Você está na página 1de 90

Giuseppe E.

Santoro

Lectures on Quantum Mechanics


ICTP Diploma Course

Printed on: May 7, 2012


Preface

These are just short notes intended as a guide to the Diploma Students taking their course in
Advanced Quantum Mechanics.
Contents

1 The basics of Quantum Mechanics 6

1.1 A brief history 6

1.2 Gaussian wave-packets 7

1.3 The Schrödinger equation: free-particle 9

1.4 The Schrödinger equation 10

1.5 The time-independent Schrödinger equation 14

2 One-dimensional problems 17

2.1 Infinite square well 17

2.2 Scattering by a step 18

2.3 Finite square well 19

2.4 Delta-function potential 20

3 Harmonic oscillator 21

3.1 Commutators and Hermitean conjugates 21

3.2 Dimensionless variables 22

3.3 The algebraic solution 23

4 The hydrogen atom problem 25

4.1 Orbital angular momentum 25

4.2 Separation of variables 27

4.3 Dimensionless variables 28

4.4 The Coulomb problem bound states 29

5 Theory of angular momentum 32

5.1 Diagonalization of J 2 and Jz 32

5.2 Construction of spherical harmonics 33

6 Symmetries 36

6.1 Space transformations 36

6.2 Transformations on scalar functions 36


4 Contents

6.3 Symmetry properties of Hamiltonians 38

6.4 Commuting observables 40

6.5 Selection rules 41

7 Measurement in Quantum Mechanics 42

7.1 Measurements in QM. 43

7.2 Expectation values of measurements. 44

8 Spin 46

8.1 Stern-Gerlach experiment: the spin operators and states 46

8.2 Experiments with Stern-Gerlach apparatus 48

8.3 A single spin-1/2 in a uniform field: Larmor precession 50

8.4 Heisenberg representation 50

9 Spin-orbit and addition of angular momenta 52

9.1 Relativistic effects 52

9.2 The origin of spin-orbit 52

9.3 Construction of J-states out of p-states 53

9.4 Generalization: addition of L and S 55

10 Density matrices 56

10.1 The density matrix for a pure state 56

10.2 The density matrix for a mixed state 57

10.3 Density matrices and statistical mechanics 57

10.4 Density matrices by tracing out the universe 58

11 Time-independent perturbation theory 60

11.1 Two-state problem 60

11.2 Non-degenerate case: general solution 61

11.3 Degenerate case: first order 66

11.4 Application: the fine-structure of Hydrogen 68

12 Time-dependent perturbation theory 71

12.1 General case 71

12.2 First-order solution 72

12.3 Electromagnetic radiation on an atom 75


Contents 5
13 Many particle systems 77

13.1 Many-particle basis states for fermions and bosons. 77

13.2 Two-electrons: the example of Helium. 81

Appendix A Second quantization: brief outline. 84

A.1 Why a quadratic Hamiltonian is easy. 88


1
The basics of Quantum Mechanics

1.1 A brief history

The history of quantum mechanics starts right at the beginning of the 1900. The years from 1900
to 1930 might be really seen, as the title of a nice book by Gamow says, as the 30 years that shook
physics.

This story has to do with light (more generally, electromagnetic radiation) as well as electrons
and atoms: while a full quantum treatment of light, however, necessarily requires a relativistic
quantum theory (which we will not enter into), electrons and atoms can be very well described by
a non-relativistic quantum theory, which will be the subject of our course. I urge you to read the
first chapter of any of the many books in QM, for instance that of Gasiorowitz, for a full account
of this very interesting story. I will be here very sketchy.

The story itself, begins really with electromagnetic (e.m.) radiation, more precisely with the
puzzle of the black-body radiation, an idealized furnace (oven) where the spectrum of the radiation
inside was experimentally well known but escaping any classical physics explanation. It was Max
Planck, in 1900, who postulated (very boldly, at that time), that everything could be explained if
one assumed that the energy of an e.m. field of frequency ν was not arbitrarily and continuously
changeable, but only multiples of the fundamental energy

E = hν = ~ω , (1.1)

h being the famous Planck’s constant (ω = 2πν, and ~ = h/2π) would be admitted. This fixed all
the difficulties that physicists had encountered in the black-body radiation problem. Constructing
on this idea, soon after (in 1905) Einstein was able to explain in a very natural way the photoelectric
effect: electrons are kicked out from the metal by individual “collisions” with a single “quantum” of
radiation (a “photon”): if the energy E = hν of the photon is not larger than the work function W
(by definition, the energy you have to provide to extract an electron from a metal), then no electron
will come out, no matter how “intense” you make the field (by shining very many photons on your
metal). The final experimental confirmation of the “particle-like” nature of the e.m. radiation came
with Compton (in 1923), who performed a true scattering experiment of X-rays from electrons,
where all the results could be rationalized by applying energy and momentum conservation to the
collision event; the momentum that appeared to be associated to the X-ray had to be:
E hν h h 2π
p= = = = = ~k , (1.2)
c c λ 2π λ
where k = 2π/λ is the wave-number.

In the atomic world, the first bricks of a quantum theory came with Bohr, in 1913 (at that
time a post-doc of Rutherford), who “solved” the mystery of the atomic spectra seen in emission
and adsorption. Rutherford had put forward a “planetary model” for the atoms, with the nucleus
occupying only a small central volume and the electrons orbiting around it. However, why those
orbiting electrons would not “emit radiation” (being accelerated) as the electrons in a synchrotron
do, was a mastery. Bohr elaborated a model for “quantifying” the allowed orbits, and their energies
Gaussian wave-packets 7
En (which were, once again, discrete) in such a planetary system, with results which were in
surprising agreement with spectral observations. In particular, the various radiation lines observed
could be accounted for as differences between the quantized energy values En that Bohr had
calculated, as:
hνn,m = En − Em . (1.3)
In 1925 de Broglie realized that the orbits of Bohr’s atom appeared to possess a nice wave-like
feature: an orbit of energy En (with n = 1, 2, · · · ) had a radius rn and a circumference 2πrn which
appear to “contain” exactly n wavelengths λ = h/p (p being the momentum of the electron in the
orbit):
2πrn = nλ . (1.4)
This mysterious wave-like nature of electrons was revealed later on (before 1930) in diffraction
experiments, by Davisson and Gerner, and, independently, by G. Thompson (son of the Thompson
who discovered the electron!). Notice that electrons of energy E = 160 eV have momenta p ≈
6.6 × 10−19 (in c.g.s), and therefore the resulting λ ≈ 10−8 cm. de Broglie was, however, too slow
in elaborating a full mathematical treatment of his electron waves. In 1926, Schrödinger came out
with a beautiful theory that put de Broglie insight on a firm ground, and explained in a very
natural way the results of Bohr.

1.2 Gaussian wave-packets

Question: how are we supposed to explain the motion of a material particle with a wave? A single
plane-wave eikx cannot certainly be representing anything localized, because it is periodic in space,
with a period λ = 2π/k. Therefore, in order to look for something “localized”, which might be
exchanged for a “point-particle”, at least from reasonably far away, we must superimpose many
plane waves and create a wave-packet. We do that, by taking a superposition of the form:
Z +∞
dk
ψ(x) = √ g(k)eikx . (1.5)
−∞ 2π

g(k) can in principle be any weight function representing how much of the wave eikx is included in
the superposition, for each k. In order to carry out the calculations explicitly, we will assume that
g(k) has the simple Gaussian form:
2
g(k) = Ce−α(k−k0 ) , (1.6)

where C is a normalization constant which we choose so that dk|g(k)|2 = 1. The reason why we
R

impose that the square of g(k) is normalized, and not g(k) itself, will be clear in a while.

Now, the simple Gaussian integral that you certainly have encountered many times is:
Z
dk − ak2 1
√ e 2 =√ . (1.7)
2π a
Based on this, it is simple to show that:
 1/4

C= .
π

By defining σk2 = 1/(4α), it is also quite simple to show that σk is the width of the function |g(k)|2 ,
i.e., its second moment is:
Z Z (k−k0 )2
2 2 2 2 2 − 2σ 2
h(k − k0 ) i = dk (k − k0 ) |g(k)| = C dk (k − k0 ) e k = σk2 . (1.8)

Prove, as an exercise, the latter identity.


8 The basics of Quantum Mechanics

Still based on the Gaussian integral in Eq. (1.7), by completing the square, it is a simple matter
to show the following generalization:
Z
dk − ak2 kJ 1 J2
√ e 2 e = √ e 2a . (1.9)
2π a

There is no restriction on J being a real quantity, in this integral. So, if we substitute J = ix,
where x is real, the equality still holds, and reads:
Z
dk − ak2 ikx 1 x2
√ e 2 e = √ e− 2a . (1.10)
2π a

This very useful identity, expressed in words by saying that the Fourier transform of a Gaussian
in k-space is a Gaussian in x-space, with an inversely proportional width, will be used now to
calculated ψ(x). Indeed, a straightforward calculation leads to:

Z +∞  1/4
dk 2 1 x2
ψ(x) = Ceik0 x √ e−αk eikx = eik0 x e− 4α , (1.11)
−∞ 2π 2πα

i.e., ψ(x) is a plane-wave eik0 x with a Gaussian envelope. If we consider |ψ(x)|2 , it will be a simple
Gaussian with a width σx2 = α. Notice that the widths of the distributions in k- and x-space are
related by a simple inverse proportionality, i.e.:

√ 1 1
σx · σk = α · √ = . (1.12)
2 α 2

We will see in a while that this is a realization of a minimum Heisenberg-uncertainty-principle


wave-packet (we will see that Heisenberg uncertainty principle requires ∆x∆p ≥ ~/2).

Let us now consider the time dependence of this wave-packet. For a single plane wave eikx the
time-dependence has the form:
ψk (x, t) = ei(kx−ωk t) ,

where ωk is the angular frequency of the wave. Such a form guarantees that the wave moves with
a phase velocity
ωk
vk = ,
k

since, indeed, ψk (x, t) = eik(x−vk t) , i.e., the wave amplitude depends only on x − vk t. Now consider
the full time evolution of ψ(x), which naturally reads:
Z +∞ Z +∞
dk dk
ψ(x, t) = √ g(k)ei(kx−ωk t) = √ g(k)ψk (x, t) . (1.13)
−∞ 2π −∞ 2π

The crucial question is the choice of the k-dependence of ωk . For the case of an electromagnetic
wave in vacuum, we have ωk = kc, and the phase velocity of each plane-wave component is identical
to the speed of light: vk = c. As a consequence, it is very simple to prove that, in such a case,
the wave would move rigidly, undistorted, with the speed of light, i.e., ψ(x, t) = ψ(x − ct, 0). One
refers to such a case as non-dispersive. On the contrary, whenever each k-component moves with a
different phase-velocity vk , the resulting wave has a dispersive behaviour: we will soon show that
this implies a rigid motion with a group velocity vg = dωk /dk plus an inevitable spreading of the
envelope. Clearly, to possibly describe the motion of a non-relativistic particle, we have to chose
this second route. In order to carry on the calculation, we will need to Taylor expand ωk (which is in
general a non-linear function of k) around the peak wave-vector k0 . To fully justify this expansion,
The Schrödinger equation: free-particle 9
let us assume that α is large, so that g(k) is quite narrow around its peak at k = k0 . The Taylor
expansion of ωk around k = k0 , up to second order, reads:

ωk = ωk0 + vg (k − k0 ) + β(k − k0 )2 + · · · , (1.14)

where
dωk
vg = , (1.15)
dk k0
and β = (1/2)d2 ωk /dk 2 |k0 . By substituting these expressions in ψ(x, t) we easily get:
Z +∞ (x−vg t)2
dk 2 C
ψ(x, t) = Cei(k0 x−ωk0 t)
√ e−(α+iβt)k eikx = p ei(k0 x−ωk0 t) e− 4(α+iβt) , (1.16)
−∞ 2π 2(α + iβt)

where we have used that fact that the Fourier transform in Eq. (1.10) holds even when a is complex,
as long as with a positive real part (so that the integrand does not blow up), which is the case here,
since α > 0. Consider now |ψ(x, t)|2 . It reads:
α(x−v t)2
r
2 1 α − 2(α2 +βg2 t2 )
|ψ(x, t)| = √ e . (1.17)
2π α2 + β 2 t2
Notice that, since β 6= 0, there is a spreading of the wave-packet, i.e., its width inevitably increases:
this is completely independent of the sign of β. Now, the velocity with which the center of the
wave-packet moves should be related to the classical momentum in the usual non-relativistic way:

dωk p0 ~k0
vg = = = , (1.18)
dk k0
m m

where we have used de Broglie’s identification of the momentum with ~k0 . It is then clear that the
only way this identity holds is that:

~2 k 2 p2
Ek = ~ωk = = , (1.19)
2m 2m
which is the standard non-relativistic kinetic energy of a particle.

Exercise 1.1 Consider the problem of spreading of a Gaussian wave-packet for a free non-relativistic
particle. Calculate how much the width of the wave-packet changes in t = 1 second in the following
two cases: a) the particle is an electron (m = 9 × 10−31 kg) and the initial wave-packet has dimensions
∆x = 10−5 m, or b) ∆x = 10−10 m; c) the particle has mass m = 10−3 kg and initial size ∆x = 10−3 m.
It is convenient to calculate, in all cases, the fractional change of the width given by:

∆xt − ∆x
.
∆x

1.3 The Schrödinger equation: free-particle

Let us now verify that a wave-packet of the form given in Eq. (1.16) – or more generally one
obtained from Eq. (1.13), as long as ωk has the non-relativistic form in Eq. (1.19) – satisfies the
following Schrödinger equation

∂ψ(x, t) ~2 ∇2
i~ =− ψ(x, t) , (1.20)
∂t 2m
where ∇2 = ∂ 2 /∂x2 + ∂ 2 /∂y 2 + ∂ 2 /∂z 2 is the Laplacian (in three-dimension, in one-dimension only
the x-component survives). Verifying this statement is very simple. We just prove that it is true
10 The basics of Quantum Mechanics

for a single plane-wave ψk (x, t) = ei(k·x−ωk t) , and then, by linearity of the Schrödinger equation, it
will be true also for any combination of plane-waves, as in Eq. (1.5). Verifying that ψk (x, t) solves
Eq. (1.20) is very simple and is left as an exercise (done in class).

Moreover, if we define the momentum operator as:

~ ,
p̂ = −i~∇ (1.21)

it is simple to show that the plane-waves are eigenstates of the momentum operator (remember de
Broglie):
p̂ eik·x = (~k) eik·x . (1.22)

It is also simple to show that the ∇2 term appearing in Eq. (1.20) can be also written as:

~2 ∇2 p̂2
− = , (1.23)
2m 2m

i.e., the same form as the kinetic energy of a free non-relativistic particle.

1.4 The Schrödinger equation

In general, for a particle which is subject to a potential V (x), the Schrödinger equation (SE) reads:

p̂2
 
∂ψ(x, t)
i~ = Ĥψ(x, t) = + V (x) ψ(x, t) . (1.24)
∂t 2m

The operator Ĥ appearing in this equation is the Hamiltonian of the system, which has the standard
non-relativistic form, except that it is an operator, acting of functions (wave-functions).

We must think of the problem in terms of a space of functions (the ψ’s) on which we act with
linear operators (p̂, x̂, Ĥ).

The SE is first-order in time. So, given any initial wave-function ψ(x, t = 0), we can in principle
calculate ψ(x, t) at any later time t > 0 in a perfectly deterministic way. There is a formal way of
expressing this solution, in terms of the so-called evolution operator. Let us consider for simplicity
the case in which the Hamiltonian is time-independent. (If the Hamiltonian is time-dependent,
as in the presence of external varying fields, the evolution operator can be still defined, but its
expression is more complicated.) In general, the exponential of an operator Ô is defined as:

def 1 2 X 1 n
eÔ = 1 + Ô + Ô + · · · = Ô , (1.25)
2! n=0
n!

where 1 is the identity operator, i.e., multiplying by 1. One should be careful, because Some of the
usual rules of ordinary exponential functions do not hold for exponential of operators, for instance:

eÔ1 eÔ2 6= eÔ1 +Ô2 , (1.26)

unless the two operators commute (i.e., Ô1 Ô2 = Ô2 Ô1 ). However, some other rules do apply here
too. For instance:
d Ôt
e = ÔeÔt = eÔt Ô . (1.27)
dt
The proof of this equality (done in class) is very simple, starting from the Taylor expansion of the
exponential.
The Schrödinger equation 11
Consider now the following operator:
iĤt
Û (t) = e− ~ . (1.28)
It is known as evolution operator, for a reason that will be clear in a moment. First of all, one
immediately has Û (t = 0) = 1, and:
d − iĤt iĤt
i~ e ~ = Ĥ e− ~ . (1.29)
dt
As a consequence, if we define the time-evolved state
iĤt
ψ(x, t) = Û (t)ψ(x, t = 0) = e− ~ ψ(x, t = 0) , (1.30)
it is a very simple matter to prove (done in class) that it satisfies correctly the initial-time condition
at t = 0 and the SE.

We ask now how to define, given the fact that our basic objects are wave-functions, the prob-
ability density P (x, t) for finding the particle at some position x at time t. Obvious requirements
for P (x, t) are:
a) P (x, t) ≥ 0
Z 
d
b) dx P (x, t) = 0
dt
Z
c) dx P (x, t) = 1 . (1.31)

The first requirement is common to all probability densities (real and non-negative functions); the
second tells us that the total probability is conserved in time. As a consequence of this conservation,
we will be able to rescale P in such a way to impose the standard normalization of the total
probability given in c) above. How is P (x, t) related to ψ(x, t)? For one thing, ψ(x, t) is inevitably
a complex wave-function! Even if we start from a real ψ(x, t = 0) the SE-evolution will make ψ(x, t)
complex. A possible guess would be to take P = |ψ(x, t)|, but we will show that any such guess will
violate the conservation requirement in b), except for the following one suggested by Max Born
(in strict analogy with the energy density of the electromagnetic field, in terms of the field itself):
P (x, t) = |ψ(x, t)|2 = ψ ∗ (x, t)ψ(x, t) . (1.32)
To show this in a nice way, it is better to introduce a few formal tools, which will be anyway useful
in the following. Notice that this form immediately implies interference between waves, because
|ψ1 (x, t) + ψ2 (x, t)|2 6= |ψ1 (x, t)|2 + |ψ2 (x, t)|2 .

1.4.1 A few formal tools

First of all, we introduce in the space of wave-functions ψ a structure of a linear space, which is
rather obvious: if c1 and c2 are arbitrary constants, and ψ1 , ψ2 wave-functions, then c1 ψ1 + c2 ψ2 is
also a possible wave-function. Moreover, we introduce in this space a scalar products, by analogy
with the usual euclidean scalar product of ordinary vectors:
Z
def
hψ1 |ψ2 i = dx ψ1∗ (x)ψ2 (x) = hψ2 |ψ1 i∗ . (1.33)

In this way, the norm ||ψ|| of a function is defined as:


Z
hψ|ψi = dx |ψ(x)|2 = ||ψ||2 . (1.34)

The mathematicians call this norm (there are many other possible scalar products and norms that
you can introduce) the L2 -norm. In this way, the linear space of all possible normalizable wave-
functions (i.e., with ||ψ|| < ∞), forms a space known as Hilbert space, which we will denote by H
12 The basics of Quantum Mechanics

(to distinguish it from the Hamiltonian H or Ĥ). Notice immediately that our beloved plane-waves
ψk (x) = eik·x are immediately rather bizarre, being ||ψk || = ∞. We will keep using them because
they are useful as states on which we expand physical normalizable wave-functions (remember
the wave-packet discussion, and the Fourier expansion); we will also soon learn how to regularize
them properly on a finite-volume, through a procedure very familiar in condensed-matter (periodic
boundary conditions).

Of all the linear operators that one can imagine to act on the wave-functions, of particular
importance are the so-called Hermitean operators, which are defined as those operators such that,
for any ψ1 and ψ2 , satisfy:
hψ1 |Ôψ2 i = hÔψ1 |ψ2 i , (1.35)

i.e., the operator Ô is Hermitean if it can move back-and-forth on the two sides of the scalar
product. Many (indeed all!) of the operators encountered so far are Hermitean. To start with, the
operator x̂, which simply multiplies by x a ψ, is Hermitean because x is real, and therefore (try
writing both terms explicitly to fully convince yourself):

hψ1 |x̂ψ2 i = hx̂ψ1 |ψ2 i . (1.36)

Similarly, any real potential V (x) corresponds to an Hermitean operator. Perhaps surprisingly (at
least at a first sight, because of the −i), the momentum operator is Hermitean: the very simple
proof (done in class) uses integration by parts and the extra minus sign in it makes for the necessary
complex conjugate of −i:
hψ1 |p̂ψ2 i = hp̂ψ1 |ψ2 i . (1.37)

Powers of an Hermitean operators are Hermitean (just move one operator at a time from the right
to the left of the scalar product). Therefore:

hψ1 |p̂2 ψ2 i = hp̂ψ1 |p̂ψ2 i = hp̂2 ψ1 |ψ2 i . (1.38)

Also, the sum of two Hermitean operators is Hermitean:

hψ1 |(Ôa + Ôb )ψ2 i = h(Ôa + Ôb )ψ1 |ψ2 i , (1.39)

as the linearity of the scalar product immediately shows. Finally, the Hamiltonian Ĥ of a particle
in a potential is Hermitean because it is the sum of two Hermitean operators, p̂2 and V (x).

Exercise 1.2 Of the following operators acting on functions, tell which ones are linear operators, and, out
of these, which ones are Hermitean operators: 1) O1 ψ(x) = x3 ψ(x),R x 2) O2 ψ(x) = xdψ/dx, 3) O3 ψ(x) =
λψ ∗ (x), 4) O4 ψ(x) = eψ(x) , 5) O5 ψ(x) = dψ/dx + a, 6) O6 ψ(x) = −∞ dx0 (x0 ψ(x0 )).

With these definitions and with the use of the SE and of its conjugate:

∂ψ ∗ (x, t)  ∗
−i~ = Ĥψ(x, t) , (1.40)
∂t

it is very simple to show that:


Z
d
i~ dx ψ ∗ (x, t)ψ(x, t) = hψ|Ĥψi − hĤψ|ψi = 0 , (1.41)
dt

where the last equality follows from the fact that Ĥ is Hermitean.
The Schrödinger equation 13
Before leaving this formal section, let us introduce the concept of expectation value of an operator.
By analogy with classical probability, a natural choice for the average position at time t < x > (t)
is:
Z Z
< x > (t) = dx xP (x, t) = dx ψ ∗ (x, t)xψ(x, t) = hψ(t)|x̂ψ(t)i .

Similarly, and more generally, the mean value (or expectation value) of an operator Ô on a state
ψ(t) is defined as:
Z
def
< O > (t) = hψ(t)|Ôψ(t)i = dx ψ ∗ (x, t)Ôψ(x, t) . (1.42)

Notice that Ô might involve derivatives, which should be applied only to the state to its immediate
right, and therefore, generally speaking, expectations values of operators cannot be expressed as
integrals where the probability P (x, t) appears!

Exercise 1.3 Prove that the expectation value of an Hermitean operator is a real quantity (done in class).

Exercise 1.4 Write down the expectation value of the momentum operator, and observe explicitly that:
1) it is real even though there is an imaginary unit i in its expression; 2) it cannot be expressed in terms
of P (x, t).

A final word on the notation. The Dirac’s notation is strictly related to what we have written
above, except for adding an extra |:

hψ1 |Ôψ2 i = hψ1 |Ô|ψ2 i , (1.43)

where the right hand side is in the Dirac’s notation, with the operator sitting between a bra hψ1 |
and the ket |ψ2 i.

Exercise 1.5 Define the uncertainty in x and p on a state ψ with the usual rules:

(∆x)2 = hψ|x2 |ψi − hψ|x|ψi2 , (∆p)2 = hψ|p2 |ψi − hψ|p|ψi2 .

Next define the following quantity:


Z
D(α) = dx |(x − hxi)ψ(x) + iα(p − hpi)ψ(x)|2 ≥ 0 ,

which is non-negative, by inspection, for every value of α. Expand the modulus square appearing in D(α)
to show that:
D(α) = (∆x)2 + α2 (∆p)2 − ~α ≥ 0 .
From the fact that this quadratic function of α is always non-negative, deduce something on the discrimi-
nant of the parabola, and show that Heisenberg’s uncertainty principle follows:

~
∆x ∆p ≥ .
2
14 The basics of Quantum Mechanics

1.4.2 The continuity equation and the current

To convince ourself that ψ(x, t) is the fundamental quantity of Quantum Mechanics, and not
P (x, t), let us write down the partial differential equation that P obeys. By using the SE and its
conjugate it is immediate to show that:
∂ ∂ 1  ∗ 
P (x, t) = (ψ ∗ (x, t)ψ(x, t)) = ψ (Ĥψ) − (Ĥψ)∗ ψ . (1.44)
∂t ∂t i~
Now, you can verify directly that the potential V (x) disappears completely from this expression,
and only the kinetic term survives:
∂ i~
ψ ∗ (x, t)(∇2 ψ(x, t)) − (∇2 ψ(x, t))∗ ψ(x, t) .

P (x, t) = (1.45)
∂t 2m
It takes now a small calculus exercise, based on identities like
∇ ~
~ · (ψ ∗ ∇ψ) ~ ∗ · ∇ψ
= ∇ψ ~ + ψ ∗ ∇2 ψ ,

to show that the right hand side of the latter equation is nothing but the divergence of a current
J(x, t):
∂ ~ · J(x, t)
P (x, t) = −∇
∂t
i~  ∗ ~ ~

J(x, t) = − ψ (x, t)(∇ψ(x, t)) − (∇ψ(x, t))∗ ψ(x, t)
2m
1
= (ψ ∗ (x, t)(p̂ψ(x, t)) + (p̂ψ(x, t))∗ ψ(x, t)) . (1.46)
2m
In this form, the equation for P expresses the differential form of the conservation of probability,
and is nothing but the familiar continuity equation which you encounter in electrodynamics.

1.5 The time-independent Schrödinger equation

Suppose the system is isolated, so that its Hamiltonian is time independent, Ĥ. (If the system is
subject to external fields, for instance an external electromagnetic field, then the Hamiltonian is
explicitly time-dependent, Ĥ(t), and the evolution operator is not the simple exponential we have
considered previously.). We already know that the formal solution of the SE, given an initial-time
ψ(x, t = 0), can be written as:
iĤt
ψ(x, t) = Û (t)ψ(x, t = 0) = e− ~ ψ(x, t = 0) . (1.47)
Our question, now, is if and when such a wave-function can factorized into a space and time
function:
ψ(x, t) = f (t)φ(x) . (1.48)
It is now difficult to show (done in class) that the only possibility this can happen is that φ(x) =
ψ(x, t = 0) is one of the eigenstates of Ĥ
Ĥφn (x) = En φn (x) , (1.49)
−iEn t/~
and, as a consequence, the time-dependent solution factorizes with f (t) = e , i.e.:
ψ(x, t) = e−iEn t/~ φn (x) . (1.50)
Notice that, for these solutions, P (x) = |ψ(x, t)|2 = |φn (x)|2 is time-independent, and, therefore,
the current must have zero divergence:
Jn (x) = 0 .

A general mathematical theorem guarantees that the full set of eigenstates φn of Ĥ provides an
orthonormal basis set for the Hilbert space: we will see several examples of this statement, with
The time-independent Schrödinger equation 15
n being a discrete set of labels, or even a continuous one. We will use a notation, for simplicity,
where n is assumed to be discrete. The fact that the energy eigenstates form an orthonormal basis
means that:
hφn |φm i = δn,m ,
and that every function ψ0 (x) can be expanded in terms of the φn :
X
ψ0 (x) = cn φn (x) .
n

Indeed, the expansion coefficients cn are simply given (by using the orthonormality of the basis)
as:
cn = hφn |ψ0 i .

Armed with the knowledge of the basis set {φn }, we can readily solve any time-dependent
Schrödinger problem with assigned initial condition ψ0 (x):
∂ψ(x, t)
i~ = Ĥψ(x, t)
∂t
ψ(x, t = 0) = ψ0 (x) . (1.51)

Indeed, all we need to do is to expand the initial condition ψ0 (x) on the eigenstates φn (x), ψ0 (x) =
P
n cn φn (x), and write down the state at time t as:
X
ψ(x, t) = cn e−iEn t/~ φn (x) . (1.52)
n

1.5.1 Free-particles and plane-wave normalization.

The simplest case it that of a free particle. Here the energy eigenvalues are simply Ek = ~2 k 2 /(2m),
associated to wave-functions φk (x) = eik·x . Notice that here k stands for the index n of the previous
section, and is in general a continuous multidimensional label. The spectrum of Ĥ, that is the set
of all its possible eigenvalues, is here continuous, and extends in the region E ∈ [0, +∞).

A few observations: 1) while k univocally specifies φk ,p


the same energy Ek = E is obtained by,
in general, infinitely many k, all those with |k| = kE = 2mE/~ (a sphere in 3d, a circle in 2d,
and 2 points in 1d). 2) All the φk are not normalizable in the infinite volume case. A trick often
used in condensed matter is to consider the system as confined to a cube of side L, hence of volume
Ω = L3 in 3d (Ω = Ld , in general dimension d), with periodic boundary conditions (PBC), i.e., we
postulate that:
φ(x + Lei ) = φ(x)
where ei is any of the three coordinate versors. This requirement (done in class) restricts the
possible allowed k-vectors to a discrete (although still infinite) set:

k= (n1 , n2 , n3 ) , (1.53)
L
with ni ∈ Z (i.e., positive and negative integers, or zero). The correctly normalized free-particle
eigenstates with PBC in the cube will then be:
1
φ̃k (x) = √ eik·x , (1.54)

with the k given by Eq. 1.53. Notice that now the φ̃k form a nice orthonormal basis for periodic
functions in the cube of volume Ω, since:
Z
1 0
hφ̃k |φ̃k0 iΩ = dx ei(k −k)·x = δk,k0 .
Ω Ω
16 The basics of Quantum Mechanics

The fact that φ̃k forms a basis set, simply a restatement of the fact that any periodic function
ψ(x + Lei ) = ψ(x) can be expanded in Fourier series:
X 1 X ik·x
ψ(x) = φ̃k (x)hφ̃k |ψi = e ψk , (1.55)

k k

where the Fourier coefficient ψk is simply given by:


√ Z
ψk = Ωhφ̃k |ψi = dx e−ik·x ψ(x) . (1.56)

It is now quite simple to take the thermodynamic limit L → ∞. Consider, indeed, the wave-
functions:
1
φk = p eik·x .
(2π)d
Their overlap, restricted to the volume Ω, tends towards the Dirac’s delta function:
Ω Ω Ω→∞
hφk |φk0 iΩ = hφ̃k |φ̃k0 iΩ = δk,k0 −→ δ(k − k0 ) .
(2π)d (2π)d
R
To prove this, you should start from the natural discretization of the integral dk, which is:
 d X Z
2π Ω→∞
−→ dk ,
L
k

and show that, for any function f (k):


 d X Z
2π 0 Ω Ω→∞
f (k) = f (k ) δk,k0 −→ dk f (k0 )δ(k − k0 ) .
L (2π)d
k0

As a consequence, you can take the limit of the Fourier sum in Eq. 1.55, obtaining:
Z
1 X ik·x Ω→∞ dk ik·x
ψ(x) = e ψk −→ e ψk . (1.57)
Ω (2π)d
k

The Ω → ∞ limit of the Fourier coefficient ψk in Eq. 1.56 gives:


Z
ψk = dx e−ik·x ψ(x) , (1.58)

which we recognize to be the Fourier transform ψk of the function ψ(x). Eq. 1.57 simply defines
the inverse Fourier transform. Notice that, when L → ∞, k becomes a continuous vector.
2
One-dimensional problems

Let us begin with some general remarks on the time-independent SE. First of all, the spectrum of
the Hamiltonian (i.e., the set of all its eigenvalues) is constrained to be not below the minimum of
the potential (if one exists):
En ≥ Vmin .
The proof of this is simple (done in class). Quickly, one can show that if Hψn (x) = En ψn (x) for
an eigenstate, then by taking the scalar product with ψn (assumed to be normalized) we get:
!
 2 2
 Z ~ n |2
~2 |∇ψ
∗~ ∇
Z
2 2
En = dx −ψn ψn + V (x)|ψn (x)| = dx + V (x)|ψn (x)| , (2.1)
2m 2m

from which it is immediate to conclude that En cannot be less then Vmin , because kinetic energy
is in general positive, and the wave-function is never really localized on the minimum of V .

A second general consideration regards bound versus extended states. A bound state is associated
to a discrete eigenvalue of H, and to a normalizable wave-function, usually quite well localized in
space. On the contrary, an extended state, exemplified by the familiar plane-wave eik·x , is not
normalizable in the infinite volume, and associated to an eigenvalue belonging to a continuous part
of the spectrum. We have already seen this for the free particle. We will see more examples of both
shortly.

A third general remark has to do with the continuity conditions of a solution of the SE at the
boundary between regions were the potential has some discontinuity. Generally speaking, a solution
of the SE must be continuous, with a continuous first derivative (indeed, ψ should admit a second
derivative!). For simplicity, assume we are in one-dimension. Suppose that x0 is a point were the
potential has possibly a finite (jump) discontinuity ∆V . By integrating the SE over the interval
[x0 − , x0 + ], writing:
Z x0 + Z x0 +
~2 ~2 0
− dx ψ 00 (x) = − [ψ (x0 + ) − ψ(x0 − )] = dx [E − V (x)]ψ(x) ,
2m x0 − 2m x0 −

and taking the limit  → 0 we can quickly prove that the first derivative ψ 0 (x) is continuous at x0 ,
unless the potential V (x) has an infinite jump ∆V at x0 , or some delta-function contribution.

2.1 Infinite square well

The first problem we have solved in class is that of an infinite square-well in the region [0, L]. In
other words, the potential is infinite for x < 0 and x > L, and V = 0 in [0, L]. Due to the fact that
ψ(x) must be 0 for x < 0 and x > L and it must be continuous everywhere (no exception to this
rule), we concluded that the appropriate boundary conditions for ψ(x) are:
ψ(0) = ψ(L) = 0 .
We solved this problem by taking combinations of eikx and e−ikx (in the free-particle region [0, L])
of the form:
18 One-dimensional problems

ψk (x) = Ak sin kx .
The boundary condition ψk (0) = 0 is automatically satisfied. As for the condition ψk (L) = 0, it
implies sin kL = 0, or:

k= with n = 1, 2, · · · (2.2)
L
Negative n’s do not bring extra solutions, because sin −kx = − sin kx. The corresponding energy
eigenvalue is
~2 k 2 ~2 π 2 n2
Ek = = . (2.3)
2m 2mL2
Each solution is non degenerate, i.e., only one eigenstate is associated to each eigenvalue. Notice
that the spectrum is now purely discrete, as an effect of the confining potential. For L → ∞ we
recover a dense continuous spectrum in [0, +∞).

We discussed some properties of the box wave-functions. First of all, we showed that hψk |p̂|ψk i =
0 (because you can always take the ψk to be real). Next, we calculated hψk |p̂2 |ψk i in terms of energy
eigenvalues by using the fact that p̂2 = 2mĤ. We commented on the Heisenberg uncertainty
principle, showing that (∆x) · (∆p) grows with the quantum number n. We also commented on
the increasing number of zeroes of the eigenstates, and the corresponding kinetic energy increase,
~ 2.
since you can always re-express the kinetic energy as an integral of |∇ψ|

Exercise 2.1 Solve the same infinite square well problem by choosing the origin in the center of the region,
which is therefore [−L/2. + L/2], showing that now you have cos kx and sin kx solutions, alternating (the
ground state is even, the first excited state is odd, and so on). Plot the first eigenstates. Calculate the
eigenstate normalization factors, and show the orthogonality of different eigenstates.

Obviously, the eigenfunctions of the Hamiltonian form a orthonormal basis in the chosen in-
terval: they are the basis element of the Fourier series. For instance, with the symmetric choice
[−L/2, +L/2] the Fourier series reads:
X  nπ  X  nπ 
ψ(x) = an cos x + bn sin x . (2.4)
n=1,3,···
L n=2,4,···
L

Exercise 2.2 Write the expressions for the Fourier coefficients an and bn , taking proper care of the correct
normalization of the basis functions cos kx and sin kx.

2.2 Scattering by a step

Next, we did a scattering problem: a potential step of V0 > 0 for x > 0. There is no loss of generality
in doing so: if V0 < 0 you can reset the zero of energy and view it as a step on the left of the
origin. Since the minimum of the potential is 0, we have E ≥ 0. We should distinguish two cases: i)
E > V0 and ii) E ≤ V0 . For the case i), E > V0 we have that in both x < 0 and x > 0 the problem
is a free-particle problem, but with energy, respectively, E and E − V0 (because you can bring the
constant potential on the right-hand side of the SE). We therefore consider candidate solutions of
the form:
e + re−ikx for x ≤ 0
 ikx
ψk (x) = , (2.5)
teiqx for x ≥ 0
with positive k and q, so that the ψk describe a wave incoming from the left of the origin, partially
reflected (reflection amplitude r), and transmitted to the right of the origin (transmission amplitude
t). The overall normalization factor is not important (this is why we put the coefficient of eikx to
1) because the ψk is intrinsically not-normalizable, as any extended (unbound) state is. In order
Finite square well 19

to be a solution with energy E, we must have (verify) E = ~2 k 2√


/2m and E − V0 =p~2 q 2 /2m. So,
for any given E, k > 0 and q > 0 are univocally defined: ~k = 2mE, and ~q = 2m(E − V0 ).
The continuity of ψk (always true) and of ψk0 (because the potential has just a finite jump), lead
to the following two equations:

t = 1+r
tq = k(1 − r) , (2.6)

whose solution gives:


2k k−q
t= and r= . (2.7)
k+q k+q

The other case, ii) E < Vp


0 , is solved in a similar way, but with the choice of an evanescent wave
te−qx for x ≥ 0, with ~q = 2m(V0 − E). The calculations are very similar to case i), and simply
require substituting q → iq, obtaining:

2k k − iq
t= and r= . (2.8)
k + iq k + iq

We noted that r is a pure phase factor, and that, for V0 → +∞, we have q → +∞, r → −1 and
t → 0, and the wave-function becomes 2i sin kx for x ≤ 0, and vanishes for x ≥ 0, i.e., it is still
continuous in 0, but its derivative becomes discontinuous in the origin (an infinite V0 does not
preserve continuity of ψ 0 .

We discussed the degeneracy of solutions, the physical meaning of the reflection and transmission
amplitudes r and t, the expression for the current (which must be constant everywhere).

2.3 Finite square well

Next we considered the case of a finite, symmetrical, potential well in [−a, a]: the potential is set
to 0 in [−a, a], and is equal to V0 for |x| ≥ a. Once again we must have E ≥ 0, and we can have
two cases: i) E < V0 (search of bound states), and ii) E > V0 (scattering states, with transmission
and reflection coefficients).

In case i), E < V0 we write the wave-function as:


 qx
 ae for x ≤ −a
ψk (x) = beikx + ce−ikx for |x| ≤ a , (2.9)
 −qx
de for x≥a
√ p
with ~k = 2mE, and ~q = 2m(V0 − E). Next we discussed how to use parity, taking d = a and
c = b for even solutions, and d = −a and c = −b for odd solutions. In both cases, once continuity
of ψ and ψ 0 are imposed at point x = −a, they are automatically guaranteed at the other point
x = +a. We arrived at the (graphical) solution for the bound-state eigenvalues, proving that there
is at least one bound state independently of a and V0 , while the total number of bound states
present (both even and odd) depends on the depth of the well.

Exercise 2.3 Solve the finite symmetric square well for the scattering states with E > V0 , by setting up,
as usual in scattering problems, ψk (x)√= eiqx + re−iqx forp x ≤ −a, ψk = ae
ikx
+ be−ikx for |x| ≤ a, and
iqx
ψk (x) = te for x ≥ +a, with ~k = 2mE, and ~q = 2m(E − V0 ). Calculate the current everywhere,
and deduce a relationship between t and r. Show that there are peculiar energy values (corresponding to
resonances for the infinite square well) at which the transmission coefficient becomes 1 (resonant tunneling).
20 One-dimensional problems

Exercise 2.4 By exploiting the results of the symmetric finite potential well (in particular, the odd
solutions of it) solve the case of a square well with an infinite wall in the origin, i.e., such that the potential
is +∞ for x < 0, 0 for [0, a], and V0 for x > a.

2.4 Delta-function potential

Finally, we considered a rather singular potential, an attractive delta-function in the origin: V (x) =
−V0 δ(x) with V0 > 0. We discussed how the boundary condition for the derivate ψ 0 should be
imposed, and found the bound state of the problem.

Exercise 2.5 Discuss scattering solutions with E ≥ 0 for both the delta-function potential, in both
attractive and repulsive cases.
3
Harmonic oscillator

The problem we consider now is very important, because it arises in many contexts as an approx-
imation to the real, more complicated, potential (close to a minimum). The Hamiltonian reads:

P2 1
Ĥ = + mω 2 X2 = Hx + Hy + Hz , (3.1)
2m 2
where we have explicitly indicated the separation into three one-dimensional Hamiltonian each
referring to a given Cartesian component (α = x, y, z):

Pα2 1
Hα = + mω 2 Xα2 . (3.2)
2m 2
Obviously, these three Hamiltonians commute among each other. We now illustrate how to solve
the 3D problem by the method of separation of variables. It is easy to prove (done in class) that
if we solve the one-dimensional problems:

Hα ψn(α) (Xα ) = En(α) ψn(α) (Xα ) , (3.3)


(x) (y) (z) (x)
then Ψn,m,l (X, Y, Z) = ψn (X)ψm (Y )ψl (Z) is an eigenstates of Ĥ with energy En,m,l = En +
(y) (z)
Em + El . So, it is enough to solve a one-dimensional harmonic oscillator (indeed, one might
generalize a bit the 3D Hamiltonian and consider a different ωα for each component). Omitting all
indices, we write the Hamiltonian as:

P2 1
H= + mω 2 X 2 . (3.4)
2m 2

3.1 Commutators and Hermitean conjugates

The concept of commutator is central to QM. The order in which two operators are applied to a
wave-function matters. Even in the finite-dimensional case, when the operators are simple matrices,
this is so. For instance, consider two matrices A, and B defined as follows:
   
1 0 01
A= , B= . (3.5)
0 −1 10

Then, it is easy to show that:


   
0 1 0 −1
A·B = , B·A= . (3.6)
−1 0 1 0

We define the commutator of A and B as:


def
[A, B] = A · B − B · A . (3.7)

Consider, for instance A = X̂ and B = P̂ . When acting on a generic wave-function ψ(X):


22 Harmonic oscillator

d
X̂ P̂ ψ(X) = (−i~)X ψ(X) .
dX
However:  
d d
P̂ X̂ψ(X) = (−i~) Xψ(X) = (−i~) ψ(X) + X ψ(X) = .
dX dX
Therefore, on any ψ(X) we have that:

[X̂, P̂ ] = i~ . (3.8)

Exercise 3.1 Prove that the following equalities hold:

[AB, C] = A[B, C] + [A, C]B ,

[A, BC] = [A, B]C + B[A, C] .

With these equalities, it is simple to calculated the following commutators:

Exercise 3.2 Show that [X̂, P̂ 2 ] = 2i~P̂ , [X̂, P̂ 3 ] = 3i~P̂ 2 , and, in general, [X̂, P̂ n ] = ni~P̂ n−1 , or,
for functions f (P̂ ) which can be expanded in Taylor series: [X̂, f (P̂ )] = i~f 0 (P̂ ), where f 0 denotes the
derivative. Similarly, show that [X̂ n , P̂ ] = ni~X̂ n−1 , and [f (X̂), P̂ ] = i~f 0 (X̂).

An important mathematical concept related to operators is that of the Hermitean conjugate of


an operator A, denoted with A† . The definition is simple. For any two vectors ψ1 and ψ2 in the
Hilbert space, one has to have:
hψ1 |Aψ2 i = hA† ψ1 |ψ2 i . (3.9)
In words, you can bring an operator acting on the ket in such a way that it acts on the bra, if
you take the Hermitean conjugate. Hermitean operators, evidently, are the Hermitean conjugate
of themselves, i.e., A = A† .

Exercise 3.3 If A and B are Hermitean operators, show that:

(A + iB)† = A† − iB † = A − iB .

3.2 Dimensionless variables

It is always a good practice, both in analytical work as well as in numerical implementations on


the computer, to work with dimensionless variables, so that parameters like the mass m, or the
frequency ω, or universal constants like ~, drop out of the problem. In the present case, for instance,
it is a good idea to work with a length l such that the potential energy of the spring and the typical
confinement energy for a particle in a box of the same length are equal:

~2 1
2
= mω 2 l2 .
2ml 2
This provides a typical “oscillator length”, defined by:
The algebraic solution 23
r
~
l= .

Define now a dimensionless length x̂ = X̂/l and a dimensionless momentum p̂ = P̂ l/~. It is simple
to verify that their commutator is [x̂, p̂] = i, i.e., effectively as if ~ = 1. It is now simple to verify
that the Hamiltonian reads:
1 2
p̂ + x̂2 ,

Ĥ = ~ω ĥ with ĥ = (3.10)
2
where we have defined a dimensionless Hamiltonian ĥ = Ĥ/(~ω). It is now clear that this procedure
is totally equivalent to setting m = 1, ω = 1, ~ = 1, as often stated in books or articles. You should
always be aware, however, of the exact meaning of this statement, and be ready to reintroduce the
appropriate parameters and universal constants if asked to produce explicit numbers. Summarizing,
we measure lengths in units of l, masses in units of m, energies in units of ~ω, momenta in units
of ~/l, times in units of ~/(~ω) = 1/ω, etc.

3.3 The algebraic solution

Define now:
1
a = √ (x̂ + ip̂) . (3.11)
2
a is quite clearly not Hermitean (see exercise 3.3). Its Hermitean conjugate is:
1
a† = √ (x̂ − ip̂) . (3.12)
2
In x-representation, both are quite simple operators:
   
1 d † 1 d
a= √ x+ and a = √ x− . (3.13)
2 dx 2 dx
The commutator of a and a† is immediately calculated:
[a, a† ] = 1 .

It is simple to invert the definition of a and a† to re-express x̂ and p̂ in terms of a and a† ,


obtaining:
1 1
x̂ = √ a + a† a − a† .
 
and p̂ = √ (3.14)
2 2i

As done in class (do it as an exercise), one can then show that:


 
† 1
ĥ = a a + . (3.15)
2

Let us start finding a state |ψ0 i which is annihilated by a, i.e. such that:
a|ψ0 i = 0 .
Such a state has a simple expression in x-representation, ψ0 (x) = hx|ψ0 i, since it obeys the equa-
tion:  
1 d
√ x+ ψ0 (x) = 0 ,
2 dx
which can be rewritten as:
d
ψ0 (x) = −xψ0 (x) ,
dx
and is solved by:
2
ψ0 (x) = Ce−x /2 ,
where C a normalization constant.
24 Harmonic oscillator

Exercise 3.4 Calculate the normalization constant C appearing in the state ψ0 .

Evidently, |ψ0 i is an eigenstate of ĥ with eigenvalue 1/2:


1
ĥ|ψ0 i = |ψ0 i .
2
It is also easy to show that |ψ0 i must be the ground state of ĥ:

Exercise 3.5 Given any state |ψi, the expectation value of ĥ satisfies:
„ «
1 1 1
hψ| a† a + |ψi = + ||a|ψi||2 ≥ .
2 2 2

Construct now the state |ψ1 i = a† |ψ0 i. Using the commutator [a, a† ] = 1 it is simple to show
that |ψ1 i is an eigenstate with energy 3/2. In real space, the representation of ψ1 is simple:
 
1 d
ψ1 (x) = hx|ψ1 i = √ x− ψ0 (x) .
2 dx
Calculate and plot this wave-function.

More generally, we are now going to use an important property of the number operator

n̂ = a† a . (3.16)

Exercise 3.6 Prove (for instance by recursion) that:


“ ”n “ ”n
n̂ a† |ψ0 i = n a† |ψ0 i ,

or, in words, that the state with n operators a† applied to ψ0 is an eigenstate of the number operator
n̂ = a† a, with eigenvalue n, i.e., n̂ counts the number of times a† appears.

Based on this, and observing that ĥ = n̂+1/2, we realize that the eigenstates of ĥ are constructed
as:  
n 1
|ψn i = Cn a† |ψ0 i =⇒ ĥ|ψn i = n + |ψn i ,
2
where Cn is a normalization constant. It is easy to show that a and a† allow us to move up and
down in the ladder of states |ψn i.

Exercise 3.7 Prove that the states


(a† )n
|ψn i = √ |ψ0 i ,
n!

√ √
are correctly normalized. Deduce also that: a |ψn i = n + 1|ψn+1 i and a|ψn i = n|ψn−1 i.

Notation: You often find |ni in place of |ψn i and |0i in place of |ψ0 i. The state |0i is called the
vacuum of a.
4
The hydrogen atom problem

The hydrogen atom problem consists of a proton (more generally, a nucleus of charge +Ze), with
an electron (charge −e) orbiting around it. The Hamiltonian is:

p21 p2 Ze2
H= + 2 − , (4.1)
2m1 2m2 |r2 − r1 |

where m1 indicates the mass of the nucleus, m2 that of the electron, and (r1 , p1 ) and (r2 , p2 )
denote coordinate and momenta for nucleus and electron, respectively. This is a typical two-body
problem, and, as in classical mechanics, one solves it by separating variables into center-of-mass
and relative motion. To this end, one introduces the center of mass R = (m1 r1 + m2 r2 )/M and the
total momentum P = p1 + p2 . It is a simple matter to show that these are canonically conjugate
variables, i.e., their commutator is
[Rα , Pβ ] = i~δα,β .
Regarding the relative motion, an obvious coordinate is the relative coordinate r = r2 −r1 . However,
the relative momentum is less trivial: the naive guess p = p2 −p1 is wrong, since it is not canonically
conjugate to r, and does not commute with R either, as it should. The way to go is to pose
p = β2 p2 − β1 p1 and determine β1 and β2 requiring canonical conjugation. The solution is:
m1 m2
p= p2 − p1 . (4.2)
M M

Therefore, the problem admits now separation of variables, since the total Hamiltonian is:

P2 p2 Ze2
H= + − . (4.3)
2M 2µ |r|

The total wave-function can be written as Ψ(R, r) = Φ(R)ψ(r). The center-of-mass problem is
just a free-particle problem, with wave-function Φ(R) = eik·R and energy Ek = ~2 k 2 /2M (there is
a translation symmetry for the whole atom): we will neglect it from now on. The relative particle
wave-function ψ(r) obeys the 3D problem:
 2
Ze2

p
− ψ(r) = Eψ(r) . (4.4)
2µ |r|

This would be still a quite intricate problem (we would need a computer if the potential was not
spherically symmetric): luckily enough, we can still separate variables in spherical coordinates.
This require introduction of a very important operator: the angular momentum.

4.1 Orbital angular momentum

The classical mechanics angular momentum is defined as:

L=r×p.
26 The hydrogen atom problem

In quantum mechanics, the corresponding operator is obtained by assuming that r and p are
operators. The three Cartesian components of L are immediate:

Lx = ypz − zpy
Ly = zpx − xpz
Lz = xpy − ypx . (4.5)

Commutators between the different components of L are easy to establish, using the usual rules
for commutators, such as [AB, C] = A[B, C] + [A, C]B. The result (worked out in class) is:

[Lx , Ly ] = i~Lz ,

and similar cyclic relations, which can be all summarized as:

[Lα , Lβ ] = i~αβγ Lγ . (4.6)

where αβγ is the totally antisymmetric tensor appearing also in the vector product of two ordinary
vectors: (a×b)γ = αβγ aα bβ (sum over repeated indices are always assumed). By changing variables
from Cartesian to spherical coordinates,

x = r sin θ cos φ
y = r sin θ sin φ
z = r cos θ , (4.7)

it is relatively simple to show that all components of L do not depend on r, but only on the angles
θ and φ:
 
∂ ∂
Lx = i~ + sin φ + cos φ cot θ
∂θ ∂φ
 
∂ ∂
Ly = i~ − cos φ + sin φ cot θ
∂θ ∂φ

Lz = −i~ (4.8)
∂φ

Exercise 4.1 Verify that the three components of angular momentum are given by Eq. 4.8. To this end,
it might be useful to show that:
∂ ∂
0 1 0
− sin φ
10 1
∂x sin θ cos φ cos θ cos φ sin θ ∂r
∂ 1 ∂
@ ∂y
A = @ sin θ sin φ cos θ sin φ + cos φ A@
sin θ r ∂θ
A . (4.9)
∂ 1 ∂
∂z
cos θ − sin θ 0 r ∂φ

Another important operator is:


L2 = L2x + L2y + L2z , (4.10)

which can be expressed in spherical coordinates as:

∂2 1 ∂2
 
1 ∂
L2 = −~2 + + . (4.11)
∂θ2 tan θ ∂θ sin2 θ ∂φ2
Separation of variables 27

Exercise 4.2 Verify the expression in Eq. 4.11 for L2 .

It is simple to show (done in class) that [L2 , Lα ] = 0. Therefore, we can find simultaneous eigen-
functions of L2 and, say, Lz (we will prove this result, in general, later on). We will indeed show, in
the next lecture, that we can construct functions of θ and φ, known as spherical harmonics, which
are common eigenfunctions of L2 and Lz :

L2 Yl,m (θ, φ) = ~2 l(l + 1)Yl,m (θ, φ)


Lz Yl,m (θ, φ) = ~mYl,m (θ, φ) . (4.12)

with l integers and m = −l, · · · + l.

The reason why all this is useful is that p2 (i.e., the Laplacian) can be easily expressed in terms
of L2 and derivatives with respect to r. Classically, consider the plane containing r and p, and
convince yourself that
L2 = r2 p2⊥ = r2 (p2 − p2k ) = r2 p2 − (r · p)2 .
Quantum mechanically, things are slightly more complicated because r and p do not always com-
mute. You can still prove (done in class) that:

L2 = r2 p2 − (r · p)2 + i~r · p . (4.13)

This equality can be shown either directly, by components, or by doing the following exercise:

Exercise 4.3 Using the fact that Li = ijk xj pk , with the convention of summing on repeated indices,
construct L2 = Li Li , and use the fact that ijk ilm = (δjl δkm − δjm δkl ) to prove the result for L2 .

Moreover, if you calculate

∂ ∂x ∂ ∂y ∂ ∂z ∂ 1 ~
= + + = r·∇ r ,
∂r ∂r ∂x ∂r ∂y ∂r ∂z r

you can immediately conclude that:



r · p = −i~r . (4.14)
∂r
Therefore, using Eqs. 4.13 and 4.14 we can easily rewrite the Laplacian as:

~2 ∂ ~2 ∂
 
2 ∂ 1
p = − 2r r − 2r + L2
r ∂r ∂r r ∂r r2
∂2 2~2 ∂ 1
= −~2 2 − + 2 L2 . (4.15)
∂r r ∂r r

4.2 Separation of variables

This expression for the Laplacian clearly separates the r-variable from the θ and φ variables, which
are contained only in L2 . As a result of that, you can factorize the wave-function ψ(r) as

ψk,l,m (r) = Rk,l (r)Yl,m (θ, φ) ,

where k is an extra quantum number (still not specified) which labels the many possible solutions
of the equation for R, which reads:
28 The hydrogen atom problem

~2 ∂2 ~2 l(l + 1) Ze2
   
2 ∂
− + Rk,l + − Rk,l = Ek,l Rk,l . (4.16)
2µ ∂r2 r ∂r 2µr2 r
Notice that this is a one-dimensional differential equation, which, however, has not the standard
form of a SE. The normalization condition is also quite different from a 1D normalization, since it
reads: Z +∞
dr r2 |Rl,l (r)|2 = 1 .
0
It is therefore tempting to define a new function

uk,l (r) = rRk,l (r) ,


R +∞
which, to start with, will have a standard normalization 0 dr |uk,l (r)|2 = 1. Moreover, you can
immediately verify that the non-standard derivative terms for R become completely standard when
written in terms of derivatives of u:
 2
1 ∂ 2 uk,l

∂ 2 ∂
+ R k,l = .
∂r2 r ∂r r ∂r2

When written in terms of uk,l the SE becomes therefore a standard 1D SE in the half-line [0, ∞):

~2 ∂ 2 uk,l
 2
~ l(l + 1) Ze2

− + − uk,l = Ek,l uk,l , (4.17)
2µ ∂r2 2µr2 r
with a standard normalization condition:
Z +∞
dr |uk,l (r)|2 = 1 . (4.18)
0

4.3 Dimensionless variables

We now get rid of all the physical constants (~, e, µ, Z) by switching to dimensionless variables.
The procedure is quite similar to that for the harmonic oscillator. Let us define a length a such
that potential energy and kinetic energy of confinement are of the same order:

e2 ~2
Z = .
a µa2
Solving this equation we get:
~2
a = aB = , (4.19)
µe2 Z
which we recognize to be the effective Bohr radius of the problem. The corresponding energy is:

e2 µe4 Z 2
EH = Z = . (4.20)
aB ~2
This quantity is, for Z = 1, the Hartree and is equal to EH ≈ Z 2 (27.2)eV . Measuring lengths
in units of aB , energies in units of EH and masses in units of µ defines a system of units called
atomic units, often abbreviated as “a.u.”. If ρ = r/aB is the dimensionless length and Ẽ = E/EH

the dimensionless energy, defining ũk,l (ρ) = aB uk,l (r = aB ρ), we get the dimensionless SE in the
form:
1 ∂ 2 ũk,l
 
l(l + 1) 1
− + − ũk,l = Ẽk,l ũk,l , (4.21)
2 ∂ρ2 2ρ2 ρ
with the normalization condition:
Z +∞
dρ |ũk,l (ρ)|2 = 1 . (4.22)
0
The Coulomb problem bound states 29
Notice that this looks like a standard 1D problem in the half-line, with an effective potential given
by:
(l) l(l + 1) 1
Veff (ρ) = − , (4.23)
2ρ2 ρ
i.e., comprising an attractive Coulomb term plus a repulsive centrifugal term for l > 0, which tries
to push electrons out of the nucleus (ρ = 0) the higher is the angular momentum (which is a
conserved quantity). Notice also that exactly the same discussion would have been possible for any
central potential V (r): the explicit form of V (r) would be substituted to the 1/r in the effective
potential. Nevertheless, as we shall see, the 1/r form is rather special in being exactly solvable and
having peculiar degeneracies of levels.

4.4 The Coulomb problem bound states

First of all, we denote simply by u the ũ and omit the tilde everywhere. We also denote by u0 and
u00 the first and second derivatives with respect to ρ. The first part of the analysis would apply to
any reasonable potential, and looks for asymptotic behaviours of u for small and large ρ.

Small-ρ behaviour. Take a power law u(ρ) = A1 ρα1 + · · · where the dots indicate less relevant
powers for ρ → 0 (i.e., those with a larger exponent). Taking derivatives we have: u0 = A1 α1 ρα1 −1 +
· · · and u00 = A1 α1 (α1 − 1)ρα1 −2 + · · · . Substituting in the SE, keeping only the most relevant
terms in each piece we have:

1  l(l + 1) 
A1 α1 (α1 − 1)ρα1 −2 + · · · + A1 ρα1 −2 + · · · − A1 ρα1 −1 + · · · = E [A1 ρα1 + · · · ] .
  

2 2

The most relevant term is that coming out of u00 and the l-centrifugal term, both going as ρα1 −2 .
In order for this most relevant term to be satisfied, we need:

α1 (α1 − 1) = l(l + 1) ,

which admits two solutions: i) α1 = l + 1 > 0 and ii) α1 = −l ≤ 0. The first case is certainly
admissible: it predicts that u(ρ) ∼ A1 ρl+1 , i.e., the smaller the larger is l, and in particular
u(0) = 0 (a kind of boundary condition) for all l. The possibility ii) turns out to be physically
absurd: it predict a larger and larger u close to the origin as l increases. And even for l = 0, it
would predict a constant u(0) (non-vanishing), which in turn implies a R(r) ∼ 1/r close to the
origin. But a wave-function behaving like 1/r close to the origin, would give rise to a delta function
term when we act with the Laplacian on it, which would never be canceled by any other piece in
the SE. Therefore, we conclude that:

ukl (ρ) ∼ A1 ρl+1 + · · · for ρ → 0 . (4.24)

Large-ρ behaviour. Next consider the large ρ behaviour. There, the potential (including the
centrifugal repulsion) has decreased to zero, so that the SE becomes:

1 ∂ 2 uk,l
− ≈ Ek,l uk,l ,
2 ∂ρ2

where Ek,l < 0 because we are considering bound states. A function which satisfies this approximate
equation is the exponential:
uk,l ∼ e−λk,l ρ ,
p
with λk,l = −2Ek,l .
30 The hydrogen atom problem

Educated by these observations, let us consider the simple l = 0 case as a warm-up. Our guess
for the wave-function is:
ul=0 = Aρe−λρ .
By substitution in√the SE (done in class) we see that this is a solution provided λ = 1, which,
together with λ = −2E gives us:
1
El=0 = − .
2
This is indeed the ground state (GS) energy for the hydrogen problem, in atomic units. The fact
that it is indeed the GS, comes from the absence of nodes in the function.

Let us now set up the general solution. We write:


uk,l (ρ) = wk,l (ρ)e−λk,l ρ ,
where the function wk,l (ρ) should start as ρl+1 , and λk,l = −2Ek,l . Calculating u0 and u00 in
p

terms of w0 and w00 we see that the Ek,l -term disappears, and the equation for wk,l comes out to
be:
1 ∂ 2 wk,l
 
∂wk,l l(l + 1) 1
− + λ k,l + − wk,l = 0 . (4.25)
2 ∂ρ2 ∂ρ 2ρ2 ρ
Now, assume for wk,l the form:

 X
wk,l (ρ) = ρl+1 c0 + c1 ρ + c2 ρ2 + · · · = cq ρq+l+1 ,

(4.26)
q=0

where the coefficients of the (possibly infinite) series cq have to be determined, and c0 6= 0. Calcu-
lating w0 and w00 and substituting in Eq. 4.25 we get:
∞ ∞ ∞ ∞
1X X l(l + 1) X X
− (q + l + 1)(q + l)cq ρq+l−1 + λk,l (q + l + 1)cq ρq+l + cq ρq+l−1 − cq ρq+l = 0 .
2 q=0 q=0
2 q=0 q=0

Notice that the l(l + 1)-term cancels against a similar term coming from the first series. Therefore,
we can rewrite:
∞ ∞ ∞
1X X 0 X 0
− q(q + 2l + 1)cq ρq+l−1 + λk,l (q 0 + l)cq0 −1 ρq +l−1 − cq0 −1 ρq +l−1 = 0 . (4.27)
2 q=0 0 0
q =1 q =1

The q = 0 from the first term is actually multiplied by zero, and we can also relabel q 0 → q.
Therefore, we can finally rewrite:
∞ ∞
1X X
− q(q + 2l + 1)cq ρq+l−1 + [λk,l (q + l) − 1] cq−1 ρq+l−1 = 0 . (4.28)
2 q=1 q=1

Term by term, this implies the following recursion relation for the coefficients cq :
λk,l (q + l) − 1
cq = 2 cq−1 , (4.29)
q(q + 2l + 1)
with c0 6= 0 determined only by the normalization condition.

By looking at this recursion formula, it would appear that we have solved our problem no matter
what the value of the energy Ek,l (hence of the constant λk,l ) is. Is it really so? This would imply
that the bound states do not have a discrete spectrum. Where is the problem with this argument?
The error is here: if the series goes on forever, then, for large q’s, the ratio:
cq λk,l (q + l) − 1 2λk,l
=2 −→≈ + ···
cq−1 q(q + 2l + 1) q
But an infinite series with this behaviour for the coefficients is really representing a function w
which diverges like w ∼ e2λk,l ρ , i.e., u ∼ e+λk,l ρ which is simply the other exponential solution,
The Coulomb problem bound states 31

which must be rejected because exploding at +∞. Indeed, by writing e2λk,l ρ = q (2λk,l )q ρq /q!
P
you can simply verify that the coefficients of this Taylor series do exactly that. Therefore, and here
is the key point for energy quantization, the only acceptable solutions have a power series that
terminates in a finite number, call it k ≥ 1, of terms, i.e., ck = 0. But this, in turn, implies:
1
λk,l (k + l) − 1 = 0 −→ λk,l = , (4.30)
k+l
or, for the energy, restoring the tilde:
1 1 1
Ẽk,l = − λ2k,l = − =− 2 , (4.31)
2 2(k + l)2 2n

where we have defined the principal quantum number n = k + l, while k is known as radial
quantum number, because it determines the number of zeroes of the radial function. The ground
state corresponds to n = 1, with energy −1/2 (in a.u.), and must be associated to k = 1 and l = 0:
it is known as 1s-state (“s” because of the l = 0 and “1” to remind us of the n = 1). The next level
has n = 2 and energy −1/8 a.u.: it can be obtained either with (k = 2, l = 0) (2s-state) or with
(k = 1, l = 1) (2p-states, there are 3 of them, because I still have m = −1, 0, +1 for an l = 1).

Exercise 4.4 Enumerate the quantum numbers (and degeneracy) associated to the n = 3 and n = 4
states of the hydrogen problem, writing the corresponding energy.

Summarizing, for the radial functions we have:

Rk,l (ρ) = ρl c0 + c1 ρ + · · · ck−1 ρk−1 e−λk,l ρ ,



(4.32)

where the coefficients of the polynomial in ρ are given by the recursive formula in Eq. 4.29, and
λk,l = 1/(k + l). It is now clear what is the meaning of the quantum number k introduced at the
beginning of our discussion. Notice, however, that the hydrogen problem has extra degeneracies,
with respect to a generic central potential, for which, for instance, the 2p and 2s states would not
be degenerate.

Exercise 4.5 By calculating explicitly, in each case, the coefficients ci , write down the radial wave-
functions associated to all the energy levels up to n = 3, and draw them.

Exercise 4.6 Solve the free-particle problem in three dimensions by using spherical coordinates. (Obtain
a recursion relation for the radial function, and check that this is related to Bessel functions.)

Exercise 4.7 Solve the spherical infinite potential well in three dimensions, i.e., the potential is infinite
outside of a sphere of radius R, and zero inside. (Use again spherical coordinates, and try to find the radial
functions that vanish at r = R.)

Exercise 4.8 Solve the three-dimensional harmonic oscillator with potential mω 2 (x2 + y 2 + z 2 )/2 in
spherical coordinates by adapting the procedure we have used for the hydrogen case. Compare the results
with the solution obtained by separation of variables in Cartesian coordinates.
5
Theory of angular momentum

We devote this lecture to studying in some detail the theory of angular momentum. We will
denote by (Jx , Jy , Jz ) the three components of our angular momentum, which we do not specify in
advance: it is not necessarily the single-particle orbital angular momentum introduced in discussing
the angular motion for the hydrogen atom. The same discussion applies to any three operators
obeying (with ~ = 1):
[Jα , Jβ ] = iαβγ Jγ .

5.1 Diagonalization of J 2 and Jz

Our first aim is to construct eigenvalues and eigenvectors of J 2 = Jx2 + Jy2 + Jz2 and Jz , which com-
mute and can be therefore diagonalized together. Let us denote by |j, mi the common eigenvectors,
and fj and m the eigenvalues of J 2 and Jz :

J 2 |j, mi = fj |j, mi
Jz |j, mi = m|j, mi . (5.1)

For the time being, j and m are generic labels for the eigenvectors. We will soon show that j must
be an integer of half-integer, that m runs from −j to j in integer steps, and that fj = j(j + 1).

5.1.1 Ladder operators

Given an eigenstate of Jz with eigenvalue m, we can easily construct eigenstates with m ± 1 and
the same fj . To do that, introduce the following ladder operators:

J± = Jx ± iJy . (5.2)

You can easily verify that [J 2 , J± ] = 0 and [Jz , J± ] = ±J± . Therefore:

Jz (J± |j, mi) = ([Jz , J± ] + J± Jz )|j, mi = (m ± 1)|j, mi ,

while:
J 2 (J± |j, mi) = J± J 2 |j, mi = fj (J± |j, mi) .
Therefore, apart from a possible normalization factor Cjm (see below):

J± |j, mi = Cjm,± |j, m ± 1i .

5.1.2 Bounds on m

It is simple to prove that:


1
J 2 = Jz2 + (J+ J− + J− J+ ) . (5.3)
2
Construction of spherical harmonics 33
As it turns out, all the operators in the previous equation are positive operators (i.e., their ex-
pectation value on any state is positive). Taking the expectation value over the state |j, mi we
have:
1
hj, m|J 2 − Jz2 |j, mi = fj − m2 = ||J− |j, mi||2 + ||J+ |j, mi||2 ≥ 0 .

2
Therefore m cannot be increased (by J+ ) or decreased (by J− ) forever, since m2 ≤ fj . There must
exist a maximum mmax and a minimum mmin above and below which we cannot continue to go,
i.e. J+ |j, mmax i = 0 and J− |j, mmin i = 0. But J− J+ = J 2 − Jz2 − Jz and J+ J− = J 2 − Jz2 + Jz , as
one can simply verify. Therefore:

0 = J− J+ |j, mmax i = (J 2 − Jz2 − Jz )|j, mmax i = (fj − m2max − mmax )|j, mmax i = 0 ,

or, equivalently:
fj = mmax (mmax + 1) .
Similarly:

0 = J+ J− |j, mmin i = (J 2 − Jz2 + Jz )|j, mmin i = (fj − m2min + mmin )|j, mmin i = 0 ,

or, equivalently:
fj = mmin (mmin − 1) .
The two expressions for fj imply that mmin = −mmax . Let us call j = mmax , so that mmin = −j.
Then fj = j(j + 1) and m has to run in the interval:

−j ≤ m ≤ j .

We can now easily calculate the normalization constant Cjm,± . It is simple to show that:
p
Cjm,± = j(j + 1) − m(m ± 1) .

5.1.3 Values of j

We must reach |j, mmin = −ji by applying J− an integer number of times n to the state |j, mmax =
ji:
|j, −ji ∝ (J− )n |j, ji ,
which implies that −j = j − n, or n = 2j, or:
n
j= . (5.4)
2
For n even we have that j is integer, while for n odd j must be half-integer. The total number of
states obtained by applying J− repeatedly to |j, ji is evidently n + 1 = 2j + 1, i.e., all values of m
from +j down to −j in integer steps.

5.2 Construction of spherical harmonics

We now explicitly construct the spherical harmonics. First of all, we adopt the usual notation L
for the orbital angular momentum, and use l instead of j, accordingly. Still with ~ = 1, we need
to solve:

L2 Yl,m (θ, φ) = l(l + 1)Yl,m (θ, φ)


Lz Yl,m (θ, φ) = mYl,m (θ, φ) , (5.5)

where the spherical harmonics are Yl,m (θ, φ) = hθ, φ|l, mi. First of all, we notice that Lz = −i∂/∂φ
so that the eigenfunctions of Lz must solve:
34 Theory of angular momentum


−i Yl,m (θ, φ) = mYl,m (θ, φ) .
∂φ

The solution of this equation is evidently of the form:

Yl,m (θ, φ) = Pl,m (θ)eimφ .

Moreover, since the function Yl,m must be single-valued when φ → φ + 2π we must have that m
is integer, and therefore l must also be an integer. Half-integer values of l are not allowed for the
orbital angular momentum: we will see that they are characteristic of the spin.

Consider now the state with mmax = l, i.e. Yl,l . We must have:

L+ Yl,l (θ, φ) = L+ Pl,l (θ)eilφ = 0 .

But the differential expression for L+ = Lx + iLy is (see Eq. 4.8):


 
∂ cos θ ∂
L+ = ieiφ −i + .
∂θ sin θ ∂φ

Applying it to Yl,l we get a differential equation for Pl,l (θ):

∂Pl,l cos θ
=l Pl,l .
∂θ sin θ
It is simple to verify that the solution of this differential equation is:

Pl,l (θ) = Cl sinl θ .

The normalization constant can be calculated to be:


r
(−1)l (2l + 1)(2l)!
Cl = l ,
2 l! 4π

where the sign (−1)l is a convention. All other states Yl,m are obtained by applying L− to Yl,l a
certain number of times, where:
 
∂ cos θ ∂
L− = ie−iφ i + .
∂θ sin θ ∂φ

5.2.1 Explicit l = 1 case: the p-states

The case of l = 0 (the so-called s-states) is trivial, since the only state is:

1
Y0,0 = √ .

Let us study in more detail the case l = 1, the so-called p-states of an atom. Here:
r
iφ 3
Y1,1 (θ, φ) = C1 sin θe = − sin θeiφ .


Applying L− , and recalling that L− Y1,1 = 2Y1,0 we immediately get:


r
3
Y1,0 = − 2C1 cos θ = cos θ .

Finally, applying L− again we have:
Construction of spherical harmonics 35
r
1 3
Y1,−1 = √ L− Y1,0 = −C1 sin θe−iφ = sin θe−iφ = −Y1,1

.
2 8π

Notice that:
r
1 3
√ (Y1,−1 − Y1,1 ) = sin θ cos φ = |px i
2 4π
r
i 3
√ (Y1,−1 + Y1,1 ) = sin θ sin φ = |py i
2 4π
r
3
Y1,0 = cos θ = |pz i . (5.6)

The meaning of px , py and pz , the Cartesian expression for the p-states, should be evident, if you
recall the expression for x, y and z in spherical coordinates. We discussed in class also the meaning
of the polar plots typically used to depict them in textbooks.

Exercise 5.1 Derive and appropriately draw the five l = 2 spherical harmonics (d-states), staring from
Y2,2 and repeatedly applying L− . Refer to a textbook to construct the Cartesian form of the d-states, and
plot them with the
q usual polar plots. q q
15 ±2iφ 15 ±iφ
Result: Y2,±2 = 32π
e sin2 (θ), Y2,±1 = − 8π
e cos (θ) sin (θ), Y2,0 = 5
16π
(3 cos2 (θ) − 1).
6
Symmetries

We give here the theoretical framework of symmetries in QM, which we have already seen at play
in a few of the problems encountered.

6.1 Space transformations

Let us start with well known transformations in ordinary space of three-dimensional vectors.

We start with translations, defined as:

Ta (r) = r0 = r + a .

Rotations are slightly more complicated. For instance, for a rotation by an angle θ around the
z-axis we have:
TRθ,z (r) = r0 = Rθ,z r ,
where Rθ,z is the well known rotation matrix:
 
cos θ − sin θ 0
Rθ,z =  sin θ cos θ 0  (6.1)
0 0 1

Rotations are represented by orthogonal matrices, Rt = R−1 , whose determinant is +1.

Inversion is also simple to represent with a matrix I = −1, whose determinant is −1 in 3D. (In
2D, inversion is really not distinct from a rotation by π.)

More generally, a roto-translation is defined by a rotation matrix R (proper, if detR = +1 or


improper, that is product of a proper rotation R times the inversion I, if detR = −1) and a
translation vector a:
TR,a (r) = r0 = Rr + a . (6.2)
All roto-translations form a group (prove this very simple fact) under the product rule:

TR2 ,a2 · TR1 ,a1 = TR2 R1 ,R2 a1 +a2 .

6.2 Transformations on scalar functions

We now ask how these transformations can be seen as operators acting on wave-functions.

6.2.1 Translations

Starting with translation (for instance in a one-dimensional case) it is quite natural to say that a
translation by a of a function ψ(x) gives us (draw it to convince yourself):
Transformations on scalar functions 37

PTa (ψ(x)) = ψ(x − a) = ψ(Ta−1 (x)) .


Therefore, upon acting on a function ψ(x), I obtain a new function ψ(x − a). PTa is quite evidently
a linear operator on functions: which form has it? We now show that it is a quite familiar object!
To show this, let us expand ψ(x − a) in powers of a around point x. We have:
∂ (−a)2 ∂ 2
ψ(x − a) = ψ(x) + (−a) ψ(x) + ψ(x) + · · · .
∂x 2! ∂x2
Now, it is simple to realize that this involves just the momentum operator, so that we can write:
PTa (ψ(x)) = ψ(x − a) = ψ(Ta−1 (x))
 
a 1 a 2
= 1 + (−i p̂) + (−i p̂) + · · · ψ(x)
~ 2! ~
= e−iap̂/~ ψ(x) . (6.3)
In other words, PTa = e−iap̂/~ . More generally, for translations in 3D we have:
PTa = e−ia·p̂/~ ,
because translations along different coordinates (and corresponding components of momentum)
commute. Notice the strong similarity with the time-evolution operator:

|ψ(t)i = e−iĤt/~ |ψ(0)i .


One expresses these equalities by saying that p̂/~ is the generator of translations in space, while
Ĥ/~ is the generator of translations in time. The meaning of “generator” is that the full operator
is obtained by taking the exponential of the generator, times a parameter (a or t) specifying the
amount of translation.

6.2.2 Rotations

Slightly more complicated is the discussion of rotations. Consider the usual rotation Rθ,z . Once
again, define the action on a wave-function by:
−1
PRθ,z (ψ(r)) = ψ(Rθ,z (r)) .
−1
It is simple to calculate Rθ,z (r):
    
cos θ sin θ 0 x x cos θ + y sin θ
−1
Rθ,z (r) =  − sin θ cos θ 0   y  =  −x sin θ + y cos θ 
0 0 1 z z
Therefore:
PRθ,z (ψ(r)) = ψ(x + δx, y + δy, z) ,
where δx = x(cos θ −1) +y sin θ and δy = y(cos θ −1)−x sin θ. Assume now that θ is infinitesimally
small, and expand (cos θ − 1) and sin θ to lowest order in θ. One finds that δx = θy + O(θ2 ) and
δy = −θx + O(θ2 ). The Taylor expansion of ψ to first order in δx and δy, therefore, gives:
∂ψ ∂ψ
PRθ,z (ψ(r)) = ψ(x + δx, y + δy, z) = ψ(r) + δx + δy + ···
∂x ∂y
 
∂ψ ∂ψ
= ψ(r) + θ y −x + ···
∂x ∂y
θ
= ψ(r) − i L̂z ψ(r) + O(θ2 ) . (6.4)
~
We will soon prove that this is, once again, just the first term of the expansion of the exponential:
θ
PRθ,z (ψ(r)) = e−i ~ L̂z ψ(r) . (6.5)
38 Symmetries

6.2.3 Generalities

Let us spell out some general properties of the operators PT . If T ∈ G, where G is a group of the
transformations T , then we can prove that PT , defined by the action:

PT (ψ(r)) = ψ(T −1 (r)) , (6.6)

form a group of operators acting on scalar functions, and, moreover, such operators are unitary.
The fact that the {PT |T ∈ G} form a group is simple, if we define the product of two PT by the
rule:
PT2 · PT1 = PT1 ·T2 .
(One can easily verify associativity, that the inverse is PT−1 = PT −1 , etc.) To prove that the PT are
unitary, just write:
Z
hPT ψ1 |PT ψ2 i = dr ψ1∗ (T −1 (r))ψ2 (T −1 (r)) = hψ1 |ψ2 i ,

where in the last equality we have changed variable to r0 = T −1 (r), with a Jacobian-determinant
equal to 1, since detT = ±1. Therefore:

hψ1 |PT† PT ψ2 i = hψ1 |ψ2 i ∀ψ1 , ψ2 =⇒ PT† PT = 1 .

Consider now a general rotation by an angle θ +  around the z-axis, with θ finite but  small.
Since Rθ+,z = Rθ,z · R,z = R,z · Rθ,z , we deduce that:
  
PRθ+,z = PR,z · PRθ,z = 1 − i L̂z + O(2 ) · PRθ,z .
~
This expression allows us to write a differential equation for the operator PRθ,z . Indeed:

dPRθ,z PRθ+,z − PRθ,z i


= lim = − L̂z PRθ,z , (6.7)
dθ →0  ~
and the solution of this differential equation is given, quite evidently, by:
θ
PRθ,z = e−i ~ L̂z . (6.8)

For a general rotation by an angle θ around an axis n̂ one can show that:
θ
PRθ,n̂ = e−i ~ n̂·L̂ . (6.9)

Notice, however, that n̂ · L̂ = nx L̂x + ny L̂y + nz L̂z is a sum of three non-commuting operators
(unlike the linear momentum case) and therefore
θ θ θ θ
e−i ~ n̂·L̂ 6= e−i ~ nx L̂x · e−i ~ ny L̂y · e−i ~ nz L̂z .

In order to express PRθ,n̂ as a product of three independent rotations, one needs to resort to Euler
angles.

6.3 Symmetry properties of Hamiltonians

We say that a Hamiltonian Ĥ is invariant under a certain space transformation T if:

Ĥr0 = Ĥr , (6.10)

where Ĥr and Ĥr0 denote the Hamiltonian expressed using the variables r and r0 = T r (including
transformation of momenta, etc.). We will illustrate this definition in a moment, but we now stress
Symmetry properties of Hamiltonians 39

that all the transformation T which leave Ĥ invariant form a group GH , called invariance group
of the Hamiltonian. The fact that these transformation form a group is left as an exercise for
the reader.

Let us give a few examples, to illustrate this definition. First of all, it is simple to prove that the
kinetic energy term p2 /2m is always invariant with respect to any translation and rotation. For
translation, the proof is trivial, since from x0 = x + a you immediately deduce that d/dx0 = d/dx,
hence p̂0 = p̂ (where the prime denotes the momenta in the variables r0 ). Consider now rotations.
For a rotation around z we have:
 0   
x cos θ − sin θ 0 x
 y 0  =  sin θ cos θ 0   y  , (6.11)
z0 0 0 1 z

from which we immediately deduce that:


 ∂    ∂ 
∂x0 cos θ − sin θ 0 ∂x
∂  =  sin θ cos θ 0   ∂  .

∂y 0 ∂y (6.12)
∂ 0 0 1 ∂
∂z 0 ∂z

More generally:

~ r0 = R · ∇
Exercise 6.1 Given any rotation matrix R, and the rotated variables x0 = R · x, show that ∇ ~ r.

In words, we say that the momentum operator transforms under rotation, exactly as the vector r
does:
r0 = R · r =⇒ p0 = R · p . (6.13)
Clearly, p̂2 is trivially invariant by rotation, as any scalar product is.

Consider now more realistic Hamiltonians. Our first example is a two-dimensional infinite po-
tential well:
p2x + p2y
Ĥ1 = + V∞ (x) + V∞ (y)
2m
where V∞ (x) is zero in x ∈ [−a, a] and ∞ elsewhere. Our second example is a central potential in
3D:
p2x + p2y + p2z e2
Ĥ2 = − .
2µ |r|
As a third example, consider the hydrogen atom (proton plus electron):

Px2 + Py2 + Pz2 p2x + p2y + p2z e2


Ĥ3 = + − .
2M 2m |R − r|

Obviously, Ĥ1 is not invariant under any translations, since, for instance for x0 = x + a, we
would have V∞ (x0 ) = V∞ (x + a) 6= V∞ (x). Most of the rotations (around z axis, obviously) also
fail the test, except for R±π/2,z , which exchange the roles of x and y, and for Rπ,z which sends
x → −x and y → −y. Therefore:

GH1 = {1, PRπ/2,z , PR−π/2,z , PRπ,z } .

Regarding Ĥ2 , we know that all translations satisfy the kinetic term, but unfortunately 1/|r0 | =
6 1/|r|. All rotations, on the contrary, including inversion (hence also improper rotations)
1/|r + a| =
40 Symmetries

satisfy both the kinetic term and the 1/|r| potential. Therefore (O(3) is the standard mathematical
symbol for the group of 3D rotations):

GH2 = {All proper and improper rotations R} = O(3) .

Finally, Ĥ3 is closely connected to Ĥ2 (recall the center-of-mass transformation) but now transla-
tions of both coordinates by the same amount a does not change the potential:
e2 e2 e2
= = .
|R0 0
−r| |(R + a) − (r + a)| |R − r|
Therefore:
GH3 = {All proper and improper roto-translations TR,a } .

We now prove the following important theorem:

Theorem 6.1 Let GH be the invariance group of the Hamiltonian Ĥ, and consider the operators
{PT |T ∈ GH }. One can prove that: i) this is a group of unitary operators isomorphic to GH , and
ii) all such operators commute with the Hamiltonian, PT Ĥ = ĤPT , or:

[PT , Ĥ] = 0 .

The proof is very simple. The fact that PT are unitary operators has already been shown, and the
fact that it is a group is very simple (exercise left to the reader). The only thing we need to prove
is ii), i.e, commutation with Ĥ. Define Hr ψ(r) = φ(r). Then:

PT Hr ψ(r) = PT φ(r) = φ(T −1 r) = HT −1 r ψ(T −1 r) .

On the other hand, since T ∈ GH we have that Hr = HT −1 r :

Hr PT ψ(r) = Hr ψ(T −1 r) = HT −1 r ψ(T −1 r) .

Therefore, since the equality of the two sides holds for any wave-function ψ(r), we conclude that
PT Ĥ = ĤPT .

6.4 Commuting observables

We now explain the importance of having a set of Hermitean operators that commute with the
Hamiltonian. An important theorem (proof not given) states that a Hermitean operator can be
always diagonalized, i.e., we can find a set of states {|φa i} which forms an orthonormal basis of
the Hilbert space and such that A|φa i = a|φa i. The collection of the eigenvalues {a}, which can
be proven to be real, forms the spectrum of the operator, and can be discrete, continuum, or both.
Next, we say that two observables A and B (i.e., Hermitean operators, associated to physical
quantities that can be measured) are compatible if [A, B] = 0, and incompatible if [A, B] 6= 0. We
will fully appreciate the meaning of this when we will talk about the process of measurement in
QM. For the time being, take it as a pure definition. An important theorem (proof sketched in
class), tells us that:

Theorem 6.2 If A and B are compatible observables (i.e., if [A, B] = 0), then they can diag-
onalized simultaneously, i.e. one can find a common set of states |φa,b i which are simultaneous
eigenvectors of A and B:

A|φa,b i = a|φa,b i and B|φa,b i = b|φa,b i .


Selection rules 41
In the second example analyzed previously (hydrogen case), we saw that rotations were all
θ
commuting with Ĥ2 . But we know that rotations are represented by PRθ,n̂ = e−i ~ n̂·L̂ . From the
fact that these operators commute with Ĥ2 , taking a derivative of the commutator with respect to
θ we have:
∂  θ θ

0= Ĥ2 e−i ~ n̂·L̂ − e−i ~ n̂·L̂ Ĥ2 .
∂θ
Taking the derivative of the exponential, and setting then θ = 0, we immediately deduce that:
[Ĥ2 , n̂ · L̂] = 0 ,
i.e., Ĥ2 commutes with any component of the angular momentum operator (since you can take n̂
in any direction). Out of the three components of the angular momentum, we know that we can
simultaneously diagonalize only L2 and, say, Lz . In the end, we have proved that the following three
operators, {Ĥ2 , L2 , Lz } are all commuting among each other. By a generalization of the previous
theorem, one can show that it is possible to find common eigenvectors to all of them, and this is
just what we have done in solving the hydrogen atom problem: rotational symmetry was essential,
therefore, in diagonalizing the problem analytically.

This example is paradigmatic: symmetries imply operators which commute with the Hamilto-
nian, which, in turn, can be used to diagonalize the problem in a simpler way.

6.5 Selection rules

There are several cases where one needs to know if certain matrix elements of operators vanish
or not. Symmetry properties of the relevant states and operator are important to answer such
questions in a simple way. As a paradigmatic example, consider hψ2 |p̂|ψ1 i and the parity operator
PI (I denotes the inversion r → −r). We know that PI ψ(r) = ψ(−r), and that eigenstates of PI
are functions which are either even (PI ψ(r) = ψ(−r) = ψ(r), eigenvalue +1), or odd (PI ψ(r) =
ψ(−r) = −ψ(r), eigenvalue -1). If the Hamiltonian is invariant under inversion, it commutes with
parity: [PI , Ĥ] = 0. This, in turn, implies that we can always classify the eigenstates of Ĥ as even
or odd under parity: assume that ψ1 and ψ2 are two such states, PI |ψi i = (−1)Pi |ψi i, where Pi = 0
if parity is even, Pi = 1 if parity is odd. It is also simple to show that
PI† p̂PI = −p̂ ,
or, in words, that the momentum is odd under parity (the same is true for r, while L is even under
parity: prove all these statements). Now, inserting 1 = PI PI† we have:
hψ2 |p̂|ψ1 i = hψ2 |PI PI† p̂PI PI† |ψ1 i = −(−1)P1 +P2 hψ2 |p̂|ψ1 i ,
which clearly shows that the matrix element vanishes if the two states have the same parity.

As an example, consider hydrogen atomic states. We know that states with l even are even under
parity, while states with l odd are odd under parity. Therefore, we immediately conclude that, for
instance:
hns|p̂|1si = 0 .
Similarly, we can immediately conclude that hnd|p̂|1si = 0, while, in general, hnp|p̂|1si is not
forced to vanish (indeed, it doesn’t).

Parity is by no means the only operator useful in this game. Indeed, consider:
hnf |p̂|1si .
According to parity, this matrix element might be non-vanishing, since the two states have opposite
parity (f states have l = 3, and are therefore odd). However, angular momentum would teach us
that this matrix element vanishes. The reason which this is so has to do with a very important
theorem, Wigner-Eckart theorem, which we will not state or prove, which states that an operator
like p̂ can only connect states whose angular momentum differs by at most 1.
7
Measurement in Quantum Mechanics

Observables and Hermitean operators. One of the basic postulates of QM is that to every
physical observable A we can associate an Hermitean operator Â: example seen, include the energy,
the momentum, the position, the angular momentum.

Completeness of eigenstates. The eigenstates |φan i of an Hermitean operator  with eigen-


value an (which are real, as you should try to prove) form a complete orthonormal basis of the
Hilbert space. Allowing for a possible degeneracy gn ≥ 1 of the eigenvalue an we write:

Â|φ(i) (i)
an i = an |φan i for i = 1, · · · , gn .

Projectors. To simplify the notation, assume for a while that the complete orthonormal basis
is denoted by {φn }. We can define a family of operators P̂n such that:

P̂m |φn i = δn,m |φm i ,

whose matrix elements are therefore very simple:

hφn0 |P̂m |φn i = δn,m δn,n0 .


P
Such operators are useful because, if a general state |ψi = n φn |φn i (with φn = hφn |φi) is acted
upon with P̂m , we end up with obtaining only the φm -component of it, i.e.:

P̂m |ψi = ψm |φm i .

So, P̂m projects out all components except for the mth one. A formal but very common expression
for P̂m , in terms of Dirac’s ket and bra is:

P̂m = |φm ihφm | ,

which simply means “act on any state |ψi to the right by taking the scalar product of |ψi with the
bra hφm |”, which we can immediately verify to give the correct result. Notice that:
2
P̂m = P̂m ,

and that we can represent the identity operator 1 as:


X X
P̂m = |φm ihφm | = 1 .
m m

Position and momentum eigenstates. Examples seen until now include the eigenstates of
position |xi (for which sums are changed into integrals, and Kronecker delta δn,m is changed into
a Dirac’s delta hx0 |xi = δ(x − x0 )), with the associated projectors P̂x = |xihx|, such that:
Z Z
|ψi = dx|xihx|ψi = dx|xiψ(x) ,
Measurements in QM. 43
where we recognize the usual wave-function ψ(x) = hx|ψi. An alternative choice is to use momen-
tum eigenstates |ki, whose wave-function reads:
eik·x
φk (x) = hx|ki = p ,
(2π)3
where we used the correct normalization on the continuum, hk0 |ki = δ(k − k0 ). Using completeness
we can now write: Z
|ψi = dk|kihk|ψi ,

where we see the momentum-space wave-function hk|ψi appearing. Explicitly, inserting a com-
pleteness on the x, we can show that:
e−ik·x
Z Z
1
hk|ψi = dxhk|xihx|ψi = dx p ψ(x) = p ψ̂(k) .
(2π)3 (2π)3

where we recognize the Fourier transform ψ̂(k) = dxe−ik·x ψ(x) of the ψ(x). Substituting, we get:
R
Z
dk
|ψi = p |kiψ̂(k) .
(2π)3
Notice also that, by taking the scalar product of the last expression with hx|, we immediately
obtain the inverse Fourier transform relation:
Z
dk ik·x
ψ(x) = hx|ψi = e ψ̂(k) .
(2π)3

Expanding a general state in eigenstates of Â. Let us turn again to the general case of an
arbitrary observable A. Constructing the projectors on the corresponding eigenstates, and using
completeness,
XX gn
1= |φa(i)n ihφa(i)n | ,
n i=1
we can write, for any state |ψi:
gn
XX
|ψi = |φa(i)n ihφa(i)n |ψi . (7.1)
n i=1

Notice that, since |ψi is assumed to be normalized, we have that:


gn
(i) 2
XX
hφan |ψi = 1 . (7.2)
n i=1

7.1 Measurements in QM.

Measuring A on a state |ψi. Here comes one of the crucial postulates of the theory of mea-
surement in QM. If we perform a measurement of the observable A on a system in state |ψi, then,
the result of each individual measurement can only be, according to QM, one of the eigenvalues
an of the Hermitean operator  associated to the observable which is measured. Performing several
repeated measurements on the system prepared in the identical state |ψi I can obtain, as a result,
different eigenvalues an , which gives to QM a probabilistic aspect. According to the postulates of
QM, the probability Pan that I obtain an as a result of the measurement is given by:
gn
(i) 2
X
Pan = hφan |ψi , (7.3)
i=1
P
which is nicely normalized to unity, according to Eq. 7.2, n Pan = 1.
44 Measurement in Quantum Mechanics

Collapse of |ψi after measuring A. Another postulate of QM is that, immediately after


having measured A, obtaining an eigenvalue an of the associated Â, the wave-function is projected
into its an -component (appropriately normalized, again):
gn
Measured an 0 X
|ψi −→ |ψ i = N |φa(i)n ihφa(i)n |ψi . (7.4)
i=1

This is known as collapse of the wave-function after measurement. It contains the probabilistic
aspect of QM, since the final state |ψ 0 i depends on the eigenvalue an measured, which in turn is
obtained with a probability Pan .

Measuring A again on |ψi. If we measure again A after having measured it, obtaining an
(with probability Pan ), we in effect are performing a measurement of A on the collapsed state |ψ 0 i,
and therefore we get again an , this time with certainty.

Measuring two commuting observables A and B on |ψi, one after the other. When
two observables  and B̂ commute, one can find a common basis of eigenvectors that diagonalizes
both simultaneously (we proved this). If I measure  on |ψi, getting an with probability Pan and
collapsing the state to |ψ 0 i, a subsequent measurement of B̂ on |ψ 0 i will give me only one of the
eigenvalues of B̂ in the gn -dimensional subspace associated to an , collapsing the state to some |ψ 00 i
which is still made of an components only. A subsequent measurement of A after B will therefore
confirm the result an with certainty, i.e., the measurement of B has not disrupted that of A: we
say that two commuting observables are compatible. On the contrary, if the two observables do not
commute, measuring B disrupts completely the an -subspace, and a subsequent measurement of A
gives no longer necessarily the same an .

7.2 Expectation values of measurements.

If we repeatedly measure A on the same state |ψi, we can construct the mean value of our mea-
surements, each always given by some eigenvalue an obtained however with a probability Pan . As
we know from elementary statistics, the mean is given by:
X
hAi = an Pan .
n

Inserting the expression in Eq. (7.3) for Pan , after simple algebra (done in class), we proved that
such a mean value coincides with the expectation value hψ|Âψi:
hAi = hψ|Â|ψi . (7.5)
We stress, however, that in each individual measurement we obtain an , and not hAi.

We can also construct the mean-square:


def
X
hA2 i = a2n Pan ,
n

and show that


hA2 i = hψ|Â2 |ψi . (7.6)

Exercise 7.1 Verify Eq. (7.6).

Given the mean and the mean-square, we can construct the root-mean-square-deviation:
def p
∆A = hA2 i − hAi2 . (7.7)
Expectation values of measurements. 45

Exercise 7.2 After having proved that, in general hA2 i ≥ hAi2 , so that ∆A is well defined, prove that,
given two operators  and B̂, then

1
∆A · ∆B ≥ |hψ|[Â, B̂]|ψi| ,
2

which is a generalization of Heisenberg’s uncertainty principle to arbitrary operators. [Hint: to prove


this, use Schwartz inequality of scalar products, |hα|βi|2 ≤ hα|αihβ|βi.]
8
Spin

The spin is the new operator we are going to present. Introducing the spin is strongly demanded
by a crucial experiment done in the early days of quantum mechanics by Stern and Gerlach (in
the following, SG).

8.1 Stern-Gerlach experiment: the spin operators and states

The experiment consists in sending a collimated beam of Ag atoms (coming out of a furnace)
along a direction (call it y) inside a specially designed magnet having a magnetic field along z,
but also having ∂Bz /∂z 6= 0, and then registering the spots observed on a screen perpendicular
to the beam. Much to their surprise, two spots — call them N1 and N2 — where observed in all
circumstances, with a 50% abundance of atoms ending up in each spot. The same result if found
if the apparatus (i.e., the magnet) is rotated in any direction.

To fully appreciate why this is surprising, let us discuss a bit the theoretical background. Each
atom, as a whole, obeys to a very good approximation Newton’s equations, due to its large mass.
Classically, a magnetic field induces a potential term V (R) = −~ µ · B(R), and hence a force whose
component j (j = x, y, z) is:
∂B
Fj (R) = µ~· ,
∂Rj
where µ~ is the magnetic moment of the atom and R indicates its (center-of-mass) position. The
presence of the gradient of the magnetic field explains why the design of the poles of the magnet was
specially conceived by Stern and Gerlach so as to produce a non-uniform field with ∂Bz /∂z 6= 0: this
was achieved through a sharp corner protruding from the Nord-pole (sketched at the blackboard).
But, what is the magnetic moment of the atom? Think, for simplicity, of a beam of hydrogen
atoms (Ag atoms are not very different in that respect, since they have an electronic configuration
[Kr]4d10 5s1 with a single unpaired electron in the 5s shell). A contribution to the magnetic moment
certainly comes from the electron which orbits around the nucleus. As in classical electrodynamics
one can show that:
e e~ L L
~ =−
µ L=− = −µB .
2mc 2mc ~ ~
e~
Notice that, when L = ~, this gives a magnetic moment − 2mc = −µB , equal to the Bohr magneton.
If this was the only source of magnetic moment, then the experiment cannot be explained. For one
thing, L = 0 for an electron in the s-shell. And even if this was not the case, L would have 2l + 1
possible values, due to m = −l, · · · , l, while only two spots are seen in the experiment!

All we have said above can be quite firmly justified quantum mechanically, by resorting to a
Born-Oppenheimer approach to separate the motion of the nucleus from that of the electrons. To
give an idea of the origin of the L · B term, consider the Hamiltonian of the single electron of a
hydrogen atom in a magnetic field. So far, we would simply write:

1  e 2 e2
H= p+ A − ,
2m c r
Stern-Gerlach experiment: the spin operators and states 47
where A is the vector potential. For a uniform field Bz directed along z, we can take A =
(Bz /2)(−y, x, 0), which has zero divergence, ∇ · A = 0. Expanding the square in the kinetic
energy we can rewrite H (done in class) as:

p2 e2 e2
H= − + µB Lz Bz + B 2 (x2 + y 2 ) ,
2m r 8mc2 z
from which we see that there is an extra potential term µB Lz Bz as predicted from classical physics
(as well as a quadratic Bz term, which one can easily estimate to be of the order of (µB Bz )2 /Ry,
a very small quantity, since µB Bz ≈ 5.8 × 10−5 eV for a field Bz of 1 Tesla).

So, there is no way of having two spots only in a SG experiment without invoking some other
source of intrinsic magnetic moment, not due to the orbital motion of the electrons. Indeed, we
saw from the general theory of angular momentum, that in principle half-integer values of j are
possible in principle. So, let us postulate that there exist an intrinsic angular momentum S, called
spin, which has S = 1/2 and hence two values for Sz , Sz = ±1/2. The spin enters the Hamiltonian
in the presence of magnetic field in the following way:

1  e 2 e2
H= p+ A − + g0 µB S · B ,
2m c r
i.e., in a very similar way to what L does, but with g0 ≈ 2. A full derivation of such expression
requires taking the non-relativistic limit of the Dirac’s equation, and we will not do it. What is
important for us is to explore the experimental implications of such form. Before continuing, we
need to discuss what happens to the states of the system. The spin has its own Hilbert space,
generated, for instance, by the two eigenstates of S z , which we denote by | ± 12 iz or simply |±i if
there is no ambiguity. The most general spin state can therefore be written as

|ψispin = α+ |+i + α− |−i ,

with α± complex constants. Normalization of the state requires |α+ |2 +|α− |2 = 1. The total Hilbert
space of the system must be therefore supplemented by such a spin part, in a way that is known as
tensor product of Hilbert spaces: H = Horb ⊗ Hspin . In practice that simply means the, if |ψn iorb
denotes a complete basis for the orbital Hilbert space (the one we have used so far), then the most
general state of the system can be written as:
X
|Ψi = |ψn iorb (cn,+ |+i + cn,− |−i) .
n

Notice that this does not imply that |Ψi = |ψiorb |ψispin , which is only a special (although, in
absence of spin-orbit coupling, see next chapter, quite frequent) case occurring when the coefficients
αn,± have the factorized form cn,± = cn α± .

At this point, we can proceed in two ways. We can go on guided by theory, and directly write
down the matrix elements of Sx , Sy , Sz in the basis of eigenstates of Sz , using the expressions for
S+ and S− . Or, alternatively, resort to the experiment.

The first route uses that S+ |−i = |+i and S− |+i = |−i (recall the expressions for J± and
specialize them to j = 1/2), and arrives at writing:
 
1 01 1
hσ|Sx |σ 0 i = = σx , (8.1)
2 10 2
 
0 1 0 −i 1
hσ|Sy |σ i = = σy , (8.2)
2 i 0 2
 
0 1 1 0 1
hσ|Sz |σ i = = σz . (8.3)
2 0 −1 2
48 Spin

Exercise 8.1 Deduce the matrix elements of the spin operators given above by using the fact that Sz |±i =
± 12 |±i and S± |∓i = |±i.

Here the notation used is that row and column indices are labeled by σ and σ 0 which take the values
± with
  the ordering convention that a state |ψispin = α+ |+i + α− |−i is denoted as a column vector
α+
. The three very important matrices σx,y,z previously defined are known as Pauli matrices.
α−

The second route requires the following series of conceptual experiments. Prepare a beam of
atoms in the eigenstate | + 21 ix by rotating the SG-apparatus with the magnet axis in the x-
direction and making a hole in the point N1 of the screen, so that only atoms directed towards
N1 pass the screen. Next, let the beam of “surviving atoms” go through a second SG-apparatus,
this time with the magnet axis along z. According to the prescription of QM, this will measure the
probability of having Sz = ± 21 in the state | + 12 ix , i.e., |h±| + 12 ix |2 . The (surprising) experimental
result would be that there is a 50% abundance in the two spots, implying that |h+| + 21 ix |2 = 12
and |h−| + 12 ix |2 = 21 . These expressions immediately imply that:

| + 12 ix = √12 |+i + eiδx |−i



 , (8.4)
| − 12 ix = √12 |+i − eiδx |−i

where δx is an arbitrary phase factor that we cannot, so far, determine (an overall phase factor in
front is, on the contrary, irrelevant, and has been not included). By a similar conceptual experiment
we can write the following expressions for the states | ± 21 iy :

| + 12 iy = √12 |+i + eiδy |−i



 , (8.5)
| − 12 iy = √12 |+i − eiδy |−i

where another unknown phase factor δy has been included. Now, let us “close the circle”, and
perform a measurement of Sy after having prepared |±1/2ix . The results of the experiments would
tell us that |y h± 12 | + 12 ix |2 = |y h± 21 | − 12 ix |2 = 12 , which, together with the explicit expressions for
the state in Eqs. (8.4,8.5), lead after simple algebra to the conclusion that:
π
cos (δy − δx ) = 0 =⇒ δy − δx = ± .
2

Exercise 8.2 Verify that the experimental finding that |y h± 12 | + 12 ix |2 = |y h± 12 | − 12 ix |2 = 1


2
, together
with the explicit expressions for the states in Eqs. (8.4,8.5) imply that cos (δy − δx ) = 0.

A possible phase convention is that δx = 0 (i.e., | ± 12 ix states have real coefficients) but that
necessarily implies that δy = ±π/2, i.e., complex numbers are mandatory in the | ± 21 iy states! The
correct choice for a right-handed system turns out to be δy = π/2.

One can easily verify (done in class) that the form of the matrix elements of Sx,y,z coming out
of this form of the states is exactly the one given above in terms of the Pauli matrices. One can
also easily verify that [Sx , Sy ] = iSz , as expected. Similarly S 2 = Sx2 + Sy2 + Sz2 = (3/4)1.

8.2 Experiments with Stern-Gerlach apparatus

One can easily prepare, experimentally, a spin-state pointing along any direction in space. If n̂
denotes a direction, associated to two angles (θ, φ), all one needs to do is to rotate the SG apparatus
Experiments with Stern-Gerlach apparatus 49
so that the magnetic field axis is along n̂ and then collect only the atoms passing through N1 (while
the other spot is blocked by the screen). The spin along n̂ is associated to the operator
cos (θ) sin (θ)e−iφ
 
1
n̂ · S = sin (θ) cos (φ)Sx + sin (θ) sin (φ)Sy + cos (θ)Sz = ,
2 sin (θ)e+iφ − cos (θ)
where the last equality applies, strictly speaking, to the matrix elements of such an operator on
the standard basis and is deduced by using the form of the Pauli matrices. By diagonalizing the
2 × 2 matrix for n̂ · S = sin (θ) it is possible to show that the eigenvalues of n̂ · S are ± 21 with
associated eigenstates given by:
| + 12 in̂ = cos θ2 |+i + sin θ2 eiφ |−i
 

| − 1 in̂ = sin θ |+i − cos θ eiφ |−i .


2 2 2
(8.6)

Exercise 8.3 Verify the expression (8.6) for the eigenstates of n̂ · S and convince yourself that the usual
form of the eigenstates of Sx and Sy are obtained as particular cases.

Think now of sending the beam of atoms which have been prepared in the state | + 21 in̂ through a
second SG-apparatus along z. The probability of measuring ± 12 for Sz are evidently given by:
   
2 θ 2 θ
P+ 12 (z) = cos , P− 12 (z) = sin ,
2 2
which are in principle subject to experimental test. Notice that this is the first time we find a result
which is not 50% abundance in each spot!

Exercise 8.4 Calculate the probability P± 1 (x) of measuring ± 12 for Sx on the state + 21 in̂ .
2

Consider now preparing | + 12 in̂ as before, and measuring each component of S (in turn, x, y, and
z) with a second SG apparatus appropriately rotated, repeatedly, over and over again, accumu-
lating statistics for the calculation of probabilities, as usual. Each individual measurement of any
component of S will always give either + 21 or − 12 . But the probabilities of each outcome is different,
and you can therefore ask what is the mean value of the spin in the ordinary statistical manner. As
discussed previously (see lecture on Measurement) this requires calculating the expectation value
of S on | + 12 in̂ . A simple calculation shows that:
1 1 1
n̂ h+ |Sx | + in̂ = sin θ cos φ , (8.7)
2 2 2
1 1 1
n̂ h+ |Sy | + in̂ = sin θ sin φ , (8.8)
2 2 2
1 1 1
n̂ h+ |Sz | + in̂ = cos θ , (8.9)
2 2 2
or, more compactly:
1 1 1
n̂ h+ |S| + in̂ = n̂ . (8.10)
2 2 2

Exercise 8.5 Verify the previous expressions for the mean values of S.

Summarizing, while individual spin measurements are quite surprising for our classical intuition,
the mean values behave in a classically well understandable way.
50 Spin

8.3 A single spin-1/2 in a uniform field: Larmor precession

Consider now the usual hydrogen atom in a magnetic field, and write the Hamiltonian as
H = H0 + g0 µB BSz ,
where H0 collects all the terms which do not depend on the spin (as we will see later on, we are
not including the spin-orbit coupling in H). Each eigenvalue En of H0 is now associated to two
possible eigenstates, |ψn i|+i, and |ψn i|−i, which are perfectly degenerate in absence of B. In the
presence of B, the two states are split in the following way:
En,± (B) = En ± µB B , (8.11)
where we have taken g0 ≈ 2. The splitting between the two levels is ∆E = 2µB B. Think now of
preparing the system, in absence of magnetic field, in the state, at t = 0:
     
1 θ0 θ0 iφ0
|ψ(t = 0)i = |ψn i| + in̂ = |ψn i cos |+i + sin e |−i .
2 2 2
At t = 0 we now turn on the magnetic field and let the state evolve. The state at time t is
|ψ(t)i = e−iHt/~ |ψ(t = 0)i which one can easily calculate to be:
     
θ0 θ0 iφt
|ψ(t)i = e−iHt/~ |ψ(t = 0)i = e−iEn t/~ |ψn ie−iµB Bt/~ cos |+i + sin e |−i ,
2 2
with
2µB Bt ∆E
φt = φ0 + = φ0 + t = φ0 + ωc t .
~ ~
This shows that, in the presence of B along z, the spin precesses around the z-axis with a frequency
∆E eB
ωc =
= . (8.12)
~ mc
Such a frequency, known as Larmor frequency (or cyclotron frequency), is entirely classical (no
~ appears) and corresponds to the frequency with which a classical charged particle rotates in a
uniform magnetic field.

If we measure S at time t we get:


1
hψ(t)|S|ψ(t)i = n̂(t) , (8.13)
2
where n̂(t) denotes the direction (θ0 , φt ).

8.4 Heisenberg representation

Given an arbitrary operator Â, the expectation value on the state |ψ(t)i is
hψ(t)|Â|ψ(t)i = hψ(t = 0)|eiHt/~ Âe−iHt/~ |ψ(t = 0)i . (8.14)
The expression on the right-hand side defines the Heisenberg representation of the operator Â:
ÂH (t) = eiHt/~ Âe−iHt/~ , (8.15)
in terms of which we have:
hψ(t)|Â|ψ(t)i = hψ(t = 0)|ÂH (t)|ψ(t = 0)i . (8.16)
One can write a differential equation satisfied by ÂH (t) by simply taking the derivative with respect
to time. It is simple to show that:
d
i~ ÂH (t) = eiHt/~ [Â, H]e−iHt/~ = [ÂH (t), H] , (8.17)
dt
an equation known as Heisenberg equation of motion for the operator ÂH (t).
Heisenberg representation 51
As a simple application of this approach, consider the previous problem of Larmor precession
and calculate the Heisenberg equations of motion for S. It is immediate to show that:
d
i~ Sz,H (t) = eiHt/~ [Sz , H]e−iHt/~ = 0 , =⇒ Sz,H (t) = Sz ,
dt
5 because Sz commutes with H. This illustrates the concept of constant of motion, which applies
to each operator that commutes with the Hamiltonian. A slightly more complicated calculation,
which uses the commutation rules [Sx , Sy ] = iSz , shows that:

d
Sx,H (t) = −ωc Sy,H (t) (8.18)
dt
d
Sy,H (t) = +ωc Sx,H (t) . (8.19)
dt
Similar equations are obeyed by the expectation values of the spin, and clearly describe the uniform
rotation around the z-axis of hSH (t)i.
9
Spin-orbit and addition of angular
momenta

Until now, an hydrogen atom in a magnetic field is described by:

1  e 2 e2
H= p+ A − + g0 µB B · S ,
2m c r
where µB = e~/2mc, and g0 ≈ 2. The spin appears only due to the Zeeman term (last term in H),
and is totally decoupled from the orbital variables (r, p, L). As a consequence of such decoupling,
wave-functions can be written, at this level, as products of the orbital wave-function times the
spin wave-function, |ψorb i ⊗ |ψspin i, where, e.g., ψorb (r) = Rnl (r)Ylm (θ, φ) (in absence of B), and
|ψspin i = α+ |+i + α− |−i.

9.1 Relativistic effects

However, there are important relativistic effects that should be added to the previous H. A first
relativistic
p contribution comes from kinetic energy. By expanding the relativistic dispersion E(p) =
p c + m2 c4 we have:
2 2

 2 2
p p2 1 p
E(p) = p2 c2 + m2 c4 = mc2 + − + ··· .
2m 2mc2 2m

Notice that the last term, considering that p2 /2m ≈ Ry, is smaller with respect to the ordinary
non-relativistic kinetic energy by a factor of order Ry/mc2 . Such a dimensionless factor is very
important:
 2 2
2Ry Hartree e
= = = α2 ,
mc2 mc2 ~c
where α = e2 /(~c) ≈ 1/137 is known as fine-structure constant. There are other corrections to the
same order in α. These can be derived by taking the non-relativistic limit of the (fully relativistic)
Dirac’s equation (including the B · S term).

9.2 The origin of spin-orbit

We concentrate here only on a term that couples S and L. The rigorous derivation is complicated,
but the order-of-magnitude and form of the term can be understood easily. Suppose there is no
B-field. The electron travels in the E-field due to the nucleus, E = (e/r3 )r. However, in the frame
at rest on the electron, the nucleus moves with velocity −v, and this generates a current  = −ev
which in turn generates a B-field:
v
Beff = − × E .
c
This B-field seen by the electron in its rest-frame affects its spin with a term
Construction of J-states out of p-states 53
µB µB
µB Beff · S = (E × v) · S = (E × p) · S .
c mc
Notice that we have not included the factor g0 here! (This is just an argument, not a rigorous
derivation.) In conclusion, we can anticipate a term
µB µB e
Hso = (E × p) · S = (r × p) · S = λso (r) L · S ,
mc mc r3
where we have extracted ~ from the definition of L and defined:
e~ e~ 1 µ2
λso (r) = 3
=2 B .
2mc mc r r3
Let us estimate the order of magnitude of this term by using r ∼ aB = ~2 /(me2 ). We get:
µ2B me4 e4
λso ∼ 2 = = (Ry) α2 .
a3B 2~2 c2 ~2
Once again, we see that the order of magnitude is the same as the correction to the kinetic energy:
order α2 . Numerically, one might expect this small effect to be not very important. However, effects
of this order are spectroscopically measurable and determine, for instance, the fine structure of the
p-levels of hydrogen. Moreover, it is simple to realizes that the spin-orbit coupling becomes more
and more effective as the nuclear charge Ze is increased (i.e., for heavier atoms). Indeed, the
electric field due to a nucleus of charge Ze is E = (Ze/r2 ), and the factor 1/r3 appearing in λso (r)
is now important at r ∼ aB /Z (the effective Bohr radius is reduced by Z). Therefore, we anticipate
(Z)
λso ≈ Z 4 (Ry)α2 .

Summarizing, even in absence of an external magnetic field we should write a Hamiltonian for
the hydrogen atom where the spin operator explicitly appears through the spin-orbit term Hso :
p2 e2
H= − + λso (r)L · S + Hother relativistic .
2m r

9.3 Construction of J-states out of p-states

The full treatment of a Hamiltonian including relativistic effects like the previous one requires
using perturbation theory (which will be covered later on in the course). We only concentrate here
on the symmetry aspects of the problem and learn how to construct the appropriate combinations
of orbital and spin states which make the effect of the L · S term simple. Consider, to fix the ideas,
the 2p-states (l = 1) of the unperturbed hydrogen problem. We classify them (including spin) in
the following way:
1 1 1 1
|n = 2, l = 1, m; s = , sz = ± i = |n = 2, l = 1, mi ⊗ |s = , sz = ± i .
2 2 2 2
From now on we will omit the ⊗ sign, and simply indicate the spin states with |±i or with |sz i.
Similarly, we indicate the orbital part of the state with |1, m = 0, ±1i. There are therefore a total
of 3 × 2 = 6 states, which we list below:


 |1, +1i|+i
|1, 0i|+i , |1, +1i|−i

(9.1)
 |1, 0i|−i , |1, −1i|+i

|1, −1i|−i

Observe that
1
L · S = Lz Sz +
(L+ S− + L− S+ ) ,
2
which implies that, when applied to the states we have listed, L · S mixes them (through the
off-diagonal term L+ S− + L− S+ ).
54 Spin-orbit and addition of angular momenta

The idea, now, is that L · S can also be viewed as the cross term appearing in (L + S)2 . Indeed,
let us introduce the operator J = L + S, which represents the total angular momentum, sum of
the orbital plus spin parts. Notice that the two terms of J act on different Hilbert spaces and,
more formally, we should write J as L ⊗ 1spin + 1orb ⊗ S, but we will not insist on this notation.
First of all J is a legitimate “angular momentum”, as you can easily verify (done in class) that
[Jx , Jy ] = iJz (and so on for the cyclic combinations). Next, consider

J2 = (L + S)2 = L2 + S2 + 2L · S .

As a consequence:
1 2
J − L2 − S2 .

L·S=
2
Therefore, if we were able to construct common eigenstates of J2 , L2 and S2 , the effect of L · S
on such states would be very simple to calculate. The product states |1, mi|sz i are obviously
eigenstates of L2 (with eigenvalue l(l + 1) = 2), of S2 (with eigenvalue s(s + 1) = 43 ), and of
Jz = Lz + Sz (with eigenvalue jz = m + sz ).

Consider now the top-most state with highest eigenvalue of jz = + 23 , associated to the product
state |1, +1i|sz = + 12 i. Question: Is this an eigenstate of J2 , and, if so, with what eigenvalue?
It is not difficult to show that the answer is positive: |1, +1i|sz = + 21 i is an eigenstate of J2 with
a value of j = 32 . This can be seen in at least two ways. First, by direct calculation, by writing
J2 = L2 + S2 + 2Lz Sz + (L+ S− + L− S+ ). Second, by the following argument: j cannot be less
than 3/2 because otherwise jz = 3/2 would not be possible; j cannot be larger than 3/2 because
otherwise there should be some value of jz = j > 3/2, which is evidently impossible (the highest
jz , by direct inspection, is 3/2). In conclusion we have the first eigenstate of J2 , Jz , L2 , and S2 in
the simple form:
3 3 1 1
|j = , jz = + ; l = 1, s = i = |1, +1i|sz = + i . (9.2)
2 2 2 2
To calculate the other states belonging to the j = 32 multiplet (2j + 1 = 4 state in total), we just
need to apply J− repeatedly. A simple calculation gives:
3 1 1 1 3 3 1
|j = , jz = + ; l = 1, s = i = √ (L− + S− )|j = , jz = + ; l = 1, s = i
2 2 2 3 2 2 2
r r
2 1 1 1
= |1, 0i| + i + |1, +1i| − i . (9.3)
3 2 3 2
Similarly, we get:
r r
3 1 1 1 1 2 1
|j = , jz = − ; l = 1, s = i = |1, −1i| + i + |1, 0i| − i , (9.4)
2 2 2 3 2 3 2
3 3 1 1
|j = , jz = − ; l = 1, s = i = |1, −1i| − i . (9.5)
2 2 2 2
Notice that the two states |j = 32 , jz = ± 12 ; l = 1, s = 12 i are not simple product states, but sum of
two different product states!

In this way we have constructed 4 states, out of the 6 product states available. Where are the
missing 2 states? Obviously, by combining |1, 0i| + 12 i and |1, +1i| − 12 i we can form two states,
while Eq. (9.3) represents only a single combination. There is a further combination which we can
form, still with the same jz = 21 , which we write as:

1 1 1 1
|j =?, jz = + ; l = 1, s = i = α|1, 0i| + i + β|1, +1i| − i ,
2 2 2 2
with |α|2 + |β|2 = 1. This new combination must be chosen to be orthogonal to the previous one,
which requires
Generalization: addition of L and S 55
r r
2 1
α+ β=0.
3 3
Normalizing, we get:
r r
1 1 1 1 2 1
|j =?, jz = + ; l = 1, s = i = |1, 0i| + i − |1, +1i| − i .
2 2 3 2 3 2
The only remaining point is the value of j, which we have indicated by j =?. It is not hard to show
that j = 12 in this case.

Exercise 9.1 Apply J+ to the state |j = 21 , jz = + 21 ; l = 1, s = 21 i showing that the result vanishes.

The final state missing is the second state of the multiplet j = 12 , with jz = − 21 , which is
r r
1 1 1 2 1 1 1
|j = , jz = − ; l = 1, s = i = |1, −1i| + i − |1, 0i| + i .
2 2 2 3 2 3 2
The coefficients appearing in the previous expressions are known as Clebsch-Gordan coefficients.

9.4 Generalization: addition of L and S

Exercise 9.2 Construct J-states out of L = 2 and S = 1/2, as you would need for the fine structure of
d-states.

Exercise 9.3 For more then one electron, you find cases where you need to add the two spins S1 and S2
of each electron. Determine the eigenstates (triplet and singlet, so called) and eigenvalues of S 2 and Sz
where S = S1 + S2 is the total spin of the two electrons in terms of the basis elements | ± 12 i1 | ± 21 i2 .
10
Density matrices

Let us ask ourselves the following question: what is the spin state |ψspin i of the silver atoms coming
out of the furnace before the Stern-Gerlach (SG) measuring apparatus? Unfortunately, none of the
states
|ψspin i = α+ |+iz + α− |−iz
describes the experiment. Indeed, I can always write one such state as a spin pointing into an
arbitrary direction n̂:
θ θ
|+in̂ = cos |+iz + sin eiφ |−iz .
2 2
This, in turn, would imply that by rotating the SG apparatus in the direction n̂ I should be able
to measure 100% of atoms in one spot, which is, experimentally, not the case: no matter how the
SG apparatus is rotated, you always get 50% of atoms in each of the two spots.

How to describe such a state entering the SG apparatus?

10.1 The density matrix for a pure state

Let us go back to the expectation value of an operator  and reinterpret it, in terms of the projector
Pψ = |ψihψ|, after having inserted to identities (in terms of a general basis)::
def
X
hÂi = hψ|Â|ψi = hψ|φα ihφα |Â|φα0 ihφα0 |ψi
α,α0
X X
= hφα |Â|φα0 ihφα0 |ψihψ|φα i = hφα |Â|φα0 ihφα0 |Pψ |φα i
α,α0 α,α0

= Tr[Â][Pψ ] . (10.1)

In words, the expectation value of any operator  on |ψi can be expressed as the trace of the
product of the matrix representing  with that representing the projector Pψ . It is therefore useful
to say that the state is univocally assigned once we know the operator ρ̂ = Pψ , whose matrix
elements on a given basis |φα i = |αi are:

hα|ρ̂|α0 i = hα0 |ψihψ|αi = ψα0 ψα∗ .

Examples: In real space |αi = |xi, and:

ρ(x0 , x) = ψ(x0 )ψ ∗ (x) .


2
/(2l2 )
For a one-dimensional harmonic oscillator, for instance, ψ(x) = Ce−x , and therefore:
x2 +x02
ρ(x0 , x) = |C|2 e− 2l2 .

For a plane-wave state: ψ(x) = √1 eik·x so that:


V
The density matrix for a mixed state 57
1 ik·(x0 −x)
ρ(x0 , x) = e .
V

For a spin-1/2 state |ψspin i = α+ |+iz + α− |−iz , we have:


∗ ∗
ρ̂ = |ψspin ihψspin | = |α+ |2 |+ih+| + |α− |2 |−ih−| + α+ α− |+ih−| + α− α+ |−ih+| ,

so that the matrix elements of ρ̂ are:



|α+ |2 α+ α− cos2 θ2 cos θ2 sin θ2 e−iφ
   
[ρ̂] = ∗ = , (10.2)
α− α+ |α− |2 θ θ iφ
cos 2 sin 2 e sin2 θ2

where the last equality assumes that the spin is directed along n̂, parameterized by θ and φ.

Notice the following two properties of ρ̂ = Pψ : i) Trρ̂ = 1, and ii) ρ̂2 = Pψ2 = Pψ = ρ̂. The
first property is a consequence of the normalization of ψ, while ii) follows from the properties of
projectors.

10.2 The density matrix for a mixed state

Suppose that the system we are trying to describe P is found in any of the pure states |ψi i, with
i = 1, · · · , Nρ , with a certain probability wi (clearly, i wi = 1). What is the natural probabilistic
way of calculating the mean value of any operator Â? Simply sum all averages with pure states
with their weight wi , obtaining:
Nρ Nρ Nρ
X X X X
hÂi = hψi |Â|ψi i = wi hφα |Â|φα0 ihφα0 |Pψi |φα i = Tr[Â][ wi Pψi ] .
i=1 i=1 α,α0 i=1

It is therefore clear that, the density matrix describing such a state (called mixed state, as opposed
to a pure state described by a single ψ) is:


X
ρ̂ = wi |ψi ihψi | , (10.3)
i=1

in terms of which averages are expressed, once again, as:

hÂi = Tr[Â][ρ̂] . (10.4)

Simple properties of ρ̂ are that, once again, Trρ̂ = 1. However, it is simple to show (done in class)
that: X
ρ̂2 = wi2 |ψi ihψi | =
6 ρ̂ ,
i

unless the state is pure, i.e., only one of the wi is equal to 1, all the others being zero. In particular:
X
Tr ρ̂2 = wi2 ≤ 1 .
i

10.3 Density matrices and statistical mechanics

Suppose you have a quantum system which is enclosed into a thermostat which holds it at temper-
ature T . This is the so-called canonical ensemble in statistical mechanics (no exchange of particles
58 Density matrices

is possible, but heat can flow in and out of the system to maintain the constant temperature). The
quantum system is not described by a pure state, but rather by a density matrix which is
X
ρ̂T = pn |ψn ihψn | , (10.5)
n

where pn = e−βEn /Z, β = (kB T )−1 , Z = n e−βEn is the partition function, and {|ψn i} denote
P

the energy eigenstates of the system Hamiltonian Ĥ (which form a basis of the Hilbert space of
the system). It is very simple to show that you can also rewrite:
1 −β Ĥ
ρ̂T = e . (10.6)
Z

As a simple application, consider a one-dimensional quantum oscillator of Hamiltonian


~ω 2
Ĥ = (p + x2 ) ,
2
(p and x being, as usual, dimensionless momentum and coordinate) in thermal equilibrium with
a reservoir at temperature T .Expressing Ĥ in terms of the annihilation and creation operators a
and a† , as Ĥ = (~ω) a† a + 12 , one can easily calculate many averages, as you will learn by doing
the following exercise.

Exercise 10.1 Thermal averages for an harmonic oscillator. (1) Calculate the thermal average ha† ai =
Tr[ρ̂T a† a]. [Hint: Calculate first the partition function Z.] (2) Next, calculate the average potential energy
~ωx2 /2 and the average kinetic energy ~ωp2 /2 and plot these quantities as a function of kB T /(~ω). (3)
Finally, calculate the specific heat CV = ∂U/∂T , U being the total internal energy, and plot it as a function
of kB T /(~ω). (4) At what temperatures strong deviations from the expected classical result (state what
it is that you expect, classically) are seen, due to quantum effects?

10.4 Density matrices by tracing out the universe

Suppose that you have a system A (Hilbert space HA with a basis set {|φα iA }) surrounded by the
rest of the universe B (Hilbert space HB with basis set {|Φβ iB }). The total wave-function |Ψi of
the combined system will leave in the tensor space HA ⊗ HB , which by definition has a basis set
{|φα iA |Φβ iB } (we omit the ⊗ sign between the states of A and B, as notation is unambiguous).
The total wave-function, therefore, is generally a combination of those basis states
X
|Ψi = cα,β |φα iA |Φβ iB , (10.7)
α,β

with appropriate coefficients cα,β . Notice that, in general, the state |Ψi cannot be written as a
separable state, i.e. a single product of a state of A times a state of B (as the basis states are):

|Ψi =
6 |φiA |ΦiB .

This occurs only for some special choices of the coefficients cα,β , for instance cα,β = aα bβ . Whenever
the state |Ψi cannot be written as a separable state one says that the state is entangled. A simple
example of such a case is, for two spin-1/2 particles, a singlet state.

Suppose I want to calculate the expectation value of an operator  which involves only system
A. Then the result is (simple proof, which uses orthogonality of |Φβ iB state):
 
X X
hΨ|Â|Ψi =  c∗α0 ,β cα,β  hφα0 |Â|φα i , (10.8)
α,α0 β
Density matrices by tracing out the universe 59
which you immediately recognize it can be expressed as:

hΨ|Â|Ψi = Tr[A][ρ] ,

where [A]α0 ,α = hφα0 |Â|φα i, and: X


[ρ]α,α0 = c∗α0 ,β cα,β . (10.9)
β

One can easily prove that the matrix ρ so defined is indeed a density matrix. To prove it, observe
that it is an Hermitean matrix, and hence can be diagonalized by an appropriate change of basis.
Moreover, its trace is obviously 1:
X X
Tr[ρ] = c∗α,β cα,β = |cα,β |2 = 1 .
αβ αβ

Let |ψn i be the orthonormal basis (in the space HA ) which diagonalizes [ρ]. Therefore, we can
write: X
ρ̂ = wn |ψn ihψn | ,
n
P
where, evidently, the eigenvalues wn are real, and n wn = 1, since the trace has to be 1. To
conclude that ρ̂ is indeed a density matrix you only have to prove that its eigenvalues are non-
negative: wn ≥ 0. Indeed, consider the operator  = Pφn = |ψn ihψn | (the projector on state φn ).
Then, on one hand you can write:

hΨ|Â|Ψi = Tr[A][ρ] = wn .

But, on the other hand, from the original expression of |Ψi you can easily show that:
2
X X
wn = hΨ|Â|Ψi = hψn |φα icα,β ≥ 0 ,

α
β

which clearly shows that wn ≥ 0.


11
Time-independent perturbation theory

Until now, we have dealt only with exactly solvable models. The list of such models is, however,
rather short. It is therefore very important to learn approximate techniques to tackle more compli-
cated problems which cannot be solved exactly. We begin with the treatment of time-independent
perturbation theory. That is, we assume we know how to solve:

H0 |φ(0) (0) (0)


n i = En |φn i , (11.1)
(0)
where the unperturbed states |φn i could be degenerate or non-degenerate, and aim at solving the
perturbed problem
(H0 + λV ) |φn (λ)i = En (λ)|φn (λ)i . (11.2)
where V is a time-independent Hermitean operator, known as the perturbation. The coupling
coefficient λ is included only for convenience, so that for λ = 0 one recovers the unperturbed
problem, and also as a way of accounting for different orders in the perturbation V .

11.1 Two-state problem

To get a feeling for the general theory, let us start considering a two-state problem, i.e., a two-
(0) (0
dimensional Hilbert space with two states |φ1,2 i with unperturbed energies E1,2 . The system is
(0) (0)
degenerate if E1 = E2 , and non-degenerate otherwise.

Solving this problem exactly amounts to diagonalizing the 2 × 2 matrix


!
(0)
E1 + λV11 λV12
[H0 + λV ] = (0) (11.3)
λV21 E2 + λV22

which represents the Hamiltonian H0 + λV in the unperturbed basis. We find the two perturbed
(λ)
eigenvalues E1,2 by imposing det(H0 + λV − E1) = 0, which gives, after very simple algebra:
v"
u (E (0) + λV ) − (E (0) + λV ) 2
u #
(0) (0)
E1 + λV11 + E2 + λV22 2 22 1 11
E± = ± t + λ2 |V12 |2 . (11.4)
2 2

11.1.1 Non-degenerate case

(0) (0)
Consider first the non-degenerate in which E2 > E1 . Rewrite it as:
(0) (0)
(E2 + λV22 ) + (E1 + λV11 )
E± = ±
2 s
(0) (0)
(E2 + λV22 ) − (E1 + λV11 ) 4λ2 |V12 |2
1+ (0) (0)
.
2 [(E2 + λV22 ) − (E1 + λV11 )]2
Non-degenerate case: general solution 61

Then, expanding the square-root up to second-order in λ, i.e., 1 + αλ2 ≈ 1 + αλ2 /2 + · · · , and
neglecting terms of order λ3 , we have:
(0) |V12 |2
E2 (λ) = E+ ≈ E2 + λV22 + λ2 (0) (0)
E2 − E1
(0) |V12 |2
E1 (λ) = E− ≈ E1 + λV11 − λ2 (0) (0)
. (11.5)
E2 − E1
These results anticipate those we will derive from the general theory, up to second order.

11.1.2 Degenerate case

(0) (0)
In the case E2 = E1 = E (0) the previous expressions are not well defined because of a zero in
the denominator of the second-order term. Luckily, we can solve the problem exactly, in this case,
the solution being:
 s 
2
V 11 + V 22 V 22 − V 11
E± (λ) = E (0) + λ  ± + |V12 |2  . (11.6)
2 2

So, in this case what we have to do is to diagonalize the perturbation V in the basis of the
degenerate unperturbed eigenstates of H0 . We will see that this is general.

11.2 Non-degenerate case: general solution

To solve the general non-degenerate case, write:


En (λ) = En(0) + λEn(1) + λ2 En(2) + · · ·
|φn (λ)i = |φ(0) (1) 2 (2)
n i + λ|φn i + λ |φn i + · · · , (11.7)
(i) (i)
where En and |φn i denote, respectively, the i-th order correction to the energy eigenvalue and
eigenstate, and impose that the perturbed Schrödinger problem is solved when terms of a given
order in λ are considered. To zeroth-order in λ, the problems is solved by Eq. 11.1.

11.2.1 First-order correction

Collecting all the first-order terms in λ we get:


H0 |φ(1) (0) (1) (0) (0) (1)
n i + V |φn i = En |φn i + En |φn i ,

or, equivalently:
(H0 − En(0) )|φ(1) (1) (0)
n i + (V − En )|φn i = 0 . (11.8)
(0)
Multiplying to the left by hφn |, we get:
En(1) = hφ(0) (0)
n |V |φn i =⇒ En (λ) = En(0) + λhφ(0) (0) 2
n |V |φn i + O(λ ) . (11.9)
(0)
As for the correction to the eigenstate, multiply to the left by a general hφm |, to get:
(0)
(Em − En(0) )hφ(0) (1) (0) (0) (1)
m |φn i + hφm |V |φn i − En δm,n = 0 . (11.10)
If m = n, this equation gives exactly (11.9). When m 6= n, we can deduce:
(0) (0) (0) (0)
hφm |V |φn i X hφm |λV |φn i
hφ(0) (1)
m |φn i = (0) (0)
=⇒ |φn (λ)i = |φ(0)
n i+ |φ(0)
m i (0) (0)
+ O(λ2 ) , (11.11)
En − Em m6=n En − Em
that is the nth state gets a small component of all other states with a coefficient given by the
matrix element of the perturbation divided by the energy gap.
62 Time-independent perturbation theory

11.2.2 Second-order correction

Collecting all the second-order terms in λ we get:

H0 |φ(2) (1) (2) (0) (1) (1) (0) (2)


n i + V |φn i = En |φn i + En |φn i + En |φn i ,

or, rearranging the terms:

(H0 − En(0) )|φ(2) (1) (1) (2) (0)


n i + (V − En )|φn i = En |φn i .

(0)
Multiplying to the left by hφn |, we get:

En(2) = hφ(0) (1)


n |V |φn i ,

(1)
and, upon inserting the expression (11.11) for |φn i,

X hφ(0) (0) (0) (0)


n |V |φm ihφm |V |φn i
En(2) = hφ(0) (1)
n |V |φn i = (0) (0)
. (11.12)
m6=n En − Em

Therefore, up to second-order we have:

X |hφ(0) (0) 2
m |λV |φn i|
En (λ) = En(0) + hφ(0) (0)
n |λV |φn i + (0) (0)
+ O(λ3 ) . (11.13)
m6=n En − Em

This is a very important expression, often used in calculations. The second-order correction to the
eigenstate, on the contrary, is very seldom used, and we will not derive it here.

11.2.3 Harmonic oscillator plus a linear potential

Consider, as a warm-up exercise, a one-dimensional harmonic oscillator under the action of a


uniform electric field E (imagine the particle has a charge q). The extra potential felt by the
particle is V (x) = −qEx, and the Hamiltonian reads:

p2 1
H = H0 + V = + mω 2 x2 − qEx .
2m 2
Luckily, this problem can be solved exactly in a very simple way, by completing the square:
2
q2 E 2

1 1 qE
mω 2 x2 − qEx = mω 2 x − 2
− .
2 2 mω 2mω 2

The exact eigenvalues are given by:

q2 E 2
 
1
En = ~ω n + − .
2 2mω 2
(0)
If we denote by φn (x) the eigenstates of the unperturbed harmonic p oscillator (recall that they
2 2
are Hermite polynomials times the Gaussian e−x /(2l ) , where l = ~/(mω) is the oscillator
characteristic length-scale), the exact eigenstates in the presence of the field are given by:
 
(0) qE
φn (x) = φn x − ,
mω 2
qE
i.e., they are the same functions, but shifted by a quantity − mω 2 . To do the perturbative calcula-

tion, go to dimensionless quantities, and express


Non-degenerate case: general solution 63
x qEl
V = −qEl = − √ (a + a+ ) .
l 2
(0) (0)
Next, observe that the expectation value of the perturbation is vanishing to first order: hφn |V |φn i
because both a and a+ change n to n−1 and n+1, respectively, and the bra then gives zero overlap.
For the second-order calculation you need:

qEl √ √ 
hφ(0) (0)
m |V |φn i = − √ δm,n+1 n + 1 + δm,n−1 n .
2

The second-order expression for the energies gives:

X |hφ(0) (0) 2
m |V |φn i|
En = En(0) + (0) (0)
+ ···
m6=n En − Em
q 2 E 2 l2 [(n + 1) − n]
 
1
= ~ω n + − + ···
2 2~ω
q2 E 2
 
1
= ~ω n + − ,
2 2mω 2

which coincides with the exact result: evidently, all the higher order corrections cancel exactly in
this case, at least for the eigenvalues. Indeed, verify that this is not the case for the eigenstates:

Exercise 11.1 Calculate the correction up to first order for the ground state of the harmonic oscillator
in the electric field. Compare the approximate expression with the exact one.

11.2.4 An-harmonic oscillator

Consider now a perturbation to the usual one-dimensional harmonic oscillator that is of the form:
 x 3
V = −λ~ω . (11.14)
l

Cubic terms come on quite general ground when you expand a generic potential close to a minimum
(example, the inter-atomic potential of a diatomic molecule). The peculiarity of a potential that
stops at this level, however, is that it is necessarily unbounded in one direction (x → +∞ if λ > 0),
and therefore, strictly speaking, it always provides a dramatic perturbation to the spectrum, no
matter how small the coefficient λ is. The perturbation series is, in this case, not a regular power
series (it is only an asymptotic series, but don’t worry about this): no bound state survives in the
presence of V , and the particle is always able to tunnel away (to x → +∞) from close to the origin.
Nevertheless, if λ is small enough, the time it takes for a particle to tunnel away is large enough
that, for all practical purposes, the perturbative calculation makes sense.

Working with the a, a+ , and n̂ = a+ a, we have:

λ~ω λ~ω
V = − √ (a + a+ )3 = − √ a3 + (a+ )3 + 3n̂a+ + 3(n̂ + 1)a ,

2 2 2 2

where we have carefully expanded the cube (a + a+ )3 paying due attention to the fact that the
terms do not commute, and used that an̂ = (n̂ + 1)a and a+ n̂ = (n̂ − 1)a+ . Once again, we have
no effect to first-order in λ, because V changes the n in the ket and the bra gives zero overlap. For
the second-order calculation you need:
64 Time-independent perturbation theory

(0) λ~ω p
hφn+3 |V |φ(0)
n i = − √ (n + 3)(n + 2)(n + 1)
2 2
(0) λ~ω p
hφn−3 |V |φ(0)
n i = − √ n(n − 1)(n − 2)
2 2
(0) λ~ω p
hφn+1 |V |φ(0)
n i = − √ 3 (n + 1)
3
2 2
(0) λ~ω √ 3
hφn−1 |V |φ(0)
n i = − √ 3 n .
2 2
(0) (0)
The energy denominators are very simple: En − Em = ~ω(n − m). Putting all the ingredients
together we get:
X |hφ(0) (0) 2
m |V |φn i|
En = En(0) + (0) (0)
+ ···
En − Em
m6=n
2
15λ2 ~ω 7λ2 ~ω
  
1 1
= ~ω n + − n+ − + ··· .
2 4 2 16
Notice that the energy separation between two neighboring levels is now:
 
15
En − En−1 = ~ω 1 − λ2 n + · · · ,
2
i.e., it decreases with increasing n from the equispaced harmonic levels, and shows a clear signal
of the fact that the perturbative series is dangerous, here, for high values of n: it changes sign for
large n, somewhat worrying for a quantity which is positive by definition!

Exercise 11.2 Calculate the correction up to first order for the ground state of the harmonic oscillator
in presence of the x3 correction. Show that the average value of x on such a perturbed state is now not
vanishing, but equal to hφ0 |x|φ0 i = 3λl/2.

11.2.5 Van-der-Waals attraction between two H-atoms

Consider two hydrogen atoms far away from each other, with the two nuclei at a distance R. What
will be the potential V (R) between them? Each of the two atoms being neutral, there cannot be
any 1/R term in V . The goal of the present exercise is to show that the first contribution to V (R)
comes from dipole-dipole terms, and leads to a V (R) having an attractive 1/R6 tail, known as van
der Walls attraction, i.e.,
C
V (R) = − 6 .
R

The Hamiltonian of the two atoms, with protons sitting at positions R1 and R2 , is:
 2
p2 e2
  2
p2 e2

P1 P2
H= + 1 − + + 2 − + VCoulomb
2M 2m r1 2M 2m r2
| {z } | {z }
H1 H2
2 2
e e
VCoulomb = +
|R2 − R1 | |(R2 + r2 ) − (R1 + r1 )|
e2 e2
− − .
|(R2 + r2 ) − R1 | |R2 − (R1 + r1 )|
We will consider here H1 + H2 , the Hamiltonian of the two isolated atoms, as our unperturbed
Hamiltonian H0 = H1 + H2 . The last four terms, accounting for the Coulomb repulsion between
Non-degenerate case: general solution 65
the two protons and the two electrons, and the attraction of each electron from the proton of
the other atom, will be treated as a perturbation. Such a perturbation is small if the separation
R = R2 − R1 between the two protons is much larger than the typical scale, the Bohr radius aB ,
over which each electronic coordinate r1 and r2 varies. It is convenient to expand the last four
Coulomb term assuming |r1 |, |r2 |, |r2 − r1 |  R. The typical expression appearing is:
−1/2
d2 + 2R · d

1 1 1
=√ = 1+ .
|R + d| R2 + d2 + 2R · d R R2

Using the Taylor expansion (1 + x)−1/2 ≈ 1 − (1/2)x + (3/8)x2 + · · · , and neglecting terms of order
d3 and higher, we get:

1 d2 + 2R · d 3 (2R · d)2
 
1 1
≈ 1− + + · · · .
|R + d| R 2 R2 8 R4

Applying this expansion to VCoulomb we get:

e2   e2
VCoulomb ≈ Vdip−dip = r1 · r2 − 3(R̂ · r1 )(R̂ · r2 ) = (x1 x2 + y1 y2 − 2z1 z2 ) ,
R3 R3

where we have assumed that the direction between the two atoms to be in the z-direction, R̂ =
R/R = ẑ.
(0)
The ground state of H0 is given by φ0 (r1 , r2 ) = ψ100 (r1 )ψ100 (r2 ), with unperturbed energy
(0)
E0 = 2E1 , where E1 = −e2 /(2aB ) = −Ry is the ground state energy of a single hydrogen atom.
This is the first time we encounter a problem with more than one electron. The wave-function
written above is symmetric under exchange of the two electrons, and we will see later on that a
fundamental principle of physics demands that the wave-function of a system of identical fermions
(as the electrons are) should be anti-symmetric under exchange of the particles. We will also see
that the way out is given by the fact that there is actually a spin part to the wave-function, which
is anti-symmetric in the present case, and makes the whole wave-function anti-symmetric:

(0) 1
φ0 = ψ100 (r1 )ψ100 (r2 ) √ (| ↑↓i − | ↓↑i) .
2
Such a spin-part is, however, inessential in the present case, because the Hamiltonian never acts
upon it, and the corresponding spin-bra makes the spin-ket to disappear. So, let us neglect this
potential complication for the time being.

As is often the case, the first-order contribution of Vdip−dip to the ground-state energy vanishes
here, because of parity reasons. For instance, the x1 x2 term would give a contribution:

hψ100 ψ100 |x1 x2 |ψ100 ψ100 i = hψ100 |x1 |ψ100 ihψ100 |x2 |ψ100 i = 0 ,

because hψ100 |x1 |ψ100 i = hψ100 |x2 |ψ100 i = 0 by parity (ψ100 (r) is even, which x is odd under
parity). A similar fate occurs to y1 y2 and z1 z2 .

The second order contribution requires, in principle, all the excited states (treat the two electrons
as distinguishable, for simplicity)
(0)
φn0 l0 m0 ,n00 l00 m00 (r1 , r2 ) = ψn0 l0 m0 (r1 )ψn00 l00 m00 (r2 )

of unperturbed energy En0 +En00 , where En = E1 /n2 is the unperturbed energy of a single hydrogen
atom:
6=(100,100)
(0) e4 X |hψn0 l0 m0 ψn00 l00 m00 |(x1 x2 + y1 y2 − 2z1 z2 )|ψ100 ψ100 i|2
E0 = E0 − .
R6 En0 + En00 − 2E1
(n0 l0 m0 ,n00 l00 m00 )
66 Time-independent perturbation theory

This immediately leads us to conclude that V (R) = −C/R6 , as anticipated. The matrix elements
which now appear are, for instance:

hψn0 l0 m0 ψn00 l00 m00 |z1 z2 |ψ100 ψ100 i = hψn0 l0 m0 |z1 |ψ100 ihψn00 l00 m00 |z2 |ψ100 i ,

and similar ones for x1 x2 and y1 y2 . Parity immediately tells us that l0 and l00 should be both odd, in
order for the matrix elements not to vanish. This is an example of a selection rule (i.e., a vanishing
of a matrix element) due to a symmetry (parity, in the present case). There is, however, another
symmetry in the present case, i.e., rotational symmetry, which provides a further selection rule.
Indeed, remember that ψ100 (r) = R10 (r)Y00 (θ, φ) is rotationally symmetric, because Y00 = √14π
p
does not depend on (θ, φ). Also, recall that z = r cos θ = r 4π/3 Y10 , and that hYl0 m0 |Ylm i =
δl,l0 δm,m0 . Therefore:
Z ∞ r
4π π
Z Z 2π
hψn l m |z|ψ100 i =
0 0 0
2
drr Rn l (r)R10 (r)r
0 0 dθ dφ Yl∗0 m0 Y10 Y00
0 3 0 0
Z ∞
r
= drr2 Rn0 l0 (r)R10 (r) √ hYl∗0 m0 |Y10 i
0 3
Z ∞
r
= δl0 ,1 δm0 ,0 drr2 Rn0 l0 (r)R10 (r) √ .
0 3
A similar argument applies to the x and y matrix elements, which can be expressed in terms of
Y1,+1 and Y1,−1 . Therefore, rotational symmetry implies a further, stronger, selection rule telling
us that:
l0 = 1 and l00 = 1 . (11.15)
So, only p-states need to be included in the second-order calculation. The calculation might proceed
now with the evaluation of all the necessary integrals. It is, however, rather long and boring.
An approximate estimate for this second-order contribution is obtained by noticing that |En0 | =
|E1 |/(n0 )2  |E1 |, so that we can approximately neglect En0 + En00 in the energy denominator.
The advantage of this is that one can then perform the sum over (n0 l0 m0 , n00 l00 m00 ) exactly, because
it leads to appearance of the identity:
X
|ψn0 l0 m0 ψn00 l00 m00 ihψn0 l0 m0 ψn00 l00 m00 | = 1 .
(n0 l0 m0 ,n00 l00 m00 )

Therefore:

(2) e4 hψ100 ψ100 |(x1 x2 + y1 y2 − 2z1 z2 )2 |ψ100 ψ100 i


E0 ≈−
R6 −2E1
6e2 aB 2
hψ100 |x21 |ψ100 i ,

≈− (11.16)
R6
where the factor 6 comes from 1 + 1 + 4, due to the terms x21 x22 , y12 y22 and 4z12 z22 , all giving the same
contribution by rotational symmetry, and we used the fact that −2E1 = e2 /aB . The necessary
integral can be easily evaluated:
hψ100 |x21 |ψ100 i = a2B ,
so that, finally:
(2) 6e2 a5B
V (R) ≈ E0 ≈− . (11.17)
R6

11.3 Degenerate case: first order

Until now we have considered only the case in which the level we are considering is non-degenerate,
(0) (0)
i.e., there is a unique eigenstate |φn i with energy En . Consider now the case in which there are
Degenerate case: first order 67
(0) (0)
gn > 1 unperturbed states {|φn,α i, α = 1 · · · gn } with energy En . Clearly, this brings a potential
problem in all formulas that have energy denominators, because different states with the same
energy would give a divergent contribution.

On the other hand, the situation in which gn > 1 is quite common: think, for instance, of the
hydrogen atom, where the first excited state at energy E2 = E1 /22 is four-fold degenerate (g2 = 4,
without accounting for the spin), the second excited state at E3 = E1 /32 has g3 = 9, and so on.
(You can show that the the level En has gn = n2 .)
(0)
Clearly, you can mix the levels {|φn,α i, α = 1 · · · gn } with a unitary gn × gn matrix Uα,α0 ,
obtaining states:
gn
(0)
X
0
|φ(0)
n,α i = Uα,α0 |φn,α0 i ,
α0 =1

(0)
which still have the same unperturbed energy En , and are as good as the original ones. Which,
among the infinite many possible choices obtained with arbitrary unitary rotations U , is the “best”
(0)
possible choice? The answer is that the optimal choice is given by those states {|φn,α i, α = 1 · · · gn }
which diagonalize the gn × gn matrix representing the perturbation V in the degenerate subspace,
i.e., such that:
(0)
hφ(0) (1)
n,α |V |φn,α0 i = En,α δα,α0 .

To prove this, imagine that we pretend that all quantities are continuous as a function of λ for
small λ, writing, as usual:

En,α = En(0) + λEn,α


(1)
+ ···
|φn,α i = |φ(0) (1)
n,α i + λ|φn,α i + · · · , (11.18)

and substituting in the Schrödinger equation (H0 + λV )|φn,α0 i = En,α0 |φn,α0 i. The zero-th order
terms are automatically satisfied. The first-order terms read (just add the index α to Eq. (11.8):

(1) (1) (0)


(H0 − En(0) )|φn,α0 i + (V − En,α0 )|φn,α0 i = 0 . (11.19)

(0) (0)
Multiply on the left by hφn,α |, and again observe that the first term (H0 − En ) vanishes. The
(0)
second term, on the other hand, using the fact that the {|φn,α i, α = 1 · · · gn } are assumed to be
orthonormal, gives:
(0) (1) (0)
hφ(0) (0) (1)
n,α |V |φn,α0 i = En,α0 hφn,α |φn,α0 i = En,α δα,α0 ,

i.e., the selected states should diagonalize the corresponding matrix [V ] in order for the expansion
to be continuous for small λ. In words, when you turn on an infinitesimal λ, the perturbation V
“prefers” immediately the states that diagonalize it, out of the infinite possible combinations.

11.3.1 Linear Stark effect

Let us start applying this ideas to the problem of an hydrogen atom in the presence of an electric
field in the z-direction. The Hamiltonian is:
 2
e2

p
H= − + eEz . (11.20)
2m r
| {z }
H0

(0)
Consider first the ground state |φ1 i = |ψn=1,l=0,m=0 i. Clearly, there is no first-order contribution
to the ground state energy, because, by parity, hψ100 |z|ψ100 i = 0.
68 Time-independent perturbation theory

Exercise 11.3 Calculate (approximately) the second-order shift (down) of the ground state energy in
presence of the electric field E, by applying the second-order formula. The calculations are similar to those
occurring in the van der Walls problem for two distant hydrogen atoms.

Consider now the 4 levels corresponding to the first excited states, which are the 2s (n = 2, l =
0, m = 0), and the three 2p (n = 2, l = 1, m = 0/ + 1/ − 1). It is simple to show that the matrix
representing the perturbation V = eEz in this 4-dimensional space is:
 
hψ200 |z|ψ200 i hψ200 |z|ψ210 i hψ200 |z|ψ21+1 i hψ200 |z|ψ21−1 i
 hψ210 |z|ψ200 i hψ210 |z|ψ210 i hψ210 |z|ψ21+1 i hψ210 |z|ψ21−1 i 
[V ] = eE  
 hψ21+1 |z|ψ200 i hψ21+1 |z|ψ210 i hψ21+1 |z|ψ21+1 i hψ21+1 |z|ψ21−1 i 
hψ21−1 |z|ψ200 i hψ21−1 |z|ψ210 i hψ21−1 |z|ψ21+1 i hψ21−1 |z|ψ21−1 i
 
0 ∆00
∆ 0 0 0
= eE 
 0 0 0 0 .

0 0 00

Exercise 11.4 Calculate


“ explicitly
” the quantity ∆ appearing“in the
” matrix [V ] above, using the fact that
ψ200 (r) = √ 1 3 1 − 2arB e−r/(2aB ) and ψ210 (r) = √ 1 3 azB e−r/(2aB ) , to show that ∆ = 3aB .
8πaB 32πaB

So, the levels ψ21±1 are not affected by the perturbation, to first order, while the 2s state and the
2p state with m = 0 are mixed by an off-diagonal term. By diagonalizing the small 2 × 2 block of
the matrix [V ] we realize that the correct states to consider are:
1
|ψ2,± i = √ (|ψ200 i ± |ψ210 i) , (11.21)
2
which are split linearly in the field E:
(0)
E2,± = E2 ± eE∆ . (11.22)

This linear splitting is also known as linear Stark effect. The corresponding optical transition line
towards the ground state is consequently split, in presence of an electric field.

11.4 Application: the fine-structure of Hydrogen

There are several relativistic corrections to the hydrogen atom Hamiltonian H0 which we have
encountered so far:
 2  2 2
e2 e2 π~2 e2

p 1 p
H= − + L · S − + δ(r) + · · · . (11.23)
2m r 2m2 c2{z
r3 2
} |2mc {z2m } |2m {z
2 c2
| {z } | }
H0 Hspin−orbit Hkinetic−energy HDarwin

The origin of the spin-orbit term has already been discussed. The kinetic energy correction comes
from the expansion of the relativistic energy dispersion:
p p p2 p4
p2 c2 + m2 c4 = mc2 1 + p2 /m2 c2 ≈ mc2 + − + ··· ,
2m 8m3 c2
where the first constant term, mc2 , is never included in condensed matter, p2 /2m is the usual
non-relativistic expression, and the third term is the promised relativistic correction. The final
Application: the fine-structure of Hydrogen 69
term, called Darwin correction, is very difficult to justify: it comes from the limit of the Dirac’s
relativistic equation, precisely from terms of the form ∇2 (1/r), that is all I can tell you.

Let us estimate the order of magnitude of these three terms. As for the spin-orbit term we have:

hHspin−orbit i e2 ~2 /(m2 c2 a3B ) ~2


≈ = = α2 ,
hH0 i e2 /aB m2 c2 a2B

where α = e2 /(~c) ≈ 1/137 is the so-called fine-structure constant. Similarly:

hHkinetic−energy i h(p2 /2m)i e2


≈ 2
≈ = α2 .
hH0 i 2mc aB mc2

Finally, since |ψ(0)|2 ≈ 1/a3B :

hHDarwin i e2 ~2 |ψ(0)|2 /(m2 c2 )


≈ ≈ α2 .
hH0 i e2 /aB

So, all three terms are of the same order of magnitude, i.e. α4 mc2 = 2α2 Ry.

Let us consider the n = 2 multiplied of the 2s-2p states. These are 4 orbitals, i.e., 4 × 2 = 8
states, including spin. The Darwin term shifts only the 2s state, because ψ(0) = 0 for the 2p states.
We can calculate the effect of HDarwin to first-order:

(1,Darwin) π~2 e2 π~2 e2


E2s = hψ200 | δ(r)|ψ 200 i = |ψ200 (0)|2 .
2m2 c2 2m2 c2
Using |ψn00 (0)|2 = 1/(πn3 a3B ), we get:

(1,Darwin) 1
E2s = mc2 α4 .
16

Let us continue with the 2s states. The effect of the spin-orbit term is obviously zero, because
Lψ200 = 0 for the 2s (l = 0) state. As for the kinetic-energy correction we have, in general:
2 2
p2 e2
 
(1,kinetic−energy) 1 1
Enlm =− hψnlm | |ψnlm i = − hψ nlm | H 0 + |ψnlm i .
2mc2 2m 2mc2 r
Expanding the square:
2
e2 e4 e2 e2

H0 + = H02 + + H 0 + H0 ,
r r2 r r
and noticing that H0 |ψnlm i = En |ψnlm i (even when H0 acts to the left, on the bra), we can easily
calculate:
e2 e4
 
(1,kinetic−energy) 1 2
Enlm =− En + 2En hψnlm | |ψnlm i + hψnlm | 2 |ψnlm i .
2mc2 r r

So, all we need is h1/ri and h1/r2 i. Let us give the general results for these averages:
1 1
hψnlm | |ψnlm i = 2 (11.24)
r n aB
1 1
hψnlm | 2 |ψnlm i = 3 2 (11.25)
r n aB (l + 1/2)
1 1
hψnlm | 3 |ψnlm i = 3 3 . (11.26)
r n aB l(l + 1/2)(l + 1)

(The last term will be useful in calculating the spin-orbit term.) Putting all the terms together,
we get, after simple algebra:
70 Time-independent perturbation theory
2
e2

(1,kinetic−energy) 1 16 − 3(2l + 1)
Enlm =− . (11.27)
mc2 aB 128(2l + 1)

For the 2s state, therefore:


(1),kinetic−energy 13
E2s =− mc2 α4 ,
128
and, summing up all three terms we get:

(1) (1) 5
E2s = E2j= 1 l=0s= 1 = − mc2 α4 .
2 2 128
1
where we have labeled the state with its total j = 0 + 2 = 12 .

Consider now the 2p states. The Darwin term, as mentioned, has no effect. The kinetic-energy
correction term gives:
(1,kinetic−energy) 7
E2p =− mc2 α4 . (11.28)
384
The spin-orbit term is the most delicate one. As previously discussed, the spin-orbit term (and,
indeed, the full H) commutes with L2 , S 2 , J 2 and Jz , where J = L + S. Indeed, recall that:

J 2 − L2 − S 2
L·S= .
2
Therefore, rather then using the 3 × 2 = 6 product states |ψ21m i|s = 12 , ± 21 i, the most convenient
basis to use is that of the eigenstates of L2 , S 2 , J 2 and Jz : |l = 1, s = 21 , j, jz i. As discussed
previously, the possible values of J are j = l + 21 = 32 (4 states), usually denoted by 4 p 32 , and
j = l − 12 = 21 (2 states), denoted by 2 p 21 . Therefore, for general l:

(1,spin−orbit) e2 1 j(j + 1) − l(l + 1) − s(s + 1)


Enjls = .
2m2 c2 n3 a3B l(l + 1/2)(l + 1) 2

In particular, for the 2p states, where we have l = 1 and s = 12 , we get:

e2
 
(1,spin−orbit) 1 11
E2jl=1s= 1 = j(j + 1) − .
2 m c 96a3B
2 2 4

Summing together the kinetic-energy correction and the spin-orbit term we get, explicitly:
 
(1) 1 7 5
E2j= 1 l=1s= 1 = − − mc2 α4 = − mc2 α4
2 2 48 384 128
 
(1) 1 7 1
E2j= 3 l=1s= 1 = + − mc2 α4 = − mc2 α4 .
2 2 96 384 128

Notice that the two j = 12 doublets, i.e., the 2 s 12 (derived from l = 0 and s = 21 ) and the 2 p 12
(derived from l = 1 and s = 12 ) are perfectly degenerate after accounting for the relativistic
corrections. This degeneracy survives all the relativistic corrections, to all orders in perturbation
theory, and is really due to the fact that the eigenvalues of the Dirac’s equation depend on j, rather
than l or s. The degeneracy between 2 s 12 and 2 p 12 is only removed (the so-called Lamb shift) by
quantum electrodynamics corrections. Comment on the spectroscopic fingerprints of the resulting
fine-structure spectrum, which is experimentally measurable with high accuracy.
12
Time-dependent perturbation theory

Until now, we have dealt with time-independent perturbations V . There are many cases, however,
where the perturbation one considers represents an external time-dependent field (for instance
an electromagnetic field). Therefore, we have to learn how to solve a time-dependent Schrödinger
problem
d
i~ |ψ(t)i = [H0 + λV (t)] |ψ(t)i. (12.1)
dt
developing a systematic expansion in λ.

12.1 General case


(0)
Let {|φn i} represent the unperturbed eigenstates associated to H0 , with energies {En0 }, which
form a complete orthonormal basis for expanding a generic state. The state |ψ(t)i can then be
written as: X 0
|ψ(t)i = cn (t)e−iEn t/~ |φ(0)
n i. (12.2)
n

where we have explicitly included the time-dependence associated to H0 , in such a way that for
λ = 0 the coefficients cn would not depend on time, and we imagine them to be “slowly varying”
for small λ. (Technically, this is equivalent to the so-called interaction picture representation.)
By substituting this expansion in the time-dependent Schrödinger problem (12.1), and taking the
scalar product with hφn | we get:

dcn (t) X 0 0 X
i~ =λ hφn |V (t)|φm i ei(En −Em )t/~ cm (t) = λ Vnm (t) eiωnm t cm (t) , (12.3)
dt m m

where, as a shorthand, we have defined

En0 − Em
0
ωnm = , (12.4)
~
to be the frequency associated to the difference between the unperturbed energies relative to levels
n and m, and
Vnm (t) = hφn |V (t)|φm i , (12.5)
to be the corresponding matrix element of the perturbation V (t). As it is, this is a complicated
set (infinite-dimensional, in general) of time-dependent, first-order differential equations for the
{cn (t)}. Assume, as usual, a series expansion in powers of λ:

X
cn (t) = c(0) (1) 2 (2)
n (t) + λcn (t) + λ cn (t) + · · · = λq c(q)
n (t) , (12.6)
q=0

and substitute this series in Eq. (12.3). Noticing that the right-hand side of it has an extra factor
λ, we get:
72 Time-dependent perturbation theory
(0)
dcn (t)
i~ =0
dt
(1)
dcn (t) X
i~ = Vnm (t) eiωnm t c(0)
m
dt m
··· = ···
(q)
dcn (t) X
i~ = Vnm (t) eiωnm t c(q−1)
m (t) . (12.7)
dt m

In words, one can solve the problem iteratively.

12.2 First-order solution

Suppose that the system is found in the initial state with n = i for t ≤ t0 (very often, the ground
state, with n = i = 0), and that we switch-on the perturbation abruptly at t = t0 . The zeroth-order
(0) (0)
term is clearly time-independent: cn (t) = cn = δn,i . Up to first order, we get:
(1) Z t
dcn (t) 1 0
i~ = Vni (t) eiωni t =⇒ c(1)
n (t) = dt0 Vni (t0 ) eiωni t . (12.8)
dt i~ t0

So, the transition probability for the system to go from state i at t0 to a state n 6= i at time t is:
(1)
Pi→n6=i (t) = |c(0) (1) 2
n + λcn (t) + · · · | ≈ Pi→n6=i (t)
Z t 2
(1) (1) 2 1 0 0 iωni t0

Pi→n6=i (t) = |λcn (t)| = 2 λ dt Vni (t ) e . (12.9)
~ t0

Let us consider the case in which H0 is the usual one-dimensional harmonic oscillator. At time
(0)
t0 = −∞ the harmonic oscillator is in its ground state |φ0 i; then we switch on an electric field E
which is subsequently switched-off as t → +∞:
2
/τ 2
λV (t) = qExe−t .
(0)
We ask what is the probability that the system is left in the n-th excited state, |φn i, at t = +∞.
We have here:
2 2 l
+ (0)
λVni (t) = qEe−t /τ √ hφ(0) n |a + a |φ0 i .
2
2 2
Calculating the matrix element and the time-integral (Fourier transform of a Gaussian, dt e−t /τ eiωt =
R
2 2
(πτ 2 )1/2 e−ω τ /4 ) we get:
(1) πq 2 E 2 τ 2 −ω2 τ 2 /2
P0→n (t) = δn,1 e . (12.10)
2m~ω
There is a non-zero probability of exciting the system to the first-excited state (the only one which
the perturbation couples the ground state with); such a probability of excitation goes to zero both
in the extreme adiabatic limit τ → ∞, and in the fast switching-on-and-off limit τ → 0.

As a second application, imagine that, starting from any H0 , we turn on a perturbation λV


slowly in time from t0 = −∞ to time t = 0 in the following way:

λV (t) = λV et/τ ,
(0)
so that the Hamiltonian at time t = 0 is H(t = 0) = H0 + λV . Every initial state |φn i is mixed
(0)
with other states |φm i, for an amount which is exactly given by cm (t = 0). To first order in λ the
resulting state at time t = 0 is:
First-order solution 73
X
|φn (t = 0)i = |φ(0)
n i+ λc(1) (0)
m (0)|φm i ,
m6=n

where
1 0
Z
0 0 1 1
c(1)
m (0) = dt0 Vmn et /τ eiωmn t = Vmn .
i~ −∞ i~ iωmn + τ −1
In the limit in which the perturbation is turned on infinitely slowly (τ → ∞) we see that we recover
exactly the same first-order perturbation result obtained previously:
(0) (0)
X hφm |λV |φn i
|φn (λ)i = |φ(0)
n i+ |φ(0)
m i (0) (0)
+ ··· . (12.11)
m6=n En − Em

12.2.1 Periodic perturbation

A particularly important case is represented by a perturbation which is periodic in time, with a


frequency ω:
eiωt + e−iωt
Vnm (t) = Vnm cos (ωt) = Vnm .
2
The time integrals can now be done explicitly:
2
|Vni |2 1 − ei(ωni +ω)t 1 − ei(ωni −ω)t

(1)
Pi→n6=i (t) = + , (12.12)
4~2 ωni + ω ωni − ω
where we have dropped any reference to λ, by setting λVni → Vni . Now suppose that ωni > 0
(as indeed is the case if the systems starts from the ground state i = 0). Then, quite clearly,
the second term is resonant for ω ≈ ωni , and dominates over the first term. Indeed, if we have a
frequency ω = ωni + ∆ω, the first term is proportional to 1/(2ωni + ∆ω), the second to 1/(−∆ω)
and dominates over the first if |∆ω|  2|ωni |. (For ωni < 0 the resonance occurs in the first term,
but the conclusions will be identical.) Keeping only the resonant term, we have:
2
|Vni |2 ei(ωni −ω)t/2 − e−i(ωni −ω)t/2 |Vni |2 sin2 (ωni − ω)t/2

(1,res)
Pi→n6=i (t) ≈ = . (12.13)
4~2 ωni − ω 4~2 [(ωni − ω)/2]2
Notice that the last expression is a function of ω which is strongly peaked at ω = ωni , with a width
in frequency of ±|∆ω| = ±2π/t. As mentioned above, the validity of the resonance approximation
requires
2π 2π 2π
2|ωni |  |∆ω| = =⇒ t ≈ ,
t |ωni | ω
i.e., the perturbation should act for a time t which is long enough that the system experiences
many cycles, 2π/ω, of the perturbation, in order to be able to do a “Fourier analysis” of it. Next,
at resonance, i.e., for ω = ωni , the transition probability behaves as:
(1,res) |Vni |2 t2 ~
Pi→n (t) ≈ 1 =⇒ t , (12.14)
4~2 |Vni |
1,res)
where the inequality Pi→n (t)  1 is justified because the t2 behaviour would lead to a divergence
for large t, which is clearly an artifact of our first-order approximation. Therefore, both inequalities
imply:
2π 2π ~
≈ t =⇒ ~|ωni |  |Vni | , (12.15)
|ωni | ω |Vni |
that is, in words, that the perturbation matrix elements |Vni | should be much smaller than the
bare excitation energy ~|ωni |, i.e.:

hφ(0) |λV |φ(0) i
n i
1,

(0)
En − E (0)
i
which is indeed the requirement that the first-order correction to the initial eigenstate is “small”
(see Eq. 12.11).
74 Time-dependent perturbation theory

12.2.2 Fermi Golden Rule

Let us consider again the expression for the transition probability in the resonance approximation.
The resonance approximation is equivalent, for ωni > 0, to assuming that the time-dependence of
the perturbation is simply V e−iωt /2, neglecting the other term V e+iωt /2. It is therefore customary
to assume from the beginning a time-dependence

H(t) = H0 + V e−iωt ,

without any factor 1/2. Repeating the steps done in the previous section, we would find and
identical result (without any need to drop the non-resonant term), without the factor 4 in the
denominator (coming from 1/22 ):

(1,res) |Vni |2 sin2 (ωni − ω)t/2


Pi→n6=i (t) ≈ πt , (12.16)
~2 πt[(ωni − ω)/2]2

where we have multiplied and divided by a factor πt, which will be useful in a second. Define now
x = (ωni − ω)/2. The sine-factor reads:
sin2 tx
,
πtx2
which is a function whose value in x = 0 is t/π, with a width |x| = π/t, and whose integral over
all x is 1:
Z +∞
sin2 tx
dx =1.
−∞ πtx2
Therefore, as t grows towards +∞, the peak of the function is sharper and higher, but the integral
remains 1, i.e., the function approaches a delta function:

sin2 tx t→+∞
−→ δ(x) .
πtx2
Now, let us define the transition rate
(1,res)
Pi→n6=i (t)
Ri→n = ,
t
to be the transition probability per unit time. Then, for long times we have:

|Vni |2 2π (0)
Ri→n ≈ 2
πδ((ωni − ω)/2) = |Vni |2 δ(En(0) − Ei − ~ω) , (12.17)
~ ~

where we have used the property that δ(αx) = (1/α)δ(x) to put the factors ~ and 2π in the proper
place. Notice that, strictly speaking, the expression involving the delta-function, expressing in a
strict way the requirement of “energy conservation” between the two energy levels involved in the
transition when the energy ~ω of the perturbing field is taken into account, is justified only if we
have a continuum of states with a given density of states: for any discrete spectrum, it is necessary
to leave a certain width to the energy conservation requirement (i.e., leave the sine-factor). For
instance, when we will deal with an electromagnetic field impinging on an hydrogen atom, the
situation in which ~ω > 13.6eV, which would lead to exciting the electron to the continuum of
E > 0 states, is quite appropriately described by the delta-function expression: this is the so-called
photoelectric effect, in which the impinging electromagnetic field takes off the electron from the
atom.
(0)
In a situation in which you have a continuum of energy states En towards which we are exciting
(0)
the system, it is meaningful to integrate over all final states En having an energy in the interval
Electromagnetic radiation on an atom 75
[Ef − ∆E/2, Ef + ∆E/2], in the limit in which ∆E → 0. Since there are ρ(Ef ) × ∆E such states,
by the very definition of the density of states ρ(Ef ), we can calculate:

Ef +∆E/2 Z Ef +∆E/2
X
Ri→f = Ri→n → dEf ρ(Ef ) Ri→n
(0) Ef −∆E/2
En =Ef −∆E/2
2π (0)
≈ |Vf i |2 ρ(Ei + ~ω) , (12.18)
~
(0)
where we have set Ef = Ei + ~ω, as dictated by the energy-conserving delta function, and
(0)
assumed that the matrix element Vni does not depend much on the wave-function |φn i of the
final state, but only on the final energy, and can be therefore written in terms of a representative
(0)
state |φf i at the final energy.

12.3 Electromagnetic radiation on an atom

Suppose that you have an hydrogen atom in a region of space in which there is a monochromatic
electromagnetic field of frequency ω. We will describe the em-field classically, and in the Coulomb
gauge (∇~ · A = 0, and φ = 0), as a plane wave described by a vector potential:

1h i
A(r, t) = A0 ei(k·r−ωt) + c.c. ,
2

where the fact that ∇~ · A = 0 implies k · A0 = 0, i.e., the vector A0 is in a plane perpendicular
to the direction of propagation k, which we will assume to be in the y-direction, k = kŷ. Taking
A0 = A0 ẑ, with A0 real, we get the following expression for the electric and magnetic field in the
wave:

A = = A0 ẑ cos (ky − ωt)


1 ∂
E=− A = −Eẑ sin (ky − ωt)
c ∂t
B = ∇ × A = (kA0 )x̂ sin (ky − ωt) , (12.19)

where E = A0 ω/c. The Hamiltonian for the hydrogen atom (neglecting all relativistic corrections)
reads:
1  e 2 e2 2µB
H= p+ A − + B·S
2m c r ~
e 2µB e2
= H0 + A·p+ B·S+ A2 . (12.20)
mc ~ 2mc2
It is easy to estimate the order of magnitude of these three perturbing terms. We have, for instance:

h 2µ~B B · Si e~
mc kA0 aB
e = e
= kaB = 2π 1,
h mc A · pi ~
mc A0 aB
λ

where we used k = 2π/λ, and assumed an em-field with a wavelength λ ≈ 10−6 m, i.e., in the
visible range. In a similar way, one can show that the A2 term is negligible with respect to the
A · p term. We are therefore lead to considering the following perturbation problem:
e A0  iky −iωt
+ e−iky eiωt ẑ · p ,

H(t) = H0 + e e (12.21)
mc 2
where we easily recognize the structure we have studied so far, with the e−iωt -term being resonant
(0) (0)
when we consider the excitation process (absorption), in which En > Ei , and the eiωt -term being
76 Time-dependent perturbation theory
(0) (0)
resonant when the de-excitation process (emission), En < Ei , is considered. Both processes lead
(0) (0)
to an identical result, in this approximation. Considering the excitation process, En > Ei , and
keeping only the resonant term, we have:

|Vni |2 sin2 (ωni − ω)t/2


Ri→n =
4~2 [(ωni − ω)/2]2 t
eA0 (0) iky (0)
Vni = hφ |e pz |φi i . (12.22)
mc n
There is now an important approximation (the so-called dipole approximation) we can use in
calculating the last matrix element. Indeed, for visible light eiky ≈ 1, because kaB  1, and
therefore:
eA0 (0) (0)
Vni ≈ hφ |pz |φi i .
mc n
Next, by using the fact that the commutator

p2z i~
[z, H0 ] = [z, ] = pz ,
2m m
and that H0 when applied to its unperturbed eigenstates gives the corresponding eigenvalues, we
get:
(0) (0)
eE En − Ei (0)
Vni ≈ − hφ(0)
n |z|φi i . (12.23)
i ~ω
where we have used E = A0 ω/c. Notice that at resonance:
(0)
En(0) − Ei = ~ω ,

and therefore:
res (0)
|Vni | ≈ eE hφ(0) |z|φ i . (12.24)

n i

In summary, the relevant matrix element for the transition is given by the same dipole matrix
element considered in the Stark effect. In particular, the same selection rules apply:

l0 = l ± 1
m0 = m (12.25)

where the m0 = m is due to our choice of A along the ẑ direction: had we chosen A = A0 x̂ we
would get m0 = m ± 1. This selection rule is known as dipole selection rule and applies to optical
transition in which a photon is absorbed or emitted by an atom. For an hydrogen atom in the 1s
ground state, l = m = 0, and therefore only p-states can be reached.

A few comments are in order. First, we have treated the em-field as classical; a proper quantum
treatment would require quantization of A in terms of photon creator and annihilation operators;
in that case, one would get that the emission process has a slightly different rate (∝ n + 1) from
the absorption one (∝ n), because of spontaneous emission (in absence of any field). Second, as a
function of ω, the absorption rate from the ground state (1s) would show several sharp peaks in
correspondence with the 1s → 2p, 1s → 3p, etc. transitions. Third, one can calculate the photo-
ionization process when the excitation occurs towards E > 0 final states and the electron is taken
away from the nucleus, i.e. for ~ω > 13.6 eV, by using the Fermi golden-rule formula: this shows-up
as a continuum contribution in the absorption spectrum.
13
Many particle systems

Dealing with a system of N interacting quantum particles, either fermions or bosons, is a com-
plicated business. The required extension of the quantum mechanical formalism, i.e., learning the
basic properties of the many-particle states which are physically allowed, will be treated next, with
an application to a variational approach to the Helium atom. A sketch of the second-quantization
formalism is included in the appendix, for those of you who are interested in getting deeper into
the subject.

13.1 Many-particle basis states for fermions and bosons.

Suppose we have a system of N identical particles. Consider an appropriate basis set of single-
particle states {|αi}. Generally speaking, each state is a so-called spin-orbital, composed of an
orbital part |αio and a spin part |σα is

|αi = |αio |σα is (13.1)

which one could express in terms of real-space wave-functions by taking the scalar product with
the position-spin eigenstates |ri = |rio |σis , 1
def
ψα (r) = hr|αi = φα (r)χσα (σ) . (13.2)

Quite clearly, φα (r) = hr|αio is simply the orbital part of the wave-function, and χσα (σ) = hσ|σα i =
δσ,σα is the spin part. 2 The single particle spin-orbitals are often assumed, as we will henceforth,
to form an orthonormal basis of the single particle Hilbert space. 3

Examples of such single particle spin-orbitals might be the plane waves, α = (k, σα ) and φk (r) =
eik·r , the hydrogenoid atomic orbitals, α = (nlm, σα ) and φnlm (r) = R(r)Ylm (Ω̂), the Bloch or the
Wannier states of a crystalline solid, etc. whichever is more appropriate to describe the system
under consideration.

A state for N identical particles may be written down as product of N single particle states

|α1 , α2 , · · · , αN ) = |α1 i · |α2 i · · · |αN i , (13.3)

or, in terms of wave-functions,


Product states
(r1 , r2 , · · · , rN |α1 , α2 , · · · , αN ) = ψα1 (r1 )ψα2 (r2 ) · · · ψαN (rN ) .

However, such product wave-functions are not allowed for systems of identical particles.
1 We will often use the symbol r to mean the combination of r, the position of a particle, and σ, its spin projection
along the z-axis, i.e., whenever you see r, just read r = r, σ.
2 We will often omit the subscripts “o” or “s” for the orbital and spin parts of a state whenever there is no
ambiguity.
3 The requirement of orthonormality might be relaxed, if necessary, at the price of introducing the so-called
overlap matrix.
78 Many particle systems

Consider a system consisting of N = 2 identical particles. Its Hamiltonian will read something
like:
p2 p2
H = 1 + V (r1 ) + 2 + V (r2 ) + Vint (r1 , r2 ) = H1 + H2 + Vint ,
2m 2m
where V (ri ) and Vint (r1 , r2 ) are the most generic external and interaction potentials, which could
depend on the spins of the particles. The fact that the two particles are identical implies that the
masses are the same, m1 = m2 = m, that the form of V (r) is identical for both particles, and that
Vint (r1 , r2 ) = Vint (r2 , r1 ). If we denote by H(1, 2) the Hamiltonian we just wrote, then evidently
H(2, 1) = H(1, 2). Formally, we can introduce the operator P12 that exchanges the two particles,
an operator whose definition on a generic wave-function Ψ(r1 , r2 ) is given by:
def
P12 Ψ(r1 , r2 ) = Ψ(r2 , r1 ) .
2
Quite clearly, the square of P12 is just the identity: P12 = 1. Moreover, using H(2, 1) = H(1, 2),
you can easily convince yourself that:

Exercise 13.1 By applying the definition of P12 on the action of H on a generic wave-function:

P12 H(1, 2) Ψ(r1 , r2 ) = H(2, 1) Ψ(r2 , r1 ) ,

show that P12 commutes with H, i.e.,

P12 H(1, 2) = H(1, 2) P12 .

Now, the possible eigenvalues of P12 are just ±1, since their square should be 1, and therefore, the
corresponding eigenstates are, quite clearly, even or odd under exchange of the two particles:
def
P12 Ψeven (r1 , r2 ) = Ψeven (r2 , r1 ) = Ψeven (r1 , r2 )
def
P12 Ψodd (r1 , r2 ) = Ψodd (r2 , r1 ) = −Ψodd (r1 , r2 ) .

Moreover, since [H, P12 ] = 0 we can diagonalize simultaneously H and P12 , i.e., it is possible to
classify all eigenstates of H as even or odd under exchange of the two particles. Suppose now
that the two particles do not interact, i.e., Vint = 0. Then, if ψa (r) and ψb (r) are eigenstates of
p2 /(2m) + V (r), the product state ψa (r1 )ψb (r2 ) is formally a good candidate eigenstate of the
non-interacting Hamiltonian H1 + H2 , with eigenvalue Ea + Eb , but certainly not an eigenstate of
P12 . How can we construct good eigenstates of P12 , out of ψa and ψb ? Consider the “symmetrized
states”:
1
Ψa,b (r1 , r2 ) = √ [ψa (r1 )ψb (r2 ) ± ψb (r1 )ψa (r2 ] .
2
Evidently, they both correspond to the same “energy” Ea + Eb , and they are good eigenstates
of P12 . Which one is the correct physical eigenstate, depends on the statistics of the particles
under consideration. Physics restricts the possible eigenstates to be totally antisymmetric under
permutations of the particle labels, for fermions, or totally symmetric for bosons (see below).
Therefore, the - sign applies to two fermions, and the + sign is good for two bosons.

Let us extend this concept to a generic number N of identical particles. The permutation sym-
metry that physics dictates is easy to enforce by applying, essentially, a projection operator to the
product states. Consider a permutation P : (1, · · · , N ) → (P1 , · · · , PN ), and define a permutation
operator P to act on product states in the following way:
def
P |α1 , α2 , · · · , αN ) = |αP1 , αP2 , · · · , αPN ) . (13.4)

Define now (−1)P to be the parity of the permutation P . This will appear shortly in constructing
the antisymmetric states for fermions. In order to have a notation common to both fermion and
Many-particle basis states for fermions and bosons. 79
boson cases, we introduce the symbol ξ = −1 for fermions and ξ = +1 for bosons, and form the
correctly symmetrized states 4
Symmetrized
def 1 X P 1 X P
|α1 , α2 , · · · , αN i = √ ξ P |α1 , α2 , · · · , αN ) = √ ξ |αP1 , αP2 , · · · , αPN ) , (13.5) states
N! P N! P

where the sum is performed over the N ! possible permutations of the labels. Indeed, it is very
simple to verity that

Exercise 13.2 The symmetrized states |α1 , α2 , · · · , αN i satisfy the relationship

P|α1 , α2 , · · · , αN i = |αP1 , αP2 , · · · , αPN i = ξ P |α1 , α2 , · · · , αN i . (13.6)

This is the precise formulation of what one means in saying that a state is totally antisymmetric (for
fermions, ξ = −1) or totally symmetric (for bosons, ξ = 1).

Perhaps a more familiar expression is obtained by writing down the corresponding real-space
wave-functions: 5
def
ψα1 ,α2 ,··· ,αN (r1 , r2 , · · · , rN ) = (r1 , r2 , · · · , rN |α1 , α2 , · · · , αN i
1 X P
= √ ξ ψαP1 (r1 )ψαP2 (r2 ) · · · ψαPN (rN ) . (13.7)
N! P

For the case of fermions, this is the so-called Slater determinant


Slater
1 X
ψα1 ,α2 ,··· ,αN (r1 , r2 , · · · , rN ) = √ (−1)P ψαP1 (r1 )ψαP2 (r2 ) · · · ψαPN (rN ) determinant
N! P
 
ψα1 (r1 ) ψα2 (r1 ) · · · ψαN (r1 )
1  ψα1 (r2 ) ψα2 (r2 ) · · · ψαN (r2 ) 
= √ det  .  . (13.8)
 
..
N!  .. ··· . 
ψα1 (rN ) ψα2 (rN ) · · · ψαN (rN )
The corresponding expression for Bosons is called a permanent, but it is much less nicer then the
determinant, computationally.

For Fermions, it is very simple to show that the same label cannot appear twice in the state,
an expression of the Pauli principle. Indeed, suppose α1 = α2 , for instance. Then, by applying Pauli principle
the permutation P : (1, 2, 3, · · · , N ) → (2, 1, 3, · · · , N ) which transposes 1 and 2, we have, on one
hand, no effect whatsoever on the labels
P|α1 , α1 , α3 , · · · , αN i = |αP1 , αP2 , αP3 , · · · , αPN i = |α1 , α1 , α3 , · · · , αN i ,
but, on the other hand, the action of P must result in a minus sign because the state is totally
antisymmetric
P|α1 , α1 , α3 , · · · , αN i = (−)P |α1 , α1 , α3 , · · · , αN i = −|α1 , α1 , α3 , · · · , αN i ,
and a state equal to minus itself is evidently zero. Perhaps more directly, we can see the result from
the expression of the corresponding Slater determinant, which has the first two columns which are
identical, and therefore vanishes, by a known properties of determinants. So, if we define
nα = number of times the label α appears in (α1 , · · · , αN ) , (13.9)
then Pauli principle requires nα ≤ 1. For Bosons, on the contrary, there is no limit to nα .
4 We will consistently use the symbol | · · · ), with round parenthesis, for the product states, and the ket notation
| · · · i for symmetrized states.
5 Observe that we take the scalar product with the product state |r , · · · , r ).
1 N
80 Many particle systems

A few examples should clarify the notation. Consider two particles occupying, in the plane-wave
basis, α1 = k1 , ↑, and α2 = k2 , ↓ or α2 = k2 , ↑ (we will do both spin calculations in one shot). The
correctly symmetrized states are:
1  
(r1 , r2 |k1 ↑, k2 ↓ (↑)i = √ φk1 (r1 )φk2 (r2 )χ↑ (σ1 )χ↓(↑) (σ2 ) ∓ φk2 (r1 )φk1 (r2 )χ↓(↑) (σ1 )χ↑ (σ2 ) ,
2
where the upper sign refers to fermions, and φk (r) = eik·r . Notice that, in general, such a state is
not an eigenvector of the total spin. Exceptions are: i) when both spins are ↑, in which case you
obtain a maximally polarized triplet state:
1
(r1 , r2 |k1 ↑, k2 ↑i = √ [φk1 (r1 )φk2 (r2 ) ∓ φk2 (r1 )φk1 (r2 )] χ↑ (σ1 )χ↑ (σ2 ) , (13.10)
2
or when the two orbital labels coincide. In general, good eigenstates of the total spin are ob-
tained only by linear combination of more than one symmetrized state. For the example above, for
instance, it is simple to show that
1
√ [(r1 , r2 |k1 ↑, k2 ↓i ± (r1 , r2 |k1 ↓, k2 ↑i] (13.11)
2
is a triplet (upper sign) or a singlet (lower sign). The triplet (S = 1) and singlet (S = 0) spin states
are the result of the addition of two s = 1/2 spin states for the two particles, with a procedure
that is virtually identical (but much simpler) to that used to construct j = 3/2 and j = 1/2 states
out of l = 1 and s = 1/2. This is just another elementary example of the general addition of two
angular momenta.

As a further example, consider three identical particles occupying three states ψa , ψb and ψc .
The resulting physically allowed state is:
1
Ψa,b,c (r1 , r2 , r3 ) = √ [ + ψa (r1 )ψb (r2 )ψc (r3 ) + ψb (r1 )ψc (r2 )ψa (r3 ) + ψc (r1 )ψa (r2 )ψb (r3 )
6
± ψb (r1 )ψa (r2 )ψc (r3 ) ± ψc (r1 )ψb (r2 )ψa (r3 ) ± ψa (r1 )ψc (r2 )ψb (r3 )] ,

where the upper (+) sign applies to three bosons, the lower (-) sign to three fermions.

The normalization, or more generally the scalar product, of symmetrized states involves a bit
Normalization of permutation algebra.

Exercise 13.3 Show that the scalar product between two correctly-symmetrized states |α1 , α2 , · · · , αN i
and |α10 , α20 , · · · , αN
0
i is non-zero only if the set of labels (α10 , α20 , · · · , αN 0
) is a permutation of (α1 , α2 , · · · , αN ).
Verify then that: Y
hα10 , α20 , · · · , αN
0
|α1 , α2 , · · · , αN i = ξ P nα !
α

where P is a permutation that makes the labels to coincide, and nα is the number of times a certain label
α appears.

As a consequence, fermionic symmetrized states (Slater determinants) are normalized to 1, since


nα ≤ 1, while bosonic ones not necessarily. One can easily normalize bosonic states by defining
1
|{nα }i = pQ |α1 , · · · , αN i (13.12)
α nα !

where we simply label the normalized states by the occupation number nα of each label, since the
order in which the labels appear does not matter for Bosons.

We conclude by stressing the fact that the symmetrized states constructed are simply a basis set
for the many-particle Hilbert space, in terms of which one can expand any N -particle state |Ψi
General |Ψi
Two-electrons: the example of Helium. 81
X
|Ψi = Cα1 ,α2 ,··· ,αN |α1 , α2 , · · · , αN i , (13.13)
α1 ,α2 ,··· ,αN

with appropriate coefficients Cα1 ,α2 ,··· ,αN . The whole difficulty of interacting many-body systems is
that the relevant low-lying excited states of the systems often involve a large (in principle infinite)
number of symmetrized states in the expansion.

13.2 Two-electrons: the example of Helium.

Suppose we would like to tackle the He atom problem, which has N = 2 electrons. More generally,
we can address a two-electron ion like Li+ , Be2+ , etc. with a generic nuclear charge Z ≥ 2. The
Hamiltonian for an atom with N electrons is evidently:

N  2
Ze2 e2

X p i 1X
H= − + , (13.14)
i=1
2m ri 2 |ri − rj |
i6=j

where we have assumed the N interacting electrons to be in the Coulomb field of a nucleus of charge
Ze, which we imagine fixed at R = 0. We have neglected spin-orbit effects and other relativistic
corrections. We plan to treat here the N = 2 case. Evidently, a good starting basis set of single-
particle functions is that of hydrogenoid orbitals which are solutions of the one-body Hamiltonian
ĥ = p2 /2m − Ze2 /r. We know that such functions are labeled by orbital quantum numbers (nlm),
and that
 2
Ze2

p
− φnlm (r) = n φnlm (r) . (13.15)
2m r
The eigenvalues n are those of the Coulomb potential, and therefore independent of the angular
momentum quantum numbers lm:
Z 2 e2 1
n = − , (13.16)
2 aB n2
where aB = ~2 /(me2 ) ≈ 0.529rA is the Bohr radius, and e2 /aB = EHartree ≈ 27.2 eV. The
wave-functions φnlm (r) are well known. For instance, the 1s orbital (n = 1, l = 0, m = 0) is:
 3/2
1 Z
φ1s (r) = √ e−Zr/aB .
π aB

It is very convenient here two adopt atomic units, where lengths are measured in units of the Bohr
length aB , and energies in units of the Hartree, EH = e2 /aB .

If we start ignoring the Coulomb potential, we would put our N = 2 electrons in the lowest two
orbitals available, i.e., 1s ↑ and 1s ↓, forming the Slater determinant:

1
Ψ1s↑,1s↓ (r1 , r2 ) = φ1s (r1 )φ1s (r2 ) √ [χ↑ (σ1 )χ↓ (σ2 ) − χ↓ (σ1 )χ↑ (σ2 )] ,
2

i.e., the product of a symmetric orbital times a singlet (antisymmetric) spin state. The one-body
part of the Hamiltonian contributes an energy

E (one−body) = h1s ↑, 1s ↓ |ĥ|1s ↑, 1s ↓i = 21s = −Z 2 (a.u.) . (13.17)

For He, where Z = 2, this is −4 (a.u.) a severe underestimate of the total energy, which is known
to be E (exact,He) ≈ −2.904 (a.u.). This can be immediately cured by including the average of the
Coulomb potential part, which is simply the direct Coulomb integral K1s,1s of the 1s orbital, often Coulomb
integral
82 Many particle systems

indicated by U1s in the strongly correlated community,

e2
Z
def
h1s ↑, 1s ↓ |V̂ |1s ↑, 1s ↓i = dr1 dr2 |φ1s (r2 )|2 |φ1s (r1 )|2 = K1s,1s = U1s . (13.18)
|r1 − r2 |

The calculation of Coulomb integrals can be performed analytically by exploiting rotational invari-
ance. We simply state here the result:

5
U1s = Z (a.u.) . (13.19)
8
Summing up, we obtain the average energy of our Slater determinant as
 
2 5
E1s↑,1s↓ = −Z + Z (a.u.) = −2.75 (a.u.) , (13.20)
8

where the last number applies to He (Z = 2).

To improve over this result, let us use the variational method, which states that the true ground
state energy is the minimum of the average energy over all allowed states:

hΨ|H|Ψi
Eg.s. = Min|Ψi . (13.21)
hΨ|Ψi

The validity of this statement can be easily proven by expanding |Ψi in the basis of eigenfunctions
of H, and calculating the resulting average energy. We will write a guess for the wave-function
made by choosing the same orbital wave-function for the two opposite spin states

ψαo ↑ (r) = φαo (r)χ↑ (σ) =⇒ ψαo ↓ (r) = φαo (r)χ↓ (σ) . (13.22)

What is the best possible choice for such an orbital φαo (r)? The proper answer to such a question
requires going through the Hartree (or, more generally, Hartree-Fock) mean-field approach. We
will simply assume here this Hartree approach as very reasonable, without deriving the optimality
condition. So, going back to our He exercises, we will put two electrons, one with ↑ spin, the other
with ↓ spin, in the lowest orbital solution of the Hartree equation

~2 ∇2 Ze2
 
− − + Vdir (r) φ1s (r) = ˜1s φ1s (r) , (13.23)
2m r

where now the 1s label denotes the orbital quantum numbers of the lowest energy solution of a
rotationally invariant potential Vdir (r). The form of the direct (Hartree) potential Vdir (r) is:

e2
Z
Vdir (r) = dr0 |φ1s (r0 )|2 , (13.24)
|r − r0 |

because the electron in φ1s feels the repulsion of the other electron in the same orbital. We im-
mediately understand that the Hartree potential partially screens at large distances the nuclear
charge +Ze. Indeed, for r → ∞ we can easily see that

e2 e2
Z
Vdir (r) ∼ dr0 |φ1s (r0 )|2 = , (13.25)
r r

so that the total effective charge seen by the electron at large distances is just (Z − 1)e, and not
Ze. This is intuitively obvious. The fact that we chose our original φ1s orbital as solution of the
hydrogenoid problem with nuclear charge Ze gives to that orbital a wrong tail at large distances.
This is, primarily, what a self-consistent mean-field calculation needs to adjust to get a better
form of φ1s . Indeed, there is a very simple variational scheme that we can follow instead of solving
Two-electrons: the example of Helium. 83
(Z )
Eq. (13.23). Simply consider the hydrogenoid orbital φ1s eff which is solution of the Coulomb
potential with an effective nuclear charge Zeff e (with Zeff a real number),
 2
Zeff e2

p (Z ) (Z )
− φ1s eff (r) = ˜1s φ1s eff (r) , (13.26)
2m r
(Z )
and form the Slater determinant |1s ↑, 1s ↓; Zeff i occupying twice φ1s eff , with Zeff used as a single
variational parameter. The calculation of the average energy E1s↑,1s↓ (Zeff ) is quite straightforward,
6
and gives:

2 5
E1s↑,1s↓ (Zeff ) = h1s ↑, 1s ↓; Zeff |Ĥ|1s ↑, 1s ↓; Zeff i = Zeff − 2ZZeff + Zeff . (13.27)
8
opt
Minimizing with respect to Zeff this quadratic expression we find Zeff = Z − 5/16 = 1.6875 and
opt
E1s↑,1s↓ (Zeff ) = −(Z − 5/16)2 (a.u.) ≈ −2.848 (a.u.) ,

where the numbers are appropriate to Helium. This value of the energy in not too far form the
exact Hartree-Fock solution, which would give

EHF ≈ −2.862 (a.u.) (for Helium).

In turn, the full Hartree-Fock solution differs very little from the exact non-relativistic energy
of Helium, calculated by accurate quantum chemistry methods (Configuration Interaction). A
quantity that measures how far the Hartree-Fock mean-field energy differs from the exact answer
is the so-called correlation energy, defined simply as:
def
Ecorr = E exact − EHF , (13.28)

which, for Helium amounts to Ecorr ≈ −0.042 (a.u.), a bit more than 1% of the total energy.

6 You (Z )
simply need to calculate the average of p2 /2m and of 1/r over φ1s eff .
Appendix A
Second quantization: brief outline.

It is quite evident that the only important information contained in the symmetrized states
|α1 , α2 , · · · , αN i is how many times every label α is present, which we will indicate by nα ; the
rest is automatic permutation algebra. The introduction of creation operators a†α is a device by
which we take care of the labels present in a state and of the permutation algebra in a direct way.

Given any single-particle orthonormal basis set {|αi} we define operators a†α which satisfy the
Creation following rule: 1
operators def
a†α |0i = |αi
def . (A.1)
a†α |α1 , α2 , · · · , αN i = |α, α1 , α2 , · · · , αN i

The first is also the definition of a new state |0i, called vacuum or state with no particles, which
The vacuum should not be confused with the zero of a Hilbert space: in fact, we postulate h0|0i = 1. It is formally
required in order to be able to obtain the original single-particle states |αi by applying an operator
that creates a particle with the label α to something: that “something” has no particles, and
obviously no labels whatsoever. The second equation defines the action of the creation operator a†α
on a generic correctly-symmetrized state. Notice immediately that, as defined, a†α does two things
in one shot: 1) it creates a new label α in the state, 2) it performs the appropriate permutation
algebra in such a way that the resulting state is a correctly-symmetrized state. Iterating the creation
rule starting from the vacuum |0i, it is immediate to show that

a†α1 a†α2 · · · a†αN |0i = |α1 , α2 , · · · , αN i . (A.2)

We can now ask ourselves what commutation properties must the operators a†α satisfy in order to
enforce the correct permutation properties of the resulting states. This is very simple. Since

|α2 , α1 , · · · , αN i = ξ|α1 , α2 , · · · , αN i

for every possible state and for every possible choice of α1 and α2 , it must follow that

a†α2 a†α1 = ξa†α1 a†α2 , (A.3)

i.e., creation operators anti-commute for Fermions, commute for Bosons. Explicitly:

{a†α1 , a†α2 } = 0 for Fermions (A.4)


 †
aα1 , a†α2 = 0

for Bosons , (A.5)

with {A, B} = AB + BA (the anti-commutator) and [A, B] = AB − BA (the commutator).

Destruction The rules for a†α clearly fix completely the rules of action of its adjoint aα = (a†α )† , since it
operators
1 It might seem strange that one defines directly the adjoint of an operator, instead of defining the operator a
α
itself. The reason is that the action of a†α is simpler to write.
Second quantization: brief outline. 85
must satisfy the obvious relationship

hΨ2 |aα Ψ1 i = ha†α Ψ2 |Ψ1 i ∀Ψ1 , Ψ2 , (A.6)

where Ψ1 , Ψ2 are correctly-symmetrized many-particle basis states. First of all, by taking the
adjoint of the Eqs. (A.4), it follows that

{aα1 , aα2 } = 0 for Fermions (A.7)


[aα1 , aα2 ] = 0 for Bosons . (A.8)

There are a few simple properties of aα that one can show by using the rules given so far. For
instance,
aα |0i ∀α , (A.9)

since hΨ2 |aα |0i = ha†α Ψ2 |0i = 0, ∀Ψ2 , because of the mismatch in the number of particles. 2 More
generally, it is simple to prove that that an attempt at destroying label α, by application of aα ,
gives zero if α is not present in the state labels,

aα |α1 , α2 , · · · , αN i = 0 if α ∈
/ (α1 , α2 , · · · , αN ) . (A.10)

Consider now the case α ∈ (α1 , · · · , αN ). Let us start with Fermions. Suppose α is just in first
position, α = α1 . Then, it is simple to show that

aα1 |α1 , α2 , · · · , αN i = |α̂1 , α2 , · · · , αN i = |α2 , · · · , αN i , (A.11)

where by α̂1 we simply mean a state with the label α1 missing (clearly, it is a N − 1 particle state).
To convince yourself that this is the correct result, simply take the scalar product of both sides
of Eq. (A.11) with a generic N − 1 particle state |α20 , · · · , αN
0
i, and use the adjoint rule. The left
hand side gives:

hα20 , · · · , αN
0
|aα1 |α1 , α2 , · · · , αN i = hα1 , α20 , · · · , αN
0
|α1 , α2 , · · · , αN i ,

which is equal to (−1)P when P is a permutation bringing (α2 , · · · , αN ) into (α20 , · · · , αN


0
), and 0
otherwise. The right hand side gives

hα20 , · · · , αN
0
|α2 , · · · , αN i ,

and coincides evidently with the result just stated for the left hand side. If α is not in first position,
say α = αi , then you need to first apply a permutation to the state that brings αi in first position,
and proceed with the result just derived. The permutation brings and extra factor (−1)i−1 , since
you need i − 1 transpositions. As a result:

aαi |α1 , α2 , · · · , αN i = (−1)i−1 |α1 , α2 , · · · , α̂i , · · · , αN i . (A.12)

The bosonic case needs a bit of extra care for the fact that the label α might be present more than
once.

Exercise A.1 Show that for the bosonic case the action of the destruction operator is

aαi |α1 , α2 , · · · , αN i = nαi |α1 , α2 , · · · , α̂i , · · · , αN i . (A.13)

Armed with these results we can finally prove the most difficult commutation relations, those
involving a aα with a a†α0 . Consider first the case α 6= α0 and do the following.
2A state which has vanishing scalar product with any state of a Hilbert space, must be the zero.
86 Second quantization: brief outline.

Exercise A.2 Evaluating the action of aα a†α0 and of a†α0 aα on a generic state |α1 , · · · , αN i, show that

{aα , a†α0 } = 0 for Fermions (A.14)


h i
aα , a†α0 = 0 for Bosons . (A.15)

Next, let us consider the case α = α0 . For fermions, if α ∈


/ (α1 , · · · , αN ) then
aα a†α |α1 , · · · , αN i = aα |α, α1 , · · · , αN i = |α1 , · · · , αN i ,
while
a†α aα |α1 , · · · , αN i = 0 .
If, on the other hand, α ∈ (α1 , · · · , αN ), say α = αi , then
aαi a†αi |α1 , · · · , αN i = 0 ,
because Pauli principle forbids occupying twice the same label, while
a†αi aαi |α1 , · · · , αN i = a†αi (−1)i−1 |α1 , · · · , α̂i , · · · , αN i = |α1 , · · · , αN i ,
because the (−1)i−1 is reabsorbed in bringing the created particle to the original position. Sum-
marizing, we see that for Fermions, in all possible cases
aα a†α + a†α aα = 1 (A.16)

Exercise A.3 Repeat the algebra for the Bosonic case to show that

aα a†α − a†α aα = 1 (A.17)

We can finally summarize all the commutation relations derived, often referred to as the canonical
commutation relations.
Canonical
{a†α1 , a†α2 } = 0
 †
aα1 , a†α2 = 0

commutators
Fermions =⇒ {aα1 , aα2 } = 0 Bosons =⇒ [a  α1 , aα†2 ]= 0 . (A.18)
{aα1 , a†α2 } = δα1 ,α2 aα1 , aα2 = δα1 ,α2

Before leaving the section, it is worth pointing out the special role played by the operator
def
n̂α = a†α aα (A.19)
Number often called the number operator, because it simply counts how many times a label α is present
operator in a state.

Exercise A.4 Verify that


n̂α |α1 , · · · , αN i = nα |α1 , · · · , αN i; , (A.20)
where nα is the number of times the label α is present in (α1 , · · · , αN ).

Clearly, one can write an operator N̂ that counts the total number of particles in a state by
def
X X
N̂ = n̂α = a†α aα . (A.21)
α α
Second quantization: brief outline. 87
A.0.1 Changing the basis set.

Suppose we want to switch from |αi to some other basis set |ii, still orthonormal. Clearly there is
a unitary transformation between the two single-particle basis sets:
X X
|ii = |αihα|ii = |αi Uα,i , (A.22)
α α

where Uα,i = hα|ii is the unitary matrix of the transformation. The question is: How is a†i de-
termined in terms of the original a†α ? The answer is easy. Since, by definition, |ii = a†i |0i and
|αi = a†α |0i, it immediately follows that

a†i |0i =
X
a†α |0i Uα,i . (A.23)
α

By linearity, one can easily show that this equation has to hold not only when applied to the
vacuum, but also as an operator identity, i.e.,

a†i = P α a†α Uα,i


P
∗ , (A.24)
ai = α aα Uα,i

the second equation being simply the adjoint of the first. The previous argument is a convenient
mnemonic rule for re-deriving, when needed, the correct relations.

A.0.2 The field operators.

The construction of the field operators can be seen as a special case of Eqs. (A.24), when we take
as new basis the coordinate and spin eigenstates |ii = |r, σi. By definition, the field operator Ψ†σ (r)
is the creation operator of the state |r, σi, i.e.,

Ψ†σ (r)|0i = |r, σi . (A.25)

Then, the analog of Eq. (A.22) reads:


X X
|r, σi = |αihα|r, σi = |αiδσ,σα φ∗α (r) , (A.26)
α α

where we have identified the real-space wave-function of orbital α as φα (r) = hr|αio , and used the
fact that hσα |σi = δσ,σα . Assume that the quantum numbers α are given by α = (αo , σ), where αo
denote the orbital quantum numbers. The analog of Eqs. (A.24) reads, then,
X
Ψ†σ (r) = φ∗αo ,σ (r) a†αo ,σ (A.27)
αo
X
Ψσ (r) = φαo ,σ (r) aαo ,σ .
αo

These relationships can be easily inverted to give:


Z
a†αo ,σ = dr φαo ,σ (r) Ψ†σ (r) (A.28)
Z
aαo ,σ = dr φ∗αo ,σ (r) Ψσ (r) .
88 Second quantization: brief outline.

A.0.3 Operators in second quantization.

We would like to be able to calculate matrix elements of a Hamiltonian like, for instance, that of
N interacting electrons in some external potential v(r),
N  2
e2

X p i 1X
H= + v(ri ) + . (A.29)
i=1
2m 2 |ri − rj |
i6=j

In order to do so, we need to express the operators appearing in H in terms of the creation and
destruction operators a†α and aα of the selected basis, i.e., as operators in the so-called Fock space.
Observe that there are two possible types of operators P of 2interest to us: P
1) one-body operators, like the total kinetic energy i pi /2m or the external potential i v(ri ),
One-body which act on one-particle at a time, and their effect is then summed over all particles in a totally
operators symmetric way, generally
X N
1−body
UN = U (i) ; (A.30)
i=1

2) two-body operators, like the Coulomb interaction between electrons (1/2) i6=j e2 /|ri −rj |, which
P
involve two-particle at a time, and are summed over all pairs of particles in a totally symmetric
Two-body way,
operators N
1X
VN2−body = V (i, j) . (A.31)
2
i6=j

The Fock (second quantized) versions of these operators are very simple to state. For a one-body
operator:
hα0 |U |αia†α0 aα ,
1−body
X
UN =⇒ UFock = (A.32)
α,α0

where hα0 |U |αi is simply the single-particle matrix element of the individual operator U (i), for
instance, in the examples above,
 2 2
~ ∇
Z
0 2 ∗
hα |p /2m|αi = δσα ,σα0 dr φα0 (r) − φα (r) (A.33)
2m
Z
hα0 |v(r)|αi = δσα ,σα0 dr φ∗α0 (r)v(r)φα (r) .

For a two-body operator:

1
(α20 α10 |V |α1 α2 ) a†α0 a†α0 aα1 aα2 ,
X
VN2−body =⇒ VFock = (A.34)
2 2 1
α1 ,α2 ,α01 ,α02

where the matrix element needed is, for a general spin-independent interaction potential V (r1 , r2 ),
Z
0 0
(α2 α1 |V |α1 α2 ) = δσα1 ,σα0 δσα2 ,σα0 dr1 dr2 φ∗α02 (r2 )φ∗α01 (r1 )V (r1 , r2 )φα1 (r1 )φα2 (r2 ) . (A.35)
1 2

We observe that the order of the operators is extremely important (for fermions).

The proofs are not very difficult but a bit long and tedious. We will briefly sketch that for the
one-body case.

A.1 Why a quadratic Hamiltonian is easy.

Before we consider the Hartree-Fock problem, let us pause for a moment and consider the reason
why one-body problems are considered simple in a many-body framework. If the Hamiltonian is
Why a quadratic Hamiltonian is easy. 89
PN
simply a sum of one-body terms H = i=1 h(i), for instance h(i) = p2i /2m + v(ri ), we know that
in second quantization it reads
hα0 ,α a†α0 aα ,
X
H= (A.36)
α,α0

where the matrix elements are


 2 2 
~ ∇
Z
hα0 ,α = hα0 |h|αi = δσα ,σα0 dr φ∗α0 (r) − + v(r) φα (r) . (A.37)
2m

So, H is purely quadratic in the operators. The crucial point is now that any quadratic problem can
be diagonalized completely, by switching to a new basis |ii made of solutions of the one-particle
Schrödinger equation 3
 2 2 
~ ∇
h|ii = i |ii =⇒ − + v(r) φi (r) = i φi (r) . (A.38)
2m

Working with this diagonalizing basis, and the corresponding a†i , the Hamiltonian simply reads:
X † X
H= i ai ai = i n̂i , (A.39)
i i

where we assume having ordered the energies as 1 ≤ 2 ≤ · · · . With H written in this way, we
can immediately write down all possible many-body exact eigenstates as single Slater determinants
(for Fermions) and the corresponding eigenvalues as sums of i ’s,

|Ψi1 ,··· ,iN i = a†i1 · · · a†iN |0i


Ei1 ,··· ,iN = i1 + · · · + iN . (A.40)

So, the full solution of the many-body problem comes automatically from the solution of the
corresponding one-body problem, and the exact many-particle states are simply single Slater de-
terminants.

3 Being simple does not mean that solving such a one-body problem is trivial. Indeed it can be technically quite
involved. Band theory is devoted exactly to this problem.

Você também pode gostar