
referred to as a shell, and those belonging to a particular pair of values of n, '
are usually referred to as a subshell. The energy levels are ordered as below
(with the lowest lying on the left): !
"
Subshell name 1s 2s 2p 3s 3p 4s 3d 4p 5s 4d ··· !
!
"
! !
7s 7p
! ! 7d
n= 1 2 2 3 3 4 3 4 5 4 ··· ···
!
"
!
'= 0 0 1 0 1 0 2 1 0 2 ··· 6s !
6p !
6d 6f · · ·
!
" ! ! !
Degeneracy 2 2 6 2 6 2 10 6 2 10 ··· ! 5p 5d
5s ! ! 5g
! 5f
Cumulative 2 4 10 12 18 20 30 36 38 48 ··· " ! ! ! !
!
4s 4p
! ! 4d
! 4f!
Note that the values of Z corresponding to the noble gases, 2, 10, 18, 36, at " ! ! ! !!
!
! 3p
3s ! 3d!
which the ionization energy is unusually high, now emerge naturally from this " ! ! !!
!
filling order, corresponding to the numbers of electrons just before a new shell 2s
! !2p
" ! !!
!
(n) is entered. There is a handy mnemonic to remember this filling order. By 1s
! !
writing the subshells down as shown right, the order of states can be read off !
along diagonals from lower right to upper left, starting at the bottom.
We can use this sequence of energy levels to predict the ground state
electron configuration of atoms. We simply fill up the levels starting from
the lowest, accounting for the exclusion principle, until the electrons are all
accommodated (the aufbau principle). Here are a few examples:
Z Element Configuration 2S+1
LJ Ioniz. Pot. (eV)
1 H (1s) 2
S1/2 13.6
2 He (1s)2 1
S0 24.6
3 Li He (2s) 2
S1/2 5.4
4 Be He (2s)2 1
S0 9.3
5 B He (2s)2 (2p) 2
P1/2 8.3
6 C He (2s)2 (2p)2 3
P0 11.3
7 N He (2s)2 (2p)3 4
S3/2 14.5
8 O He (2s)2 (2p)4 3
P2 13.6
9 F He (2s)2 (2p)5 2
P3/2 17.4
10 Ne He (2s)2 (2p)6 1
S0 21.6
11 Na Ne (3s) 2
S1/2 5.1
9
Note that for N = 2, the determinant reduces to the familiar antisymmetric wavefunc-
tion, √12 [ψ1 (r1 )ψ2 (r2 ) − ψ2 (r1 )ψ1 (r2 )].
Since it is generally the outermost electrons which are of most interest, con-
tributing to chemical activity or optical spectra, one often omits the inner
closed shells, and just writes O as (2p)4 , for example. However, the configura-
tion is not always correctly predicted, especially in the heavier elements, where
levels may be close together. It may be favourable to promote one or even
two electrons one level above that expected in this simple picture, in order to
achieve a filled shell. For example, Cu (Z = 29) would be expected to have con-
figuration · · · (4s)2 (3d)9 , and actually has configuration · · · (4s)1 (3d)10 . There
are several similar examples in the transition elements where the d subshells
are being filled, and many among the lanthanides (rare earths) and actinides
where f subshells are being filled.
ways of distributing the νn! electrons among the δ! individual states. The total
degeneracy, g, is then obtained from the product.
where Ĥ0 includes the kinetic energy and central field terms, Ĥ1 is the residual
Coulomb interaction, and (with ξi (ri ) = 2m12 c2 1r (∂r V (r))) Ĥ2 is the spin-orbit
interaction. We can then consider two possible scenarios:
Ĥ1 2 Ĥ2 : This tends to apply in the case of light atoms. In this situation,
one considers first the eigenstates of Ĥ0 + Ĥ1 , and then treats Ĥ2 as a
perturbation. This leads to a scheme called LS (or Russell-Saunders)
coupling.
Ĥ2 2 Ĥ1 : This can apply in very heavy atoms, or in heavily ionized light
atoms, in which the electrons are moving at higher velocities and rela-
tivistic effects such as the spin-orbit interaction are more important. In
this case, a scheme called jj coupling applies.
1. Combine the spins of the electrons to obtain possible values of total spin
S. The largest permitted value of S lies lowest in energy.
2. For each value of S, find the possible values of total angular momentum
L. The largest value of L lies lowest in energy.
3. Couple the values of L and S to obtain the values of J (hence the name
of the scheme). If the subshell is less than half full, the smallest value
of J lies lowest; otherwise, the largest value of J lies lowest.
2. Maximising L also tends to keep the electrons apart. This is less obvious,
though a simple classical picture of electrons rotating round the nucleus
in the same or different senses makes it at least plausible.
where the three 3 P states are separated by the spin-orbit interaction, and the
singlet 1 P state lies much higher in energy owing to the Coulomb interaction.
The separations between the 3 P2 and 3 P1 and the 3 P1 and 3 P0 should be in the
ratio 2:1. This is an example of the Landé interval rule, which states that
the separation between a pair of adjacent levels in a fine structure multiplet is
proportional to the larger of the two J values involved. This is easily shown
using Eq. (9.7) – the separation in energy between states J and J − 1 is
∝ J(J + 1) − (J − 1)J = 2J .
Actually in the case of helium the situation is a bit more complicated, because
it turns out that the spin-orbit interaction between different electrons makes
a non-negligible additional contribution to the fine structure. Other excited
states of helium, of the form (1s)1 (n')1 , can be handled similarly, and again
separate into singlet and triplet states.
$ Exercise. For the case of boron, with the electron configuration (1s)2 (2s)2 (2p),
use Hund’s rules to show that the ground state is 2 P1/2 .
We next consider the case of carbon, which has ground state electron
configuration (1s)2 (2s)2 (2p)2 . This introduces a further complication; we now
have two identical electrons in the same unfilled subshell, and we need to
ensure that their wavefunction is antisymmetric with respect to electron ex-
change. The total spin can either be the singlet S = 0 state, which has an
antisymmetric wavefunction √12 [| ↑1 & ⊗ | ↓2 & − | ↓1 & ⊗ | ↑2 &], or one of the triplet
S = 1 states, which are symmetric, √12 [| ↑1 & ⊗ | ↓2 & + | ↓1 & ⊗ | ↑2 &], | ↑1 & ⊗ | ↑2 &
(1) (2)
or | ↓1 & ⊗ | ↓2 &. We must therefore choose values of L with the appropriate m! m! mL
symmetry to partner each value of S. To form an antisymmetric state, the two 1 0 1
1 −1 0
electrons must have different values of m! , so the possibilities are as shown
0 −1 −1
right. Inspecting the values of mL we can deduce that L = 1.10 By contrast,
to form a symmetric total angular momentum state, the two electrons may (1) (2)
have any values of m! , leading to the possibilities shown right. Inspecting the m! m! mL
1 1 2
values of mL we infer that L = 2 or 0.
1 0 1
We must therefore take S = 1 with L = 1 and S = 0 with L = 2 or 0. 1 −1 0
Finally, to account for the fine structure, we note that the states with S = 1 0 0 0
and L = 1 can be combined into a single J = 0 state, three J = 1 states, and 0 −1 −1
five J = 2 states leading to the terms 3 P0 , 3 P1 , and 3 P2 respectively. Similarly −1 −1 −2
the S = 0, L = 2 state can be combined to give five J = 2 states, 1 D2 , while the
S = 0, L = 0 state gives the single J = 0 state, 1 S0 . Altogther we recover the E /cm−1
1 + 3 + 5 + 5 + 1 = 15 possible states (cf. Eq. (9.6) with the ordering in energy 1
S0 20649
given by Hund’s rules (shown to the right). The experimental energy values 1
D2 10195
are given using the conventional spectroscopic units of inverse wavelength.
3
P2 43
Note that the Landé interval rule is approximately obeyed by the fine structure
3
P1 16
3
P0 0
triplet, and that the separation between L and S values caused by the electron-
electron repulsion is much greater than the spin-orbit effect.
In an excited state of carbon, e.g. (2p)1 (3p)1 , the electrons are no longer
equivalent, because they have different radial wavefunctions. So now one can
combine any of S = 0, 1 with any of L = 0, 1, 2, yielding the following terms
(in order of increasing energy, according to Hund’s rules):
3
D1,2,3 3
P0,1,2 3
S1 1
D2 1
P1 1
S0 .
For nitrogen, the electron configuration is given by (1s)2 (2s)2 (2p)3 . The
maximal value of spin is S = 3/2 while L can take values 3, 2, 1 and 0. Since
10
This result would also be apparent if we recall the that angular momentum states are
eigenstates of the parity operator with eigenvalue (−1)L . Since there are just two electrons,
this result shows that both the L = 0 and L = 2 wavefunction must be symmetric under
exchange.
As a final example, let us consider the ground state of oxygen, which has
electron configuration (2p)4 . Although there are four electrons in the (2p)
subshell, the maximum value of S = 1. This is because there are only three
available values of m! = ±1, 0, and therefore one of these must contain two
electrons with opposite spins. Therefore, the maximum value of mS = 1,
achieved by having electrons with ms = + 12 in both the other m! states. By
pursuing this argument, it is quite easy to see that the allowed values of L, S Level scheme of the carbon
atom (1s)2 (2s)2 (2p)2 . Draw-
and J are the same as for carbon (2p)2 . This is in fact a general result – the
ing is not to scale. On the
allowed quantum numbers for a subshell with n electrons are the same as for left the energy is shown with-
that of a subshell with n “holes”. Therefore, the energy levels for the oxygen out any two-particle interaction.
ground state configuration are the same as for carbon, except that the fine The electron-electron interaction
structure multiplet is inverted, in accordance with Hund’s third rule. leads to a three-fold energy split-
ting with L and S remaining
good quantum numbers. Spin-
9.3.2 jj coupling scheme
orbit coupling leads to a further
When relativitic effects take precedence over electron interaction splitting of the states with J re-
6 effects, we maining a good quantum num-
must start by considering the eigenstates of Ĥ0 + Ĥ2 = Ĥ0 + i ξi (ri )L̂i · Ŝi .
ber. Finally on the right, the lev-
These must be eigenstates of Ĵ2 as before, because of the overall rotational els show Zeeman splittings in an
invariance, and also of Ĵ2i for each electron. Therefore, in this case, the cou- external magnetic field. In this
pling procedure is to find the allowed j values of individual electrons, whose case, the full set of 15 levels be-
energies will be separated by the spin-orbit interaction. Then these individual come non-degenerate.
j values are combined to find the allowed values of total J. The effect of the
residual Coulomb interaction will be to split the J values for a given set of js.
Sadly, in this case, there are no simple rules to parallel those of Hund.
As an example, consider a configuration (np)2 in the jj coupling scheme, to
be compared with the example of carbon which we studied in the LS scheme.
Combining s = 1/2 with ' = 1, each electron can have j = 1/2 or 3/2. If
the electrons have the same j value, they are equivalent, so we have to take
care of the symmetry of the wavefunction. We therefore have the following
possibilities:
$ j1 = 1/2, j2 = 3/2 ⇒ J = 2, 1.
(1/2, 1/2)0 (3/2, 1/2)1 (3/2, 1/2)2 (3/2, 3/2)2 (3/2, 3/2)0
in order of increasing energy. Note that both LS and jj coupling give the same
values of J (in this case, two states with J = 0, two with J = 2 and one with
J = 1) and in the same order. However, the pattern of levels is different; in
LS coupling we found a triplet and two singlets, while in this ideal jj scenario,
we have two doublets and a singlet. The sets of states in the two coupling
Atomic states are always eigenstates of parity and of total angular momentum,
J, so these selection rules can be regarded as absolutely valid in electric dipole
transitions. It should be emphasized again, though, that the electric dipole
approximation is an approximation, and higher order processes may occur,
albeit at a slower rate, and have their own selection rules.
In specific coupling schemes, further selection rules may apply. In the case
of ideal LS coupling, we also require:
$ ∆S = 0 and ∆MS = 0;
In LS coupling, the states are eigenstates of total spin; since the dipole operator
does not operate on the spin part of the wavefunction, the rules on ∆S and
∆MS follow straightforwardly. This, and the absolute rules relating to J,
imply the rules for L and ML . The rule for ∆'i follows from the parity change
rule, since the parity of the atom is the product of the parities of the separate
electron wavefunctions, given by (−1)!i . However, since LS coupling is only
an approximation, these rules should themselves be regarded as approximate.
With this preparation, we now turn to the consequences of the selection rules
on the atomic spectra of atoms.
$ The ground state has term 2 S1/2 . The excited states are all doublets
with J = L ± 1/2, except for the s states, which are obviously restricted
to J = 1/2.
$ As n increases, the energy levels approach (from below) those for hy-
drogen, because the nuclear charge is increasingly effectively screened
by the inner electrons. This happens sooner for the higher ' values, for
which the electron tends to lie further out from the nucleus.
$ In an absorption spectrum, the atoms will start from the ground state,
so only the 3s → np lines will be seen. In emission, the atoms are excited
into essentially all their excited levels, so many more lines will be seen
in the spectrum.
The comments above for sodium also apply for hydrogen, except that, in this
case, (2s, 2p), (3s, 3p, 3d), etc. are degenerate. One consequence is that the 2s
state in hydrogen is metastable – it cannot decay to the only lower lying level
(1s) by an electric dipole transition. In fact its favoured spontaneous decay
is by emission of two photons; a process which is described by second-order
perturbation theory. In practice, hydrogen atoms in a 2s state are more likely
to deexcite through collision processes. During an atomic collision, the atoms
are subject to strong electric fields, and we know from our discussion of the
Stark effect that this will mix the 2s and 2p states, and decay from 2p to 1s
is readily possible.
$ The ground state has term 1 S0 . The excited states are of the form
(1s)(n') (the energy required to excite both of the 1s electrons to higher
states is greater than the first ionization energy, and therefore these
form discrete states within a continuum of ionized He+ +e− states). The
excited states can have S = 0 or S = 1, with S = 1 lying lower in energy
(Hund).
$ The lines in the S = 0 system are all singlets. They can be observed
in emission, and those starting from the ground state can be seen in
absorption.
$ The lines in the S = 1 system are all multiplets. They can be observed
in emission only. Transitions of the form 3 S1 ↔3 P2,1,0 are observed as
triplets, spaced according to the Landé interval rule. Transitions of the
form 3 P2,1,0 ↔3 D3,2,1 are observed as sextuplets, as is easily seen by
application of the ∆J = ±1, 0 rule. Actually, as mentioned above, the
fine structure is a little more subtle in the case of helium.
The alkali earths follow the same principles. In the case of calcium, the
triplet 4p state is the lowest lying triplet state, and therefore metastable. In
fact a faint emission line corresponding to the 3 P1 →1 S0 decay to the ground
state may be observed; this violates the ∆S = 0 rule, and indicates that the LS
coupling approximation is not so good in this case. A more extreme example is
seen in Mercury, ground state (6s)2 (5d)10 . Excited states involving promotion
of one of the 6s electrons to a higher level can be treated just like the alkali
Figure 9.4: The well known doublet which is responsible for the bright yellow light
from a sodium lamp may be used to demonstrate several of the influences which
cause splitting of the emission lines of atomic spectra. The transition which gives
rise to the doublet is from the 3p to the 3s level. The fact that the 3s state is lower
than the 3p state is a good example of the dependence of atomic energy levels on
orbital angular momentum. The 3s electron penetrates the 1s shell more and is less
effectively shielded than the 3p electron, so the 3s level is lower. The fact that there
is a doublet shows the smaller dependence of the atomic energy levels on the total
angular momentum. The 3p level is split into states with total angular momentum
J = 3/2 and J = 1/2 by the spin-orbit interaction. In the presence of an external
magnetic field, these levels are further split by the magnetic dipole energy, showing
dependence of the energies on the z-component of the total angular momentum.
where µB denotes the Bohr magneton. Therefore, we see that all degenerate
levels are split due to the magnetic field. In contrast to the “normal” Zeeman
effect, the magnitude of the splitting depends on '.
$ Info. If the field is strong, the Zeeman energy becomes large in comparison
with the spin-orbit contribution. In this case, we must work with the basis states
|n, ', m! , ms & = |n, ', m! & ⊗ |ms & in which both Ĥ0 and ĤZeeman are diagonal. Within
first order of perturbation theory, one then finds that (exercise)
! "4 ! "
1 Zα 3 n nm! ms
∆En,!,m! ,ms = µB (m! + ms ) + mc2 − − ,
2 n 4 ' + 1/2 '(' + 1/2)(' + 1)
the first term arising from the Zeeman energy and the remaining terms from Ĥrel. . At
intermediate values of the field, we have to apply degenerate perturbation theory to
the states involving the linear combination of |n, j = '±1/2, mj , '&. Such a calculation
reaches beyond the scope of these lectures and, for details, we refer to the literature
(see, e.g., Ref. [6]. Let us instead consider what happens in multi-electron atoms.
argument. Since 2L̂ · Ŝ = Ĵ2 − L̂2 − Ŝ2 , this operator is diagonal in the basis of
states, |J, MJ , L, S&. Therefore, the matrix element of the operator (exercise,
hint: recall that [Ŝi , Ŝj ] = i!"ijk Ŝk and [L̂i , Ŝk ] = 0),
must vanish. Moreover, from the identity [L̂ · Ŝ, Ĵ] = 0, it follows that the
matrix element of the vector product,
must also vanish. If we expand the left hand side, we thus find that the matrix
element of
L̂=Ĵ−Ŝ
(Ŝ × L̂) × Ĵ = L̂(Ŝ · Ĵ) − Ŝ(L̂ · Ĵ) = Ĵ(Ŝ · Ĵ) − ŜĴ2 ,
also vanishes. Therefore, it follows that %ŜĴ2 & = %Ĵ(Ŝ · Ĵ)&, where the expec-
tation value is taken over the basis states. Then, with Ŝ · Ĵ = 12 (Ĵ2 + Ŝ2 − L̂2 ),
we have that
J(J + 1) + S(S + 1) − L(L + 1) Figure 9.5: In the weak field
%Ŝz & = %Jˆz & . case, the vector model (top) im-
2J(J + 1)
plies that the coupling of the or-
As a result, we can deduce that, at first order in perturbation theory, the bital angular momentum L to
energy shift arising from the Zeeman term is given by the spin angular momentum S
is stronger than their coupling
∆EJ,MJ ,L,S = µB gJ MJ B , to the external field. In this
case where spin-orbit coupling is
where dominant, they can be visual-
ized as combining to form a to-
J(J + 1) + S(S + 1) − L(L + 1) tal angular momentum J which
gJ = 1 + , then precesses about the mag-
2J(J + 1) netic field direction. In the
strong field case, S and L cou-
denotes the effective Landé g-factor, which lies between 1 and 2. Note that, ple more strongly to the exter-
in the special case of hydrogen, where S = 1/2 and J = L ± 1/2, we recover nal magnetic field than to each
our previous result. The predicted Zeeman splitting for sodium is shown in other, and can be visualized as
figure 9.4. independently precessing about
the external field direction.
$ Info. In the strong field limit, where the influence of Zeeman term domi-
nates, the appropriate basis states are set by |L, ML , S, MS &, in which the operators
L̂2 , L̂z , Ŝ 2 , Ŝz , and ĤZeeman are diagonal. In this case, the energy splitting takes the
form
! "4
1 Zα nML MS
∆EL,ML ,S,MS = µB B(ML + 2MS ) + mc2 ,
2 n '(' + 1/2)(' + 1)
10.1 The H+
2 ion
The simplest system that exhibits molecular properties is the hydrogen ion
H+2 , which consists of two protons with positions Ra , Rb and one electron at
r. With the potential energy
$ %
e2 1 1 1
V (r, Ra , Rb ) = − − ,
4π#0 |Ra − Rb | |r − Ra | |r − Rb |
Although this equation can actually be solved exactly using elliptical polar
coordinates, it will be more instructive for our purposes to seek an approximate
method of solution. Since there is no obvious parameter in which to develop
a perturbative expansion, we will instead follow a variational route to explore
the low energy states of system.
If the electron is close to one of the protons, one would expect the other
proton to have a small influence on its dynamics, and that the wavefunction
in this region would be close to that of a hydrogen atomic orbital. Therefore,
in seeking the ground state of the H+ 2 ion, we may take a trial wavefunction
that is a linear combination of the ground state (1s) wavefunctions centred on
the two protons,
where ψa,b = (πa30 )−1/2 exp[−|r − Ra,b |/a0 ], represents the corresponding hy-
drogenic wavefunction with a0 the atomic Bohr radius. In this case, the co-
efficients α and β can be taken as real. The variational expression to be
minimized in order to estimate the ground state energy is given by
where Haa = #ψa |Ĥ|ψa $ = Hbb = #ψb |Ĥ|ψb $, and Hab = #ψa |Ĥ|ψb $ = #ψb |Ĥ|ψa $.
Note that the matrix elements Haa = Hbb because Ĥ is symmetric with re-
spect to Ra and Rb . Moreover, since ψa and ψb are not orthogonal we have
to introduce the overlap integral, S = #ψa |ψb $, which measures the overlap
between the two atomic wavefunctions. In fact we can simplify this expres-
sion further, because the potential is symmetric about the mid-point between
the two protons. The wavefunction must therefore be either symmetric or
antisymmetric, α = ±β, and hence,
Haa ± Hab
E0 ≤ #E$ = .
1±S
The matrix elements in this expression can be evaluated in closed form, though
the calculation is rather tedious.1
We have, therefore, found two possible wavefunctions for the H+ 2 ion,
ψa + ψb ψa − ψb
ψg = ( , ψu = ( ,
2(1 + S) 2(1 − S)
−Hab
with energies Eg = Haa1+S+Hab
, Eu = Haa1−S . The subscript g refers to the term
gerade (German for even) used in molecular physics to denote a state that is
even under the operation of inverting the electronic wavefunction through
the centre of symmetry of the molecule, without changing the positions of the
nuclei. Such an inversion changes r → Ra + Rb − r, which interchanges ψa and
ψb . Note that this is not the same as parity inversion, which would also affect
the nuclear coordinates. The ungerade (odd) state is denoted by subscript u.
Note that ψg and ψu are orthogonal, even though ψa and ψb are not. In fact,
ψg and ψu are just the orthonormal states that diagonalize the Hamiltonian,
if we limit ourselves to linear combinations of ψa and ψb . In chemistry, ψg and
ψu are called molecular orbitals and the assumption that they are linear
combinations of atomic stationary states is called the linear combination
of atomic orbitals (LCAO) approximation.
The state ψg has the lower energy, while ψu represents an excited state of
the molecular ion. Physically, the reason for this is that, in the ψg state, the E/eV
two atomic wavefunctions interfere constructively in the region between the 4!
3 Eu + Ry
protons, giving an enhanced electron density in the region where the electron #
is attracted strongly by both protons, which serves to screen the two protons 2 #
$
from each other. Conversely, in ψu we have destructive interference in the 1
region between the protons. If we plot Eg and Eu as functions of the nuclear 0 " R/pm
separation R = |Ra − Rb |, the results are as shown in the figure right. For 100 200 300
–1
both curves, we have plotted E + Ry since −Ry is the ground state energy of %
&
%
–2 %
&
the hydrogen atom. The curve of Eg + Ry in the LCAO approximation has a %% LCAO Eg + Ry
–3 % Exact E + Ry
minimum value of −1.8eV at R = R0 = 130pm, which is the predicted equi- g
where R = |Ra − Rb | and Ry is the Rydberg constant, the binding energy of the Hydrogen
atom in its ground state.
& Info. At this stage, it is helpful to introduce some notation to label the
molecular orbitals. Although the wavefunctions ψg,u that we have been discussing
are formed in the LCAO approximation from linear combinations of atomic 1s (n = 1,
' = 0) states, with no orbital angular momentum, they do not themselves necessarily
have zero orbital angular momentum. An ' = 0 state must be proportional to Y00 ,
i.e., it must have no dependence on θ and φ, giving an isotropic probability distri-
bution. But these states are certainly not isotropic: they have a ‘dumbbell’ shape,
concentrated around the two protons. They do not have unique electronic orbital
angular momentum because the operator L̂2 for the electron does not commute with
the Hamiltonian on account of the non-central terms 1/|r − Ra | and 1/|r − Rb | in the
potential.
The only component of L̂ that does commute with these terms is L̂z , provided
we choose the z-axis parallel to the internuclear axis Ra − Rb . Therefore instead of
classifying the states as s, p, d, . . . orbitals according to whether ' = 0, 1, 2, . . ., we
call them σ, π, δ, . . . orbitals according to whether Λ = 0, 1, 2, . . ., where Λ ≡ |m! |.
A subscript u or g denotes whether the state is even or odd under inversion; this
notation can be applied to all homonuclear diatomic molecules, in which the potential
is symmetric about the median plane of the molecule. Thus the ground state of
the hydrogen molecular ion is σg and the corresponding odd state is σu∗ , where the
star signifies an antibonding orbital. In the LCAO approximation used above, these
molecular orbitals are linear combinations of 1s atomic orbitals and so they can be
written as 1sσg and 1sσu∗ . To get some insight into how this denotation applies, it is
helpful to refer to the figures on the right.
Table 10.1: Binding energy and equilibrium nuclear separation of the hydrogen
molecule in various approximations.
expectation values of 1/r12 and 1/rab will be comparable and therefore the
extra term can be treated as a perturbation. Thus, as a first approximation
we neglect it and assign each electron to one of the H+ 2 molecular orbitals
defined above. There are four ways of filling the two orbitals σg and σu∗ , which
we can represent by
ψg (r1 )ψg (r2 ), ψg (r1 )ψu∗ (r2 ), ψu∗ (r1 )ψg (r2 ), ψu∗ (r1 )ψu∗ (r2 ) .
ψg (r1 )ψg (r2 ) ∝ [ψa (r1 )ψb (r2 ) + ψb (r1 )ψa (r2 )] + [ψa (r1 )ψa (r2 ) + ψb (r1 )ψb (r2 )] .
The terms in the first square bracket correspond to the two electrons being
shared between the two hydrogen atoms. This is the covalent bonding pic-
ture of the bound state. In the other square bracket, however, both electrons
are assigned to the same atom, corresponding to ionic bonding. Since all the
terms have equal coefficients, the ionic and covalent contributions are equal,
which seems rather constraining, if not implausible. For example, it means
that, when the two protons are pulled apart, the system is just as likely to be
found to consist of an H+ and an H− ion as two neutral atoms.
If we go to the pure valence bonding (VB) approximation and drop the
ionic part of the wavefunction altogether, we find that the predicted binding
energy and nuclear separation are both improved (see Table 10.2). Including
a small parameter λ for the amplitude of the ionic component, i.e., taking
ψ V B ∝ [ψa (r1 )ψb (r2 ) + ψb (r1 )ψa (r2 )] + λ[ψa (r1 )ψa (r2 ) + ψb (r1 )ψb (r2 )] ,
& Info. As mentioned above, there are four possible ways of putting two electrons
into the 1sσg and 1sσu∗ molecular orbitals. From these four two-electron states we
can make three states that are symmetric under interchange of the positions of the
electrons, all of which need to be combined with the antisymmetric spin state X0,0 ,
1
Σg : X0,0 σg (r1 )σg (r2 )
√
1
Σu : X0,0 [σg (r1 )σu∗ (r2 ) + σu∗ (r1 )σg (r2 )]/ 2
1
Σg : X0,0 σu∗ (r1 )σu∗ (r2 ) .
We have introduced to the left of these equations a new notation, called the E
molecular term, to describe the overall quantum numbers of the molecule, all of
!
which must be good quantum numbers because they correspond to operators which
commute with the molecular Hamiltonian. It derives from a historic spectroscopic
notation used in atomic physics. The term is written 2S+1 Λu/g . The prefix 2S + 1
denotes the multiplicity of the total spin state of the electrons, hence 1 for a singlet and
3 for a triplet state. The central greek capital letter represents the magnitude of the H+ +H− σ ∗2 "
total L̂z quantum number, Λ = 0, 1, 2 . . . being represented by Σ, Π, ∆ . . .. In the case u
of the molecular orbitals above based on the 1s atomic orbitals, all the wavefunctions σg2
clearly have zero orbital angular momentum about the internuclear axis and hence H+H !"
R
they are all Σ states. The g or u suffix means even or odd under inversion, and is
only meaningful for homonuclear molecules. Notice that since (±1)2 = +1 we get a
g-state by combining two g or two u-states, and a u-state by combining a g and a
u-state.
In terms of the molecular orbital approach, the VB approximation implies that
Molecular potential energy and
the ground state is not simply the σg2 configuration but rather a mixture of the two
1
Σg states, configuration mixing in H2 .
ψ V B ∝ (1 + λ)(1 + S)σg (r1 )σg (r2 ) − (1 − λ)(1 − S)σu∗ (r1 )σu∗ (r2 ) .
Thus there is configuration mixing in the hydrogen molecule. The two states
can mix because they have the same overall quantum numbers. At large nuclear
separations the energy eigenstates are clearly separated into those that are almost a
pair of neutral atoms and those consisting of an H+ and an H− ion. In the case of
the u-states, we can see by expanding the spatial wave functions that 3 Σu is of the
former type and 1 Σu of the latter. Since these configurations have different resultant
electron spins they do not mix significantly (just like singlet and triplet states of
atomic helium). Also the u- and g-states are prevented from mixing by their different
inversion symmetry. However, the two 1 Σg configurations have the same electron spin
and symmetry and can mix to give the above covalent-bonded ground state and an
orthogonal excited state that is more ionic.
Here the second term represents the interaction of the electrons with the con-
situent nuclei, and the third term involves the electron-electron interaction. In
a physical system, we would have to take into account the influence of further
relativistic corrections which would introducte additional spin-orbit couplings.
To address the properties of such a complex interacting system, we will
have to draw upon many of the insights developed previously. To begin, it is
helpful to partition the electrons into those which are bound to the core and
those which are able to escape the potential of the individual atomic nuclei
and propagate “freely” through the lattice. The electrons which are tightly
bound to the nuclei screen the nuclear charge leading to a modified nuclear
potential, Veff (r). Focussing on those electrons which are free, the effective
+ + e2 e2
Hamiltonian can be written as, Ĥ * n Ĥn + m,n 4π" 0 rmn
, where
!2 ∇2n
Ĥn = − + Veff (rn )
2me
represents the single-particle Hamiltonian experienced by each of the electrons
– i.e. Ĥn describes the motion of an electron moving in a periodic lattice
potential, Veff (r) = Veff (r + R) with R belonging to the set of periodic lattice
vectors.
Despite engineering this approximation, we are still confronted by a chal-
lenging many-particle problem. Firstly, the problem remains coupled through
the electron-electron Coulomb interaction. Secondly, the electrons move in a
periodic potential. However, if we assume that the electrons remain mobile –
the jargon is itinerant, free to propagate through the lattice, they screen each
other and diminish the effect of the Coulomb interaction. Therefore, we can
Vion (r) are associated with a set of atomic orbitals, ψq , characterized by a set of
quantum numbers, q. In the atomic limit, when the atoms are far-separated,
these will mirror the simple hydrogenic states. To find the variational ground
state of the system, we can then build a trial state from a linear combination
of these atomic orbitals. Taking only the lowest orbital, q = 0, into account
we have,
!
ψ(r) = αR ψ(r − R) ,
R
where, as before, αR represent the set of variational coefficients, one for each
site (and, in principle, each atomic orbital if we had taken more than one!).
Once again, we can construct the variational state energy,
+ ∗
#ψ|Ĥ0 |ψ$ R,R" αR HRR" αR"
E= = + ∗ ,
#ψ|ψ$ R,R" αR SRR" αR"
,
where, as before, HRR" = dd rψ ∗ (r − R)Ĥ0 ψ(r − R% ) denote the,matrix ele-
ments of the orbital wavefunction on the Hamiltonian and SRR" = dd rψ ∗ (r−
R)ψ(r − R% ) represent the overlap integrals. Then, varying the energy with
respect to αR∗ , we find that the coefficients obey the secular equation (ex-
ercise)
!
[HRR" − ESRR" ] αR" = 0 .
R"
Note that, if the basis functions were orthogonal, this would just be an eigen-
value equation.
& Exercise. To develop this idea, let us first see how the method relates to
back to the problem of H+
2 : In this case, the secular equation translates to the 2 × 2
matrix equation,
$ %
Haa − E Hab − ES
α = 0,
Hab − ES Haa − ES
where the notation and symmetries of matrix elements follow from section 10.1. As
a result, we find that the states α divide into even and odd states as expected.
for each n, where Hnn = ε denotes the atomic orbital energy, Hn,n+1 =
Hn+1,n = −t is the matrix element between neighbouring states, Sn,n = 1
and Sn+1,n = Sn+1,n = S. All other matrix elements are exponentially small.
Here we consider a system with N lattice sites and impose the periodic bound-
ary condition, αn+N = αn . The periodicity of the secular equation suggests
a solution of the from αn = √1N eikna , where k = 2πm/N a denote a dis-
crete set of N reciprocal lattice vectors with m integer lying in the range
−N/2 < m ≤ N/2. Substitution confirms that this is a solution with energies
(exercise),
ε − 2t cos(ka)
Ek = .
1 + S cos(ka)
& Exercise. Consider how the calculation above can be extended to a two-
dimensional square lattice system. What would be the corresponding band dispersion?
& Info. To add flesh to these ideas, let us then consider a simple, but prominent
problem from the realm of quantum condensed matter physics. In recent years, there
has been great interest in the properties of graphene, a single layer of graphite.
Remarkably, high quality single crystals of graphene can be obtained by running
graphite – a pencil! – over an adhesive layer. The resulting electron states of the
single layer compound have been of enormous interest to physicists. To understand
why, let us implement the LCAO technology to explore the valence electron structure
of graphene.
Graphene forms a periodic two-dimensional honeycomb lattice structure with two Atomically resolved STM image
atoms per each unit cell. With an electron configuration 1s2 2s2 2p2 , the two 1s of graphene sheet at 77K.
electrons are bound tightly to the nucleus. The remaining 2s electrons hybridize
with one of the p orbitals to form three sp2 hydridized orbitals. These three orbitals
form the basis of a strong covalent bond of σ orbitals that constitute the honeycomb
lattice. The remaining electron, which occupies the out-of-plane pz orbital, is then
Figure 10.1: Left: dispersion relation of graphene Ek obtained with the LCAO
approximation. Notice that near the centre of the band, the dispersion becomes
point-like. Around these points, the dispersion takes form of a linear (Dirac) cone.
Right: band structure of a sample of multilayer epitaxial graphene. Linear bands
emerge from the K-points in the band structure. Three “cones” can be seen from
three layers of graphene in the MEG sample.
capable of forming an itinerant band of electron states. It is this band which we now
address.
Once again, let us suppose that the wavefunction of this band involves the basis
of single pz orbital states localized to each lattice site,
!
ψ(r) = [αR ψ1 (r − R) + βR ψ2 (r − R)] .
R
Here the pz orbital wavefunction ψ1 is centred on one of the atoms in the unit cell, and
ψ2 is centred on the other. Once again, taking into account matrix elements involving
only nearest neighbours, the trial wavefunction translates to the secular equation,
(ε − E)αR − (t + ES)(βR + βR−a1 + βR−a2 ) = 0
(ε − E)βR − (t + ES)(αR + αR+a1 + αR+a2 ) = 0 ,
√ √
where the lattice vectors a1 = ( 3/2, 1/2)a and a2 = ( 3/2, −1/2)a, with a the
lattice spacing, are shown in the figure to the right. Note that the off-diagonal matrix
elements involve only couplings between atoms on different sublattices. Once again,
we can make the ansatz that the solutions are of the form of plane waves with αR =
√αk ik·R
N
e and βR = √βN k
eik·R . Notice that, in this case, we must allow for different
relative weights, αk and βk . Substituting, we find that this ansatz is consistent if
(ε − E)αk − (t + ES)fk βk = 0
(ε − E)βk − (t + ES)fk∗ αk = 0
√
where fk = 1+2e−i 3kx a/2 cos(ky a/2). Although this equation can be solved straight-
forwardly, it takes a particularly simple form when the overlap integral between neigh-
bouring sites, S, is neglected. In this case, one obtains,
Ek = ε ± |fk |t .
The corresponding band structure is shown right as a function of k. In particular, one
may note that, at the band centre, the dispersion relation for the electrons becomes
point-like (see Fig. 10.3).
In the half-filled system, where each carbon atom contributes a single electron to
the band, the Fermi level lies precisely at the centre of the band where the dispersion,
Ek is point like. Doping electrons into (or removing electrons from) the system
results in (two copies) of a linear dispersion, Ek * c|k|, where is c is a constant
(velocity). Such a linear dispersion relation is the hallmark of a relativistic particle
(cf. a photon). Of course, in this case, the electrons are not moving at relativistic
velocities. But their properties will mirror the behaviour of relativistic particles. And
herein lies the interest that these materials have drawn.
where µ is the reduced mass of the two atoms. In this case, E0 (R) acts as a
central potential, and the usual separation into angular and radial equations
can be carried out.
The simplest solutions are the purely rotational states, in which the whole
molecule rotates around its centre of mass. The solutions will be the spherical
harmonic functions YJ,mJ . Conventionally the quantum numbers are labelled
J and mJ (mJ taking values J, J − 1, · · · , −J), and the corresponding energy
is given by
!2
EJ = J(J + 1) ,
2I
where I = µR02 is the moment of inertia of the molecule about an axis through
the centre of mass orthogonal to the bond, and R0 is the equilibrium bond
length.
As mentioned earlier, the typical energies of rotational states of molecules
are much smaller than those of electronic excited states. Since molecular
dimensions are determined by the electronic wavefunction, their scale is set by
the Bohr radius a0 . Thus moments of inertia are of order mN a20 and the scale
of rotational energies is !2 /mN a20 . For the electronic states, the uncertainty
relation implies momenta of order !/a0 and hence electron energies around
!2 /me a20 , a factor of mN /me ∼ 104 greater.
To bring about a radiative rotational transition, an emitted or absorbed
photon must interact with the electric dipole moment of the molecule. Since
the initial and final electronic states are the same, this state needs to have
a permanent electric dipole moment. Thus we can have purely rotational
radiative transitions in heteronuclear diatomic molecules like HCl and CO,
which have permanent dipole moments, but not in homonuclear ones like
H2 and O2 .4
The usual electric dipole selection rules apply; ∆J = ±1, 0 with a parity
change. In a rotational state with angular momentum quantum numbers J
and mJ , the nuclear wavefunction φ({rN }) is proportional to the spherical
harmonic YJmJ , which has parity (−1)J . (For simplicity, we consider only
molecular states in which the electronic wavefunction has zero angular mo-
mentum and even parity, such as the diatomic 1 Σg states.) Then the fact
4
However, we can produce purely rotational transitions in all types of molecules by the
process of Raman scattering, in which a photon is effectively absorbed and then reemitted
by a molecule, since the virtual intermediate state can have a dipole moment.
that the parity must change in a radiative transition excludes the possibil-
ity ∆J = 0. Therefore the possible energy changes in emission (J + 1 → J;
J = 0, 1, 2 · · ·) are given by:
!2 !2
∆E = [(J + 1)(J + 2) − J(J + 1)] = (J + 1) .
2I I
In fact, the rate for spontaneous emission between rotational states is very
small, because of the small energy release (varying as ω 3 ), and so rotational
transitions are more conveniently studied by absorption spectroscopy. The
same formula for the energies of transition clearly applies to the J → (J + 1)
absorption case. Therefore the spectrum is expected to consist of equally
spaced lines, separated by energy !2 /I. Observation of this spacing can be
used to determine the moment of inertia and hence the bond length of the
molecule. Strictly speaking the spacing isn’t quite uniform, because the radial
Schrödinger equation acquires a centrifugal potential term !2 J(J + 1)/2µR2
which means that the equilibrium separation slightly increases with J and
consequently the moment of inertia increases and the line spacing decreases.
& Exercise. Use perturbation theory to estimate the strength of the effect of
the centrifugal term.
& Info. The intensities of rotational spectral lines show some interesting
features. Although the transition matrix element depends on the quantum numbers,
the dominant factor is usually the population of the initial state. As mentioned
earlier, non-radiative transitions due to molecular collisions bring about a thermal
distribution,
This increases with J up to some value, which depends on the temperature, and then
decreases. Thus successive spectral lines increase and then decrease in intensity.
E = E0 (R0 ) + (n + 1/2)!ω, n = 0, 1, 2, . . . .
grounds, ∂R 2 √
0 R0 will be of order ! /me a0 , and !ω ∼ ! / me mN a0 . There-
2E | 2 4 2
fore
( the vibrational energy is smaller than the electronic by a factor of order
me /mN . This puts vibrational spectra in the wavelength region around
10µm, which is in the infra-red.
We can now check explicitly that the Born-Oppenheimer approximation
is valid for nuclear vibrational states, as follows.
( The mean square nuclear
vibrational momentum
( is of order mN !ω ∼ mN /me (!/a0 )2 , which means
that ∇N φk ∼ mN /me φk /a20 , where φk is the nuclear part of the wavefunc-
2
tion. On the other hand ∇2N ψk ∼ ψk /a20 , where ψ(k is the electronic part. Thus
φk ∇2N ψk is smaller than ψk ∇2N φk by a factor of me /mN , and it is legitimate
to neglect the former.
For vibrational transitions we have the selection rule (exercise)
∆n = ±1 .
∆E = (En+1 − En ) = !ω ,
where the first term accounts for the kinetic energy of the particles whilst the
second describes their coupling.1 For convenience, we adopt periodic boundary
1
In real solids, the inter-atomic potential is, of course, more complex than our quadratic
approximation. Yet, for “weak coupling”, the harmonic contribution dominantes (cf. our
discussion of molecular vibrations). For the sake of simplicity we, therefore, neglect the
effects caused by higher order contributions.
" Exercise. Starting with the discrete form of the Lagrangian, or otherwise,
show the classical equations of motion take the form,
Remembering that the boundary conditions are periodic, obtain the normal modes. Hint: consider the ansatz,
From this result, determine the condition under which the continuum approximation φn (t) = ei(kna−ωt) .
can be justified.
ρ 2 κs a2
L(φ̇, φ) = φ̇ − (∂x φ)2 , (11.2)
2 2
denotes the Lagrangian density, ρ = m/a denotes the mass per unit length
and κs = ks /a. The corresponding classical action is given by
%
S[φ] = dt L[φ] . (11.3)
d
(∂ẋ L) − ∂x L = 0 . (11.5)
dt
Now, in Eq. (11.3), we are dealing with a system of infinitely many degrees
of freedom, φ(x, t). Yet Hamilton’s principle is general, and we may see what
happens if (11.3) is subjected to an extremal principle analogous to Eq. (11.4).
To do so, we must effect the substitution φ(x, t) → φ(x, t) + &η(x, t) into
Eq. (11.3) and demand that the contribution first order in & vanishes. When
applied to the specific Lagrangian (11.2), a substitution of the of the ‘varied’
field leads to
% % L ' (
S[φ + &η] = S[φ] + & dt dx ρ φ̇η̇ − κs a2 ∂x φ∂x η + O(&2 ).
0
Integrating by parts (with respect to time for the first term under the inte-
gral, and space in the second) and demanding that the contribution linear
3
In the mathematics and physics literature, mappings of functions into the real or com-
plex numbers are generally called functionals. The argument of a functional is commonly
indicated in rectangular brackets [ · ]. For example, in this case, S maps the ‘functions’
∂x φ(x, t) and φ̇(x, t) to the real number S[φ].
& &L
in & vanishes, one obtains dt 0 dx(ρφ̈ − κs a2 ∂x2 φ)η = 0. (Notice that the
boundary terms associated with both t and x vanish identically.4 Now, since η
was defined to be any arbitrary smooth function, the integral above can only
vanish if the term in parentheses is globally vanishing. Thus the equation of
motion takes the form of a wave equation,
The solutions
) of Eq. (11.6) have the general form φ+ (x + vt) + φ− (x − vt)
where v = a κs /ρ, and φ± are arbitrary smooth functions of their argument.
From this we can deduce that the basic low energy elementary excitations
of our model are lattice vibrations propagating as sound waves to the left or
right at a constant velocity v (see figure). The trivial behaviour of our model
is of course a direct consequence of its simplistic definition — no dissipation,
dispersion or other non-trivial ingredients. Adding these refinements leads to
the general classical theory of lattice vibrations. With this background, let us
now turn to the consider the quantization of the quantum mechanical chain.
In common with φ(x, t), the canonical momentum, π(x, t), is a continuum
degree of freedom. At each space point it may take an independent& value.
From the Lagrangian, we can define the Hamiltonian, H[φ, π] ≡ dx H(φ, π),
where H(φ, π) ≡ π φ̇−L(φ̇, φ). represents the Hamiltonian density. Applied
to the atomic chain (11.2), the canonical momentum π = ρφ̇ and H(φ, π) =
π2 κs a2
2ρ + 2 (∂x φ) .
2
1 2 κs a2
Ĥ(φ̂, π̂) = π̂ + (∂x φ̂)2 .
2ρ 2
4
If we assume that the function φ already obeys the boundary conditions, we must have
η(0, t) = η(L, t) = η(x, 0) = η(x, T ) = 0).
5
Note that the dimensionality of both the quantum and classical continuum fields is
compatible with the dimensionality of the Dirac δ-function, [δ(x − x! )] = [Length]−1 .
Promoting the displacements and momenta to operators, and applying the canonical
quanitization conditions, [p̂n , φn! ] = −i!δnn! , obtain the discrete form of the Hamilto-
nian. Taking the continuum limit, show that the Hamiltonian recovers the continuum
form derived through the Lagrangian formulation.
" Exercise. Making use of Eqs. (11.8) and (11.9) derive the canonical commu-
tation relation above.
δk+k! ,0
% , % -. /
L ! 1 L
!
!
dx (∂ φ̂)2 = (ik φ̂k )(ik " φ̂k! ) dx ei(k+k )x = k 2 φ̂k φ̂−k ,
0 !
L 0
k,k k
&L
together with the parallel relation for 0 dx π̂ 2 , the Hamiltonian assumes the
“near diagonal” form,
!" 1 1 2
#
Ĥ = π̂k π̂−k + ρωk φ̂k φ̂−k , (11.10)
2ρ 2
k
where ωk = v|k|, and v = a(κs /ρ)1/2 denotes the classical sound wave ve-
locity. In this form, the Hamiltonian can be identified as nothing more than
a superposition of independent quantum harmonic oscillators. The only dif-
ference between (11.10) and the canonical form of an oscillator Hamiltonian
p2
H = 2m + 12 mω 2 x2 is the presence of sub-indices k and −k (a consequence
†
of φ̂k = φ̂−k ). As we will show shortly, this difference is inessential. This
result is actually not difficult to understand (see figure): Classically, the sys-
tem supports a discrete set of wave-like excitations, each indexed by a wave
number k = 2πm/L. Within the quantum picture, each of these excitations is
described by an oscillator Hamiltonian with a k-dependent frequency. How-
ever, it is important not to confuse the atomic constituents, also oscillators
(albethey coupled), with the independent collective oscillator modes described
by Ĥ.
The description above, albeit perfectly valid, still suffers from a deficiency:
Our analysis amounts to explicitly describing the effective low energy excita-
tions of the system (the waves) in terms of their microscopic constituents (the
atoms). Indeed the different contributions to Ĥ keeps track of details of the
microscopic oscillator dynamics of individual k-modes. However, it would be
much more desirable to develop a picture where the relevant excitations of the
system, the waves, appear as fundamental units, without explicit account of
underlying microscopic details. (As with hydrodynamics, information is en-
coded in terms of collective density variables rather than through individual
molecules.) To understand how this programme can be achieved let us recall
the properties of the quantum harmonic oscillator.
" Info. In quantum mechanics, the harmonic oscillator has the status of a
single-particle problem. However, the fact that the energy levels, &n = !ω(n + 1/2),
are equidistant suggests an alternative interpretation: One can think of a given energy
state &n as an accumulation of n elementary entities, or quasi-particles, each having
energy !ω. What can be said about the features of these new objects? First, they
are structureless, i.e. the only ‘quantum number’ identifying the quasi-particles is
their energy !ω (otherwise n-particle states formed of the quasi-particles would not
be equidistant). This implies that the quasi-particles must be bosons. (The same
state !ω can be occupied by more than one particle — see figure.) This idea can be
formulated in quantitative terms by employing the formalism of ladder operators in
which )the operators p̂ and )
x̂ are traded for the pair of Hermitian adjoint operators
a ≡ mω 2! (x̂ + mω
i
p̂), a†
≡ 2! (x̂ − mω p̂). Up to a factor of i, the transformation
mω i
(x̂, p̂) → (a, a† ) is canonical, i.e. the new operators obey the canonical commutation
relation, [a, a† ] = 1. More importantly, in the a-representation, the Hamiltonian
takes the simple form, Ĥ = !ω(a† a + 1/2), as can be checked by direct substitution.
The complete hierarchy of higher energy states can be generated by setting |n& ≡
√1 (a† )n |0&.
n!
While the a-representation provides another way of constructing eigenstates of
the quantum harmonic oscillator, its real advantage is that it naturally affords a
many-particle interpretation. Temporarily forgetting about the original definition of
the oscillator, we can declare |0& to be a ‘vacuum’ state, i.e. a state with no particles
present. a† |0& then represents a state with a single featureless particle (the operator
a† does not carry any quantum number labels) of energy !ω. Similarly, (a† )n |0& is
considered as a many-body state with n particles, i.e. within the new picture, a† is
an operator that creates particles. The total energy of these states is given by !ω ×
(occupation number). Indeed, it is straightforward to verify that a† a|n& = n|n&, i.e.
the Hamiltonian basically counts the number of particles. While, at first sight, this
may look unfamiliar, the new interpretation is internally consistent. Moreover, it
fulfils our objective: it allows an interpretation of the excited states of the harmonic
oscillator as a superposition of independent structureless entities.
With this background, we may return to the harmonic atomic chain (11.10)
Eqs. (11.11) and (11.12) represent the final result of our analysis: The low-lying
elementary excitations of the discrete atomic chain are described by oscillator
Figure shows a typical measured
wave-like modes – known as phonons – each characterised by a wavevector k phonon dispersion of an ordered
and a linear dispersion, ωk = v|k|. A generic state of the system is given by crystalline solid obtained by neu-
1 tron scattering. The x-axis in-
|{nk } = (n1 , n2 , · · ·)& = )3 (a†k1 )n1 (a†k2 )n2 · · · |0& . dexes wavenumbers along a lat-
i ni ! tice direction (specified in units
of π/a). Three generic aspects
The representation derived above illustrates the capacity to think about
are visible: (1) near k = 0, the
quantum problems in different complementary “pictures”, a principle that dispersion is, as expected, linear.
finds innumerable applications. The existence of different interpretations of a (2) The several branches are as-
given system is by no means heretic but, rather, is consistent with the spirit sociated with different “polariza-
of quantum mechanics. Indeed, it is one of the prime principles of quan- tions” of the lattice fluctuations.
tum theories that there is no such thing as ‘the real system’ which underpins (3) For wavelengths comparable
the phenomenology. The only thing that matters is observable phenomena. to the lattice spacing, k ∼ π/a,
non-universal features specific to
For example, the ‘fictitious’ quasi-particle states of the harmonic chain, the
the particular material become
phonons, behave as ‘real’ particles, i.e. they have dynamics, can interact, be visible.
detected experimentally, etc. From a quantum point of view there is actually
no fundamental difference between these objects and ‘real’ particles.
" Example: Debye theory of solids: Our analysis above focussed on the
longitudinal vibrations of the one-dimensional atomic chain. In three-dimensions,
each mode is associated with three possible polarizations, λ: two transverse and one
longitudinal. Taking into account all degrees of freedom, it is straightforward to show
Peter Josephus Wilhelmus De-
that the generalized Hamiltonian takes the form, bye 1884-1966:
! 1 2 Dutch-American
† 1
Ĥ = !ωk ak,λ ak,λ + , physicist
2 renowned for his
kλ
work on molec-
ular structure,
where, for simplicity, we assume that the dispersion, ωk = v|k| is independent of
especially dipole
polarization. Let us use this result to obtain the internal energy and specific heat moments and
due to phonons. Now, for an equilibrium thermal distribution, the average phonon the diffraction
occupancy of state (k, λ) is given simply by the Bose-Einstein distribution, nB (!ωk ) ≡ of X-rays and
electrons in
1
e!ωk /kB T −1
. The internal energy is therefore given by gases. Debye was awarded a Nobel
! ! " # Prize in Chemistry, 1936, “for his
1 1 contributions to our knowledge of
E= !ωk (nB (!ωk ) + 1/2) = !ωk + .
kλ kλ
e!ωk /kB T − 1 2 molecular structure through his
investigations on dipole moments
and on the diffraction of X-rays and
In the thermodynamic limit, where N ≡ V /a3 → ∞, we may replace the sum over electrons in gases”.
+ & kD d3 k & kD 2
modes by an integral, k → V 0 (2π)3 = 2π 2 0 k dk, where kD , denotes the
V
6
As for the consistency of these definitions, recall that φ̂†k = φ̂−k and π̂k† = π̂−k . Under
these conditions the second of the definitions below indeed follows from the first upon taking
the Hermitian conjuate.
largest wave vector accessible in the crystal. We can fix kD by ensuring that the
total number of modes (for each polarization) matches the total number of degrees
2 1/3
of freedom, i.e. (2π)13 /V 43 πkD
3
= N , i.e. kD = (6π a) . The corresponding frequency
scale, ωD = vkD , is known as the Debye frequency. In this limit, dropping the tem-
perature independent contribution from zero point fluctuations, the internal energy
per particle is given by
%
E 3a3 kD 2 !vk
ε≡ = k dk !vk/k T .
N 2π 2 0 e B −1
& T /T z 3 dz
Then, defining the Debye temperature, TD = !vkD /kB , we have ε = 9kB T ( TTD )3 0 D ez −1 .
The corresponding specific heat per particle can be obtained from the temperature
derivative and leads to
1 23 % TD /T
T z 4 dz
cV = 9kB .
TD 0 (ez − 1)2
In particular, at high temperatures, we recover the Dulong-Petit law, cV = 3kB
following from the equipartition theorem – each degree of freedom is associated with
an energy kB T /2. At low temperatures, T # TD , we may replace the upper limit on
the integral by ∞ from which we find that cV ∼ T 3 . Both limits compare well with
experiment (see figure).
This completes our discussion of the classical and quantum field theory of
the harmonic atomic chain. In this example, we have seen how we can effect a
quantum formulation of a continuum system. Using the insights obtained in
this example, we now turn to consider the quantization of the electromagnetic
field.
uum permeability,
0 −Ex /c −Ey /c −Ez /c
Ex /c 0 Bz −By
Fµν = ∂µ Aν − ∂ν Aµ =
Ey /c
−Bz 0 Bx
Ez /c By −Bx 0 µν
1
Ä = ∇2 A .
c2
The structural similarity between the EM field and the continuous formu-
lation of the harmonic chain is clear. By analogy with our discussion above,
we should now switch to the Fourier representation and quantize the classical
field. However, in contrast to our analysis of the chain, we are now dealing (i)
with the full three-dimensional Laplacian acting upon (ii) the vector field A
that is (iii) subject to the constraint ∇ · A = 0. It is these aspects which lead
to the complications outlined above.
We can circumvent these difficulties by considering cases where the geom-
etry of the system reduces the complexity of the eigenvalue problem while still
retaining the key conceptual aspects of the problem. This restriction is less
artificial than it might appear. For example, just as the field φ in the classical
atomic chain can be expanded in Fourier harmonics, in long waveguides, the
EM vector potential can be expanded in solutions of the eigenvalue equation9
where xb parameterize points on the boundary of the system, and E$ (B⊥ ) is the
parallel (perpendicular) component of the electric (magnetic) field. Applied to the
problem at hand, let us consider a long cavity with uniform rectangular cross-section
Ly × Lz . To conveniently represent the Lagrangian of the system, we wish to express
the vector potential in terms of eigenfunctions uk that are consistent with the bound-
ary conditions (11.15). A complete set of functions fulfilling this condition is given
by
c1 cos(kx x) sin(ky y) sin(kz z)
uk = Nk c2 sin(kx x) cos(ky y) sin(kz z) .
c3 sin(kx x) sin(ky y) cos(kz z)
1 22
2 π
uk = √ sin(πy/Ly ) sin(kx) êz , λk = k +
2
, (11.16)
V Ly
Returning to the problem posed by (11.13) and (11.14), one + can expand
the vector potential in terms of eigenfunctions uk as A(x, t) = k αk (t)uk (x),
where the sum runs over all allowed values of the index parameter k. (In a
waveguide of length L, k = πn/L with n integer.) Substituting this expan-
sion into (11.13), and using the normalization properties of uk , we obtain the
Lagrangian,
" #
1 ! 1 2
L[α̇, α] = α̇ − λ 2
k k ,
α
2µ0 c2 k
k
where ωk2 = c2 λk .
Then, guided by the analysis of the atomic chain, we now introduce the
ladder operators,
0 1 2 0 1 2
&0 ωk i † &0 ωk i
ak = α̂k + π̂k , ak = α̂k − π̂k ,
2! &0 ωk 2! &0 ωk
whereupon the Hamiltonian assumes the now familiar form
! 1 2
† 1
Ĥ = !ωk ak ak + . (11.17)
2
k
For the specific problem of the first excited mode in a waveguide of width
Ly , !ωk = c[k 2 + (π/Ly )2 ]1/2 . Eq. (11.17) represents our final result for the
quantum Hamiltonian of the EM waveguide. Before concluding this section
let us make a few comments on the structure of the result.
" Firstly, notice that the construction above almost completely paralleled
our previous discussion of the atomic chain.10 The structural similarity
between the two systems finds its origin in the fact that the free field
Lagrangian (11.13) is quadratic in the fields and, therefore, bound to
map onto an oscillator-type Hamiltonian. That we obtained a simple
one-dimensional superposition of oscillators is due to the boundary con-
ditions specific to a narrow waveguide. For less restrictive geometries,
e.g. free space, a more complex superposition of vectorial degrees of free-
dom in three-dimensional space would have been obtained (see below).
However, the principal mapping of the free EM field onto a superposition
of oscillators is independent of geometry.
" Physically, the quantum excitations described by (11.17) are, of course,
the photons of the EM field. The unfamiliar appearance of the dis-
persion ωk is again a peculiarity of the waveguide. However, in the
limit of large longitudinal wave numbers k - L−1 y , the dispersion ap-
proaches ωk . c|k|, i.e. the familiar linear (relativistic) dispersion of
the photon field. Also notice that, due to the equality of the Hamilto-
nians (11.12) and (11.17), all that has been said about the behavior of
the phonon modes of the atomic chain carries over to the photon modes
of the waveguide.
10
Technically, the only difference is that, instead of index pairs (k, −k), all indices (k, k)
are equal and positive. This can be traced back to the fact that we have expanded in terms
of the real eigenfunctions of the closed waveguide instead of the complex eigenfunctions of
the periodic oscillator chain.
" As with their phonon analogue, the oscillators described by (11.17) ex-
hibit zero-point fluctuations. It is a fascinating aspect of quantum elec-
trodynamics that these oscillations, caused by quantization of the most
relativistic field, surface at various points of non-relativistic physics, e.g.
the attraction of two conducting plates in vacuum – the Casimir effect.
With the analysis of the waveguide complete, let us go back and consider
the quantitization of the full three-dimensional system. For the waveguide, we
have found
+ that the vector potential can be expanded in modes of the cavity as
= k α̂k uk where, rearranging the expressions for the ladder operators,
Â(x) :
! †
α̂k = 2'0 ωk (ak + ak ). More generally, in a fully three-dimensional cavity, one
may show that11
;
! ! < =
Â(x) = êkλ akλ eik·x + ê∗kλ a†kλ e−ik·x ,
2&0 ωk V
kλ=1,2
where V denotes the volume of the system, ωk = c|k|, and the two sets of
polarization vectors, êkλ , are in general complex and normalized to unity,
ê∗kλ · êkλ = 1. To ensure that the vector potential satifies the Coulomb gauge
condition, we require that êkλ · k = ê∗kλ · k = 0, i.e. the two polarization
vectors are orthogonal to the wave vector. Two real vectors, êkλ correspond
to two linear polarizations while, for circular polarization, the vectors are
complex. It is also convenient to assume that the two polarization vectors are
mutually orthogonal, êkλ · êkµ = δµν . The corresponding operators obey the
commutation relations,
With these definitions, the Hamiltonian then takes the familiar form
! < =
Ĥ = !ωk a†kλ akλ + 1/2 , (11.18)
kλ
while, defining the vacuum, |Ω&, the eigenstates involve photon number states,
11 P R
In the infinite system, the mode sum becomes replaced by an integral, k → V
(2π)3
d3 k.
Time-dependent perturbation
theory
where VI (t)!= eiĤ0 t/!V e−iĤ0 t/!. Then, if we form the eigenfunction expansion,
|ψ(t)!I = n cn (t)|n!, and contract the equation of motion with a general
state, "n|, we obtain
"
i!ċm (t) = Vmn (t)eiωmn t cn (t) , (12.2)
n
1
Note how this definition differs from that of the Heisenberg representation, |ψ!H =
iĤt/!
e |ψ(t)!S in which all time-dependence is transferred into the operators.
where the matrix elements Vmn (t) = "m|V (t)|m!, and ωmn = (Em − En )/! =
−ωnm . To develop some intuition for the action of a time-dependent potential,
it is useful to consider first a periodically-driven two-level system where the
dynamical equations can be solved exactly.
$ Info. The two-level system plays a special place in the modern development
of quantum theory. In particular, it provides a platform to encode the simplest
quantum logic gate, the qubit. A classical computer has a memory made up of
bits, where each bit holds either a one or a zero. A quantum computer maintains a
sequence of qubits. A single qubit can hold a one, a zero, or, crucially, any quantum
superposition of these. Moreover, a pair of qubits can be in any quantum superposition
of four states, and three qubits in any superposition of eight. In general a quantum
computer with n qubits can be in an arbitrary superposition of up to 2n different
states simultaneously (this compares to a normal computer that can only be in one
of these 2n states at any one time). A quantum computer operates by manipulating
those qubits with a fixed sequence of quantum logic gates. The sequence of gates to
be applied is called a quantum algorithm.
An example of an implementation of qubits for a quantum computer could start
with the use of particles with two spin states: |↓! and |↑!, or |0! and |1!). In fact any
system possessing an observable quantity A which is conserved under time evolution
and such that A has at least two discrete and sufficiently spaced consecutive eigenval-
ues, is a suitable candidate for implementing a qubit. This is true because any such
system can be mapped onto an effective spin-1/2 system.
Specifying the wavefunction by the two-component vector, c(t) = (c1 (t) c2 (t)), Eq. (12.2)
translates to the equation of motion (exercise)
# $
0 ei(ω−ω21 )t
i!∂t c = δ c(t) ,
e−i(ω−ω21 )t 0
where ω21 = (E2 − E1 )/!. With the initial condition c1 (0) = 1, and c2 (0) = 0, this
equation has the solution,
δ2
|c2 (t)|2 = sin2 Ωt, |c1 (t)|2 = 1 − |c2 (t)|2 ,
δ 2 + !2 (ω − ω21 )2 /4
where Ω = ((δ/!)2 +(ω −ω21 )2 /4)1/2 is known as the Rabi frequency. The solution,
which varies periodically in time, describes the transfer of probability from state 1 to
state 2 and back. The maximum probability of occupying state 2 is a Lorentzian with
γ2
|c2 (t)|2max = ,
γ 2 + !2 (ω − ω21 )2 /4
taking the value of unity at resonance, ω = ω21 .
$ Exercise. Derive the solution from the equations of motion for c(t). Hint:
eliminate c1 from the equations to obtain a second order differential equation for c2 .
$ Info. The dynamics of the driven two-level system finds practical application
in the Ammonia maser: The ammonia molecule NH3 has a pryramidal structure
with an orientation characterised by the position of the “lone-pair” of electrons sited
on the nitrogen atom. At low temperature, the molecule can occupy two possible
states, |A! and |S!, involving symmetric (S) or an antisymmetric (A) atomic con-
figurations, separated by a small energy splitting, ∆E. (More precisely, along the
axis of three-fold rotational symmetry, the effective potential energy of the nitrogen
Charles Hard Townes 1915-
atom takes the form of a double-well. The tunneling of the nitrogen atom through the
(left)
double well leads to the symmetric and asymmetric combination of states.) In a time- is an American
dependent uniform electric field the molecules experience a potential V = −µd · E, Nobel prize-
where E = Eêz cos ωt, and µd denotes the electric dipole moment. Since µd is odd un- winning physicist
and educator.
der parity transformation, P µd P = −µd , and P |A! = −|A! and P |S! = |S!, the ma-
Townes is known
trix elements of the electric dipole moment are off-diagonal: "S|µd |S! = "A|µd |A! = 0 for his work on
and "S|µd |A! = "S|µd |A! = & 0. the theory and
If we start with all of the molecules in the symmetric ground state, we have application of the maser – microwave
amplification by stimulated emission
shown above that the action of an oscillating field for a particular time can can drive of radiation, on which he got the
a collection of molecules from their ground state into the antisymmetric first excited fundamental patent, and other work
in quantum electronics connected
state. The ammonia maser works by sending a stream of ammonia molecules, traveling with both maser and laser devices.
at known velocity, down a tube having an oscillating field for a definite length, so the He received the Nobel Prize in
Physics in 1964.
molecules emerging at the other end are all (or almost all, depending on the precision
of ingoing velocity, etc.) in the first excited state. Application of a small amount of
electromagnetic radiation of the same frequency to the outgoing molecules will cause
some to decay, generating intense radiation and therefore a much shorter period for
all to decay, emitting coherent radiation.
cn (t) = c(0)
n + cn (t) + cn (t) + · · · ,
(1) (2)
(m) (0)
where cn ∼ O(V m ) and cn is some (time-independent) initial state. The
programme to complete this series expansion is straightforward but technical.
This result provides a self-consistent equation for UI (t, t0 ), i.e. if we take this expres-
sion and substitute UI (t" , t0 ) under the integrand, we obtain
% t # $2 % t % t!
i i
UI (t, t0 ) = I − dt" VI (t" ) + − dt" VI (t" ) dt"" VI (t"" )UI (t"" , t0 ) .
! t0 ! t0 t0
"∞ # $n % t % tn−1
i
UI (t, t0 ) = − dt1 · · · dtn VI (t1 )VI (t2 ) · · · VI (tn ) , (12.3)
n=0
! t0 t0
where the term n = 0 translates to I. Note that the operators VI (t) are organised in
a time-ordered sequence, with t0 ≤ tn ≤ tn−1 ≤ · · · t1 ≤ t. With this understanding,
we can write this expression more compactly as
& i Rt ! ! '
− dt VI (t )
UI (t, t0 ) = T e ! t0 ,
where “T” denotes the time-ordering operator and its action is understood by Eq. (12.3).
cn (t)
" ( )* +
|i, t0 , t! = UI (t, t0 )|i! = |n! "n|UI (t, t0 )|i! .
n
!
Making use of Eq. (12.3), and the resolution of identity, m |m!"m| = I, we
obtain
c(0)
n c(1)
n c(2)
n
()*+ ( )* +( )* +
% t % t % t! "
i 1
cn (t) = δni − dt" "n|VI (t" )|i! − 2 dt" dt"" "n|VI (t" )|m!"m|VI (t"" )|i! + · · · .
! t0 ! t0 t0 m
where Vnm (t) = "n|V (t)|m! and ωnm = (En − Em )/!, etc. In particular, the
probability of effecting a transition from state |i! to state |n! for n &= i is given
(1) (2)
by Pi→n = |cn (t)|2 = |cn (t) + cn (t) + · · · |2 .
From the expression for the Golden rule (12.5) we see that, for transitions to
occur, and to satisfy energy conservation:
(a) the final states must exist over a continuous energy range to match ∆E =
!ω for fixed perturbation frequency ω, or
(b) the perturbation must cover a sufficiently wide spectrum of frequency so
that a discrete transition with a fixed ∆E = !ω is possible.
For two discrete states, since |Vfi |2 = |Vif |2 , we have the semiclassical result
Pi→f = Pf→i – a statement of detailed balance.
This is a transition in which the system gains energy 2!ω from the harmonic
perturbation, i.e. two “photons” are absorbed in the transition, the first taking
the system to the intermediate energy ωm , which is short-lived and therefore
not well defined in energy – indeed there is no energy conservation requirement
for the virtual transition into this state, only between initial and final states.
Of course, if an atom in an arbitrary state is exposed to monochromatic light,
other second order processes in which two photons are emitted, or one is
absorbed and one emitted (in either order) are also possible.
Radiative transitions
p̂2
Ĥ0 = + V (r) ,
2m
denotes the usual non-interacting Hamiltonian of the isolated atom,
e
Ĥpara (t) = A(t) · p̂ ,
m
represents the time-dependent paramagnetic term arising from the coupling of
the electron to the electromagnetic field, and Ĥdia = (eA)2 /2m represents the
diamagnetic term. Since we will be interested in the absorption and emission
of single photons, we can neglect the influence of diamagnetic term which
presents only a tiny perturbation in the atomic system.
Previously, in chapter 11.2, we have see that the quantum Hamiltonian for
the electromagnetic field can be expressed as,
! " #
1
Ĥrad = !ωk a†kλ akλ + ,
2
k,λ=1,2
where the operators a†kλ and akλ create and annihilate photons with wavevec-
tor k and polarization λ, and ωk = c|k|. These ladder operators obey the
(bosonic) commutation relations, [akλ , a†k! λ! ] = δk,k! δλ,λ! , with [akλ , ak! λ! ] =
[a†kλ , a†k! λ! ] = 0, and act on photon number states as
√
akλ |nkλ " = nkλ |nkλ − 1"
√
a†kλ |nkλ " = nkλ + 1|Nkλ + 1" .
Here |nkλ " represents a photon number state with nk,λ photons in the mode
(kλ). Finally, in the Heisenberg representation, we have seen that the vector
potential can be expanded in field operators as
$
! ! % &
A(r, t) = êkλ akλ ei(k·r−ωk t) + ê∗kλ a†kλ e−i(k·r−ωk t) .
2%0 ωk V
k,λ
Taken together, Ĥ = Ĥ0 + Ĥrad + Ĥpara (t) specify the full quantum mechanical
Hamiltonian of the atom light system.
ωk ! !p
k·r& & & Zα .
c p mc
This means that, for Zα ( 1, we can expand the exponential as a power
series in k · r with the lowest terms being dominant. Taking the zeroth order
term, and making use of the operator identity, p̂ = !i [Ĥ0 , r] which follows from
the Heisenberg equations of motion for operators, the matrix element may be
written as
i Ef − Ei ∗
$f|ê∗kλ · p̂|i" = mê∗kλ · $f| [Ĥ0 , r]|i" = im êkλ · $f|r|i" = −imωk $f|ê∗kλ · r|i" .
! !
This result, which emerges from the leading approximation in Zα, is known
as the electric dipole approximation. Effectively, we have set (exercise)
e
Â(r, t) · p̂ & eÊ(r, t) · r ,
m
In particular, we find that the absorption rate increases linearly with photon
number, nkλ .
Similarly, if we now consider the emission process in which there is are
already nk,λ photons in the initial state, we have the revised transition rate,
Schematic showing the stimu-
' $ '2
' ' lated emission from an initial
2π '' e !(nk,λ + 1) −ik·r ∗ '
state at energy Ei = E2 to a final
Γi→f,kλ = ' $f| e êkλ · p̂|i"'' δ(Ef − Ei − !ωk ) .
! ' m 2%0 ωk V ' state at energy Ef = E1 .
If there are no photons present initially, this expression reduces to that ob-
tained from spontaneous emission. The nkλ -independent component of the
tion of state (kλ) is isotropic, dependent only on |k|, we find that the integrated
* k2 dk
transition rate per unit solid angle is given by dR dΩ = V
λ
Γ
(2π)3 i→f,kλ
from
which we obtain
(
dRλ 1 ω3 nλ (ω) absorption
= |$f|êkλ · d|i"|2
dΩ 4π%0 2π!c3 nλ (ω) + 1 emission
Here, in carrying out the integral, we have used the relation ωk = c|k| and
!ω = |Ef − Ei |. For a thermal distribution of photons, with the energy density
specified by the Planck formula,
!ω 3 1
u(ω) = n̄λ (ω), n̄λ (ω) = ,
πc3 e!ωk /kB T −1
this equates a stimulated absorption/emission rate,
dRλ 1 1
= |$f|êkλ · d|i"|2 u(ω)
dΩ 4π%0 2!2
From these expressions, we can obtain the power loss as Pλ = !ωRλ . Before
discussing the selection rules implied by the form of the dipolar coupling, it is
first helpful to digress and discuss connections of this result to a famous result
due to Einstein.
where u(ω), the energy density of radiation per unit ω. A and B are known as
Einstein’s A and B coefficients, and, as we have seen, are properties of the atomic
states concerned.
Now, in thermodynamic equilibrium, the rates must balance, so that
At the same time, the relative populations of the two states (assumed non-degenerate
for simplicity), are given by a Boltzmann factor
nj e−Ej /kB T
= −E /k T = e!ω/kB T .
nk e k B
Thus we have: + ,
Ak→j (ω) = Bj→k e!ω/kB T − Bk→j u(ω) . (13.1)
For a black-body, the energy density u(ω) is just given by Planck’s formula, u(ω) =
!ω 3
π 2 c3 e!ω/kB T −1 . The Ak→j coefficient in Eq. (13.1) certainly cannot depend on tem-
1
!ω 3
Bk→j = Bj→k and Ak→j (ω) = Bk→j .
π 2 c3
So, the A and B coefficients are related, and if we can calculate the B coefficient for
stimulated emission from Fermi’s golden rule, we can infer A, and vice versa.
The spin term $χf |χi " (and therefore the matrix element) vanishes unless |χi "
and |χf " are identical. This can be expressed by the selection rule
∆s = 0, ∆ms = 0 .
Similarly, since $+% , m% |[L̂z , x ± iy]|+, m" = ±$+% , m% |x ± iy|+, m", it follows that
∆m$ = 0, ±1 .
Similarly, using operator identity [L̂2 , [L̂2 , r]] = 2!2 (rL̂2 + L̂2 r) (exercise),
we have
$+% , m% |[L̂2 , [L̂2 , r]]|+, m" = [+% (+% + 1) − +(+ + 1)]2 $+% , m% |r|+, m"
= 2[+% (+% + 1) + +(+ + 1)]$+% , m% |r|+, m"
∆+ = ±1 .
One may summarize the selection rules for + and m$ is by saying that the
photon carries off (or brings in, in an absorption transition) one unit of angular
momentum. It should be noted, however, that these rules were derived for the
specific case of an electric dipole transition of the system. It is possible, though
much less likely in the case of an atom, for the electromagnetic field to interact
with some other observable such as the magnetic dipole moment or the electric
quadrupole moment. In such transitions the selection rules are different. For
example, the magnetic dipole operator is µ̂ = −µB L̂/! (or −2µB Ŝ/! for the
spin) and since the angular momentum does not change sign under the parity
transformation, there is no change of parity in a magnetic dipole transition. To
avoid confusion, we shall continue to confine the discussion to electric dipole
transitions, which are responsible for the prominent lines in atomic spectra.
For transitions with ∆m$ = 0, the dipole matrix element $f|d|i" ∼ êz and
there is no component of polarization along z-direction. Similarly, for electric
dipole transitions with m% = m ± 1, $+% , m% |x ∓ iy|+, m" = 0 = $+% , m% |z|+, m",
and $f|d|i" ∼ (1, ∓i, 0). In this case, if the wavevector of photon lies along
z, the emitted light is circularly polarized with a polarization which depends
on helicity. Conversely, if the wavevector lies in xy place, the emitted light is
linearly polarized, while in general the polarization is elliptical.
Finally, in the presence of spin-orbit coupling, stationary states are labelled
by quantum numbers J, mJ , +, s where Ĵ = L̂ + Ŝ. In this case, the selection
rules can be inferred by looking for the conditions for non-zero matrix elements
$J % , mJ ! , +% , s% |r|J, mJ , +, s". By expanding states |J, mJ , +, s" in the basis states
|+, m$ "⊗|s, ms ", one may uncover the following set of selection rules: For dipole
transitions to take place, we require that
13.3 Lasers
Finally, to close this section, we will consider a principle application of light Arthur Leonard Schawlow 1921-
matter interaction – the laser. The laser provides a light source which enables 1999
modern spectroscopy. The term “laser” is an acronym for “light amplification American
physicist and
by stimulated emission of radiation”. However, a laser not only amplifies light, corecipient, with
but it acts as a special kind of light source which is characterised by a number Nicolaas Bloem-
bergen of the US
of properties: and Kai Manne
Monochromaticity: The emission of the laser generally corresponds to Borje Siegbahn
just one of the atomic transitions of the gain medium (in contrast to discharge of Sweden, of
the 1981 Nobel
lamps, which emit on all transitions). The spectral line width can be much Prize for Physics for his work in
smaller than that of the atomic transition. This is because the emission is developing the laser and in laser
spectroscopy. In 1949 he went to
affected by the optical cavity. In certain cases, the laser can be made to Columbia University, where he began
operate on just one of the modes of the cavity. Since the Q of the cavity1 collaborating with Charles Townes on
the development of masers, lasers,
is generally rather large, the mode is usually much narrower than the atomic and laser spectroscopy. Schawlow
transition, and the spectral line width is orders of magnitude smaller than the worked on the project that led to
atomic transition. the construction of the first working
maser in 1953 (for which Townes
Coherence: In discussing the coherence of an optical beam, we must received a share of the 1964 Nobel
distinguish between spatial and temporal coherence – laser beams have a high Prize for Physics). Schawlow was a
research physicist at Bell Telephone
degree of both. Spatial coherence refers to whether there are irregularities in Laboratories from 1951 to 1961.
the optical phase in a cross-sectional slice of the beam. Temporal coherence In 1958 he and Townes published
a paper in which they outlined
refers to the time duration over which the phase of the beam is well defined. the working principles of the laser,
In general, the temporal coherence time, tcoh is given as the reciprocal of the though the first such working device
spectral line width, ∆ν. Thus the coherence length +coh is given by, was built by another American
physicist, Theodore Maiman, in
c 1960. In 1961 Schawlow became
+coh = ctcoh = . a professor at Stanford University.
∆ν He became a world authority on
laser spectroscopy, and he and
Typical values of the coherence length for a number of light sources are given Bloembergen earned their share
in the table below: of the 1981 Nobel Prize by using
lasers to study the interactions
of electromagnetic radiation with
Source ∆ν (Hz) tcoh (s) +coh (m) matter.
Na discharge lamp 5 × 1011 2 × 10−12 6 × 10−4
(D-lines at 589nm)
Multi-mode HeNe laser 1.5 × 109 6 × 10−10 0.2
(632.8nm line)
Single-mode HeNe laser 1 × 106 1 × 10−6 300
(632.8nm line)
W (N2 (n + 1) − N1 n)
where n denotes the total number of photons in the cavity, and N1,2 is the
number of atoms in states 1, 2. The first term represents the contribution
from stimulated and spontaneous emission, while the latter is associated with
absorption. Taking into account photon loss from the leaky cavity, the rate of
change of photon number is therefore given by
n
ṅ = DW n + N2 W − , (13.2)
τph
where w12 , w21 denote the “effective” transition rates between states 1 and
2 due to the pumping via the third state, and we have dropped the small
contribution from spontaneous emission. From this equation, we can deduce
that N1 + N2 = N , a constant, i.e. the decay from state 3 is so rapid that its
population is always negligible. In this case, we obtain
D0 − D
Ḋ = − 2DW n . (13.3)
T
D0 W − 1/τph
n= .
2T W/τph
When D0 > 1/W τph , the laser threshold condition, there is a rapid increase
in the number of photons in the cavity and the system starts lasing.
Although this analysis addresses the threshold conditions, it does not pro-
vide any insight into the coherence properties of the radiation field. In fact,
one may show that the radiation field generated by the laser cavity forms a
coherent or Glauber state. The proof of this statement and the coherence
properties that follow would take us too far into the realm of laser physics.
However, we can gain some insight into this statement by studying a toy ex-
ample.
To solve for the time-evolution operator, i!∂t UI (t) = VI UI (t), let us consider
the coherent state, |α" = Û (α)|0" where Û (α) = exp[αa† − α∗ a], where α =
α(t). Equivalently, making use of the BCH identity, the unitary operator may
be written as
∗ α/2 † ∗a
Û (α) = e−α eαa e−α .
† †
Then, taking the time derivative, and making use of the identity, [eαa , a]e−αa =
−α, one obtains
2 3
† ∗ i ∗
∂t Û (α) = α̇a − α̇ a + Im(α̇ α) Û (α) .
2
*t % %
Therefore, setting α(t) = −i 0 dt f˜(t ), we obtain the solution
+ ,
UI (t) = exp α(t)a† − α∗ (t)a + iϕ(t) ,
† †
2
Here we have made use of the identity (exercise), eiωta a ae−iωta a
= e−iωt a.
*t
where ϕ(t) = 0 dt% 12 Im(α̇∗ α).
*t
If the driving force f (t) = f0 e−iωt (with f0 real), we have α(t) = −i 0 dt% f0 =
−if0 t and ϕ(t) = 0 leading to the solution, UI (t) = exp[−if0 (a† + a)t]. There-
fore, if the system was prepared in the harmonic oscillator ground state |0" at
time t = 0, the solution at time t is given by |ψ(t)"I = exp[−if0 (a† + a)t]|0" =
2 †
e−(f0 t) /2 e−if0 a t |0". Then, reexpressed in the Schrödinger representation,
2 /2 −iωt a† t
|ψ(t)"S = e−iĤ0 t/!|ψ(t)"I = e−(f0 t) e−if0 e |0" .
Scattering theory
Almost everything we know about nuclei and elementary particles has been
discovered in scattering experiments, from Rutherford’s surprise at finding
that atoms have their mass and positive charge concentrated in almost point-
like nuclei, to the more recent discoveries, on a far smaller length scale, that
protons and neutrons are themselves made up of apparently point-like quarks.
More generally, the methods that we have to probe the properties of condensed
matter systems rely fundamentally on the notion of scattering. In this sec-
tion, we will provide a brief introduction to the concepts and methodology of
scattering theory.
As preparation for the quantum mechanical scattering problem, let us first
consider the classical problem. This will allow us to develop (hopefully a
revision!) some elementary concepts of scattering theory, and to introduce
some notation. In a classical scattering experiment, one considers particles
of energy E = 12 mv02 (mass m and asymptotic speed v0 ), incident upon a
target with a central potential V (r). For a repulsive potential, particles are
scattered through an angle θ (see figure). The scattering cross-section, σ,
can be inferred from the number of particles dn scattered into some element
of solid angle, dΩ, at angle (θ, φ), i.e. for an incident flux ji (number of
particles per unit time per
! unit area), dn! π = ji σ dΩ.
! 2πThe total cross-section is
then obtained as σT = dΩ σ(θ, φ) = 0 sin θdθ 0 dφ σ(θ, φ). The angle of
deflection of the beam depends on the impact parameter, b (see figure right).
We therefore have that dn = ji bdb dφ = ji σ sin θdθdφ and
b db
σ(θ, φ) = .
sin θ dθ
L2
r= ,
mκ(e cos ϕ − 1)
where r = (r, ϕ) parameterises the relative coordinates of the particle and target,1
2
and e = (1 + 2EL
κ2 m )
1/2
> 1 denotes the eccentricity. Since the potential is central, the
angular momentum L is conserved and can be fixed asymptotically by the condition
L = mv0 b.
To obtain the scattering angle, θ, we can use the relation above to find the limiting
angle, cos ϕ0 = 1/e, where ϕ0 = (π − θ)/2. We therefore have tan(θ/2) = cot ϕ0 =
1
Note that the angle ϕ is distinct form the azimuthal angle φ associated with the axis of
scattering.
√ mκ2 1/2
1/ e2 − 1 = ( 2EL2) = 2Eb .
κ
Then, from this relation, we obtain the cross-section
b db κ2 1
σ= = ,
sin θ dθ 16E 2 sin4 θ/2
known as the Rutherford formula.
14.1 Basics
Let us now turn to the quantum mechanical problem of a beam of particles
incident upon a target. The potential of the target, V (r), might represent that
experienced by a fast electron striking an atom, or an α particle colliding with
a nucleus. As in the classical problem, the basic scenario involves directing
a stream or flux of particles, all at the same energy, at a target and detect
how many particles are deflected into a battery of detectors which measure
angles of deflection. In principle, if we assume that all the in-going particles
are represented by wavepackets of the same shape and size, our challenge is
to solve the full time-dependent Schrödinger equation for such a wavepacket,
" #
!2 2
i!∂t Ψ(r, t) = − ∇ + V (r) Ψ(r, t) ,
2m
and find the probability amplitudes for out-going waves in different directions
at some later time after scattering has taken place. However, if the incident
beam of particles is switched on for times very long as compared with the time
a particle would take to cross the interaction region, steady-state conditions
apply. Moreover, if we assume that the wavepacket has a well-defined energy
(and hence momentum), so it is many wavelengths long, and we may consider
it a plane wave. Setting Ψ(r, t) = ψ(r)e−iEt/!, we may therefore look for
solutions ψ(r) of the time-independent Schrödinger equation,
" #
!2 2
Eψ(r) = − ∇ + V (r) ψ(r) ,
2m
subject to the boundary condition that the incoming component of the wave-
function is a plane wave, eik·x . Here E = p2 /2m = !2 k2 /2m denotes the
energy of the incoming particles while their flux is given by
! !k
j = −i (ψ ∗ ∇ψ − ψ∇ψ ∗ ) = .
2m m
In the one-dimensional geometry, the impact of a plane wave with the
localized target resulted in a portion of the wave being reflected and a portion
transmitted through the potential region. From energy conservation, we may
deduce that both components of the outgoing scattered wave are plane waves
with wavevector ±k, while the influence of the potential are encoded in the
amplitude of the reflected and transmitted beams, and a potential phase shift.
Both amplitudes and phase shifts are then determined by solving the time-
independent Schrödinger equation subject to the boundary conditions which
ensure energy and flux conservation. In the three-dimensional system, the
phenomenology is similar: In this case, the wavefunction well outside the
localized target region will involve a superposition of the incident plane wave
and the scattered (spherical wave),2
eikr
ψ(r) $ eik·r + f (θ, φ) ,
r
2
Here, by localized, we mean a potential which is sufficiently short-ranged. At this stage,
it is not altogether clear what constraint this implies. But it will turn out that it excludes
the Coulomb potential!
where the function f (θ, φ) records the relative amplitude and phase of the
scattered components along the direction (θ, φ) relative to the incident beam.
where P" (cos θ) = ( 2"+1 ) Y"0 (θ) denote the Legendre polynomials. If we
4π 1/2
assume that the potential perturbation, V (r) depends only on the radial coor-
dinate (i.e. that it is spherically symmetric) and that the number of particles
are conserved by the potential (the flux of incoming particles is matched by
the flux of outgoing),3 when the potential is sufficient short-ranged (decreasing
faster than 1/r), the scattering wavefunction takes the asymptotic form
∞
% &
i $ " e−i(kr−"π/2) ei(kr−"π/2)
ψ(r) $ i (2+ + 1) − S" (k) P" (cos θ) ,
2k r r
"=0
subject to the constraint |S" (k)| = 1 following from the conservation of particle
flux (i.e. S" (k) = e2iδ! (k) ). Physically, the incoming component of the spherical
wave is undisturbed by the potential while the separate components of the
outgoing spherical wave are subject to a set of phase shifts, δ" (k). Recast in
the form of a perturbation, the asymptotic form of the wavefunction can be
straightforwardly rewritten as
eikr
ψ(r) $ eik·r + f (θ) ,
r
where the second component of the wavefunction denotes the change in the
outgoing spherical wave due to the potential, and
∞
$
f (θ) = (2+ + 1)f" (k)P" (cos θ) , (14.1)
"=0
The first term represents the incident flux, while the remainder describes the
radial flux of scattered particles. In particular, the number of particles crossing
the area that subtends a solid angle dΩ at the origin (the target) is given by
!k |f (θ)|2 2
j · êr dA = r dΩ + O(1/r) .
m r2
Dropping terms of order 1/r, negligible in the asymptotic limit, one thus
obtains the differential cross-section, the ratio of the scattered flux to the
incident flux, dσ = !k
m
j · êr dA = |f (θ)|2 dΩ, i.e.
dσ
= |f (θ)|2 .
dΩ
! !
The total cross-section is then
! given by σtot = dσ = |f (θ)|2 dΩ. Then,
making use of the identity dΩP" (cos θ)P"! (cos θ) = 2"+1
4π
δ""! , and Eq. (14.1)
one obtains (exercise)
∞
4π $
σtot = 2 (2+ + 1) sin2 δ" (k) .
k
"=0
In particular, noting that P" (1) = 1, from Eq. (14.1) it follows that Im f (0) =
4π σtot , a relation known as the optical theorem.
k
At high energies, many channels contribute to the total scattering amplitude. How-
ever, at low energies, the scattering is dominated by the s-wave (+ = 0) channel.
In this case, setting u(r) = rR0 (r), the radial equation takes the simple form,
(∂r2 + U0 θ(R − r) + k 2 )u(r) = 0, with the boundary condition that u(0) = 0. We
therefore obtain the solution
)
C sin Kr r<R
u(r) = ,
sin(kr + δ0 ) r > R
1 1
k cot δ0 = − + r0 k 2 + · · · ,
a0 2
1/2
where a0 = (1 − tan γ
γ )R, with γ = U0 R, defines the scattering length, and r0 is
the effective range of the interaction. At low energies, k → 0, the scattering cross-
section, σ0 = 4πa20 is fixed by the scattering length alone. If γ ' 1, a0 is negative.
As γ is increased, when γ = π/2, both a0 and σ0 diverge – there is said to be a zero
energy resonance. This condition corresponds to a potential well that is just able Scattering wavefunction, u(r),
to support an s-wave bound state. If γ is further increased, a0 turns positive – as for three-dimensional square well
it would be for an effective repulsive interaction until γ = π when σ0 = 0 and the potential for kR = 0.1 and γ = 1
process repeats with the appearance of a second bound state at γ = 3π/2, and so on. (top), π/2 (middle) and 2 (bot-
More generally, the +-th partial cross-section tom). Note that the scattering
length, a0 changes from nega-
4π 1 tive to positive as system passes
σ" = (2+ + 1) ,
k 2 1 + cot2 δ" (k) through bound state.
takes its maximum value is there is an energy at which cot δ" vanishes. If this occurs
as a result of δ" (k) increasing rapidly through an odd multiple of π/2, the cross-section
exhibits a narrow peak as a function of energy and there is said to be a resonance.
Near the resonance,
ER − E
cot δ" (k) = ,
Γ(E)/2
where ER is the resonance energy. If Γ(E) varies slowly in energy, the partial cross-
section in the vicinity of the resonance is given by the Breit-Wigner formula,
Scattering phase shift for kR =
4π Γ (ER )/4
2 0.1 as a function of γ.
σ" (E) = (2+ + 1) . (14.2)
k2 (E − ER )2 + Γ2 (ER )/4
4
More generally, choosing the solution to be finite at the origin, we find that
R! (r) = N! (K)j! (Kr), r < R,
where N! (K) is a normalization constant. In the exterior region, the general solution can
be written as R! (r) = B! (k)[j! (kr) − tan δ! (k)η! (kr)]. Continuity of R! and the derivative
∂r R! at the boundary, r = R, lead to the following expression for the phase shifts
kj!! (kR)j! (KR) − Kj! (kR)j!! (KR)
tan δ! (k) = .
kη!! (kR)j! (KR) − Kη! (kR)j!! (KR)
Here j!! (x) = ∂x j! (x) and similarly η!! .
$ Info. Ultracold atomic gases provide a topical arena in which resonant scatter-
ing phenomena are exploited. In particular, experimentalists make use of Feshbach
resonance phenomena to tune the effective interaction between atoms. This tun-
ability arises from the coupling of free unbound atoms to a molecular state in which
the atoms are tightly bound. The closer this molecular level lies with respect to the
energy of two free atoms, the stronger the interaction between them. In the example
on the left, the two free atoms are both “spin up”, whereas the molecular state is a
“singlet”, in which the atoms have opposite spin, adding up to zero total magnetic
moment. Thus, a magnetic field shifts the energies of two free atoms relative to the
molecular state and thereby controls the interatomic interaction strength.
The interaction between two atoms can be described by the scattering length,
shown right as a function of magnetic field close to a Feshbach resonance. On the side
where the scattering length is positive, the molecular energy level is lower in energy
than the energy of two unbound atoms. The molecular state is thus “real” and stable,
and atoms tend to form molecules. If those atoms are fermions, the resulting molecule
is a boson. A gas of these molecules can thus undergo Bose-Einstein condensation
(BEC). This side of the resonance is therefore called “BEC-side”. On the side of the
resonance where the scattering length is negative, isolated molecules are unstable.
Nevertheless, when surrounded by the medium of others, two fermions can still form
a loosely bound pair, whose size can become comparable to or even larger than the
average distance between particles. A Bose-Einstein condensate of these fragile pairs
is called a “BCS-state”, after Bardeen, Cooper and Schrieffer. This is what occurs in
superconductors, in which current flows without resistance thanks to a condensate of
electron pairs (“Cooper pairs”).
Here the subscript k reminds us that the solution is for a particular incom-
ing plane wave. This integral representation of the scattering wavefunction,
known as the Lippmann-Schwinger equation, provides a more useful basis to
address situations where the energy of the incoming particles is large and the
scattering potential is weak. The elements of the derivation of this equation
are summarised in the info box below:
equation (∇2 + k 2 )ψ(r) = U (r)ψ(r), the general solution can be written formally as
,
ψ(r) = φ(r) + d3 r# G0 (r, r# )U (r# )ψ(r# ) ,
Substituted back into the expression for the scattering wavefunction, we obtain the
Lippmann-Schwinger equation (14.3).
where the vector k$ = kêr is oriented along the direction of the scattered
particle. We therefore find that the scattering wavefunction ψk (r) = eik·r +
ikr
f (θ, φ) e r can be expressed in integral form, with the scattering amplitude
given by
,
1 1 !
f (θ, φ) = − (φk! |U |ψk ) ≡ − d3 r$ e−ik ·r U (r$ )ψk (r$ ) . (14.4)
4π 4π
dσ m2
= |f |2 = |Tk,k! |2 ,
dΩ (2π)2 !4
From this equation, we can use (14.3) to obtain the next term in the series,
,
(2) (1)
ψk (r) = φk (r) + d3 r$ G0 (r, r$ )U (r$ )ψk (r$ ) ,
known as the Born series, and the leading term in known as the first Born
approximation to the scattering amplitude,
1
fBorn = − (φk! |U |φk ) . (14.5)
4π
U02
σ(θ) = |f (θ)|2 = ,
16k 4 sin4 θ/2
which is just the Rutherford formula.
where g(E) = dE dn
denotes the density of states and both states |k) and |k# ) have
2
energy E = !2 k 2 /2m = !2 k # /2m – they are said to be “on-shell”. As a result, we
k2 dΩ m
obtain the density of states g(E) = dn
dk dE = (2π/L)3 !2 k while the incident flux per
dk
dσ 1 2mV
= |(k# | 2 |k)|2 .
dΩ (4π)2 !
We can therefore recognize that Fermi’s Golden rule is equivalent to the first
order Born approximation.
* +
dσ
= |f (θ) + f (π − θ)|2 .
dΩ indist.
This makes a big difference! For example, for scattering through 90o , where f (θ) =
f (π−θ), the quantum mechanical scattering rate is twice the classical (distinguishable)
prediction.
Furthermore, if we make the standard
-∞ expansion of the scattering amplitude f (θ)
in terms of partial waves, f (θ) = "=0 (2+ + 1)a" P" (cos θ), then
∞
$
f (θ) + f (π − θ) = (2+ + 1)a" (P" (cos θ) + P" (cos(π − θ))) .
l=0
Since P" (−x) = (−1)" P" (x), the scattering only takes place in even partial wave
states. This is the same thing as saying that the overall wavefunction of two identical
bosons is symmetric. So, if they are in an eigenstate of total angular momentum,
from P" (−x) = (−1)" P" (x) it has to be a state of even +.
For fermions in an antisymmetric spin state, such as proton-proton scattering with
the two proton spins forming a singlet, the spatial wavefunction is symmetric, and the
argument is the same as for the boson case above. For parallel spin protons, however,
the spatial wavefunction has to be antisymmetric, and the scattering amplitude will
then be f (θ) − f (π − θ). In this case there is zero scattering at 90o ! Note that for
(non-relativistic) equal mass particles, the scattering angle in the center of mass frame
is twice the scattering angle in the fixed target (lab) frame.
Now, since
* +1/2
. /1/2 2r · Ri
k|r − Ri | = k r − 2r · Ri +
2
R2i $ kr 1 − $ kr − kêr · Ri ,
r2
f (θ) = f e−i∆·Ri , ∆ = k$ − k .
If we consider scattering from a crystal lattice, we must sum over all atoms. X-ray diffraction pattern of a
In this case, the total differential scattering cross-section is given by quasi-crystal.
0 02
0 0
dσ 0 $ 0
0 −i∆·Ri 0
= 0f e 0 .
dΩ 0 0
Ri
In the case of a periodic crystal, the sum over atoms translates to the Bragg
condition,
dσ (2π)3
= |f |2 3 δ (3) (k$ − k − 2πn/L) ,
dΩ L
where L represents the size of the (cubic) lattice, and n denote a vector of
integers – the Miller indices of the Bragg planes. We therefore expect that
the differential cross-section is very small expect when k$ − k = 2πn/L. These
relations can be generalised straightforwardly to address more complicated
crystal structures.
Relativistic Quantum
Mechanics
The aim of this chapter is to introduce and explore some of the simplest aspects
of relativistic quantum mechanics. Out of this analysis will emerge the Klein-
Gordon and Dirac equations, and the concept of quantum mechanical spin.
This introduction prepares the way for the construction of relativistic quantum
field theories, aspects touched upon in our study of the quantum mechanics
of the EM field. To prepare our discussion, we begin first with a survey of
the motivations to seek a relativistic formulation of quantum mechanics, and
some revision of the special theory of relativity.
Why study relativistic quantum mechanics? Firstly, there are many ex-
perimental phenomena which cannot be explained or understood within the
purely non-relativistic domain. Secondly, aesthetically and intellectually it
would be profoundly unsatisfactory if relativity and quantum mechanics could
not be united. Finally there are theoretical reasons why one would expect new
phenomena to appear at relativistic velocities.
When is a particle relativistic? Relativity impacts when the velocity ap-
proaches the speed of light, c or, more intrinsically, when its energy is large
compared to its rest mass energy, mc2 . For instance, protons in the accelera-
tor at CERN are accelerated to energies of 300GeV (1GeV= 109 eV) which is
considerably larger than their rest mass energy, 0.94 GeV. Electrons at LEP
are accelerated to even larger multiples of their energy (30GeV compared to
5 × 10−4 GeV for their rest mass energy). In fact we do not have to appeal to
such exotic machines to see relativistic effects – high resolution electron mi-
croscopes use relativistic electrons. More mundanely, photons have zero rest
mass and always travel at the speed of light – they are never non-relativistic.
What new phenomena occur? To mention a few:
Figure 15.1: Anderson’s cloud chamber picture of cosmic radiation from 1932 show-
ing for the first time the existence of the positron. A cloud chamber contains a gas
supersaturated with water vapour (left). In the presence of a charged particle (such
as the positron), the water vapour condenses into droplets – these droplets mark out
the path of the particle. In the picture a charged particle is seen entering from the
bottom at high energy. It then looses some of the energy in passing through the 6 mm
thick lead plate in the middle. The cloud chamber is placed in a magnetic field and
from the curvature of the track one can deduce that it is a positively charged particle.
From the energy loss in the lead and the length of the tracks after passing though the
lead, an upper limit of the mass of the particle can be made. In this case Anderson
deduces that the mass is less that two times the mass of the electron. Carl Anderson
(right) won the 1936 Nobel Prize for Physics for this discovery. (The cloud chamber
track is taken from C. D. Anderson, The positive electron, Phys. Rev. 43, 491 (1933).
time-like vectors. Time-like vectors can be divided into those pointing forwards in
time (x0 > 0) and those pointing backwards (x0 < 0). Lorentz transformations do
not always map forward time-like vectors into forward time-like vectors; indeed Λ
does so if and only if Λ00 > 0. Such transformations are called orthochronous.
(Since Λµ0 Λµ0 = 1, (Λ00 )2 − (Λj0 )2 = 1, and so Λ00 += 0.) Thus the group splits
into two according to whether Λ00 > 0 or Λ00 < 0. Each of these two components
may be subdivided into two by considering those Λ for which det Λ = ±1. Those
transformations Λ for which det Λ = 1 are called proper.
Thus the subgroup of the Lorentz group for which det Λ = 1 and Λ00 > 0 is
called the proper orthochronous Lorentz group, sometimes denoted by L↑+ . It
contains neither the time-reversal nor parity transformation,
−1 1
1 −1
T = , P = . (15.2)
1 −1
1 −1
We shall call it the Lorentz group for short and specify when we are including T or P .
In particular, L↑+ , L↑ = L↑+ ∪ L↑− (the orthochronous Lorentz group), L+ = L↑+ ∪ L↓+
(the proper Lorentz group), and L0 = L↑+ ∪ L↓− are subgroups, while L↓− = P L↑+ ,
L↑− = T L↑+ and L↓+ = T P L↑+ are not.
Special relativity requires that theories should be invariant under Lorentz trans-
formations xµ )−→ Λµν xν , and, more generally, Poincaré transformations xµ →
Λµν xν + aµ . The proper orthochronous Lorentz transformations can be reached con-
tinuously from identity.2 Loosely speaking, we can form them by putting together
infinitesimal Lorentz transformations Λµν = δ µν + ω µν , where the elements of ω µν - 1.
Applying the identity gαβ = Λµα Λµβ = gαβ + ωαβ + ωβα + O(ω 2 ), we obtain the
relation ωαβ = −ωβα . ωαβ has six independent components: L↑+ is a six-dimensional
(Lie) group, i.e. it has six independent generators: three rotations and three boosts.
Finally, according to the definition of the 4-vectors, the covariant and contravari-
ant derivative are respectively defined by ∂µ = ∂x∂ µ = ( 1c ∂t ∂
, ∇), ∂ µ = ∂x∂ µ =
( 1c ∂t
∂
, −∇). Applying the scalar product to the derivative we obtain the d’Alembertian
∂2
operator (sometimes denoted as !), ∂ 2 = ∂µ ∂ µ = c12 ∂t 2 − ∇ .
2
) *
∂ 2 + kc2 ψ = 0 , (15.3)
!2 ∂t (ψ ∗ ∂t ψ − ψ∂t ψ ∗ ) − !2 c2 ∇ · (ψ ∗ ∇ψ − ψ∇ψ ∗ ) = 0 ,
∂µ j µ = 0 , (15.4)
! There must exist a 4-vector current density which is conserved and whose
time-like component is a positive density.
! The components of ψ do not have to satisfy any auxiliary condition. At
any given time they are independent functions of x.
Beginning with the first of these requirements, by imposing the condition
[γ µ , p̂ν ] = γ µ p̂ν − p̂ν γ µ = 0, (and symmetrizing)
! "
1 ν µ
(γ p̂ν + m) (γ p̂µ − m) ψ =
ν µ
{γ , γ } p̂ν p̂µ − m ψ = 0 ,
2
2
the latter recovers the Klein-Gordon equation if we define the elements γ µ such
that they obey the anticommutation relation,5 {γ ν , γ µ } ≡ γ ν γ µ + γ µ γ ν = 2g µν
– thus γ µ , and therefore ψ, can not be scalar. Then, from the expansion of
Eq. (15.5), γ 0 (γ 0 p̂0 − γ · p̂ − m)ψ = i∂t ψ − γ 0 γ · p̂ψ − mγ 0 ψ = 0, the Dirac
equation can be brought to the form
γ µ† = γ 0 γ µ γ 0 , {γ µ , γ ν } = 2g µν . (15.8)
5
Note that, in some of the literature, you will see the convention [ , ]+ for the anticom-
mutator.
From this result and the continuity relation (15.4) we can identify
j µ = ψ̄γ µ ψ , (15.10)
In order that an observer in the second frame can reconstruct ψ $ from ψ there
must exist a local transformation between the wavefunctions. Taking this
relation to be linear, we therefore must have,
ψ $ (x$ ) = S(Λ)ψ(x) ,
The latter is compatible with the Dirac equation in the original frame if
To define an explicit form for S(Λ) we must now draw upon some of the
defining properties of the Lorentz group discussed earlier. For an infinitesi-
mal proper Lorentz transformation we have Λνµ = g νµ + ω νµ and (Λ−1 )νµ =
g νµ − ω νµ + · · ·, where the matrix ωµν is antisymmetric and g νµ ≡ δ νµ . Corre-
spondingly, by Taylor expansion in ω, we can define
i i
S(Λ) = I − Σµν ω µν + · · · , S −1 (Λ) = I + Σµν ω µν + · · · ,
4 4
where the matrices Σµν are also antisymmetric in µν. To first order in ω,
Eq. (15.12) yields (a somewhat unrewarding exercise!)
) *
[γ ν , Σαβ ] = 2i g να γβ − g νβ γα . (15.13)
i
Σαβ = [γα , γβ ] . (15.14)
2
In summary, if ψ(x) obeys the Dirac equation in one frame, the wavefunction
can be obtained in the Lorentz transformed frame by applying the transforma-
tion ψ $ (x$ ) = S(Λ)ψ(Λ−1 x$ ). Let us now consider the physical consequences
of this Lorentz covariance.
where ωij = $ijk nk θ, and the remaining elements Λµ0 = Λ0µ = 0. Applied to
the argument of the wavefunction we obtain a familiar result,7
6
Since finite transformations are of the form S(Λ) = exp[−(i/4)Σαβ ω αβ ], one may show
that S(Λ) is unitary for spatial rotations, while it is Hermitian for Lorentz boosts.
7
Recall that spatial rotataions are generated by the unitary operator, Û (θ) = exp(−iθn ·
L̂).
8
For finite transformations, the generator takes the form exp[−iθn · L̂].
1
[Si , Sj ] = i$ijk Sk , (Si )2 = for each i . (15.15)
4
15.2.4 Parity
So far, our discussion of the covariance properties of the Dirac equation have
only dealt with the subgroup of proper orthochronous Lorentz transformations,
L↑+ – i.e. those that can be reached from Λ = I by a sequence of infinites-
imal transformations. Taking the parity operation into account, relativistic
covariance demands
S −1 (P )γ 0 S(P ) = γ 0 , S −1 (P )γ i S(P ) = −γ i .
Thus, defining the spin elements u(p) = (ξ, η), where ξ and η represent two-
component spinors, we find the conditions, (p0 − m)ξ = σ · p η and σ · p ξ =
! E $ < −m: Finally, in this case p$2 ≡ E $2 −m2 > 0 and, depending on the
(r) (i)
sign of p$ , j3 can be greater or less than j3 . But the solution has the
" "
form e−i(p x−E t) . Since we presume the beam to be propagating to the
right, we require E $ < 0 and p$ > 0. From this result it follows that ζ < 0
(r) (i)
and we are drawn to the surprising conclusion that |j3 | > |j3 | – more
current is reflected that is incident! Since we have already confirmed
(t)
current conservation, we can deduce that j3 < 0. It is as if a beam of
particles were incident from the right.
The resolution of this last seeming unphysical result, known as the Klein
paradox,9 in fact gives a natural interpretation of the negative energy solu-
tions that plague both the Dirac and Klein-Gordon equations: Dirac particles
are fermionic in nature. If we regard the vacuum as comprised of a filled Fermi
sea of negative energy states or antiparticles (of negative charge), the Klein
Paradox can be resolved as the stimulated emission of particle/antiparticle
9
Indeed one would reach the same conclusion were one to examine the Klein-Gordon
equation.
pairs, the particles moving off towards x3 = −∞ and the antiparticles towards
x3 = ∞. What about energy conservation? One might worry that the energy
for these pairs is coming from nowhere. However, the electrostatic energy re-
covered by the antiparticle when its created is sufficient to outweigh the rest
mass energy of the particle and antiparticle pair (remember that a repulsive
potential for particles is attractive for antiparticles). Taking into account the
fact that the minimum energy to create a particle/antiparticle pair is twice
the rest mass energy 2 × m, the regime where stimulated emission is seen to
occur can be understood.
Negative energy states: With this conclusion, it is appropriate to revisit
the definition of the free particle plane wave
3 state. In particular, for energies
E < 0, it is more sensible to set p0 = + (p2 + m2 ), and redefine the plane
wave solution as ψ(x) = v(p)eip·x , where the spinor satisfies the condition
(+ p + m)v(p) = 0. Accordingly we find,
5 σ · p (r) 6
χ
v (p) = N (p) p0 + m
(r)
.
χ(r)
So, to conclude, two relativistic wave equations have been proposed. The
first of these, the Klein-Gordon equation was dismissed on the grounds that
it exhibited negative probability densities and negative energy states. By
contrast, the states of the Dirac equation were found to exhibit a positive
definite probability density, and the negative energy states were argued to
have a natural interpretation in terms of antiparticles: the vacuum state does
not correspond to all states unoccupied but to a state in which all the negative
energy states are occupied – the negative energy states are filled up by a Fermi
sea of negative energy Fermi particles. For electron degrees of freedom, if a
positive energy state is occupied we observe it as a (positive energy) electron
of charge q = −e. If a negative energy state is unoccupied we observe it as a
(positive energy) antiparticle of charge q = +e, a positron, the antiparticle
of the electron. If a very energetic electron interacts with the sea causing a
transition from a negative energy state to positive one (by communicating an
energy of at least 2m) this is observed as the production of a pair of particles,
an electron and a positron from the vacuum (pair production) (see Fig. 15.2).
1 1
L= ∂µ φ ∂ µ φ − m2 φ2 ,
2 2
(cf. our discussion of the low energy modes of the classical harmonic chain and the
Maxwell field of the waveguide in chapter 11). Defining the canonical momentum
π(x) = ∂φ̇ L(x) = φ̇(x) ≡ ∂0 φ(x), the corresponding Hamiltonian density takes the
form
1+ 2 ,
H = π φ̇ − L = π + (∇φ)2 + m2 φ2 .
2
Evidently, the Hamiltonian density is explicitly positive definite. Thus, the scalar
field is not plagued by the negative energy problem which beset the single-particle
theory. Similarly, the quantization of the classical field will lead to a theory in which
the states have positive energy.
Following on from our discussion of the harmonic chain in chapter 11, we are
already equipped to quantise the classical field theory. However, there we worked ex-
plicitly in the Schrödinger representation, in which the dynamics was contained within
the time-dependent wavefunction ψ(t), and the operators were time-independent. Al-
ternatively, one may implement quantum mechanics in a representation where the
time dependence is transferred to the operators instead of the wavefunction — the
Heisenberg representation. In this representation, the Schrödinger state vector ψS (t)
is related to the Heisenberg state vector ψH by the relation,
Similarly, Schrödinger operators ÔS are related to the Heisenberg operators ÔH (t) by
One can easily check that the matrix elements 3ψS! |ÔS |ψS 4 and 3ψH !
|ÔH |ψH 4 are
equivalent in the two representations, and which to use in non-relativistic quantum
mechanics is largely a matter of taste and convenience. However, in relativistic quan-
tum field theory, the Heisenberg representation is often preferable to the Schrödinger
representation. The main reason for this is that by using the former, the Lorentz
covariance of the field operators is made manifest.
with π̂ = ∂0 φ̂. In doing so, the Hamiltonian density takes the form
1- 2 .
Ĥ = π̂ + (∇φ̂)2 + m2 φ̂2 .
2
To see the connection between the quantized field and particles we need to Fourier
transform the field operators to obtain the normal modes of the Hamiltonian,
7
d4 k
φ̂(x) = φ̂(k)e−ik·x .
(2π)4
However the form of the Fourier elements φ̂(k) is constrained by the following con-
ditions. Firstly to maintain Hermiticity of the field operator φ̂(x) we must choose
Fourier coefficients such that φ̂† (k) = φ̂(−k). Secondly, to ensure that the field op-
erator φ̂(x) obeys the Klein-Gordon equation,10 we require φ̂(k) ∼ 2πδ(k 2 − m2 ).
Taking these conditions together, we require
) *
φ̂(k) = 2πδ(k 2 − m2 ) θ(k 0 )a(k) + θ(−k 0 )a† (−k) ,
√
where k 0 ≡ ωk ≡ + k2 + m2 , and a(k) represent the operator valued Fourier co-
efficients. Rearranging the momentum integration, we obtain the Lorentz covariant
expansion
7
d4 k + ,
φ̂(x) = 2πδ(k 2 − m2 )θ(k 0 ) a(k)e−ik·x + a† (k)eik·x .
(2π) 4
one obtains
7
d3 k ) *
φ̂(x) = a(k)e−ik·x + a† (k)eik·x .
(2π)3 2ωk
↔
where A ∂0 B ≡ A∂t B − (∂t A)B, we obtain
7
d3 k + ,
φ̂(x) = 3 a(k)fk (x) + a† (k)fk∗ (x) .
(2π)3 2ωk
10
Note that the field operators obey the equation of motion,
∂H
π̇(x, t) = − = ∇2 φ − m2 φ .
∂φ(x, t)
Similarly,
7
d3 k + ,
π̂(x) ≡ ∂0 φ̂(x) = 3 iωk −a(k)fk (x) + a† (k)fk∗ (x) .
(2π)3 2ω k
Making use of the orthogonality relations, the latter can be inverted to give
3 7 ↔ 3 7 ↔
a(k) = (2π)3 2ωk d3 xfk∗ (x)i ∂0 φ̂(x), a† (k) = (2π)3 2ωk d3 xφ̂(x)i ∂0 fk (x) ,
or, equivalently,
7 / 0 7 / 0
a(k) = d3 x ωk φ̂(x) − iπ̂(x) e−ik·x , a (k) =
†
d3 x ωk φ̂(x) + iπ̂(x) eik·x .
The field operators a† and a can therefore be identified as operators that create and
annihilate bosonic particles. Although it would be tempting to adopt a different
normalisation wherein [a, a† ] = 1 (as is done in many texts), we chose to adopt the
convention above where the covariance of the normalisation is manifest. Using this
representation, the Hamiltonian is brought to the diagonal form
7
d3 k ωk + † ,
Ĥ = a (k)a(k) + a(k)a† (k) ,
(2π) 2ωk 2
3
|k4 = a† (k)|Ω4 .
Then 3k! |k4 = 3Ω|a(k! )a† (k)|Ω4 = 3Ω|[a(k! ), a† (k)]|Ω4 = (2π)3 2ωk δ 3 (k! − k). Many-
particle states are defined by |k1 · · · kn 4 = a† (k1 ) · · · a† (kn )|Ω4 where the bosonic
statistics of the particles is assured by the commutation relations.
Associated with these field operators, one can define the total particle number
operator
7
d3 k
N̂ = 3 a† (k)a(k) .
(2π)3 2ωk
The time component P̂ 0 of this result can be compared with the Hamiltonian above.
In fact, commuting the field operators, the latter is seen to differ from P̂ 0 by an infinite
8
constant, d3 kωk /2. Yet, had we simply normal ordered11 the operators from the
outset, this problem would not have arisen. We therefore discard this infinite constant.
1 1
L= ∂µ φ∂ µ φ̄ − m2 |φ|2 .
2 2
11
Recall that normal ordering entails the construction of an operator with all the annihi-
lation operators moved to the right and creation operators moved to the left.
The latter can √be interpreted as the superposition of two independent scalar fields
φ = (φ1 +iφ2 )/ 2, where, for each (real) component φ†r (x) = φr (x). (In fact, we could
as easily consider a field with n components.) In this case, the canonical quantisation
of the classical fields is achieved by defining (exercise)
7
d3 k + ,
φ̂(x) = 3 a(k)fk (x) + b† (k)fk∗ (x) .
(2π) 2ωk
3
Thus the complex scalar field has the interpretation of creating different sorts of
particles, corresponding to operators a† and b† . To understand the physical interpre-
tation of this difference, let us consider the corresponding charge density operator,
↔ 8
ĵ0 = φ̂† (x)i ∂ 0 φ(x). Once normal ordered, the total charge Q = d3 xj0 (x) is given
by
7
d3 k + † ,
Q̂ = 3 a (k)a(k) − b† (k)b(k) = N̂a − N̂b .
(2π) 2ωk
3
From this result we can interpret the particles as carrying an electric charge, equal
in magnitude, and opposite in sign. The complex scalar field is a theory of charged
particles. The negative density that plagued the Klein-Gordon field is simply a man-
ifestation of particles with negative charge.
L = ψ̄ (iγ µ ∂µ − m) ψ ,
↔
(or, equivalently, L = ψ̄( 12 iγ µ ∂ µ −m)ψ). With this definition, the corresponding
canonical momentum is given by ∂ψ̇ L = iψ̄γ 0 = iψ † . From the Lagrangian density,
we thus obtain the Hamiltonian density,
H = ψ̄ (−iγ · ∇ + m) ψ ,
; <
{ψα (x, t), ψβ (x! , t)} = ψ̄α (x, t), ψ̄β (x! , t) = 0 .
Using the general solution of the Dirac equation for a free particle as a basis set,
together with the intuition drawn from the study of the complex scalar field, we may
with no more ado, introduce the field operators which diagonalise the Hamiltonian
density
2 7
= d3 k - .
ψ(x) = ar (k)u(r) (k)e−ik·x + b†r (k)v (r) (k)eik·x
r=1
(2π) 2ωk
3
7
d3 k - † .
=2
ψ̄(x) = a (k)ū (r)
(k)eik·x
+ b r (k)v̄ (r)
(k)e−ik·x
,
r=1
(2π)3 2ωk r
where the annihilation and creation operators also obey the anticommutation rela-
tions,
; < ; <
ar (k), a†s (k! ) = br (k), b†s (k! ) = (2π)3 2ωk δrs δ 3 (k − k! )
; < ; <
{ar (k), as (k! )} = a†r (k), a†s (k! ) = {br (k), bs (k! )} = b†r (k), b†s (k! ) = 0 .
The latter condition implies the Pauli exclusion principle a† (k)2 = 0. With this
definition, a(k)u(k)e−ik·x annilihates a postive energy electron, and b† (k)v(k)eik·x
creates a positive energy positron.
From these results, making use of the expression for the Hamiltonian density
operator, one obtains
=2 7
d3 k + ,
Ĥ = ωk a†r (k)ar (k) − br (k)b†r (k) .
r=1
(2π) 2ωk
3
Were the commutation relations chosen as bosonic, one would conclude the existence
of negative energy solutions. However, making use of the anticommutation relations,
and dropping the infinite constant (or, rather, normal ordering) one obtains a positive
definite result. Expressed as one element of the total energy-momentum operator, one
finds
2 7
= d3 k + ,
P̂ µ = k µ a†r (k)ar (k) + b†r (k)br (k) .
r=1
(2π) 2ωk
3
pµ )−→ pµ − qAµ ,
(+ p − q+A − m)ψ = 0 .
we have seen that the plane-wave solution to the Dirac equation for particles
can be written in the form
! "
χ
ψp (x) = N cσ ·p̂ ei(px−Et)/! ,
mc2 +E
χ
where we have restored the parameters ! and c. From this expression, we can
see that, at low energies, where |E − mc2 | - mc2 , the second component of
the bispinor is smaller than the first by a factor of order v/c. To obtain the
non-relavistic limit, we can exploit this asymmetry to develop a perturbative
expansion of the coefficients in v/c.
Consider then the Dirac equation for a particle moving in a potential
(φ, A). Expressed in matrix form, the Dirac equation H = cα · (−i!∇ −
c A) + mc β + eφ is expressed as
e 2
! "
mc2 + eφ cσ · (−i!∇ − ec A)
H= .
cσ · (p̂ − qc A) −mc2 + qφ
Defining the bispinor ψ T (x) = (ψa (x), ψb (x)), the Dirac equation translates to
the coupled equations,
q
(mc2 + eφ)ψa + cσ · (p̂ − A)ψb = Eψa
c
q
cσ · (p̂ − A)ψa − (mc2 − qφ)ψb = Eψb .
c
Then, if we define W = E − mc2 , a rearrangement of the second equation
obtains
1 q
ψb = cσ · (p̂ − A)ψa .
2mc2 + W − qφ c
q
Then, at zeroth order in v/c, we have ψb $ 2mc 1
2 cσ ·(p̂− c A)ψa . Substi-
tuted into the first equation, we thus obtain the Pauli equation Hnon−rel ψa =
W ψa , where
1 - q .2
Hnonr el = σ · (p̂ − A) + qφ .
2m c
Making use of the Pauli matrix identity σi σj = δij + i$ijk σk , we thus obtain
the familiar non-relativistic Schrödinger Hamiltonian,
1 q q!
Hnon rel = (p̂ − A)2 − σ · (∇ × A) + qφ .
2m c 2mc
p̂2 p̂4 1 1
Ĥnon−rel = − + (σ · p̂)V (σ · p̂) + V − (V p̂2 + p̂2 V ) .
2m 8m3 c2 4m2 c2 8m2 c2
Then, making use of the identities,
p̂2 p̂4 ! !2
Ĥnon−rel = − + σ · (∇V ) × p̂ + (∇2 V ) .
2m 8m3 c2 >4m2 c2 ?@ A > 8m2 c2
?@ A
spin−orbit coupling Darwin term
The second term on the right hand side represents the relativistic correction
to the kinetic energy, the third term denotes the spin-orbit interaction and
the final term is the Darwin term. For atoms, with a central potential, the
spin-orbit term can be recast as
!2 1 !2 1
ĤS.O. = σ · (∂r V )r × p̂ = (∂r V )σ · L̂ .
4m c
2 2 r 4m2 c2 r
To address the effects of these relativistic contributions, we refer back to chap-
ter 9.
Problem sets
Before starting these problems, you might want to revise some of the examples
from the Part IB Quantum Physics course. The examples marked with a † are
typically more challenging and are the ones to omit if your time is very short, or if
you are finding the course difficult. Some of the questions involve a routine piece
of bookwork. This is the kind of thing you will have to do in the exam. You are
strongly encouraged to do these parts and get feedback in supervisions.
where r and t are the reflection and transmission amplitudes. re−ikx teikx
eikx
�b) Show that unitarity demands that rt∗ + r∗ t = 0 and |r|2 + |t|2 = 1,
and hence that |r ± t|2 = 1. Find θ1 and θ2 in terms of r and t.
What is the difference in phase between r and t?
�c) By matching the boundary conditions, show the elements of the S-
2 k2
matrix for the scattering of particles of mass m and energy E = �2m V0
γ
from a δ-function potential, aV0 δ(x), are given by r = − γ+ik and
ik maV0
t = γ+ik , where γ = �2 . Obtain the corresponding scattering
phase shifts.
2. Operator methods: This problem addresses simple relations that follow from
the orthogonality of eigenfunctions and the time-development of wavefunc-
tions.
The Hamiltonian Ĥ has two normalized eigenstates |ψ1 � and |ψ2 � which
correspond to different eigenvalues E1 and E2 .
�a) By using the commutation relation [a� a† ] = 1, show that Hint: To prove this result most
† †
straightforwardly, consider the β
e−βa aeβa = β + a . derivative of this expression.
†
Using this result, show that |β� = N eβa |0� is a coherent state, i.e.
2
a|β� = β|β�. Finally, show that the normalization, N = e−|β| /2 .
�b) Calculate the expectation values, x0 = �x̂� and p0 = �p̂�, with Hint: Remember how creation
respect to |β� and, by considering �x̂2 � and �p̂2 �, show that and annihilation operators are
related to the phase space opera-
�2 tors x̂ and p̂. Also, note that the
(Δp)2 (Δx)2 = �
4 Hermitian conjuate of the eigen-
value equation a|β� = β|β� leads
where Δp = p̂ − �p̂� (similarly x).
to the relation �β|a† = �β|β ∗ .
�c) To determine the coordinate representation to the coherent state,
ψ(x) = �x|β�, it is helpful to revert back to the expression for
a as a differential operator. Show that the eigenvalue equation
a|β� = β|β� translates to the equation,
� �
�
x+ ∂x ψ(x) = βψ(x) .
mω
As a result, deduce expressions for x0 (t) and p0 (t) and show they
represent solutions to the classical equations of motion. How does
the width of the coherent state wavepacket evolve with time?
2
In the symmetric gauge, we therefore find that the Landau level states are localized in
both x and y directions. This contrasts with what was found from the Landau gauge condi-
tion where states were localized along only one direction. Of course, there is no contradiction
between these two representations: since the Landau levels have a huge degeneracy, we are
at liberty to reconstruct states within the basis.
5. Spin: This question develops the concept of a spinor wavefunction. Hint: You can find the relevant
spin operator by taking the dot
Using the Pauli matrices, σx , σy and σz , write down the operator corre-
product of the vector σ with a
sponding to a component of spin along the axis (θ� φ) in spherical polar unit vector in the desired direc-
coordinates. Show that the eigenvalues of spin in this direction are ±�/2 tion.
(as expected), and deduce the corresponding wavefunctions. Hence, in-
fer the wavefunctions for particles whose spins are aligned along the +x,
−x, +y and −y directions.
8. Spin: We have seen that the “ladder operator” formalism provides a frame-
work in which to define and classify the states of the quantum harmonic
oscillator. In the following, we will see that the same ladder operator for-
malism provides a representation of the quantum spin algebra. This repre-
sentation, known as the Holstein-Primakoff transformation, can be used to
where the ladder operators a and a† obey the usual commutation rela- Hint: If you find yourself
tions, [a� a† ] ≡ aa† − a† a = 1.4 Making use of these relations, show that expanding the square root,
this definition is indeed consistent with the quantum spin algebra, i.e. you should stop and consider
[Ŝ + � Ŝ − ] = 2�Ŝ z . whether there is a simpler
method...
Comments: Physically, the ladder operators simply “count” the number
of “spin deviations” away from êz . For large spin, S – the analogue the
semi-classical limit for quantum mechanical spins
√ –†the Holstein-Primakoff
transformation
√ affords the expansion, S = � 2S a +O(S −1/2 ) and S + =
−
�a) Show that the operators J± = Jx ± iJy act as raising and lower-
ing operators for the z-component of angular momentum, by first
calculating the commutator [Jz � J± ].
�b) State the allowed values of the total spin angular momentum for a
system of three electrons.
�c) The ‘coupled basis’ state |S = 3/2� mS = 3/2� (an eigenstate of
total spin) is also a state of the ‘uncoupled basis’, which may be You may note that J± |j� m� =
denoted by | ↑↑↑� ≡ | ↑� ⊗ | ↑� ⊗ | ↑�. By an application of total �[j(j +1)−m(m+1)]1/2 |j� m±1�.
spin lowering operator, show that
1
|S = 3/2� mS = 1/2� = √ (| ↓↑↑� + | ↑↓↑� + | ↑↑↓�) .
3
that jleft = jright , i.e. |A|2 + |D|2 = |C|2 + |B|2 – all of the incoming beam
is transferred to the outgoing beam, a statement of particle conservation.
This conservation of current can be written as "Ψin |Ψin # = "Ψout |Ψout #.
Therefore, using the expression for |ψout #, we have
as required. Using the result from (b), since the potential is symmetric,
we have
% &
t ik + γ k2 −ik + γ γ 2 γ2
r = ∗ (|t| − 1) = −
2 2
− 1 = =
t −ik + γ k 2 + γ 2 ik + γ k 2 + γ 2 (ik + γ)2
since E1 and E2 are real. Therefore, (E1 −E2 )"ψ1 |ψ2 # = 0 and, if E1 $= E2
then "ψ1 |ψ2 # = 0, i.e. |ψ1 # and |ψ2 # are orthogonal.
(b) If Â|ψ1 # = |ψ2 # and Â|ψ2 # = |ψ1 #, then adding them, Â(|ψ1 # + |ψ2 #) =
|ψ1 # + |ψ2 # and subtracting, Â(|ψ1 # − |ψ2 #) = |ψ2 # − |ψ1 # = −(|ψ1 # − |ψ2 #).
Hence we have an eigenvector of a = +1 corresponding to a normalized
eigenvector √12 (|ψ1 # + |ψ2 #) and an eigenvalue a = −1 corresponding to
eigenvector √12 (|ψ1 # − |ψ2 #).
3. (a) Differentiating the left hand side of the given expression with respect to β,
one obtains
† †
e−βa [a, a† ] eβa = 1 .
! "# $
=1
† †
Integrating, we therefore have that e−βa aeβa = β+“integration constant$$ .
By setting β = 0 we can deduce that the “constant” must be a yielding
the required result. Using this result, we have that
† † †
e−βa a|β# = e!−βa"#aeβa$ |0# = (β + a)|0# = β|0# .
=β+a
- -
! !mω
"x̂# = (β + β ),
∗
"p̂# = −i (β − β ∗ ) .
2mω 2
Then, using the identity (∆x)2 = "(x − "x#)2 # = "x2 # − "x#2 , we have
! !
"x2 # = "β|(a2 + aa† + a† a + (a† )2 |β# = (1 + (β + β ∗ )2 ),
2mω 2mω
!mω !
"p̂2 # = − "β|(a2 − aa† − a† a + (a† )2 |β# = − (−1 + (β − β ∗ )2 ) .
2 2mω
!
As a result, we find that ∆x = 2mω and ∆p = !mω2 leading to the required
expression.
(c) The equation follows simply from the definion of the operator a and the
solution may be checked by substitution.
(d) Using the time-evolution of the stationary states, |n(t)# = e−iEn t/! |n(0)#,
where En = !ω(n + 1/2), it follows that
2 βn
|β(t)# = e−iωt/2 e−|β| /2
√ e−inωt |n# = e−iωt/2 |e−iωt β# .
n!
Therefore, during the time-evolution, the coherent state form is preserved
but the centre of mass and momentum follow that of the classical oscilla-
tor,
x0 (t) = A cos(ϕ + ωt), p0 (t) = mωA sin(ϕ + ωt) .
The width of the wavepacket remains constant.
1 . z̄ /m 1 −z̄z/4 1
"r|0, m# = √ 2m/2 ∂z + √ e =√ z m e−z̄z/4 ,
m! 4 2π 2π2m m!
as required.
(f ) If we populate the states of the lowest Landau with electrons, starting
from states of the lowest angular momentum m, the wavefunction is a
2
Slater determinant involving entries φm (rj ) = zjm e−|zj | /4 . Taking the
determinant, and making use of the Vandemonde determinant identity,
one obtains the required many-electron wavefunction.
5. The spin operator in the (θ, φ) direction, Ŝθφ , can be found by forming the
scalar product of the spin operator Ŝ with a unit vector in the (θ, φ) direction,
(sin θ cos φ, sin θ sin φ, cos θ). Therefore
% &
! cos θ e−iφ sin θ
Ŝθφ = .
2 eiφ sin θ − cos θ
Eliminating u and v, we find λ2 = 1 and hence the eigenvalues of Ŝθφ are ±!/2,
as expected. Substituting the values λ = ±1 back into the equations relating
u and v, we can infer the ratios, uv = e−iφ cot(θ/2) and −e−iφ tan(θ/2). So, in
matrix notation, the eigenstates are
% & % &
cos(θ/2) sin(θ/2)
and ,
eiφ sin(θ/2) −eiφ cos(θ/2)
for eigenvalues +!/2 and −!/2 respectively. The spin states in the x-direction
are obtained by setting θ = π/2, φ = 0, and the spin states in the y-direction
are obtained by setting φ = π/4 in these general formulae.
we can now calculate the effect of applying Ŝ2 to each state. For example
where the three terms in square brackets correspond to the three operators on
the right hand side of Eq. (16.1). Likewise φ2 . Clearly Ŝz φ3 = 0 = Ŝz φ4 , so
only the Ŝ+ Ŝ− term need be considered. We find
√
Ŝ+ Ŝ− χ+ (1)χ− (2) = Ŝ+ !χ− (1)χ− (2) = !2 (χ+ (1)χ− (2) + χ− (1)χ+ (2)) 2!2 φ3 .
√
Similarly Ŝ+ Ŝ− χ− (1)χ+ (2) = 2!2 φ3 so that
7. Taking as a basis the states of the Ŝz operator, we can deduce the form of the
Ŝx operator most easily from the action of the spin raising and lower operators.
Noting that, for spin S = 1, Ŝ± |S = 1, m# = ![2 − m(m + 1)]1/2 |1, m ± 1#, we
can construct the matrix elements of Ŝ± . The latter are given by
√ 0 1 0
†
Ŝ+ = 2! 0 0 1 , Ŝ− = Ŝ+ .
0 0 0
Then, using the relation, Ŝx = 12 (Ŝ+ − Ŝ− ), we obtained the matrix elements
of the operator and corresponding eigenstates,
x = 1 mx = 0 x = −1
m m
0 1 0
! 1 1 1
Ŝx = √ 1 0 1 , 1 √ 1 1 √
2 2 √ 0 − 2
0 1 0 2 2 2
1 −1 1
Thus, if µBt = (2n + 1)π, with n an integer, the molecules will be in a pure
mx = −1 state,
0 and none will pass the second filter. The time is given by
t = L/v = L m/2E, where L = 20 mm and m and E are the mass and energy
of the molecules respectively. We thus have
% &1/2
(2n + 1)π! 2E
µ= = 2.84 10−26 JT−1
BL m
and hence the proton magnetic moment is 1.42 10−26 JT−1 .
Note that the result can also be obtained by treating the problem as one of
classical precession. The couple = µB = LΩ, where L = ! is the angular
momentum and Ω the angular frequency of precession. If Ωt = (2n + 1)π, the
molecules have precessed into the mx = −1 state, and the result readily follows.
8. Substituting for the definition of the spin raising and lowering operators using
the Holstein-Primakoff transformation, the commutator is obtained as
%
†
a+1%
&1/2 a #$!" &1/2 % &
1 †
a a a† a a† a
[Ŝ +
, Ŝ −
] = 1 − aa †
1 − − a †
1 − a
2S!2 2S 2S 2S
% & % &
a† a a† a a† a† aa a† a
= 1− + a† a 1 − − a† a + =1− .
2S 2S 2S S
With Ŝ z = !(S −a† a), we obtain the required commutation relation [Ŝ + , Ŝ − ] =
2!Ŝ z .
M (m1 , m2 )
9. For the case -1 = 1, -2 = 2, a table of possible ways of forming each value
3 (1, 2)
of M = m1 + m2 is shown right. The largest value of M is 3, so the largest
2 (1, 1) (0, 2)
value of L must be 3. There must also be a state with M = 2 corresponding
1 (1, 0) (0, 1) (−1, 2)
to L = 3, but we have two states with M = 2. Therefore there must be a
0 (1, −1) (0, 0) (−1, 1)
state with L = 2 as well. We need two states with M = 1, one for each of the
−1 (1, −2) (0, −1) (−1, 0)
L = 3, 2 multiplets, but we actually have three states with M = 1, so there
−2 (0, −2) (−1, 1)
must be an L = 1 state as well. All of the M states are now accounted for.
−3 (−1, −2)
For the case -1 = 3, -2 = 1, we can again for a table (see right). Following the
same logic as before, we see that states with L = 4, 3, 2 just account for all the M (m1 , m2 )
states. 4 (3, 1)
To construct the states explicitly, we start by writing the L = 3 M = 3 state, 3 (3, 0) (2, 1)
since there is only one way of forming M = 3, viz. |3, 3# = |1, 1# ⊗ |2, 2#. We 2 (3, −1) (2, 0) (1, 1)
then operate with the lowering operator L̂− , which is simply the sum of the 1 (2, −1) (1, 0) (0, 1)
lowering operators for the two separate particles. Recalling that: 0 (1, −1) (0, 0) (−1, 1)
0 −1 (0, −1) (−1, 0) (−2, 1)
L̂− |-, m# = -(- + 1) − m(m − 1)!|-, m − 1# ,
√ √ √ −2 (−1, −1) (−2, 0) (−3, 1)
we obtain 6!|3, 2# = 2!|1, 0# ⊗ |2, 2# + 4!|1, 1 ⊗ |2, 1#, where the first −3 (−2, −1) (−3, 0)
term on the right hand side comes from lowering the 0 - = 1 state and the −4 (−3, −1)
second
0 from lowering the - = 2 state. Hence |3, 2# = 1/3!|1, 0# ⊗ |2, 2# +
2/3!|1, 1#⊗|2,
0 1#. The state |2, 2#
0 must be the orthogonal linear combination,
i.e. |2, 2# = 2/3!|1, 0# ⊗ |2, 2# − 1/3!|1, 1# ⊗ |2, 1#. Further states could be
computed in the same way if required.
Notice that the total quantum number, J, is the same as the overall spin
quantum
√ number, S in this case. Therefore |S = 3/2, mS = 1/2# =
(1/ 3)J− |S = 3/2, mS = 3/2#. Using 0 the given normalization result
again for a single state, S− |1/2, 1/2# = 3/4 + 1/4!|1/2, −1/2#. This
establishes the required result.
6. (a) E1 and E2 are the ground state energies of a particle moving in Hint: Use the wavefunction of a
attractive potentials V1 (r) and V2 (r). Using the variational method, particle moving in V2 (r) as a trial
show that E1 ≤ E2 if V1 (r) ≤ V2 (r). wavefunction for potential V1 (r).
(b) Consider a particle moving in a one-dimensional attractive potential
V (x), i.e. a potential such that V (x) ≤ 0, for all x and V (x) →
0, as |x| → ∞. Use the variational principle with trial function
A exp(−λx2 ) to show that the upper bound on the ground state
energy is negative, and hence that for any such potential at least
one bound state must exist.
Show that the energy of the system is of the form E = (n21 + n22 )ε where
n1 and n2 are integers and find an expression for ε. Consider the state
with E = 5ε for each of the following three cases:
In each case, what is the symmetry of the spin and spatial parts of the
wavefunction? Hence write down the spatial wavefunction, and sketch
the probability density |ψ(x1 , x2 )|2 in the (x1 , x2 ) plane.
Describe qualitatively how the energies of these states would change if
the particles carried electric charge and hence interacted with each other
(an example of the exchange interaction).
(a) Find the wavefunctions and energies for the ground state and the
first two excited states of the system.
(b) Suppose that the two bosons interact with each other through the
perturbative “contact interaction”,
This result exposes the problem with perturbation theory: it assumes that
the ground state energy is an analytic function of λ for some sufficiently
small region around λ = 0. However, this result shows that the true
ground state solution has an essential singularity in λ and the radius of
convergence of the perturbation theory vanishes.
2
2. For a point-like nucleas, the hydrogen atom has a potential V (r) = − 4π#e
0r
. A
hollow spherical shell will have the same potential for r > b, but V (r) = V (b)
for r < b, by Gauss theorem, and thus its effect can be regarded as adding a
perturbation,
% &
e2 1 1
Ĥ (1) = − ,
4π%0 r b
to the Hamiltonian for r < b, and zero for r > b. For the 2s wavefunction, the
energy shift induced by the perturbation is
! b % &% &2
1 e2 1 1 r
∆E = !ψ|Ĥ |ψ" =
(1)
4πr dr
2
− 1− e−r/a0 .
8πa30 4π%0 0 r b 2a0
Since b ' a0 , the terms involving r/a0 are negligible in the region of integration,
so we can simplify the integral to
! b % &
e2 1 1 b2 e2
∆E = r 2
dr − = R ∞ , R ∞ = .
8π%0 a30 0 r b 6a20 8π%0 a0
Likewise for the 2p0 wavefunction, making the same approximation, we obtain
! b % &! π
e2 1 1 b4
∆E = r 4
− dθ2π sin θ cos 2
θ = R∞ .
128π 2 %0 a0 0
5 r b 0 240a40
Both energy shifts are very small, but that for the 2p state is much smaller,
because the 2p wavefunction vanishes at the origin.
This is not a good method to explore the nucleus because other effects, such
as spin-orbit interaction and other relativistic corrections would swamp the
nuclear size effect. It is more effective for heavy atoms, with larger nuclei and
smaller Bohr radii, and especially for “muonic” atoms, where the greater mass
of the muon again reduces the Bohr radius.
3. If, without loss of generality, we take the field to lie along z, the perturbation is
given by Ĥ $ = −eEz. At first order in perturbation theory, ∆E = −!0|eEz|0"
vanishes since the ground state of the hydrogen atom |0" is an eigenstate parity.
The leading contribution to ∆E is therefore the second order term ∆E =
. |&k|eEz|0'|2
k%=0 E0 −Ek . If the induced dipole moment is d = α%0 E, its energy of
interaction with the electric field is given by ∆E = − 12 d · E = − 12 α%0 E 2 . So,
by comparing with our perturbation theory result we obtain the required result.
An alternative derivation of this result starts from the first order perturbation
.
theory expression for the perturbed wavefunction: |ψ" = |0" + k%=0 ck |k",
where ck = &k|−eEz|0'
E0 −Ek . The dipole moment operator for the electron is ez, and
its expectation value in this state is (neglecting small terms of order (c2k )),
/
!ψ|ez|ψ" = !0|ez|0" + [ck !0|ez|k" + c∗k !k|ez|0"] +O(c2k ) = α%0 E
( )* +
k%=0
=0 ( )* +
P |"k|ez|0#|2
2E k!=0 Ek −E0
! π ! ∞
!0|z 2 |0" = !0|r2 cos2 θ|0" = 2π sin θ cos2 θdθ r2 drr2 e−2r/a0 = a20 .
0 0
e2
We also need the energy difference, E1 − E0 = (1 − 14 )R∞ = 3
4 8π#0 a0 from
64πa30
which we obtain α ≤ 3 = 9.9 × 10 −30
m , a figure that is not too far from
3
experiment.
4. From
0 ∞ the trial wavefunction,
0a we can obtained A from the normalization, 1 =
−∞
|ψ|2 dx = A2 −a (x4 − 2a2 x2 + a4 )dx = 16 15 A a . Moreover, using the
2 5
identity,
% & , 2 -
!2 2 1 ! 1
Ĥψ = − ∂ + mω x ψ = A
2 2
+ mω (a x − x )
2 2 2 4
2m x 2 m 2
0∞ 4πA2 β3
5. From the normalization, 1 = A2 0
4πr2 e−2βr dr = 4β 3 ⇒ A2 = π . From
2β −βr
the identity ∇2 ψ = r12 ∂r (r2 ∂r ψ) = r12 ∂r (−βr2 e−βr ) = β 2 e−βr − r e , we
have
! ∞ , % & -
!2 2β −2βr e2 −2βr
!ψ|Ĥ|ψ" = A2 4πr2 dr − β 2 e−2βr − e − e
0 2m r 4π%0 r
!2 β 2 e2 β
= − .
2m 4π%0
me2
Minimising with respect to β we obtain β = 4π#0 !2 = a−1
0 , which is the inverse
e2 2
of the Bohr radius and thus !ψ|Ĥ|ψ" = m
− 2!2 ( 4π# )
= −13.6 eV. This is the
0
correct value for the ground state energy of the Hydrogen atom, as expected,
because we chose the correct functional form for the trial function.
6. (a) Suppose that the two Hamiltonians are Ĥ1 and Ĥ2 with ground state
wavefunctions ψ1 and ψ2 , i.e. Ĥ1 ψ1 = E1 ψ1 , and Ĥ2 ψ2 = E2 ψ2 . Given
that V1 ≤ V2 , we have Ĥ1 = Ĥ2 − V2 (r) + V1 (r) = Ĥ2 + ∆V (r). From the
variational principle,
E1 ≤ !ψ2 |Ĥ1 |ψ2 " = !ψ2 |Ĥ2 |ψ2 " − !ψ2 |∆V |ψ2 " = E2 + !ψ2 |∆V |ψ2 " ≤ E2 ,
where the last inequality follows because ∆V (r) ≤ 0. Thus E2 ≥ E1 .
2
!
(b) The Hamiltonian is given by Ĥ = − 2m ∂x2 + V (x), and the normalized trial
1/4 −λx2
function is given by ψ =$ (2λ/π) e . Using standard integrals, we
!2
0
2λ ∞ 2
!2
obtain !ψ|Ĥ|ψ" = 2m λ + π −∞ dxV (x)e−2λx = 2m λ + I. Minimsing
with respect to λ, we obtain,
1 !
!2 I I 2λ 2
0= + + + V (x)(−2x2 )e−2λx ,
2m 2λ 2λ π
where the second term arises from differentiating the normalization in I,
and the third term from differentiating the integrand. This is an implicit
equation for λ and if we solve for I and substitute into the equation from
above, we obtain
1 !
!2 !2 2λ 2
!ψ|Ĥ|ψ" = λ+I =− λ + 2λ dxV (x)(2x2 )e−2λx .
2m 2m π
This is our upper bound on the ground state energy, and since V (x) ≤ 0,
both terms are manifestly negative. Hence the ground state energy is
negative, and at least one bound state must exist.
7. A single particle in the potential well has the (unnormalized) wavefunction and
!2 π 2 2
energy, ψn (x) = sin(nπx/L) and E = 2mL 2n ≡ %n2 . The wavefunction for a
system of two identical particles must be either symmetric or antisymmetric,
i.e.
ψ(x1 , x2 ) = sin(n1 πx1 /L) sin(n2 πx2 /L) ± sin(n2 πx1 /L) sin(n1 πx2 /L) ,
with energy (n21 + n22 )%. If E = 5%, we must have n1 = 1, and n2 = 2 (or vice
versa).
(a) Spin-zero particles are bosons and must have a symmetric wavefunction,
2 2
π !
8. For the single-particle states, E = n2 8ma 2 = %n
2
Since this well is not centred
on zero, the single-particle eigenstates are all just proportional to sin(nπx/2a) ≡
|n".
(a) If we write the two-particle states as |n1 , n2 ", the ground state is |1, 1"
(E = 2%). The first excited states are |2, 1" and |1, 2" (E = 5%). The second
excited state is |2, 2" (E = 8%). The overall wavefunction needs to be sym-
metric for bosons, which |1, 1" and |2, 2" are already. These therefore pair
with a symmetric spin wavefunction, which is always possible, whether or
not the bosons have spin zero. For the first excited state, both symmetric
√
and antisymmetric combinations are possible: (|2, 1" ± |1, 2")/ 2; these
would need to pair with spin wavefunctions that are respectively symmet-
ric and antisymmetric. If S > 0, both are possible; if S = 0, only the
symmetric space state is allowed.
The (normalized) ground state wavefunction is given by
1
ψ(x1 , x2 ) = !x1 , x2 |1, 1" = sin(πx1 /2a) sin(πx2 /2a) .
a
(b) According to first order perturbation theory, the change in the ground state
energy caused by Ĥ $ is given by ∆E = !Ĥ $ ", where0 0 the expectation value
involves the unperturbed
00 eigenfunctions, ∆E = ψ ∗
(x
0 1 , x2 )Ĥ ψ(x1 , x2 )dx1 dx2 .
$
Using the identity, f (x1 , x2 )δ(x1 − x2 )dx1 dx2 = f (x1 , x2 )dx1 , for
any function f , we have
! ! ! 1
2V0 2a 4
∆E = −2aV0 |ψ(x, x)| dx = −2
sin (πx/2a) = −4V0 sin4 (πy)dy .
a 0 0
The sin4 (πy) looks nasty, but written as sin2 (πy)×sin2 (πy), with sin2 (πy) =
(1 − cos(2πy))/2, it is easily evaluated and gives ∆E = −3V0 /2.
where a0 denotes the Bohr radius and Gn! is a polynomial function of its
argument. Show that the expectation values of the energies associated
with the three terms listed above all have the same dependence on Z.
(Part IB Advanced Physics 1993.)
5. Zeeman effect: This problem involves the study of the infleunce of a weak
magnetic field on the spectrum of a multielectron atom. Here we are in-
terested in exploring the interplay between LS coupling and the influence of
the external field.
For an atom characterised by LS coupling, and subject to a weak uniform
magnetic field, derive the expression for the Landé g-factor,
6. Atomic structure: As well as the interaction between the spin and orbital
degrees of freedom of the electron which follow from the relativistic cor-
rections, the magnetic field generated by the nuclear magnetic momentum
also lead to corrections to the electron Hamiltonian. These corrections are
known as hyperfine coupling. The following problem involves exploring the
where the superscripts e and p refer to the electron and proton, the vector
components of σ are the Pauli spin operators, µe,p are the respective
magnetic dipole moments, and W is a constant.
where ψa and ψb are the (real) ground state wavefunctions of the two
hydrogen atoms.
" " # $
Z dr Zr
!ψ| Ĥ2 |ψ" ∼ r2 dr|ψ|2 3 ∼ Z 4 Gn# e−2Zr/na0 .
r r a0
The integral is independent of Z, as can easily be seen by making a change
of variables u = Zr/a0 , and hence !ψ| Ĥ2 |ψ" ∝ Z 4 .
(c) Making the usual substitution u = Zr/a0 , we have δ 3 (r) = ( aZ0 )3 δ 3 (u) and
thus the expectation value of this term in the Hamiltonian is
" " 2 # $3 # $3
u du Z Z
!ψ| Ĥ3 |ψ" ∼ r dr Zδ (r)|ψ| ∼
2 (3) 2
Z δ (u)
3
G2 (u)e−2u/n ,
Z3 a0 a0
which is proportional to Z 4 again.
The given energy levels clearly form two doublets (a,b) and (c,d), so the jj
coupling scheme is the better approximation in this case. The electric dipole
selection rules (∆J = ±1, 0; but 0 → 0 forbidden) tell us than only the J = 1
levels can decay to the ground state. We also expect the (1/2, 1/2) level to
lie lower, since !L̂ · Ŝ" ∝ (J(J + 1) − L(L + 1) − S(S + 1)), so this suffices to
identify the levels: (a) is (1/2, 1/2)0 ; (b) is (1/2, 1/2)1 ; (c) is (3/2, 1/2)2 ; (d) is
(3/2, 1/2)1 .
(i) As an alkali atom, sodium has a single electron outside closed shells.
The low energy excited states involve excitation of this electron. The
appropriate quantum numbers are $, s and j (the quantum numbers for
the whole atom being the same as for the unpaired electron). The al-
lowed states for the electron are [3s(ground state), 3p, 3d], [4s, 4p, 4d, 4f ],
[5s, 5p, 5d, 5f, 5g], etc.
(ii) The spin-orbit interaction splits each level into a doublet according to
j = $±1/2, except for the s-states for which j = 1/2 is the only possibility.
(iii) The strength of the spin-orbit interaction decreases with n, since the
electrons in higher energy levels experience a smaller magnetic field as
the nucleus is better screened.
(iv) Finally, the dipole selection rules require ∆J = ±1, 0, parity change, and
∆$ = ±1.
Since the ns states are singlets with term 2S1/2 while the np states are doublets
with terms 2P3/2,1/2 , we can deduce that all the doublets must involve s ↔ p
type transitions. Since those in group II involve the same doublet spacing, they
must all involve the same p state. They are therefore likely to involve 4s → 3p,
5s → 3p, 6s → 3p and 7s → 3p transitions respectively.
By contrast, those in group I are likely to involve p states decaying to the same
s state. Noting that the first one has the same splitting as group II, we can
deduce that the they are likely to involve 3p → 3s, 4p → 3s, 5p → 3s and
6p → 3s transitions respectively, with the spin-orbit splitting decreasing with
n as expected.
The nd states have terms 2D5/2,3/2 . However, since the selection rules prohibit The Grotrian diagram for
transitions between J = 5/2 and J = 1/2, we can deduce that the transitions sodium showing (some) of
between d ↔ p states must involve triplets, as observed in group III. Since the the dipole allowed transitions.
splittings are the same as in group II, we can deduce that 3p is involved again. The two sodium D-lines are
They must be 3d → 3p, 4d → 3p, 5d → 3p and 6d → 3p transitions respectively. emphasized.
We can compare these predictions with the diagram (right) showing the spectra
of sodium (known as a Grotrian diagram).
(a) The 5p5/2 and 5p3/2 states are involved in the transitions of frequency
1.05086 and 1.05079 × 1015 Hz, which differ by 7 × 1010 Hz. The energy
splitting is thus 7 × 1010 h = 4.6 × 10−23 J = 0.29 meV.
(b) We expect the sodium energy levels to converge to the hydrogen levels
for large n, and thus to scale like 1/n2 . To test this, note that the ratio
of the energy differences (6s − 5s)/(7s − 6s) = 1.9, compared with the
expected ratio ( 25
1 1
− 36 )/( 36
1 1
− 49 ) = 1.65, which is not too bad. The energy
difference between the 7s state and the ionisation energy may be estimated
as the energy difference between 7s and 6s (Planck’s constant h times
0.049×1015 Hz) times ( 49 1
−∞ 1
)/( 36
1 1
− 49 ), which yields h×0.135×1015 Hz.
We add this to the energy difference between 7s and 3s, inferred fron the
sum of the 0.63142 × 1015 Hz (7s → 3p) and 0.50899 × 1015 Hz (3p → 3s)
transitions to obtain h × 1.27 × 1015 Hz, i.e. 5.2 eV. You could use other
states instead in a similar way.
(c) The spin-orbit energy in the p-states can be estimated from the splittings
of the corresponding doublets in group I. The Coulomb effect is given by
the difference between the p- and s-levels for a given n. The ratios in the
n = 3 and n = 6 cases are:
spin − orbit 0.00052
n=3: = = 1.02 10−3
coulomb 0.51
spin − orbit 0.00003
n=6: = = 1.07 10−3
coulomb 0.028
i.e. both effects decrease with n at about equal rates.
6. (i) First two terms represent interaction between the magnetic moments of
electron and proton respectively with the external field B; the final term
represents the spin-spin (hyperfine) interaction between the electron and
proton.
(ii) Neglecting the term in µp , and making use of the identity σ (e) · σ (p) =
(e) (p) (e) (p) (e) (p)
σz σz + 12 (σ+ σ− + σ+ σ− ), where
# $ # $ # $
0 1 0 2 0 0
σz = , σ+ = , σ− = .
1 0 0 0 2 0
We therefore have
etc. From these results, we can deduce the matrix elements of the Hamil-
tonian and confirm the expression given in the problem.
(iii) By inspection, the states | ↑e " ⊗ | ↑p " and | ↓e " ⊗ | ↓p " are energy eigen-
states with energies W + b and W − b respectively. The other two energy
eigenvalues are given by the solutions of
% %
% b−W −E 2W %
% % = 0,
% 2W −b − W − E %
√
which leads to E = −W ± 4W 2 + b2 . In the case where b . W , i.e.
where the external magnetic field is very weak, these reduce to W +b2 /2W
and −3W − b2 /2W , so there is a triplet of S = 1 states with energy close
to W , and the singlet S = 0 state with energy close to −3W . In the other
limit where b / W , i.e. where the external magnetic field is very strong,
this expression reduces to ±b, so we have two states with energy close to
+b and two close to −b, corresponding to the two possible orientations of
the electron spin.
7. Taking the 1s hydrogen wavefunctions as our basis, the matrix elements of the
Hamiltonian are defined in the question. Neglecting the overlap integrals, the
upper bound E on the energy levels is given, according to the Rayleigh-Ritz
variational method, by setting the determinant of the matrix Hij − Eδij to
zero. Multiplying out, and factorizing, this condition translates to the relation
* +
(α − E − γβ) (α − E)2 + γβ(α − E) − 2β 2 = 0 .
8. Defining (with obvious notation) |VB" = C[|a1 " ⊗ |b2 " + |b1 " ⊗ |a2 "], we have:
(i) the normalization |VB": !VB|VB" = C 2 [!a1 |a1 "!b2 |b2 " + 2!a1 |b1 "!b2 |a2 " + !b1 |b1 "!a2 |a2 "] =
C 2 (2 + 2SS ∗ ), and hence C 2 = 12 (1 + SS ∗ ).
(ii) Defining
- the normalized bonding and-antibonding orbitals, |g" = (|a" +
|b")/ 2(1 + S) and |u" = (|a" − |b")/ 2(1 − S), we have
/ / / /
1+S 1−S 1+S 1−S
|a" = |g" + |u", |b" = |g" − |u" .
2 2 2 2
As a result, we find that |VB" = C[(1 + S)|g1 " ⊗ |g2 " − (1 − S)|u1 " ⊗ |u2 "].
(iii) The state orthogonal to |VB" is then given by |ψ⊥ " = C[(1 − S)|g1 " ⊗
|g2 " + (1 + S)|u1 " ⊗ |u2 "]. Rewriting this in the |a" and |b" basis, after a
little algebra we arrive at:
C(1 + S 2 ) 2CS
|ψ⊥ " = {|a1 " ⊗ |a2 " + |b1 " ⊗ |b2 "} − {|a1 " ⊗ |b2 " + |b1 " ⊗ |a2 "} ,
(1 − S )
2 (1 − S 2 )
of which the first term represents the ionic component and the second the
covalent, i.e.
(1 + S 2 ) 2S
|ψ⊥ " = |IB" − |VB" .
(1 − S 2 ) (1 − S 2 )
(iv) Inserting the given value of ρ gives S = 0.697, and hence the ratio of
2
IB/VB in the state |ψ⊥ ": IB/VB = [ (1−S ) 2
2S ] = 0.14.
9. For the rotation/vibration system, the energy levels have the form
!2 J(J + 1)
E = E0 + (n + 1/2)!ω + ,
2I
and the selection rules are ∆n = ±1 and ∆J = ±1. Thus, neglecting any
change in the moment of inertia, we have the following energy changes:
!2
R branch J → J + 1 J = 0, 1, 2 · · · ∆E = !ω + (J + 1)
I
!2
P branch J → J − 1 J = 1, 2 · · · ∆E = !ω − J
I
(i) The above formulae imply equally spaced lines above and below !ω, but
the transition at !ω (∆J = 0, the Q-branch) is not permitted by the
parity change selection rule.
(ii) The double peaks are an isotope effect. Cl has two isotopes 35 Cl and
37
Cl with abundances in the ratio of roughly 3:1. The less abundant 37 Cl
will have a larger reduced mass (by about 1 part in 1000) and thus a
lower vibration frequency. Thus there are two separate spectra slightly
displaced in frequency.
(iii) The uneven spacing results from the change in moment of inertia. If
!2 /2I ≡ B initially and B + δB finally (with δB < 0), the formulae for
the energy changes become:
From the central vibrational frequency ω we can-infer the force constant of the
bond k: ω = 2πν 1 5.4 × 1014 Hz. Then, ω = k/µ, where µ is the reduced
mass (approximately equal the mass of the Hydrogen atom in this case), so we
deduce that k = 490 Nm−1 .
α , Ŝ β ] =
where J > 0, and the spins obey the quantum spin algebra , [Ŝm n
γ
i!δmn "αβγ Ŝm .
(a) Making use of the spin commutation relations and Ehrenfest’s the-
orem, show that, in the Heisenberg representation, the spins obey
the equations of motion,
dŜm
! = J Ŝm × (Ŝm+1 + Ŝm−1 ) ,
dt
where we suppose that the boundary conditions are periodic, i.e.
Sm+N = Sm .
(b) For large spin S, we may take the spin expectation values to be
defined by their classical values, #Ŝm $ = Sm . Moreover, if we are
interested in low-energy excitations of the spin chain, only modes
of long wavelength contribute (cf. the vibrational modes of the
harmonic chain). In this case, we may develop the Taylor series ex-
pansion, Sm+1 = Sm +∂Sm + 12 ∂ 2 Sm +· · ·, where the lattice spacing
is taken as unity. In doing so, show that the leading contribution
to the equations of motion in the gradient expansion is given by,
Ṡ = JS × ∂ 2 S .
chain. Notice that the dispersion relation in this case has a quadratic
dependence on k (cf. non-relativitic particles) as opposed to the lin-
ear dependence of the phonon spectrum. If you are feeling energetic,
you might contemplate the spin wave modes for the antiferromag-
netic chain where the dispersion becomes linear. Let us now consider
an alternative approach to the spin wave spectrum which follows an
operator-based formalism.
(d) Defining the spin raising and lowering operators, Ŝ ± = Ŝ x ± iŜ y ,
show that the ferromagnetic Heisenberg model can be written as
"# 1$ + − %&
− +
Ĥ = −J Ŝm+1 +
z z
Ŝm Ŝm Ŝm+1 + Ŝm Ŝm+1 .
m
2
Then, making use of the Holstein–Primakoff spin representation In lectures, we have discussed the
√ †
− = ! 2S a† (1 − am am )1/2 , Ŝ + = (Ŝ − )† , Fourier series expansion,
(defined on page 192), Ŝm m 2S m m
z = !(S −a† a ) where [a , a† ] = δ
and Ŝm m m m n mn , show that the Hamil- 1 "
f (x) = √ fk eikx
tonian can be expanded as a bilinear (i.e. to quadratic order) in L k
the raising and lowering operators, ' L
1
" † fk = √ f (x)e−ikx
Ĥ = −JN S 2 + S (am+1 − a†m )(am+1 − am ) + O(S 0 ) . L 0
m
In the following, we are deal-
(e) Being bilinear in operators (i.e. quadratic), the Hamiltonian can ing with a discrete lattice where
be diagonalized by discrete Fourier transformation. With periodic we have to consider the discrete
Fourier representation,
boundary conditions, a†m+N = a†m , defining
1 "
N
" fn = √ fk eikn
1 1 " −ikm † N k
a†k = √ eikm a†m , a†m =√ e ak ,
N m=1
N k 1 "
fk = √ fn e−ikn .
N n
where the sum on k = 2πn/N , runs over the N integers n =
−N/2 + 1, −N/2 + 2 · · · N/2, show that the transformed operators
obey the commutation relations, [ak , a†k! ] = δkk! . In the Fourier
representation, show that
"
Ĥ = −JN S 2 + !ωk a†k ak + O(S 0 )
k
4. Scattering theory: The following problem revises the derivation of the Born
approximation and then applies it to the high energy scattering from an
attractive square well potential.
Show that the Born approximation yields the following expression for
the elastic scattering of a particle of mass m and momentum !k from a
spherically symmetric potential V (r),
( ) *' *2
dσ 2m 2 ** ∞ *
* ,
= * V (r)r sin (Kr)dr *
dΩ !2 K 0
where ∆E = E2p − E1s = 3R∞ /4. Putting all this together we obtain the
probability of being in the 2p0 state after a long time as
e2 E02 a20 215 1
|c2p0 (∞)|2 = · .
3 10 ∆E + !2 /τ 2
2
3. From the lecture notes, the decay rate for unpolarized light is given by,
ω 3 |dkj |2
A= ,
3π!0 c3 !
and the lifetime is thus τ = 1/A. Take for example the 2p0 state of Hydrogen
decaying to 1s (the other 2p states must have the same lifetime, but this one
depends on the same matrix elements that we computed in in the previous
question. Only the z-component of d is non-zero for this transition, (the φ
integral yields zero if you compute the matrix elements of x or y) giving,
256ea0
%2p0 | ez |1s& = √ = 6.31 × 10−30 Cm .
243 2
The energy of the emitted photon is
3 3 me4
!ω = R∞ = · ⇒ ω = 1.56 × 1016 Hz .
4 4 2(4π!0 )2 !2
Hence, the lifetime of the state is τ = 1.56 × 10−9 s.
The only lower lying state to which 3s can decay is 2p according to the selection
rules. We can expect the matrix element %3s| ez |2p& ∼ ea0 on dimensional
grounds, and thus not very different from %2p| ez |1s&. The main difference
between the lifetimes of the 3s and 2p levels will arise from the difference in
ω 3 . For the 3s→2p transition,
1 1 5
!ω = ( − )R∞ = R∞ .
4 9 36
The ratio of the lifetimes is therefore approximately
+ ,3
τ (3s) 3 36
∼ · ∼ 150 .
τ (2p) 4 5
The only state lying below 2s is 1s, but the decay 2s→1s is not allowed by
the electric dipole selection rules. The 2s state is “metastable”. The dominant
decay is actually via two-photon emission, a process which can arise through
second order perturbation theory, and occurs very slowly. In practice, atoms
may well make transitions from 2s to 2p (for example) before decay takes place
as a result of collision processes. Alternatively, decay of the 2s state may be
induced by the application of an external electric field, which mixes 2s and 2p
through the Stark effect.
and hence
+ ,2 / * /2
dσ 2m / /
/ V (r)rdr sin(∆r)/ .
= / /
dΩ ∆!2
Taking V (r) = −V0 for r ≤ a, and V (r) = 0 otherwise, the integral becomes
(integrating by parts),
* a 23 4a * a 5
cos(∆r) cos(∆r)
−V0 r sin(∆r)dr = −V0 −r + dr
0 ∆ 0 0 ∆
V0
= − 2 (sin(∆a) − ∆a cos(∆a)) ,
∆
and thus
3 42
dσ 2mV0
= (sin(∆a) − ∆a cos(∆a)) .
dΩ !2 ∆ 3
In the low energy limit, ∆ → 0,
1
sin(∆a) − ∆a cos(∆a) ≈ ∆a − (∆a)3 − ∆a(1 − (∆a)2 /2) = (∆a)3 /3 ,
3!
and hence
+ ,2
dσ 2mV0 a3
= .
dΩ 3!2
This is independent of ∆ and hence independent of θ, so isotropic, as required.
The total cross-section is obtained by integrating over solid angles, which simply
involves multiplying by 4π in this case
+ ,2
2mV0 a3
σtot = 4π .
3!2
A sin(kR) = sin(KR + δ0 )
kA cos(kR) − k cos(kR + δ0 ) = U0 sin(kR + δ0 ) .
(c) Now supose that tan(kR) is small. In this case, we have a resonance when
k − U0 tan(kR) = 0, i.e. tan(kR) = Uk0 - 1, and
π π
δ0 = − kR " .
2 2
The cross-section σ0 = 4π
k2 sin δ0 " k2 . The resonance is near tan(kR) =
2 4π
[Ĥ, L̂i ] = !ijk [α · p̂, x̂j p̂k ] = !ijk (αl p̂l x̂j p̂k − x̂j p̂k αl p̂l )
= !ijk (−iαl δlj p̂k ) = −iα × p̂ .
1
[Ĥ, S] = [α · p̂, S] = (αi p̂i σj − σj αi p̂i )
3+ 2 , + ,4 + ,
1 0 σ · p̂ σ 0 1 0 1
= , = [σ · p̂, σ]
2 σ · p̂ 0 0 σ 2 1 0
+ ,
0 1
= −i p̂ × σ = −ip̂ × α.
1 0
8. Applying the plane wave solution of the Dirac equation ψ(p) = e−p·x u(p) (de-
fined in this form for positive and negative energy states) to the two edges of
the potential step, we obtain the boundary conditions
1 1
0 −ipa/2 0 ipa/2
p e +r p e
E+m − E+m
0 0
1 1
0 ! 0 ! ip! a/2
= t" p! e−ip a/2 + r" e
E ! +m − E !p+m
0 0
1 1 1
0 ! 0 ! −ip a/2
! 0
t" p! eip a/2 + r" e = t p eipa/2 ,
E ! +m − E !p+m E+m
0 0 0
where the reflection and transmission coefficients are defined in the figure.
From these equations we obtain
! !
2e−ipa/2 = t" (1 + ζ)e−ip a/2 + r" (1 − ζ)eip a/2
! !
2reipa/2 = t" (1 − ζ)e−ip a/2 + r" (1 + ζ)eip a/2
! !
teipa/2 = eip a/2 t" + e−ip a/2 r"
- ! !
.
teipa/2 = ζ eip a/2 t" − e−ip a/2 r" .
From this result, we obtain the expression for the transmitted current shown
in the question.
For energies E " > m, the particles traverse the barrier as a plane wave. In
particular, when p" a = nπ there is perfect transmission. For m > E " > −m,
p" is imaginary and exchange of particles occurs by resonant tunnelling across
the barrier. For energies E " < −m, the Klein paradox regime, p" is real and
positive, and there is again perfect transmission when p" a = nπ. Here the
transmission is mediated by negative energy states under the barrier.