You are on page 1of 84

Lecture notes on Quantum Physics I (Phys.

591)
Kirill Tuchin1
1

Department of Physics and Astronomy, Iowa State University, Ames, IA 50011


(Dated: November 5, 2014)

Contents

I. Basic principles of Quantum Mechanics


1. The uncertainty principle
2. Wave function
3. Wave function of a free particle
A. Wave packet
B. Time evolution of wave packet
C. Particle in a box
4. Operators
A. Expectation values of coordinate and momentum
B. Hermitian operators
C. Examples
5. Eigenfunctions and eigenvalues of operators
A. Properties of discrete spectrum
B. Properties of continuous spectrum
6. Examples of eigenvalue problems

A. Operator of momentum p.
B. Operator of position r.

C. Operator of orbital angular momentum L.


7. Description of quantum states
8. Heisenberg relations
A. Example
9. Classical limit of quantum mechanics

3
3
4
4
4
6
6
7
7
8
9
10
11
13
14
14
14
14
16
16
17
18

II. Schr
odinger equation
1. Hamiltonian
2. Stationary states
3. Time derivative of operators
4. Schr
odinger and Heisenberg pictures of time-evolution
A. Schr
odinger picture
B. Heisenberg picture
C. Interaction picture
5. Symmetries and conserved quantities
A. Translations
B. Rotations
C. Time evolution
D. Space inversion

19
19
20
22
24
24
25
26
26
27
27
29
29

III. Motion in one-dimension


1. Delta-potential
A. Continuity of the wave function
B. Discrete spectrum
C. Continuous spectrum
2. Rectangular potential well
A. Discrete spectrum
B. Three-dimensional well
C. Continuous spectrum

30
30
30
30
31
32
32
34
34

Kirill Tuchin

Phys. 591 Lecture Notes

3. Infinite one-dimensional crystal


4. Harmonic oscillator
A. Energy spectrum
B. Ladder operators
5. General properties of one-dimensional motion
6. Integral form of Schr
odinger equation
A. Discrete spectrum
B. Continuous spectrum
IV. Representation theory
1. Representations of states
A. Hilbert space
B. Discrete (energy) representation
C. Continuous (momentum) representation
D. General case
2. Representations of operators
3. Eigenvalue problem in matrix form
4. Unitary operators
5. Schr
odinger equation in momentum representation
6. Occupation number representation of harmonic oscillator
V. Motion in central potential
1. General properties of motion in central potential
2. Spherical waves
3. Spherical potential well
4. Spherical harmonic oscillator
5. Coulomb potential
A. Discrete spectrum
B. Continuous spectrum
6. Effective electric potential of the hydrogen atom
7. Origin of degeneracy of the energy spectrum
VI. Angular momentum
1. Angular momentum operator
2. Spin
3. Addition of angular momenta
4. Matrix elements of vector operators
VII.

2
35
37
37
39
41
43
43
44
46
46
46
46
47
47
48
51
52
54
56
59
59
61
62
63
64
64
66
67
68
69
69
72
75
77
80

A. Fourier analysis

81

B. Dirac delta function

81

C. Levi-Civita symbol

82

Kirill Tuchin

Phys. 591 Lecture Notes


I.

BASIC PRINCIPLES OF QUANTUM MECHANICS


1.

The uncertainty principle

Quantum mechanics considers microscopic systems with length scales . 106 cm. Physical laws that govern such
systems are very different form those that govern macroscopic systems, which are described by the classical mechanics.
Moreover, the microscopic phenomena can be observed only by means of a macroscopic apparatus that translates the
action of microscopic objects into the macroscopic language (examples: Geiger counter, bubble chamber). Quantum
mechanics is a coherent mathematical framework that describes the laws of microscopic systems and translates them
into the classical language. The ultimate success of quantum mechanics is evident in a great number of experiments.
However, its interpretation is still being debated. The goal of this course is to develop the mathematical framework
of quantum mechanics and to illustrate how it can be applied to number of important problems.
Historically, the first indications that the classical theory is not adequate for description of microscopic systems
came in the early 20th century. It was experimentally established that electromagnetic radiation possesses both wave
and corpuscular character. In particular, it is absorbed and emitted in separate portions, quanta, which we now call
photons. Photon energy E turned out to be proportional to its frequency as
E = ~ ,

(1.1)

where ~ = 1.054 1027 ergsec is Plancks constant. In free space photons move with the velocity of light c and have
the momentum
p = ~k ,

(1.2)

where k is the wave-vector with the length k = |k| = 2/ = 1/. We know from the electromagnetic theory that
= kc. It follows that p = ~/c = E/c, which is the relation between the energy and momentum of a relativistic
particle of zero mass.
Similarly to the electromagnetic radiation, matter particles also posses both wave and corpuscular character. This
was most clearly seen in the electron diffraction experiment where a homogeneous beam of electrons pass through a
crystal. The emergent beam exhibits a pattern of alternate maxima and minima of intensity similar to the diffraction
of electromagnetic waves. In the two-slit experiment one considers two screens: one impermeable to electrons in which
to slits are cut and behind it another continuous screen. When electrons pass through the first slit, while the second
one is covered, one observes a certain distribution of intensity on the continuous screen. A different distribution is
obtained when the first slit is opened and the second is covered. However, the intensity distribution that is obtained
when both slits are open is different from a simple sum of the two intensity distribution in contrast to expectation from
the classical physics. One consequence of this observation is that electrons have no trajectory. Using this diffractive
experiments one can assign to a free particle a wave length that is uniquely determined by its momentum
k=

p
,
~

k=

2
,

(1.3)

in analogy with the photon. In this context is called the de Broglie wavelength.
An important consequence of the electron diffraction experiment is that the classical notion of trajectory does not
apply to microscopic objects. Had electron beams moved along a certain trajectory they would have not interfered.
This absence of a trajectory is the essence of the uncertainty principle. Later in this course we will give a mathematical
formulation of this fundamental principle. Suppose now that we measure the electron position with an apparatus
(a classical devise). The more accurate is the measurement, the stronger apparatus affects the electron, so that
consecutive measurements would give a discontinuous and disorderly results indicating absence of trajectory. Only at
very low accuracy (e.g. in bubble chamber), the results can be approximated by a smooth curve classical electron
trajectory. As a result, the classical definition of velocity as a time derivative of the particle position is meaningless
in quantum theory. Later we will construct a reasonable definition of velocity in quantum mechanics that has correct
classical limit.
In classical mechanics a full description of a system is achieved by providing the coordinates and velocities of all
particles at a given time. The uncertainty principle implies that such description is impossible in quantum mechanics.
Quantum description is less detailed than the classical one, and its predictions are less certain. For instance, given
an electron in a certain initial state, a subsequent measurement can yield various results. The problem in quantum
mechanics consists in determining the probability of obtaining of these different results in a measurement.
Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, 1, Merzbacher, Quantum Mechanics 3rd edition, Ch. 1.

Kirill Tuchin

Phys. 591 Lecture Notes


2.

Wave function

The properties of microscopic systems are described in quantum mechanics by means of an auxiliary quantity the
wave function . The wave function is a complex continuous function of coordinates and time and determines the
probabilities of a different outcomes of a measurement. This is known as the statistical or Copenhagen interpretation
of the wave function. Let q denote the set of coordinates of a quantum system (often referred to as the configuration
space) and dq the product of differentials of these coordinates (i.e. element of volume of the configuration space).
For example, for one particle q = (x, y, z) and dq = dx dy dz, for two particles q = (x1 , y1 , z1 , x2 , y2 , z2 ) and dq =
dx1 dy1 dz1 dx2 dy2 dz2 , etc. According to the statistical interpretation, quantity
|(q)|2 dq = (q) (q)dq

(1.4)

is proportional to the probability that we find as the result of our measurement that values of coordinates lie within
the interval [q, q + dq]. The sum of the probabilities of all possible value of the coordinates of the system must be
equal to unity implying that
Z
|(q)|2 dq = 1 .
(1.5)
This equation is called the normalization condition. If a wave function is normalized according to (1.5), then = |(q)|2
is known in mathematics as the probability density. In some idealized systems integral over the configuration space
does not converge, viz.
Z
|(q)|2 dq = .
(1.6)
In this case |(q)|2 cannot be interpreted as the probability density. Instead, it is meaningful to talk about the relative
probabilities at different qs.
Wave function satisfies the following fundamental superposition principle: if a system can be in states described
by wave functions 1 and 2 , it can also be in all states described by the linear combination a1 1 + a2 2 , for any
time-independent complex numbers a1 and a2 . In particular, in view of the normalization condition, wave functions
and a, with non-zero a, describe the same state. Note, that (1.5) requires a = ei with any real . This
ambiguity does not affect any of the physical results, because does not depend on it. Note that the superposition
principle in quantum mechanics is essentially different from the superposition of vibrations in classical physics, where
a superposition of a vibration onto itself gives a new vibration with a larger or smaller amplitude.
Apart of its dependence on coordinates, the wave function also depends on time. Its time-dependence, or evolution,
is determined by a differential equation that we will discuss in the next chapter. That equation must be linear so that
its solutions satisfy the superposition principle.
Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, 2, Merzbacher, Quantum Mechanics 3rd edition, Ch. 1.
3.

Wave function of a free particle


A.

Wave packet

It is instructive to consider a simple example before we start developing mathematical framework of quantum
mechanics. Consider a free non-relativistic particle of mass m. Its energy E = p2 /2m and momentum p are conserved,
which reflects uniformity of space and time (will discuss this in more detail later). Therefore, its wave function must
depend on E and p. According to our discussion in I 1 this particle behaves in diffractive experiments as a wave with
wave-vector k = p/~ and energy = E/~. In optics, a wave with given k and is a plane wave
(r, t) = N ei(krt) ,

(1.7)

where N is a constant. Therefore, it is reasonable to assume that (1.7) describes motion of a free particle. This
argument is apparently an educated guess, which cuts many corners, but leads to the right answer as will be shown
later. A state of free particle of mass m, momentum p0 and energy E 0 is described by
0

0 (r, t) = N ei(k r t) .

(1.8)

Kirill Tuchin

Phys. 591 Lecture Notes

According to the superposition principle = a + a0 0 also describes a possible state of a particle. In this state
however, the particle does not have a definite momentum.
Generally, one can consider a linear combination of any number of states of form (1.7) with different momenta.
Suppose for simplicity that particle moves along the z-axis. Then its state is described by the following wave function
Z +
(z, t) =
ak ei(kzt) dk
(1.9)

with an arbitrary a(k). In practice, particle momentum, say k0 , is known with a certain accuracy k, where k  k0
(otherwise we cannot say that particle momentum is known). In this case, (1.9) is called the wave packet and can be
written down as
Z k0 +k
ak ei(kzt) dk .
(1.10)
(z, t) =
k0 k

To estimate this integral, introduce a new variable l = k k0 , and expand as a function of k near k = k0 :
 
d
l,
(k) 0 +
dk 0

(1.11)

where 0 = (k0 ). In this approximation integration in (1.10) produces


(z, t) =

+k

ak0 ei(l+k0 )zi0 ti( dk )0 lt dl


d

Z +
d
= ak0 ei(k0 z0 t)
ei[z( dk )0 t]l dl


 

sin z d
dk 0 t k
i(k0 z0 t)

= 2ak0
|e {z }
z d
dk 0 t
|
{z
} fast-oscillating

(1.12)

2
.
k

(1.13)

amplitude

The wave packet is a product of two factors: slowly changing amplitude and rapidly oscillating plane wave. At
t = 0 the wave packet has global maximum at z = 0 and then decreases as 1/z while oscillating with period 2/k.
Zeros of the amplitude are located at zn = n/k, where n = 1, 2, . . .. The spatial extent of the packet can be
estimated as a distance between the two zeros around the maximum at z = 0:
z =
In view of (1.3) we conclude that
zp = 2~ .

(1.14)

In other words, uncertainty in the particle positions is inversely proportional to the uncertainty of its momentum.
This is consistent with the uncertainty principle that a microscopic particle does not have a trajectory.
According to (1.12) the maximum of the wave packet moves along the z-axis with group velocity
 
d
vg =
.
(1.15)
dk 0
Since = p2 /2m~ = ~k 2 /2m we find
vg =

p0
,
m

(1.16)

as it should be for a free particle. Note, that the phase velocity vp = /k = E/p = p/2m is not velocity of particle.

Kirill Tuchin

Phys. 591 Lecture Notes


B.

Time evolution of wave packet

Consider a wave packet that at t = 0 is given by


2

(z, 0) = N e

z
2
2

(1.17)

Here is a parameter that specifies the width of the Gaussian (1.17). Using (1.5) we find the normalization constant
(z, 0) =

e
1/4

z
2
2

(1.18)

Using (1.9) we can write


Z

ak eikz dk =

1/4

z
2
2

(1.19)

Inverting the Fourier transform (see App. A) we get


Z +

z2
k 2 2

dz ikz
1
2
2

e
e 2 .
ak =
e
=

1/4
3/4

2
2
Employing (1.9) again and recalling that = ~k 2 /2m we have
Z +

z2
k 2 2

p
ei(kzt) e 2 dk =
(z, t) =
e 2(2 +i~t/m) .
2 3/4
1/4 + i~t/m

(1.20)

(1.21)

The probability density reads

z2
1

|(z, t)|2 = p
e 2 +~2 t2 /m2 2 .
m2 4 + ~2 t2

(1.22)

This shows that the width of the wave packet grows as a function of time, so that the particle position becomes less
certain at later time. This can be easily understood as follows. At t = 0 the width of the packet is z and
according to (1.14) p ~/. Components of the wave packet with different k move with different velocities, the
uncertainty in velocity being v = p/m ~/m. Thus, the packet width grows as z(t) ~t/m in agreement
with (1.22). This effect is important at t > t0 = m2 /~, whereas at t  t0 the diffusion of the wave packet can
be neglected. To put this into perspective, take = 108 cm as a typical atomic scale. Then for electron we get
t0 = 1016 sec. In contrast, for m = 1 gr, t0 = 104 years.
C.

Particle in a box

Wave function of a free particle (1.7) is an example of a wave function that cannot be normalized with condition
(1.5), but rather its normalization integral is divergent as in (1.6). This divergence appears because the notion of free
particle is an idealization. In reality, we can only talk about a particle that can be considered approximately free in
a sufficiently large box, beyond which its interactions cannot be neglected. So, consider a free particle confined in a
cubic box of edge L  106 cm. On the surface of the box must satisfy certain boundary conditions. Since L is
much larger than the particle de Broglie wavelength, these boundary conditions have very little effect on the particle
motion. Therefore, lets set the periodic boundary conditions because they are the simplest to deal with:
(x, y, z) = (x + L, y, z) = (x, y + L, z) = (x, y, z + L) .

(1.23)

A wave function that has the same form as (1.7) and satisfies the boundary condition (1.23) is (at t = 0)
1
k (r) = eikr ,
L3

kx =

2nx
,
L

ky =

2ny
,
L

kz =

2nz
,
L

where nx ,ny ,nz are non-zero integers. This wave function satisfies the following relation
Z
k 0 (r)k (r)d3 r = k0 k ,

(1.24)

(1.25)

Kirill Tuchin

Phys. 591 Lecture Notes

where the integration extends to the entire volume(1 . Indeed, if k 6= k0 the integrand is rapidly oscillating function
that averages to zero. Since momentum of a free particle is conserved, each function (1.24) corresponds to motion
with definite momentum p = ~k.
Consider now a particle in a box in an arbitrary state (i.e. not necessarily free) described by a wave function (r).
We know from the Fourier analysis (see Appendix A) that the set of wave functions (1.24) is a complete set. Namely,
(r) can be expanded as follows
X
(r) =
ak k (r) .
(1.26)
k

This formula tells us that if a particle is in a state (r) and we measure its momentum, the result can be any number
p = ~k from a discrete infinite set specified in (1.24). Multiplying this formula by k 0 (r) on both sides, integrating
over the volume of the box and using (1.25) yields the Fourier coefficients
Z
ak = (r)k (r)d3 r .
(1.27)
Notice, that it follows from (1.5),(1.26) and (1.25) that
Z
Z X
X
|ak |2 .
1 = (r)(r)d3 r =
ak ak0 k (r)k 0 (r)d3 r =
k0 ,k

(1.28)

It seems reasonable to interpret |ak |2 as the probability that measurement of momentum of a particle in state (r)
gives p = ~k. Taking L corresponds to the free space limit.
A simple example developed in this section illustrates many important features of the general theory. I will use it
as a reference in the forthcoming sections.
Additional reading: Merzbacher, Quantum Mechanics 3rd edition, Ch. 2.
4.
A.

Operators

Expectation values of coordinate and momentum

Suppose now that a particle is in a state described by (r, t). Average (or expectation) value of the position vector
is
hri =

(r, t)r(r, t)d3 r .

(1.29)

hF (r)i =

(r, t)F (r)(r, t)d3 r .

(1.30)

If F (r) is a function of r, than

To determine the expectation value of momentum, we place the particle into a big box. We can always expand the
wave function into a complete set of states with definite momentum, see (1.26). The probability to find the value
of particle momentum p = ~k is given by |ak |2 , where ak is given by (1.27). Therefore, according to the statistical
interpretation,
Z
X
XZ
hpi = ~
ak ak k = ~
(r)k (r)d3 rk (r 0 )k (r 0 )d3 r0
(1.31)
k

(1

Kronecker delta ab is defined to be equal unity when a = b and zero otherwise.

Kirill Tuchin

Phys. 591 Lecture Notes

It follows from (1.24) that ik (r) = kk (r), so that (1.31) can be written as
Z
XZ
hpi = i~
(r)k (r)d3 r (r 0 )k (r 0 )d3 r0

(1.32)

Integrating by parts and using the periodic boundary conditions (1.23) implies
Z
Z
Z
(r 0 )k (r 0 )d3 r0 = [(r 0 )k (r 0 )]d3 r0 (r 0 )k (r 0 )d3 r0 .
{z
}
|

(1.33)

=0

Substituting (1.33) into (1.32) and employing the completeness relation (see home assignments)
X
k (r)k (r) = (r r 0 )

(1.34)

we obtain
hpi =

(r)(i~)(r)d3 r .

(1.35)

Taking limit L we derive that (1.35) is valid also for a particle in entire space. Generalization of (1.35) for any
rational function f (p) is straightforward:
Z
hF (p)i = (r)F (i~)(r)d3 r .
(1.36)
For example, expectation value of kinetic energy is
 2 2
 2 Z
~
p

= (r)
(r)d3 r .
2m
2m

(1.37)

We can now generalize (1.30) and (1.37) as follows. Let F be a function of r and p such that
F (r, p) = F1 (r) + F2 (p) .
Expectation value of F can be written in a symbolic form
Z
hF i = (r)F (r)d3 r ,

(1.38)

(1.39)

where
F = F1 (r) + F2 (i~)

(1.40)

is a differential operator.
B.

Hermitian operators

Eq. (1.40) is an example of an operator corresponding physical quantity F . We can can further generalize this
definition to include any physical quantity, not necessarily depending on r and/or p. (Quantities that have no
classical analogue cannot be expressed as a function of r and p, e.g. spin). Namely, an operator F can be put into
correspondence with any physical quantity F . By convention, F means that operator F acts to the right on function
. Most of operators in quantum mechanics are linear, which means that
F (1 + 2 ) = F 1 + F 2 ,

F (a) = aF ,

for any functions 1 , 2 and for any constant a.

Physical quantities are real, hence hF i = hF i . In view of (1.39) this means that
Z
Z

(q)F (q)dq = (q)F (q)dq .

(1.41)

(1.42)

Kirill Tuchin

Phys. 591 Lecture Notes

In linear algebra operator F is called adjoint of Hermitian conjugate of F if


Z
Z

(q)F (q)dq = (q)(F (q)) dq .


Linear operator is Self-adjoint or Hermitian if
Z
Z

(q)F (q)dq = (q)(F (q)) dq .

(1.43)

(1.44)

A short-hand operator notation of (1.44) is


F = F .

(1.45)

Clearly (1.42) is a particular case of (1.44) with = . Thus we can state that to any physical quantity there
corresponds a linear Hermitian operator.
of two operators F and K
as a successive action of operator K
and then of operator F .
Define the product F K
6= K
F ; in operator notation: F K
6= K
F . Even though F and K
represent physical
Note, that generally, F K
may also
quantities, their product does not necessarily do so. We are interested to find a condition under which F K
when is the operator
represent a physical quantity. In mathematical language, given Hermitian operators F and K,
= F K
also Hermitian? To answer this question use (1.44) twice in the following integral
L
Z
Z
Z
Z
Z

F ) dq .
Ldq = F (K)dq = (K)F dq = (F )Kdq = (K
(1.46)
On the other hand, according to (1.43)
Z

Ldq
=

) dq .
(L

(1.47)

Comparing (1.47) with (1.46) we can write down an operator equation(2


= (F K)
=K
F .
L

(1.48)

is Hermitian iff operators F and K


commute with one another, viz.
We see that operator F K
=K
F .
F K

(1.49)

= F K
K
F .
[F , K]

(1.50)

Commutator of two operators is defined as

= 0.
With this notation (1.49) reads [F , K]
we can always construct Hermitian operators S = K
F + F K
and
Out of any two Hermitian operators F and K
= i(K
F F K)
(see homework assignments). In particular, [F , K]
= iG.

G
C.

Examples

1) It is convenient to denote components of vector A by Ai , where i = 1, 2, 3. For example, components of r are ri


such that r1 = x, r2 = y, r3 = z; components of p are pi such that p1 = px , p2 = py , p3 = pz . We have the following
commutators
[
ri , rk ] = 0 ,

(2

[
pi , pk ] = 0 ,

and B
(1.46) and (1.47) imply (A
B)
=B
A
.
For any two operators A

(1.51)

Kirill Tuchin

Phys. 591 Lecture Notes

10

where pi = i~ r i i~i . From the following identities




x
px (x) = x
i~
x
(x)

px x
(x) = i~
= i~
i~
x
x

(1.52)
(1.53)

we derive
[
x, px ] = i~

[
x, px ] = i~ .

(1.54)

In general,
[
ri , pk ] = i~ik .

(1.55)

2) Orbital angular momentum of a particle is L = r p. The corresponding operator:


= i~(r ) .
L

(1.56)

It is Hermitian as can be seen from (1.55) employing a theorem that we proved in the previous subsection. Cross
products can be written using the Levi-Civita symbol ijk defined in Appendix C: Li = ijk rj pk , where summation
over the repeated indices is implied. One can prove that (homework assignment)
i, L
k ] = i~ikl L
l ,
[L
2, L
i] = 0 ,
[L

(1.57)

i , rk ] = i~ikl rl ,
[L
i , pk ] = i~
[L
pl .

(1.59)

(1.58)
(1.60)

Additional reading: Merzbacher, Quantum Mechanics 3rd edition, 4.11.4, 3.4, 11.311.5, Landau and Lifshitz,
Quantum Mechanics: Non-relativistic theory, 3,4.
5.

Eigenfunctions and eigenvalues of operators

As we have seen in I 3 a particle can be in a state with definite value of momentum p = ~k. In general, a system
can be in a state with definite value of a certain physical quantity F . In such a state variation from the average
hF i must vanish. A measure of such variation is variance, defined as the expectation value of the square of operator
d = F hF i. Namely,
F
Z
Z
Z 
Z 
2



d dq ,
d dq =
d F
d dq = F
d F
d F
d dq =
(1.61)
F
F
(F )2 = F
d = 0. Since in such
where I used that fact the operator F is Hermitian. Variance vanishes for states satisfying F
states hF i of course coincides with F we can write this condition as
F = F ,

(1.62)

Eq. (1.62) is usually supplemented by a boundary condition on and is called the eigenvalue problem for operator F
with a particular boundary condition. In mathematics it is a subject of the Sturm-Liouville theory. One of the results
of this theory is that the eigenvalue problem has solutions only at certain discrete values of parameter F . These
particular values are called the eigenvalues of the operator F and the corresponding solutions the eigenfunctions. A
set of all eigenvalues is called its spectrum. I will denote eigenfunctions corresponding to the eigenvalue F as F .
For example, functions (1.24) are eigenfunctions of operator of momentum corresponding to momentum p = ~k. As
we have already seen before, it is often convenient to consider unbounded systems, in which case the spectrum is
continuous.
The physical meaning of the spectrum is that measurement of physical quantity F can yield only it eigenvalues.
If a system is in a state described by one of the eigenfunctions F , than the result of the measurement definitely

Kirill Tuchin

11

Phys. 591 Lecture Notes

produces the corresponding eigenvalue F . Sometimes, several linear independent eigenvalues will correspond to one
eigenvalue of operator. The number of eigenvalues is called the degeneracy of the eigenvalue.
Since F corresponds to a physical quantity, its eigenvalues must be real numbers. To prove that this is indeed the
case, write F = F and its complex conjugate F = F and consider the following integrals
Z
Z
( F F )dq = ( F F )dq .
(1.63)
Since F is Hermitian we can write this as
(F F )

||2 =

( F F )dq = 0 .

(1.64)

Thus, Im F = 0. Now let us discuss properties of discrete and continuous spectra.


Additional reading: Arfken et. al. Mathematical methods for physicists, 7th edition, Ch.8.
A.

Properties of discrete spectrum

1. Consider operator F that has a non-degenerate discrete spectrum Fn , i.e.


F n = Fn n ,

were I used (1.64). It follows that


Z
Z

m
F n dq = Fn m
n dq ,

F m
= Fm
m = Fm m
,

and

n F m
dq = Fn

(1.65)

n m dq .

(1.66)

Subtracting these two equations yields


(Fn Fm )

m
n dq = 0 .

If m 6= n we arrive at the orthogonality property of the wave functions


Z

m
n dq = 0 , m 6= n .

(1.67)

(1.68)

It physical meaning is that measurement of F gives definitely either Fn in n state or Fm in m state for any
m 6= n. In other words, measurement of F cannot result in a system being in a superposition of eigenfunctions.
2. As I mentioned above (see discussion after (1.62)), an eigenvalue problem has a discrete spectrum
R if it is restricted
to a finite domain. In this case wave functions n rapidly decrease at q implying that |n |2 dq is finite.
Thus, we can normalize these wave functions to unity. Together with (1.68) we obtain the orthonormality
property
Z

m
n dq = mn .
(1.69)
3. An important result of the Sturm-Liouville
theory is that eigenfunctions {n }n form a complete set. Namely,
R
any function (q) such that |(q)|2 dq is finite can be expanded as follows
X
(q) =
an n (q) .
(1.70)
n

Using the orthonormality property we can calculate the coefficients an :


Z
an = (q)n (q)dq .
These equations generalize (1.25),(1.26),(1.27) for a particle in a box.

(1.71)

Kirill Tuchin

12

Phys. 591 Lecture Notes

4. Eigenfunctions satisfy the completeness relation that reads (cp. (1.34))


X
n (q 0 )n (q) = (q q 0 ) .

(1.72)

Indeed, using (1.70) and (1.71) we can write


0

(q q ) =

an (q )n (q) ,

where

an (q ) =

(q q 0 )n (q)dq = n (q 0 ) .

(1.73)

5. Suppose now that Fn is N -fold degenerate, i.e. functions nl with l = 1, . . . , N are eigenfunctions corresponding
to the same eigenvalue Fn . Eigenfunctions with different n are still orthogonal, which can be proved as in (1.65)
(1.68). However, eigenfunctions with the same n, but different l are not necessarily so. Since it is convenient
to deal with a set of orthogonal functions we can replace a set of N independent functions nl by another set
of functions which are eigenfunctions of F and orthogonal. Let me show how this can be done in the simplest
case N = 2.
We start with two normalized eigenfunctions n1 and n2 . Define two new functions n1 = n1 and n2 =
a(n1 + n2 ). These two are also eigenfunctions of F . The idea is two choose such that
Z
Z
Z

n1 n2 dq = 0 a n1
(n1 + n2 )dq = 0 1 = n1
n2 dq ,
(1.74)
and choose a such that

n2 n2 dq = 1 .

(1.75)

This procedure is called orthogonalization or Gram-Schmidt process. It can be easily generalized to arbitrary
N . The resulting spectrum satisfies the following relations:
Z

ml
(q)nk (q)dq = mn kl ,
(1.76)
Z
X

(q) =
anl nl (q) , where anl = (q)nl
(q)dq ,
(1.77)
X
n,l

n,l

nl
(q 0 )nl (q) = (q q 0 ) .

(1.78)

6. Expectation value of F can be expressed through an as follows


!
Z
Z X
Z
X
XX

hF i = F dq =
an n F
am m dq =
an am n Fm m
=

XX
m

Fm mn an am =

X
n

|an |2 Fn .

(1.79)

(1.80)

In particular, for F = 1 we have

|an |2 = 1 .

Compare this with (1.28). As in I 3 we can interpret |an |2 as probability to find a system in state n.
It is important to note that knowledge of probabilities |an |2 does not completely specify the wave function of a
quantum state. The wave function is not known unless all relative phases between different states n are also known.
States for which only probabilities are known are called the mixed states, whereas states with a wave function are
called the pure states. Instead of wave function, the mixed states are described by the density matrix. Such description
is not full even in quantum mechanical sense. Mixed states play in quantum statistical physics the same role as the
wave function in quantum mechanics.
Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, 3.

Kirill Tuchin

13

Phys. 591 Lecture Notes


B.

Properties of continuous spectrum

1. Consider
operator F that has continuous spectrum: F F = F F , where F is continuous. In this case
R
|F |2 dq = because the system (or its part) can go to infinity. We can derive properties of continuous
spectrum by taking the limit L as we did in the case of a particle in box. As a result we obtain that
{F }F is a complete set, i.e. any (q) can be expanded as
Z
(q) = aF F (q)dF .
(1.81)
If (q) is normalized, then aF must be normalized as well due to (1.80). Namely
Z
Z
(q)(q)dq = aF aF dF = 1 .
Let us show that this allows us to invert (1.81) similarly to (1.71). Write
Z

Z
Z
Z
Z

dqF (q)
(q)(q)dq = dq(q) aF F (q)dF = dF aF

(1.82)

(1.83)

and subtract from it (1.82). This gives


0=

dF aF

Z

dqF (q)

aF =

F (q)dq .

aF

(1.84)

which is true if
(1.85)

2. Substitute (1.81) into (1.85) to find


aF =

dqF (q)

aF 0 F 0 (q)dF 0 .

This equation is self-consistent if the following delta-function normalization holds


Z
F (q)F 0 (q)dq = (F F 0 ) .
Similarly, we can start with (1.81), substitute into it (1.85) and obtain that
Z
F (q 0 )F (q)dq = (q 0 q) .

(1.86)

(1.87)

(1.88)

Thus, we learned how to normalize the wave functions of continuous spectrum.


Finally, some operators have both discrete and continuous spectrum. In this case the complete set of states includes
eigenfunctions from both parts of the spectrum. Any function (q) can be expanded as
Z
X
(q) =
an n (q)dq + aF F (q)dq .
(1.89)
n

Normalization

|(q)|2 dq = 1 implies that


X
n

|an |2 +

|aF |2 dF = 1 .

Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, 5.

(1.90)

Kirill Tuchin

Phys. 591 Lecture Notes


6.

14

Examples of eigenvalue problems

Operator of momentum p.

A.

1. Eigenvalue problem for the operator px = i~x reads


i~

(x)
= px (x) .
x

(1.91)

Solution to this equation is


px (x) = N eipx x/~ .

(1.92)

This is the eigenfunction of px corresponding to eigenvalue px . Spectrum of this operator is continuous <
px < . px (x) describes motion of a particle along x-axis with definite momentum px .
The normalization constant N can be fixed using (1.87) and (B11) to be N = 1/(2~)1/2 . Indeed,
Z
Z
0
1
ei(px px )x/~ dx = (px p0x ) .
px (x)p0x (x0 )dx =
2~

(1.93)

2. In three dimensions p = i~. The corresponding eigenvalue problem is solved by eigenfunctions


p (r) =

1
eipr/~
(2~)3/2

(1.94)

corresponding to eigenvalues p. Normalization is similar to the one-dimensional case if one uses (B16).
B.

Operator of position r.

This is another example of continuous spectrum. The eigenvalue problem reads


rr0 (r) = r0 r0 (r) .

(1.95)

Action of r on a function of r is a simple product. If r = r0 , then (1.95) is identity, if r 6= r0 , then (1.95) is satisfied
only if r0 = 0. Therefore, the properly normalized solution to (1.95) is
r0 (r) = (r r0 ) .
We can expand any wave function in complete set of eigenfunctions r0 (r) as follows
Z
(r) = ar0 r0 (r)d3 r0 ,

(1.96)

(1.97)

where
ar0 =

(r)(r r0 )d3 r = (r0 ) .

(1.98)

We observe that |ar |2 = |(r)|2 is the probability density to find a particle at position r. This agrees with the
interpretation we gave in the previous section.
C.

Operator of orbital angular momentum L.

z projection of orbital angular momentum on z-axis. In Cartesian coordinates it has form


1. Consider operator L
(see (1.56))
z = i~ (xy yx ) .
L

(1.99)

Kirill Tuchin

Phys. 591 Lecture Notes

15

In spherical coordinates
z = i~ ,
L

(1.100)

where 0 2. The eigenvalue problem reads


i~ Lz () = Lz Lz () ,

(1.101)

Lz () = Lz ( + 2) .

(1.102)

with the periodic boundary condition

Since this eigenvalue problem is restricted to unit circle, its eigenvalues Lz are discrete. Solution to the differential
equation (1.101) is
Lz () = N eiLz /~ .

(1.103)

Boundary condition implies that Lz = m~, with any integer m. We can thus label eigenvalues and eigenfunctions
z by m. Imposing the normalization condition
of L
Z 2

m
m d = 1
(1.104)
0

we finally obtain the orthonormal set of eigenfunctions


1
m () = eim .
2

(1.105)

2 = P3 L
2
2. Another operator that we will use extensively in this course is L
k=1 k . In spherical coordinates it reads




1

1 2
2
2

sin
+
.
(1.106)
L = ~
sin

sin2 2
2 lm = L2 lm becomes
The eigenvalue equation L





1 2
L2
1
2
sin
+
+ 2 lm (, ) = 0 .
~
sin

~
sin2 2

(1.107)

One of the boundary conditions is periodicity in , as in (1.102). Another requires that the eigenfunction be
finite at = 0 and = . The resulting eigenfunctions are spherical harmonics Ylm (, ) and the corresponding
eigenvalues are L2 = ~2 l(l + 1), where
l = 0, 1, 2, . . .

and m = 0, 1, 2, . . . , l .

(1.108)

(2l + 1)(l m)!


Plm (cos )eim ,
4(l + m)!

(1.109)

Explicit form of the spherical harmonics is


Ylm (, ) =

where Plm are associated Legendre polynomials. A list of spherical harmonics can be found in all books on
quantum mechanics. Normalization condition
Z

Ylm
(, )Yl0 m0 (, )d = ll0 mm0 .
(1.110)
Here d = sin dd is an element of solid angle.
2 , spherical harmonics are also eigenfunctions of L
z:
In addition to being eigenfunctions of L
z Ylm = i~ Ylm = ~mYlm .
L

(1.111)

2 is 2m + 1-fold degenerate because 2m + 1


We make two important observations: (i) Eigenvalue l of operator L
different values of Lz correspond to the same l, see (1.108). (ii) Two physical quantities can have definite values
in the same quantum state, i.e. two operators can share the same eigenfunction.
Merzbacher, Quantum Mechanics 3rd edition, 11.3,11.4.

Kirill Tuchin

Phys. 591 Lecture Notes


7.

16

Description of quantum states

Since, as we have seen in the previous section, two different physical quantities can simultaneously have definite
that have definite
values, let us find a relationship between the corresponding operators. Consider operators F and G
values in state n . Formally, this means that
F n = Fn n ,

n = Gn n .
G

(1.112)

we get
Acting on this equations with F and G
F n = Fn G
n = Fn GN n ,
G
n = Gn F n = Gn Fn n .
F G

(1.113)

F F G)
n = 0.
(G

(1.115)

(1.114)

Subtracting (1.113) from (1.114) gives

Now consider any function :


F F G)
= (G
F F G)

(G

an n =

X
n

F F G)
n = 0.
an (G

(1.116)

The conclusion is
= 0.
[F , G]

(1.117)

Converse statement is also true, viz. if two operators commute, they have the same eigenfunctions. Indeed, let
= 0 and G
n = Gn n . Then
[F , G]
F n ) = F (G
n ) = Gn (F n ) .
G(

(1.118)

n corresponding to eigenvalue Gn . If n is not degenerate,


This implies that function F n is also an eigenfunction of G

than F n n . Denoting the proportionality


factor by Fn we obtain F n = Fn n . If n is degenerate, we can
P
construct linear combinations nl = k alk nk that are eigenfunctions of F .
In summary, two physical quantities can have simultaneously definite values iff the corresponding operators commute. Based on this result, one can decide which set of physical quantities can be chosen for a description of a
quantum system. If we know the values of all independent quantities with definite values in a given state, then the
wave function in that state must be an eigenfunction of all corresponding operators. We have already discussed two
examples: states with definite value of momentum p (its three projections commute with each other), see (1.94),
(1.51) and states with definite values of square of orbital angular momentum and its projection along some direction,
see (1.109), (1.58). More examples will be considered in forthcoming chapters.
Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, 4.
8.

Heisenberg relations

and F be Hermitian operators. Their commutator can be written as [K,


F ] = iG,
where G
is also a Hermitian
Let K
operator (see below (1.50)). Define
d = F hF i ,
F

It is easy to see using this definition that

d =K
hKi .
K

F F K
= K
d F
d F
d K
d = iG
.
K

Consider now the following auxiliary function I() of a real parameter :


Z
2
d
d ) dq 0 .
I() = (K
iF

(1.119)

(1.120)

(1.121)

Kirill Tuchin

Phys. 591 Lecture Notes

17

are Hermitian operators, and hence obey (1.44), we can cast this function into a different
Using the fact that F and K
(3
form :
Z
Z

d
d
d
d
d + iF
d )(K
d iF
d )dq
I() = (K + iF ) (K iF ) = (K
Z


d 2 + (F
d )2 + G]dq

= [2 (K)
= 2 (K)2 + (F )2 + hGi 0 .
(1.122)
I() is a parabola with its branches pointing up (because 2 > 0). Therefore, (1.122) is satisfied iff the minimum
value of I is positive. Minimum of I occurs at = 0 given by
0 =

hGi
,
2 h(K)2 i

(1.123)

hence the minimum of I is


I(0 ) = (F )2

hGi
0.
4 h(K)2 i

(1.124)

We arrived at the Heisenberg uncertainty relation between the uncertainties of two Hermitian non-commuting operators:

hGi2
(F )2 (K)2
.
4

(1.125)

A system can be in a certain state that minimizes the uncertainty relation; for such a state inequality (1.121)
becomes an equation. This happens when
 2
Z 


hGi
d
d
dq = 0 ,

(1.126)
K

i
F

I(0 ) =

2 h(K)2 i
implying that


hGi
d + iF
d = 0.
K
2 h(K)2 i

(1.127)

States that satisfy this equation are called the coherent states.
A.

Example

=x
= ~ since [
Let K
, F = px , then G
x, px ] = i~. The uncertainty relation (1.125) reads

Denote hxi = x0 and hpx i = p0 , so that


Plugging into (1.127) yields


(x)2

~2
(p)2
.
4

d = x x0 ,
K

d = i~x p0 .
F


~(x x0 )
+ i(i~x p0 ) (x) = 0 .
2 h(x x0 )2 i

(1.128)

(1.129)

(1.130)

Look for a solution in form (x) = N eS(x) , where N is a constant. We obtain an equation for S:
(x x0 )
dS
ip0
+

= 0.
2
2 h(x) i
dx
~

(3

d is not Hermitian.
Caution: operator iF

(1.131)

Kirill Tuchin

Phys. 591 Lecture Notes

18

Its solution (modulo an additive constant)


S=

ip0 x (x x0 )2

.
~
4 h(x)2 i

Constant N is fixed by the normalization condition. The final result is




ip0 x (x x0 )2
1
exp
(x) =

.
~
4 h(x)2 i
(2 h(x)2 i)1/4

(1.132)

(1.133)

It is not difficult to verify that for this state (1.128) turns into equation.
Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, 16.
9.

Classical limit of quantum mechanics

We know that macroscopic objects consists of microscopic ones: molecules, atoms, electrons etc. Microscopic
systems are governed by quantum mechanics, while the macroscopic ones by classical physics. Therefore, one should
be able to obtain classical equations of motion as a limiting case of quantum mechanics. This statement is known as the
correspondence principle. Consider application of this principle to the wave packet discussed in I 3. The uncertainties
of particle position and its momentum (1.14) decouple (i.e. become independent as in classical mechanics) in the formal
limit ~ 0. In other words, in this limit one can in principle measure the position and momentum with any accuracy.
To better understand what is meant by this limit (after all ~ is a physics constant) note that ~ has units of action S.
So when for a given system S  ~, quantization can be neglected.
Consider the wave function of a free particle with given energy and momentum (1.7):
(r, t) = N ei(prEt)/~ .

(1.134)

In classical mechanics, action of a free particle S = p r Et, so we can write


cl (r, t) = N eiS/~ .

(1.135)

The classical limit corresponds to large values of the phase in (1.135). This is similar to the geometrical optics limit
of electromagnetic theory. There any component of electromagnetic wave can be written as u = N ei with real N
and . In geometrical optics limit phase becomes large corresponding to the limit of very short wave length 0.
In such a case propagation of electromagnetic wave can approximated by straight lines (rays). Moreover, both S and
satisfy the variational principle that says that they must take the least possible value. Since any wave function can
be expanded in plane waves (1.134), it seems reasonable that (1.135) represents the form of the wave function in the
classical limit. We will discuss a more accurate derivation later in this course.
For future reference, recall a few facts form the classical mechanics. A mechanical system described by action S(q, t)
satisfies the Hamilton-Jacobi equation

S
=H,
t

(1.136)

where H = H (q, p, t) is the Hamiltonian function with the canonical momenta given by
p=

S
.
q

(1.137)

If the Hamiltonian function does not explicitly depend on time, then energy E = H is a conserved quantity. Dependence of E on coordinates and momentum of a free particle follows from general physical requirements of uniformity
and isotropy of space and time and invariance under Galileo transformations, which leads to
Hfree =
for one free particle.

p2
2m

(1.138)

Kirill Tuchin

Phys. 591 Lecture Notes


II.

19

SCHRODINGER
EQUATION
1.

Hamiltonian

Wave function contains all information about a given physical system at any time (see for example I 2). Therefore,
/t at any given time must be determined by the value of at the same time. The most general such relation
reads
i~

.
= H
t

(2.1)

must be a linear operator and i~ is extracted from H


for convenience. Eq. (2.1)
Due to the superposition principle H
describes the time-evolution of the wave function.(4
let us take the classical limit by substituting the wave function (1.135) into
To find out the physical meaning of H
(2.1). We have
cl = S cl = Hcl ,
H
t

(2.2)

in the classical limit becomes the Hamiltonian function, see (1.136). On account of this correspondence
i.e. operator H

operator H is called the Hamiltonian operator or simply the Hamiltonian. Its operator form for a free particle follows
from (1.138):
2
2
free = p = ~ 2 .
H
2ma
2m

(2.3)

Interaction of a system of non-relativistic particles is described by potential energy U which is a function coordinates
of particles. Thus, the Hamiltonian of a system of non-relativistic particles reads
=
H

X p2
a
+ U (r1 , . . . , ra ) .
2m
a

(2.4)

In obtaining (2.4) we relied on the correspondence principle, which is supported by a great number of experiments.
Substituting (2.4) into (2.1) we obtain the Schr
odinger equation for a single particle
i~

~2 2
=
+ U (r) ,
t
2m

(2.5)

where U is the potential energy of the particle in an external field. For a free particle it has solution
(r, t) = N ei(prEt)/~ ,

(2.6)

which agrees with (1.7).


A possible time-dependence of the wave function is restricted by the requirement of the total probability conservation
Z
d
|(q, t)|2 dq = 0 .
(2.7)
dt
Let us verify that Schr
odinger equation satisfies (2.7). From (2.1) it follows that
i~

(4

= H
t

i~


.
= H
t

(2.8)

This time-evolution of the wave function is a result of the influence of forces acting on the system. It has nothing to do with the changes
in the system introduced by the measurement process. The later replaces one wave function by another, which is known as the wave
function collapse.

Kirill Tuchin

Phys. 591 Lecture Notes

20

Subtracting these equations gives


i~


H
.
( ) = H
t

(2.9)

is a Hermitian operator we derive (2.7).


Integrating over q and using the fact that H
While the total probability is conserved, the probability density = |(q, t)|2 is not. Indeed, substituting (2.4) (for
a single particle) into (2.9) yields
i~
Introduce the probability current density


~2

=
2 2 .
t
2m
j=

~
( ) .
2mi

(2.10)

(2.11)

Then, (2.10) can be cast in the form of the continuity equation:

= j .
t

(2.12)

Since is a complex function, it can represented in term of two real functions f and as
(r, t) = f (r, t)ei(r,t) .
In this notation we have
2

=f ,

j =

~
m

(2.13)

(2.14)

In particular, if the phase is independent of coordinates, then the probability current vanishes j = 0.
Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, 8,17, Merzbacher,
Quantum Mechanics 3rd edition,
2.

Stationary states

If the Hamiltonian function of a classical system is not explicitly time-dependent, its energy is conserved. To see
how energy conservation is expressed in quantum mechanics, consider a system describe by a Hamiltonian that does
not explicitly depend on time, i.e.

H
= 0.
t

(2.15)

For such a system time-dependence of the wave function can be computed from (2.1) employing the method of
separation of variables:
(q, t) = T (t)E (q) .

(2.16)

This yields
i~

E (q)
H
T (t) 1
=
=E,
t T (t)
E (q)

(2.17)

where E is a constant. Indeed, a function of t can be equal a function of q only if both are constants. As a result we
have two equations
E (q) = EE (q) ,
H
T (t)
i~
= ET (t)
t

(2.18)
(2.19)
(2.20)

Kirill Tuchin

21

Phys. 591 Lecture Notes

Eq. (2.19) immediately gives


T (t) = eiEt/~ .

(2.21)

States with definite energy are called the stationary states. The time dependence of their wave functions reads
(q, t) = E (q)eiEt/~ .

(2.22)

Eq. (2.19) becomes the time-independent Schr


odinger equation :

~2 2
+ (U E) = 0 .
2m

(2.23)

In fact, since one is usually interested in stationary states, (2.23) as simply referred to as the Schrodinger equation .
Consider properties of the stationary states.
1. Time-dependence of a stationary state is uniquely determined by its energy.
2. = const.(t) and j = const.(t).
3. If F does not explicitly depend on time, then its expectation value in a stationary state is also time-independent:
Z
hF i = (q, t)F (q, t)dq = const.(t) ,
(2.24)
which means that F is a conserved.
Note, that although hF i is time-independent, the wave function of a stationary state does depend on time
according to (2.21).
4. Consider a system in P
a stationary state (q, t) and let {n (q)} be a complete set of eigenfunctions of F . We
can expend (q, t) = n an (t)n (q). The probability to find the system in the nth eigenstate is
2
Z


|an |2 = (q, t)n (q)dq = const.(t)

(2.25)

5. Functions {E (q)}, being eigenfunctions of Hamiltonian, formP


a complete set. Therefore, any solution of the
Schr
odinger equation can be expanded in them as (q, t) =
E TE (t)E (q). Owing to (2.21) we obtain a
general form of the expansion of an arbitrary wave function (q, t) in a complete set of stationary states:
X
(q, t) =
an n (q)eiEn t/~ ,
(2.26)
n

for discrete spectrum and


(q, t) =

aE E (q)eiEt/~ dE ,

(2.27)

for the continuous one.


Eqs. (2.26),(2.27) indicate that in order to calculate the time-evolution of a wave function one needs to (i) find
the spectrum of the Hamiltonian, (ii) fix coefficients an using the initial condition at t = 0 and (iii) sum or
integrate over the energy spectrum according to (2.26), (2.27).
6. We proved in (1.64) that values of E are real. This is also seen in (2.23) which is a real differential equation.
Moreover, the wave functions of the stationary states can also be chosen to be real (except for a system in
magnetic field). Indeed, and satisfy the same equation. For a non-degenerate energy level it implies that
they are the same functions up to a non-important phase. For a degenerate level, one can form real linear
combinations of the wave functions.
Additional reading: Landau and Lifshitz, Quantum Mechanics:
Merzbacher, Quantum Mechanics 3rd edition, 3.1.

Non-relativistic theory, 8,10,17,19,

Kirill Tuchin

Phys. 591 Lecture Notes


3.

22

Time derivative of operators

Time derivative of a physical quantity F cannot be defined as in classical mechanics since in quantum mechanics the
value of F may not be known at two arbitrarily close time instants. Instead we will consider an observable quantity:
the expectation value of F . I pointed out in (2.24) that if F does not explicitly depend on time, its expectation
value in a stationary state isR time-independent. Consider now the expectation value of a Hermitian operator F in

an arbitrary state : hF i = F dq. Taking its time derivative and using the Schrodinger equation = H/i~,

/i~ we derive
= H
)
Z (

d hF i

dq
(2.28)
=
+
F + F

dt
t
t
t
)
Z (

1
1
F
=

(H )F + F H dq
(2.29)
t
i~
i~
(
)
Z
F
1

=
+ [F , H] dq ,
(2.30)
t
i~
Define now an operator dF /dt such
where in the last step I used (1.44) with the replacements F and F H.
that its expectation value obeys the following condition
*
+ Z
d hF i
dF
dF
=
=
dq .
(2.31)
dt
dt
dt
In view of (2.30) we obtain an operator expression of the time derivative
dF
F
1
.
=
+ [F , H]
dt
t
i~

(2.32)

= 0, then F does not change with time, i.e.


This implies that if tF = 0 (no explicit time dependence) and [F , H]
F is a conserved quantity. We also conclude that any conserved quantity can be measured concurrently with the
Hamiltonian.

 Consider application of (2.32) to the motion of a single particle. Replacing F first by p and then by r we get two
operator equations
dp
1
,
H]
= [p,
dt
i~
d
r
1
.
= [
r , H]
dt
i~

(2.33)
(2.34)

Commutators are computed as follows:


= [p,
H]
U ] = i~[, U ] = i~[(U ) U ] = i~U ,
[p,
1
p
= 1 [
[
r , H]
r , p2k ] =
{ei [
ri , pk ]
pk + pk [ei ri , pk ]} = i~ ,
2m
2m
m

(2.35)
(2.36)

where ei is a unit vector. In the last line summation over k and i is implied and (1.55) is used. Plugging (2.35),(2.36)
into (2.33) and (2.34) gives
dp
= U ,
dt
d
r
p
v =
=
,
dt
m

(2.37)
(2.38)

in agreement with the correspondence principle (which justifies our definition (2.31) a posteriori). These equations
are knows as the Ehrenfest theorem. Taking another time derivative of (2.38) we derive using (2.32)
d
v
1 dp
1
=
= U .
dt
m dt
m

(2.39)

Kirill Tuchin

Phys. 591 Lecture Notes

23

Formally, this formula has the same form as the classical equation of motion. However, it must be understood as a
relationship between the corresponding expectation values:
Z
Z
d2

3
rd r = U d3 r .
m 2
(2.40)
dt

To better understand the difference between the operator formula (2.39) and its classical counterpart, it is worth
examining the classical limit at this point. Namely, we would like to know what are the conditions that allow us to
replace operators in (2.39) with the corresponding classical quantities. For the sake of simplicity we consider only the
x-components. Denote x
= hxi, x = x x
and expand at small x (in the classical limit x vanishes):
1 3U
U (x)
U (
x) 2 U
x
+
(x)2 + . . .
=
+
x
x

x
2
2 x
3
R
Substituting into (2.40) and noting that ||2 xdx = hxi h
xi = 0 we have

d2 x

U (
x) 1 3 U

(x)2 + . . .

2
3
dt
x

2 x

In the classical approximation we must require




3
U




 1 U (x)2
x

2 x
3
m

(2.41)

(2.42)

(2.43)

In addition to the requirement that uncertainty in x be small, we also have to find a similar requirement for the
momentum. Consider the kinetic energy


 2  

2
(px )2
px
[hpx i + (px hpx i)]2
hpx i
2 hpx i hp hpx ii

=
+
+
.
(2.44)
2m
2m
2m
2m
2m
The second term obviously vanishes. In order that the third term be small the following condition must be satisfied


~2
2
hpx i  (px )2 =
.
(2.45)
4 h(x)2 i

Inequalities (2.43) and (2.45) indicate that the classical limit applies when external field is slowly varying and momenta
are large. We can combine (2.43) and (2.45) into one condition:

2
~2 U 000
px 
.
(2.46)
8 U0
 As in classical mechanics, we can prove the quantum virial theorem. On one hand, in a stationary state expectation
value of p r is time-independent, see (2.24), i.e.
d
hp ri = 0 .
dt

(2.47)

On the other hand,


2


1
d
= 1 p [
+ [p,
r = p rU .
H]
(p r) = [p r, H]
r , H]
dt
i~
i~
m
Employing the definition (2.31) we cast this in form
 2
d
p
hp ri =
hrU i
dt
m

We thus arrive at the required relation

p2
m

= hrU i .

(2.48)

(2.49)

(2.50)

For illustration, suppose U rn . Then, hrU i = n hU i and the virial theorem implies a relation between the
expectation values of kinetic and potential energies:
 2
p
n
= hU i .
(2.51)
2m
2
Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, 9,19, Merzbacher,
Quantum Mechanics 3rd edition, 3.3,3.5,3.6.

Kirill Tuchin

24

Phys. 591 Lecture Notes


4.

Schr
odinger and Heisenberg pictures of time-evolution

In practical application it is sometimes convenient to use different representations or pictures of time-evolution.


Observed quantities do not depend on a choice of a picture, of course.
A.

Schr
odinger picture

Thus far we have followed the Schr


odinger picture of time-evolution in which wave functions depend on time,
whereas the operators corresponding the physical quantities usually do no explicitly depend on time. We can express

the time-evolution of a wave function (q, 0) (at t = 0) as a result of action of the time-evolution operator S(t)
as
follows

(q, t) = S(t)(q,
0) .

(2.52)

The normalization of the wave function should not change with time, i.e.
Z
Z

(q, t)(q, t)dq = (q, 0)(q, 0)dq = 1 .

(2.53)

On the other hand,


Z

(q, t)(q, t)dq =

[S (q, 0)][S(q,
0)]dq =

(q, 0)S S(q,


0)dq ,

(2.54)

We conclude that
where I used the definition (1.43) with F S and S.
S S = 1 .

(2.55)

substitute (2.52) to the


Operators satisfying (2.60) are unitary operators. To find the explicit form of operator S,
Schr
odinger equation :
i~

S(t)
S(t)(q,

(q, 0) = H
0) .
t

(2.56)

S(t)
S(t)
.
=H
t

(2.57)

In operator form:
i~

For closed systems Hamiltonian does not explicitly depend on time. In such a case one can integrate (2.57) to obtain

= e ~i Ht
S(t)
.

(2.58)

How does the operator (2.58) act on a wave function? Since S is a function of the Hamiltonian, it is helpful to
Namely, using
expand (q, 0) in a set of stationary states {n (q)}n and act on the result by the evolution operator S.
(2.52) we have
!k
X 1

i
Ht
(q, t) =S
an n (q) =
an

n (q)
k!
~
n
n
k
X
X X 1  iEn t k
=
an

n (q) =
an n eiEn t/~ ,
k!
~
n
n
X

(2.59)

in agreement with (2.26). One can think of (2.58) as a compact way of expressing the time evolution of (2.26) and
(2.27).

Kirill Tuchin

Phys. 591 Lecture Notes


B.

25

Heisenberg picture

In the Heisenberg picture wave functions do not evolve with time, but the operators do. Let (q, t) be a wave
function in the Schr
odinger picture describe a certain state and H (q, t) a wave function in the Heisenberg picture
describing the same state. According to (2.52), in order that H (q, t) be time-independent it must satisfy
H (q, t) = S1 (q, t)

(2.60)

Consider now an expectation value of a physical quantity F . It must be the same regardless of the picture we choose
to work in. Thus,
Z
Z
Z

hF i = F dq = (SH ) F SH dq = H
S F SH dq .
(2.61)
Let us call F an operator in the Schr
odinger picture and append a subscript S to it. Then the same operator in the
Heisenberg picture reads
FH = S1 F S ,

(2.62)

where I used (2.55) to write S = S1 .


Suppose that at some initial time t FH (t) = F . After evolving over a short time-interval t with (2.62) we have

FH (t + t) = S1 (t)FH (t)S(t)
.

(2.63)

Expanding in Taylor series we get


dFH
t =
FH (t) +
dt




i
i

1 + Ht FH (t) 1 Ht .
~
~

(2.64)

Keeping only linear terms in t we obtain


dFH
1
.
= [FH , H]
dt
i~

(2.65)

This formula is similar to (2.32), but has a different meaning. Eq. (2.32) is a definition of the time-derivative of
F , while in dFH /dt in (2.65) is the derivative of the physical quantity (whose operator is time-dependent in the
Heisenberg picture).
= p2 /2m. In the Schrodinger picture
As an example, consider motion of a free particle. Its Hamiltonian is H
p = i~. In the Heisenberg picture
i

pH = e ~ Ht (i~)e ~ Ht = p +

it
= p ,
[H, p]
~

(2.66)

where I used the following identity:


A

B]
+ 1 [A,
[A,
B]]
+ ...
+ 1 [A,
eA Be
=B
1!
2!

(2.67)

Position operator can be calculates similarly using (2.36):


i

rH = e ~ Ht re ~ Ht = r +

it
(it)2
p
[H, r] +
[H, [H, r]] + . . . = r + t .
2
~
2~
m

(2.68)

Alternatively, we can solve equations of motion


d
i
p
rH = [H,
rH ] =
,
dt
~
m
which leads to (2.68).

d
i
pH = [H,
pH ] = 0
dt
~

(2.69)

Kirill Tuchin

Phys. 591 Lecture Notes


C.

26

Interaction picture

Another popular representation of the time-evolution is the interaction picture. Suppose we can split the Hamiltonian into two parts:
=H
0 + V ,
H

(2.70)

0 describes a system without taking interactions into account, while V describes interactions. Introduce a
where H
unitary operator

S = eiH0 t/~

(2.71)

and define the wave function in this picture as follows

I (q, t) = S(q,
t) .

(2.72)

yields
Substituting this into the Schr
odinger equation i~ = H
i~ I = SV S1 I .

(2.73)

VI = SV S = eiH0 t/~ V eiH0 t/~ .

(2.74)

i~ I = VI I .

(2.75)

Define operator VI in the interaction picture

Then we can rewrite (2.73) as

In other words, the wave function I evolves in time only according to the interaction part of the Hamiltonian. At
the same time an operator F evolves with the non-interacting part of the Hamiltonian as in (2.74):

FI = eiH0 t/~ F eiH0 t/~ .

(2.76)

Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, 13, Merzbacher,
Quantum Mechanics 3rd edition, Ch.14.
5.

Symmetries and conserved quantities

In classical mechanics uniformity of space implies momentum conservation, its isotropy implies the orbital angular
momentum conservation and uniformity of time intimates energy conservation. Since the Hamilton function of a
system carries all information about its properties, it must exhibit the same symmetry properties as the system. The

same is true in quantum mechanics were we have to replace the Hamilton function with the corresponding operator H.

Let S be an operator performing a symmetry transformation. Invariance of Hamiltonian under such transformation
H]
= 0 (5 .
means that [S,

(5

Although this is pretty evident, let me illustrate this in the case of infinitesimal continuos transformation.
Such
can
R
R a transformation
Sdq

be written as S = 1 + i
s, where s is Hermitian operator. This follows from the requirement that dq = (S)
up to small
we need to require that [
= 0, implying
terms of order s2 . It is easy to check that in order that hHi stay invariant under S
s, H]
H]
= 0. This equation is true for discrete transformations as well.
that [S,

Kirill Tuchin

27

Phys. 591 Lecture Notes


A.

Translations

Consider translation (displacement as a whole) of a closed system of particles by a small constant vector , viz.
ra ra0 = ra + . Under this transformation the wave function transforms as follows
X
X

(r1 + , r2 + . . .) = (r1 , r2 , . . .) +
a (r1 , r2 , . . .) = 1 +
a (r1 , r2 , . . .) .
(2.77)
a

Operator 1 + a a is an operator of an infinitesimal translation. Since the Hamiltonian is invariant (i.e. do not
change) under translations we arrive at the statement
X

= 0.
a , H
(2.78)
a

Recalling
that pa = i~a , we conclude that uniformity of space implies the total momentum conservation
P
= 0. In fact, we guessed the form of the momentum operator in I 4. Instead, we could have defined
[ a pa , H]
momentum as the infinitesimal translation operator (up to a constant factor). Consider the operator of finite translations
Ta (r) = (r + a) ,

(2.79)

where a is a finite constant. Expanding the right-hand-side in Taylor series we get


#
"

2
i
1
i

a p + . . . (r) = eiap/~
(r) .
Ta (r) = 1 + a p +
~
2 ~

(2.80)

Thus,

Ta = eiap/~
.

(2.81)

The generator of translations is the following operator


GT =
Notice that the generator does not depend on a.



i
Ta
= p = .
a
~
a=0
B.

(2.82)

Rotations

Invariant of a closed system under rotations follows from the isotropy of space. The corresponding operator must
be proportional to the operator of the total angular momentum. To see that this is indeed the case, consider an
infinitesimal rotation by angle :
ra ra0 = ra + ra .

(2.83)

Under this rotation the wave function (r) changes as follows



ri + [ r]i = (ri + ijk j rk ) = (ri ) + i (ri )ijk j rk = (ri ) + (r )(ri ) ,
(r1 + r1 , r2 + r2 , . . .) = (1 +

X
a


ra a ) (r1 , r2 , . . .)

(2.84)

(2.85)

P
Operator 1 + a ra a ) is the operator of infinitesimal rotations. Since the closed system is invariant under
rotations it commutes with the Hamiltonian. In particular,
X
= 0.
ra a ), H]
(2.86)
a

Kirill Tuchin

Phys. 591 Lecture Notes

28

P
It follows that a ra a ) must be proportional to the total angular momentum of the system, which justifies our
definition (1.56).
Operator of rotation by a finite angle around axis with direction n can be computed by analogy with (2.79)-(2.81)
with the result

n = eiLn/~
R
.

(2.87)

Generator of rotations around the n-direction is



i
n
n

= L
R
n.
G =
=0
~

If represent vectors as columns

Ax
A = Ay ,
Az

(2.88)

(2.89)

then rotation operators can be represented as 3 3 matrices. For example, operator of rotation around x-axis reads

1 0
0
x = 0 cos sin
(2.90)
R
0 sin cos
The corresponding generator of rotations around x-axis is


0 0 0

x =
x
G
R
= 0 0 1
=0
0 1 0

By the same token,

0 0 1
y = 0 0 0 ,
G
1 0 0

0 1 0
z = 1 0 0
G
0 0 0

(2.91)

(2.92)

x, G
y] = G
z etc. Since G
x = iL
x /~, G
y = iL
y /~, G
z = iL
z /~, we confirm the commuIt is easy to verify that [G
tation relations (1.57). We will see later that this only one of infinite number of possible representations of angular
momentum.
Operators of all physical quantities can be classified according to their transformational properties under rotations.
Scalar operator S is an operator that commutes with the operator of rotation:
R
n ] = 0 ,
[S,

L
i] = 0 .
[S,

(2.93)

2 are all scalar operators.


For example, r2 , p2 , p r, L

Vector operator V is an operator whose components satisfy


i , Vk ] = iikl Vl .
[L

(2.94)

(p r)p.

Examples of vector operators: r, L,

Tensor operator of second rank Tik is an operator whose components satisfy


i , Tkl ] = i(ikp nl + iln kp )Tpn .
[L
Examples of tensor operators: ri rk , ri pk .

(2.95)

Kirill Tuchin

29

Phys. 591 Lecture Notes


C.

Time evolution

For infinitesimal time shift t t0 = t +

(t + ) = (t) + (t) =
t

1+
t

(t) .

(2.96)

Due to uniformity of time, the Hamiltonian of a closed system must commute with the operator t . Therefore we
identify the time derivative operator as the energy operator (up to a multiplicative factor). But energy operator is
the Hamiltonian, see (2.1). It is clearly commutes with itself and so energy is conserved.

Operator of finite time shift can is defined and calculated as in (2.79) and (2.80). Using t = iH/~
we have

S (t) = (t + ) = eiH /~ (t) .

(2.97)

Operator S evolves quantum system forward in time. This agrees with (2.58).
Translation, rotation and time evolution are examples of continuous transformations. Invariance of a system
(and hence the Hamiltonian) under such transformations leads to conservation laws both in classical and quantum
mechanics.
D.

Space inversion

Unlike the previous continuous symmetry transformations, space inversion is a discrete transformation. In classical
mechanics it does not lead to any conservation laws, however in quantum mechanics it does. Space inversion reflects
the position vector r r 0 = r, i.e. the right-handed coordinates system goes into the left-handed one. The
corresponding operator is called the parity operator P and is defined as
P (r) = (r) .

(2.98)

Among systems conserving parity are those that interact through electromagnetic and strong forces, and systems in
P ] = 0.
central potentials. Conservation of parity means that [H,
Consider the eigenvalue problem for the parity operator:
P (r) = P (r)

(2.99)

Applying P on both side again and noting that P 2 = 1 (unit operator) we derive (r) = P 2 (r), so P 2 = 1 and
P = 1

(2.100)

are the eigenvalues of the parity operator. The corresponding eigenstates satisfy
P = .

(2.101)

State + ( ) is said to be an even (odd ) parity state.


Suppose that the Hamiltonian of a system is an even function of its coordinates, viz. H(r) = H(r). Then,
P ](r) = H(r)
P (r) P H(r)(r)

[H,
= H(r)(r)
H(r)(r)
= 0,

(2.102)

i.e. the parity operator commutes with the Hamiltonian implying that that the parity is conserved. Wave functions
of the stationary states of a system invariant under the space inversion can be chosen to have a definite parity.
Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, 15,26,30, Merzbacher,
Quantum Mechanics 3rd edition, 17.1,17.2,17.9.

Kirill Tuchin

Phys. 591 Lecture Notes


III.

30

MOTION IN ONE-DIMENSION
1.
A.

Delta-potential

Continuity of the wave function

Wave function must be a continuous function of coordinates in order that its probabilistic interpretation make
sense. Moreover, its first derivative must also be a continuous function, provided that the potential is a non-singular
function. To prove this statement, consider a particle of mass m moving in the following singular one-dimensional
potential
U (x) = V (x) (x x0 ) ,

(3.1)

where and x0 are real numbers, V (x) is a non-singular function. Schrodinger equation reads

~2 00
(x) + [V (x) (x x0 )]E (x) = EE (x) .
2m E

Integrating over x0 x x x0 + x:

Z x0 +x 
Z x0 +x
~2 00
dx
E (x) + [V (x) (x x0 )]E (x) = E
dxE (x) .
2m
x0 x
x0 x

(3.2)

(3.3)

The discontinuity of the wave function at x is


0
0
0
E
(x) = E
(x + x) E
(x x) .

(3.4)

Taking x 0 we derive from (3.3)


0
E
(x0 ) =

2m
E (x0 ) .
~2

(3.5)

This indicates, in particular, that if the potential is continuous, the first derivative of the wave function is continuous
as well.
B.

Discrete spectrum

Consider a particle moving in the delta potential U = (x). The corresponding Schrodinger equation

~2 00
(x) (x)(x)E (x) = EE (x) .
2m E

Suppose E < 0 and introduce a parameter that has dimension of inverse distance (same as wavenumber):
r
2mE
= 2 > 0.
~

(3.6)

(3.7)

Solution to the Schr


odinger equation that is finite at x is
(x) = Aex
(x) = Bex ,

x > 0,
x < 0.

(3.8)
(3.9)

To fix constants A and B we need two conditions. First, we use (3.5):


0 (0) = A() B =

2m
A.
~2

(3.10)

Second, the continuity of the wave function itself at x = 0 implies A = B. Thus,


=

m
~2

(3.11)

Kirill Tuchin

Phys. 591 Lecture Notes

31

There are two possibilities: (i) if < 0, then according to (3.11) < 0 which contradicts (3.7). In this case there are
no bound states. (ii) If > 0, there is one bound state
E0 =

m2
2 ~2
= 2 .
2m
2~

(3.12)

The corresponding wave function is


0 (x) =

C.

0 e0 |x| ,

0 =

m
.
~2

(3.13)

Continuous spectrum

Continuous spectrum lie at E > 0. It is convenient to introduce


r
2mE
k=
> 0.
~2

(3.14)

Solution to the Schr


odinger equation is
E (x) = A1 eikx + A2 eikx ,
E (x) = B1 e

ikx

+ B2 e

ikx

x > 0,

(3.15)

x < 0.

(3.16)

It corresponds to a free particle traveling in positive and negative directions along the x-axis. A usual problem is that
given a flux of particles incident on the potential from, say, x = , find what fraction of particles is reflected from
the potential and what fraction is transmitted. In such a case at x > 0 there are only particles that move toward
x = +, meaning that A2 = 0. It is also convenient to normalize the incident flux of particles to one particle per
unit volume. To this end, we compute the probability current density of incident particles using (2.11):
ji = |B1 |2

~k
= |B1 |2 v .
m

(3.17)

The required normalization is B1 = 1.


Using the continuity of and (3.5) at x = 0 we can now compute A1 and B2 :
A1 =

1
1+

B2 =

m
i~2 k

1
1+

i~2 k
m

(3.18)
.

(3.19)

Fluxes of transmitted and reflected particles are


jt = |A1 |2 v ,
2

jr = |B2 | v .

(3.20)
(3.21)
(3.22)

The corresponding fractions are transmitted and reflected


1
jt
= |A1 |2 =
,
m2
ji
1 + 2~
2E
jr
1
R=
= |B2 |2 =
2E .
ji
1 + 2~
m2

T =

(3.23)
(3.24)

T and R are called the transmission and reflection coefficients respectively. The probability conservation implies
R + T = 1, which is indeed seen in (3.23),(3.24).
Notice that the transmission and reflection coefficients do not depend on the sign of .
Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, 25, Merzbacher,
Quantum Mechanics 3rd edition, 6.1.

Kirill Tuchin

32

Phys. 591 Lecture Notes

U (x)
U0
1

a2

a
2

FIG. 1:

2.

Rectangular potential well


A.

Discrete spectrum

2In the
case of the rectangular potential shown in Fig. 1 all solutions have E 0. Indeed, equation hEi =
p /2m + hU i implies that hEi hU i 0. Thus, we can define

2mE
k=
(3.25)
~
in place of E. Furthermore, bound states correspond to energies E U0 , at which the particle cannot escape to
x .
Another important observation is that the potential is invariant under the inversion U (x) = U (x) implying that

H(x)
= H(x).
In this case the stationary states can be chosen to have definite parity either even or odd, which will
be denoted by plus or minus superscript, see (2.102). As a consequence, it is sufficient to find the wave functions only
at x 0. Schr
odinger equation in the region 2 reads
200 (x) + k 2 2 = 0 ,

0 x a/2 .

(3.26)

It has two independent solutions with definite parity


2+ = B cos(kx) ,

P = +1 ,

(3.27)

= C sin(kx) ,

P = 1 .

(3.28)

300 2 3 = 0 ,

x a/2 ,

(3.29)

In the region 3,

where
p
2m(U0 E)
=
.
~

(3.30)

Its solution finite at x is


3 (x) = Aex ,

x a/2 .

(3.31)

Matching (boundary) conditions for even states:


2+ (a/2) = 3 (a/2) ,
B cos(ak/2) = Ae

20+ (a/2) = 30 (a/2) ,

a/2

kB sin(ak/2) = Ae

(3.32)
a/2

(3.33)

Taking the ratio gives


k tan(ka/2) = =

2mU0
k2 .
~2

(3.34)

Kirill Tuchin

33

Phys. 591 Lecture Notes

Bound states exist only for those values of k = kn , n = 0, 1, . . . , N that satisfy (3.34). There is a finite number of
solutions because according to (3.34)

2mU0 a2
0 ka
.
(3.35)
~
The larger is U0 a2 , i.e. the deeper and wider the the well, the more bound states it contains. This can be seen in
Fig. 2 where I illustrated a numerical solution to this equation. Intersection of blue and red lines give the values of k
that solve (3.34). The red line terminates at the maximal ka satisfying (3.35). Using (3.25) they can be converted to
the corresponding energy levels.

5
5

ka

ka

-5

-5

2
FIG.
p 2: Values of parameters are a = 1 and mU0 /~ = 100. Left panel (even solutions):
p red line is tan(k/2), blue panel
2
2
is 2mU0 /k ~ 1. Right panel (odd solutions): red line is cot(k/2), blue panel is 2mU0 /k2 ~2 1. Intersections are
solutions to (3.34) and correspondingly. In this figure there are three even and two odd states.

Similarly, for odd states, boundary conditions yield


C sin(ak/2) = Aea/2 ,

C cos(ak/2) = Aea/2 .
k

(3.36)

Thus,
k cot(ka/2) = .

(3.37)

Solution to this equation is illustrated in the right panel of Fig. 1.


 Consider the limit U0 a2 at fixed ka, which corresponds to infinitely deep well. In such case, (3.34) for even
states becomes
r
2mU0
tan(ka/2)
 1.
(3.38)
k 2 ~2
The corresponding solutions
kn a = (2n + 1) ,

En+ =

2 (2n + 1)2 ~2
~2 kn2
=
.
2m
2ma2

(3.39)

Even eigenfunctions:
n+ (x)

2
cos
a

(2n + 1)x
a

0 x a/2 .

At x > a/2 eigenfunction vanishes because its argument is

2mU0
n a
a .
~

(3.40)

(3.41)

For odd states


cot(ka/2)  1

kn a = 2n

En =

2 (2n)2 ~2
.
2ma2

(3.42)

Kirill Tuchin

Phys. 591 Lecture Notes

34

The corresponding eigenfunctions are


n (x)

2
sin
a

(2n)x
a

0 x a/2 .

(3.43)

Thus, there is an infinite number of even and odd levels in the infinitely deep potential well.
Notice that (a/2) = 0, which means that particle cannot penetrate the region |x| a/2 because of the infinite
potential. At x = a/2 the derivative 0 is discontinuous.

 In the opposite
limit a U0  ~/ m, which corresponds to shallow and/or narrow well there is only one even
level because k 2mU0 /~ and ka  2. The corresponding energy is
E=

B.

~2 k 2
= U0 .
2m

(3.44)

Three-dimensional well

In three dimensional rectangular potential well U (x, y, z) with sides a, b, c, solution to the Schrodinger equation
can be found employing separation of variables:
n1 ,n2 ,n3 (r) = n1 (x)n2 (y)n3 (z) .
In the infinite well the corresponding energy levels are


n22
n23
2 ~2 n21
+ 2 + 2 ,
En1 ,n2 ,n3 =
2m
a2
b
c
and the wave functions

q
2 cos
qa
n (x) =
2 sin
a

nx
a
nx
a

n1 , n2 , n2 = 1, 2, 3, . . .

n = 1, 3, 5, . . .

n = 2, 4, 6, . . .

(3.45)

(3.46)

(3.47)

If a 6= b 6= c each energy value corresponds to one wave function, i.e. there is no degeneracy. Potential U is invariant
under rotations by around the three coordinate axes and inversion. If a = b = c, then
En1 ,n2 ,n3 =

2 ~2 2
(n + n22 + n23 ) ,
2ma2 1

n1 , n2 , n2 = 1, 2, 3, . . .

(3.48)

Additional symmetry of this case is the symmetry of a cube. This results in degeneracy of the energy levels. For
2 2
~
example, quantum numbers (n1 , n2 , n3 ) = (5, 1, 1), (1, 5, 1), and (1, 1, 5) correspond to the same level 27
2ma2 . Interestingly, there is another level (3, 3, 3) that has the same energy. This additional degeneracy is caused by the particular
way in which potential depends on the coordinates, but not on the symmetry of the field and is called the accidental
degeneracy.
C.

Continuous spectrum

Continuous spectrum lies at E > U0 . In this case we are interested to know which fraction of the flux incident
on the well is reflected and which one is transmitted. Accordingly we write down the solution to the Schr
odinger
equation in the following way
p
1 = Aeik0 x + Beik0 x , x < a/2 , k0 = 2m(E U0 )/~ ,
(3.49)

ikx
ikx
2 = e + e
, a/2 < x < a/2 , k = 2mE/~ ,
(3.50)
3 = Ceik0 x ,

x > a/2 .

(3.51)

Matching conditions at x = a/2 are


Ceik0 a/2 = eika/2 + eika/2 ,

k0 ik0 a/2
Ce
= eika/2 eika/2 .
k

(3.52)

Kirill Tuchin

35

Phys. 591 Lecture Notes

Solving for and we find


(

C
2

1+

C
2

k0
k


k0
k

Matching conditions at x = a/2 are

which together with (3.53) yield

ei(k0 k)a/2

(3.53)

ei(k0 +k)a/2

Aeik0 a/2 + Beik0 a/2 = eika/2 + eika/2 ,



k  ika/2
e
eika/2 ,
Aeik0 a/2 Beik0 a/2 =
k0








C ik0 a
k
k0
k
k0
e
1+
1+
eika + 1
1+
eika
4
k0
k
k0
k







C
k
k
k
k0
B=
1
1+
eika + 1 +
1
eika .
4
k0
k0
k0
k

A=

(3.54)
(3.55)

(3.56)
(3.57)

Transmission coefficient
T =
1+

1
4

1
k0
k

k
k0

2

(3.58)

sin2 (ka)

and the reflection coefficient R = 1 T .


When sin(ka) = 0, i.e. ka = n, the flux of particles is completely transmitted T = 1, R = 0. The corresponding
values of energy
2 2 2
n = ~ n ,
E
2ma2

where n is integer,

(3.59)

n U0
are called the resonance energies or virtual levels. Since E
n2

2a2 mU0
.
2 ~2

(3.60)

Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, 22,25,50. Merzbacher,
Quantum Mechanics 3rd edition, 6.4,7.4. Messiah, Quantum mechanics, VI.10.
3.

Infinite one-dimensional crystal

30

25

20

15

10

-3

-2

-1

FIG. 3: a = 1. Delta functions are approximated by (B12) with  = 0.01.

Consider motion of a particle in a periodic one-dimensional potential


U (x) =

n=

(x na) .

(3.61)

Kirill Tuchin

Phys. 591 Lecture Notes

36

that serves as an elementary model of an ideal infinite crystal. Points xn = na correspond to the lattice sites. To
determine the wave function we need to solve the Schrodinger equation between any two lattice sites, say an < x <
a(n + 1) and apply the matching conditions at the points x = a(n 1) and x = an. Solution to the Schr
odinger
equation reads

2mE
.
(3.62)
(x) = An eik(xna) + Bn eik(xna) , k =
~
Here An , Bn are constants and constant phases eina are written for further convenience (they could have been
absorbed into the An and Bn ). Since the potential is a periodic function of x with the period a, the wave function
must also be periodic with the same period:
(x + a) = (x) ,

(3.63)

where is a complex constant. Suppose that ||2 > 1, then diverges at x and cannot be normalized. Similarly,
if ||2 < 1, it diverges at x . Therefore, ||2 = 1, i.e. constant is a phase. We can parametrize it as
= eiqa ,

qa .

(3.64)

States with fixed q are called Bloch states, while ~q is called the quasi-momentum (dont confuse it with the particle
momentum ~k).
From (3.62) and (3.63) it follows that
An = An1 ,

Bn = Bn1 .

(3.65)

Now we apply the boundary conditions at x = na. For 0 we have


(na + ) = (na ) ,
0 (na + ) 0 (na ) =

(3.66)
2m
(na) .
~2

(3.67)

The wave functions on the each side of x = na are


(na ) = An1 eik(x(n1)a) + Bn1 eik(x(n1)a) ,

(na + ) = An eik(xna) + Bn eik(xna) ,

(3.68)
(3.69)

Using (3.66),(3.67) we get


An1 eika + Bn1 eika = An + Bn ,

(3.70)

2m
ik(An1 eika Bn1 eika ) ik(An Bn ) = 2 (An + Bn ) ,
~

(3.71)

Replacing An1 and Bn1 with An and Bn according to (3.65) we arrive at the following system of equations
An (eika ) + Bn (eika ) = 0 ,






2mi
2mi
An eika 1 + 2
Bn eika 1 2
= 0.
~ k
~ k
Dividing the second of these equations by the first one we obtain a condition on :
h
i
m
2 2 cos(ka) + 2 sin(ka) + 1 = 0 .
~ k

Being a quadratic equation it has two solutions 1 and 2 :


p
m
1,2 = f (E) f 2 (E) 1 , f (E) = cos(ka) + 2 sin(ka) .
~ k

(3.72)
(3.73)

(3.74)

(3.75)

In view of (3.64) f 2 1 for otherwise is not a pure phase. It is easy to verify that 1 2 = 1 1 = 1.
We can rewrite (3.74) using (3.64) as follows
f=

2 + 1
= cos(qa) .
2

(3.76)

Kirill Tuchin

37

Phys. 591 Lecture Notes


fHEL
3

cos(qa)

1st zone

ka

3rd zone

2nd zone

-1

FIG. 4: Solution to (3.77) with cos(qa) = 0.6 (red line) and m/~2 = 2. f (E) is a blue line. Intersection of blue and red lines
are solutions to (3.77). Physical values of f satisfy f 2 1 and lie between the two horizontal black lines which divide possible
values of ka into the Brillouin zones. In each zone there is one solution kn a (red circles) corresponding to the energy level En .

Now employing (3.75) we obtain


cos(qa) = cos(ka) +

m
sin(ka) .
~2 k

(3.77)

Solutions to this equations is a set of discrete values kn and the corresponding energy levels En . They are depicted
in Fig. 4.
Solution to the Schr
odinger equation have form (3.61) even at x indicating that a particle is not localized
and can move in the entire crystal with quasi-momentum ~q.
Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, , Merzbacher, Quantum Mechanics 3rd edition,
4.

Harmonic oscillator
A.

Energy spectrum

In many physical problems a physical system moves near a minimum of potential well U (x). Suppose that this
minimum is at x = 0 and expand
1
U (x) U (0) + U 00 (0)x2 .
2

(3.78)

The corresponding force is U 00 (0)x. In classical mechanics motion of a particle under such force is governed by the
equation m
x = U 00 (0)x describing periodic motion with frequency = (U 00 (0)/m)1/2 . In quantum mechanics we
are interested to solve the following Schr
odinger equation


~2 d2
2 m 2

+
x E (x) = 0 .
(3.79)
2m dx2
2
Introduce dimensionless variables
=x

m
,
~

=

2E
.
~

(3.80)

Kirill Tuchin
We have

Phys. 591 Lecture Notes

38


d2
2
+  () = 0 .
d 2

(3.81)

At , we can neglect . The resulting equation 00 = 2 has asymptotic solution e


look for a solution in the form
() = v()e

/2

/2

. Thus, we can
(3.82)

Substituting and denoting  1 = 2n we get


v 00 2v 0 + 2nv = 0 .

(3.83)

We can find a solution to this equation in the form of the power series
v() =

ak k .

(3.84)

k=0

Upon substitution

k=2

ak k(k 1)

k2

kak + 2n

k=0

ak k = 0 .

(3.85)

k=0

Changing the summation variable in the first sum as k 0 = k 2 and then renaming it again to k 0 k we obtain

k=0

ak+2 (k + 2)(k + 1) k 2

kak k + 2n

ak k = 0 .

(3.86)

2(k n)
ak .
(k + 2)(k + 1)

(3.87)

k=0

k=0

This equations is satisfied if coefficients in front of same powers vanish, viz.


ak+2 (k + 2)(k + 1) 2kak + 2nak = 0 ,

ak+2 =

If we start with a finite a0 and a1 = 0 we get a polynomial containing only even powers of . However, if we start
with a finite a1 and a0 = 0 we get an odd polynomial. This reflects the fact that the Hamiltonian of the harmonic
oscillator is invariant under the spatial inversion, and thus we can choose its stationary states to have definite parity.
It can be shown (see Merzbachers book) that () diverges at unless the polynomials are finite, which
requires n to be integer. Since E 0, n 1/2 implying that n = 0, 1, . . .. For illustration consider n = 2:
a2 =

2 (2)
a0 = 2a0 ,
21

a4 = a6 = . . . = 0 .

(3.88)

The corresponding polynomial is H2 () = (2 2 1)a0 . Consider n = 3.


a3 =

2 (1 3)
2
a1 = a1 ,
32
3

a5 = a7 = . . . = 0 .

(3.89)

The corresponding polynomial is H3 () = ( 32 3 + 1)a1 . Coefficients a0 and a1 are chosen for each n in such a way
that the resulting polynomials satisfy the following formula
2

Hn () = (1)n e

dn e
,
d n

(3.90)

and are known as the Hermit polynomials. The lowest polynomials are
H0 = 1 ,

H1 = 2 ,

H2 = 4 2 2 ,

H3 = 8 3 12 ,

(3.91)

Let me note for the future reference the following recurrence relations
1
Hn () = nHn1 () + Hn+1 () ,
2
dHn
= 2nHn1 () .
d

(3.92)
(3.93)

Kirill Tuchin

Phys. 591 Lecture Notes

To wave functions normalized as

39

n (x)m (x)dx = mn

(3.94)

 r

2
1
m

emx /(2~) .
Hn x
n
~
2 n!

(3.95)

read
n (x) =
The corresponding energy levels

 m 1/4
~

1
E = ~ n +
2

(3.96)

There is only one eigenfunction for each n, so there is no degeneracy. Also note that at n = 0 the oscillator has energy
E0 = ~/2 known as the zero-point energy and is interpreted as the energy of vacuum fluctuations.
Since the Hamiltonian of the harmonic oscillator is invariant under spatial inversion, the stationary states have
definite parity, viz. states with even n are even functions of x, while states with odd n are odd.
Knowledge of the stationary states of the harmonic oscillator allows us to calculate various observable quantities.
As an example, let us compute the expectation value of the particle position in the nth stationary state.
Z
(3.97)
n2 xdx = 0
hxin =
because x is an odd function, while

n2

is even. Variance of the position is


Z


x n = x2 n =
n2 x2 dx .

(3.98)

The easiest way to take this integral is to employ the orthonormality condition (3.94). To this end we substitute
(3.95) into (3.92) and obtain after some simple algebra
r
r
n
n+1
n =
n1 +
n+1 .
(3.99)
2
2

Using this equations again we get


2 n =
Thus,



1p
1
1p
n(n 1)n2 + n +
(n + 1)(n + 2)n+2 .
n +
2
2
2

2
~
x =
m

n2 2 dx

~
=
m



1
n+
.
2

For future reference let me record a useful relation that follows from (3.93):


n
1
=
nn1 n + 1n+1 .

2
B.

(3.100)

(3.101)

(3.102)

Ladder operators

Eigenvalue problem for the harmonic oscillator can be reformulated in terms of the ladder operators (6
r


m
i
a
=
x
+
p ,
2~
m
r


m
i
x

p .
a
=
2~
m

(3.103)
(3.104)

Properties of the ladder operators:

(6

In fact, this problem can be solved entirely employing these operators without solving first the Schr
odinger equation , see Merzbachers
Quantum Mechanics, 10.6.

Kirill Tuchin

40

Phys. 591 Lecture Notes

1. a
and a
are not Hermitian, i.e. they do not correspond to any physical quantities. Only their certain combinations, such as (3.108),(3.109) and (3.110) are Hermitian.
2. The ladder operators satisfy the following commutation relations:
[
a, a
] = 1 ,

[
a , a
] = 0 ,

[
a, a
] = 0 ,

(3.105)

which can be derived using (3.103),(3.104) and the commutators [


x, px ] = i~ etc.
3. To verify the physical meaning of the ladder operators, apply them to the wave functions of the stationary
states. Using (3.95) and (3.93) we derive
r



m
~ d
x+
n = nn1 ,
(3.106)
a
n =
2~
m dx
r



m
~ d

(3.107)
a
n =
x
n = n + 1n+1 .
2~
m dx
The raising operator a
moves the particle to the next higher excited state, while the lowering operator a
moves
the particle one state lower.
Equations (3.103),(3.104) can be used to express x
and px through the ladder operators:
r
~
x
=
(
a+a
) ,
2m
r
m~
(
a a
) ,
px = i
2

(3.108)
(3.109)

In terms of the ladder operators the Hamiltonian reads


2
2
2
 2
 ~  2

= px + mx = ~ a
H
+ (
a )2 + a
a
+ a
a
+
a
+ (
a )2 a
a
a
a
=
2m
2
4
4



1
a
a
+
~ ,
2

(3.110)

where I used (3.105). Evidently, operator a


a
commutes with the Hamiltonian. Employing (3.106) abd (3.107) we
find its eigenvalue in the nth stationary state:
a
a
n = nn .
(3.111)

Thus we reproduce the spectrum (3.96): E = ~ n + 12 .
The ladder operator method can be applied to calculate the eigenfunctions n as follows. Let 0 be the wave
function of the ground state. Then, according to (3.106)
r


m
~ d
a
0 =
x+
0 = 0 .
(3.112)
2~
m dx
Solution to this differential equation is
2

0 = N0 emx

/2~

(3.113)

Constant N0 is fixed by the normalization condition. It is easy to see from (3.107) that the wave functions of the
excited states can be obtained as follows
1
n = (
a )n 0 .
n!

(3.114)

You can verify that it gives the same result as (3.95).


Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, 23, Merzbacher,
Quantum Mechanics 3rd edition, Ch. 5, 10.6.

Kirill Tuchin

41

Phys. 591 Lecture Notes


5.

General properties of one-dimensional motion

We can now generalize a few examples that we considered in the previous sections and derive the general properties
of one-dimensional motion. We start with the Schrodinger equation
d2 2m
+ 2 [E U (x)] = 0 .
dx2
~

(3.115)

1. Energy levels of discrete spectrum are non-degenerate. Indeed, suppose the contrary is true, viz. there are two
linear independent eigenfunctions corresponding to the same energy levels:
100
00
2m
= 2 [U (x) E] = 2 ,
1
~
2

100 2 = 200 1 .

(3.116)

Integating by parts
Z
Z
Z
(100 2 200 1 )dx = 10 2 10 20 dx 20 1 + 20 10 dx = 10 2 20 1 .
{z
}
|

(3.117)

const

Recall that for a bound state 1,2 0 as x , hence const=0 and


10
0
= 2.
1
2

(3.118)

Integrating again
const =

Z 

10
0
2
1
2

dx =

Z 


d
d
ln 1
ln 2 dx = ln 1 ln 2 .
dx
dx

(3.119)

Thus, 1 2 and upon normalization we obtain that 1 = 2 .


2. Let the discrete spectrum of the Hamiltonian be E0 < E1 < . . . and the corresponding eigenfunctions 0 , 1 , . . .
correspondingly. The oscillation theorem states that eigenfunction n has n zeros, i.e. it vanishes n times, not
including the boundary values. You can find a proof in the book of Messiah. As an illustrate to this theorem I
plotted in Fig. 5 a few wave functions of the infinitely deep potential well.
n

1.0

0.5

-0.4

-0.2

0.2

0.4

-0.5

-1.0

FIG.
5: Wave functions of thepthree lowest states in the infinitely deep potential well: 0 =
p
2/a sin(2x/a) (blue), 2 = 2/a cos(3x/a) (red). a = 1.

p
2/a cos(x/a) (black), 1 =

3. Consider potentail well of a general form shown in Fig. 6.


(a) Discrete spectrum corresponds to motion in a finite region of space. For a potential in Fig. 6 this is motion
with energies
in the interval Umin < E < 0. (Motion with energies below Umin is not possible because
hEi = hU i + p2 /2m hU i Umin ).

Kirill Tuchin

42

Phys. 591 Lecture Notes


UHxL

U0

1.0

0.5

-3

-2

-1

Umin

-0.5

FIG. 6: An arbitary potential well U (x).

(b) In the region 0 < E < U0 the spectrum is continuous because particle can go to x . The eigenstates
are not degenerate, which can be proven as before using the fact 1,2 0 as x .

2m
2mE
at x + : 00 + 2 E = 0 , = A cos(kx + ) , k =
.
(3.120)
~
~
p
2m(U0 E)
2m
at x : 00 2 (U0 E) = 0 , = Bex + Cex , =
.
(3.121)
~
~
The boundary condition requires that C = 0. Therefore, there is only one function that satisfies the
boundary conditions at given E.
(c) At E > U0 the spectrum is continuous and two-fold degenerate:
ikx

ikx

at x + :

= Ae

at x :

= Ceik x + Deik x ,

+ Be

2mE
,
~
p
2m(E U0 )
k0 =
.
~

k=

(3.122)
(3.123)

There are two solutions that satisfy the boundary conditions one with particles incident from the left
and another from the right.
(x) = U (x) + U (x) with U (x) 0. Let En and n () be eigenvalues and
4. Consider two potentials U (x) and U
eigenfunctions of the discrete spectrum of the following Hamiltonian
p2

H()
=
+ U (x) + U (x) ,
2m
where is a parameter. In one of the home assignments you proved the following theorem:
Z

En ()
H
= n ()
n ()dx

(3.124)

(3.125)

Plugging (3.124) into (3.125) we have


En ()
=

|n ()|2 U (x)dx 0 .

(3.126)

= p2 /2m+U , while En (1) = E


n is an eigenvalue of H
= p2 /2m+ U
.
Note that En (0) = En is an eigenvalue of H

We conclude that En En . This means that a more shallow potential has higher lying energy levels. This
statement can be generalized to a multidimensional potentials.
As U increases, energy levels go up, so that eventually the highest energy level (the one that has the smallest
|En |) becomes En = 0. A further increase of U will push this level into the continuum spectrum. Thus, in
order to determine at what values of parameters of the Hamiltonian a new bound states appears/disappears one
has to solve the Schr
odinger equation with E = 0.

Kirill Tuchin

Phys. 591 Lecture Notes

43

Example: rectangular potential well shown in Fig. 1. Note, that the y-axis in Fig. 1 is shifted by U0 as
compared to
Fig. 5. This is why the continuous spectrum starts at E = U0 . The corresponding values of k and
are k0 = 2mU0 /~ and 0 = 0. Consider even states, (3.34) implies


k0 a
k0 tan
= 0 = 0 , k0 a = 2n , n = 0, 1, 2, . . .
(3.127)
2
The ground state exists if k0 a 0, which is always satisfied. The first excited even state exists if k0 a > 2, etc.
The nth excited even state exists if k0 a > 2n. The total number of even states


ak0
+ 1.
(3.128)
N+ =
2
Square brackets indicate the integer part. Similarly, odd states exist when k0 a = (1 + 2n), n = 0, 1, . . .. If
k0 a < there are no odd states. The total number of odd states is




ak0
1
ak0
1
N =

+1=
+
(3.129)
2
2
2
2
Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, 21, Messiah, Quantum Mechanics, Sec. 12.
6.

Integral form of Schr


odinger equation

Eigenvalue problem for the Hamiltonian can be written down as an integral equation for the wave function. I will
consider separately the discrete and continuous spectra.
A.

Discrete spectrum

We start with the Schr


odinger equation

~2 00
(x) + U (x) = E(x) ,
2m

(3.130)

and assume that U is a potential well such that U 0 as x and consider the bound states corresponding to
E < 0. Introduce the Greens function GE (x, x0 ) satisfying

~2 d2
GE (x, x0 ) EGE (x, x0 ) = (x x0 ) .
2m dx2

Suppose that we solved (3.131) and found GE (x, x0 ). Then a solution to (3.130) reads
Z
(x) = Aex + Bex
GE (x, x0 )U (x0 )(x0 )dx0 ,

(3.131)

(3.132)

where A, B are constants and


=

2mE
.
~2

(3.133)

Indeed,




~2 d 2
~2 d 2

E (x) =
E (Aex + Bex )
2m dx2
2m dx2

Z + 
~2 d2 GE (x, x0 )
0

EG
(x,
x
)
U (x0 )(x0 )dx0
E
2m
dx2

Z +
=
(x x0 )U (x0 )(x0 )dx0

= U (x)(x) .

(3.134)

Kirill Tuchin

44

Phys. 591 Lecture Notes

Note, that although (3.132) is a formal solution of the Schrodinger equation , it is still an equation for . Since we
are interested in bound states which have wave functions localized at x we set A = B = 0.
The Greens function can be calculate as described in III 1B (7 . We write

0
C(x0 )e(xx ) , x < x0 ,
0
GE (x, x ) =
(3.135)
0
D(x0 )e(xx ) , x > x0 .
and use boundary conditions (continuity of GE (x, x0 ) at x = x0 and discontinuity its derivative given by (3.5)) to fix
C and D:
C = D,

(D + C) =

2m
,
~2

C=D=

m
~2

(3.136)

Thus,
GE (x, x0 ) =

m |xx0 |
e
.
~2

Finally, Schr
odinger equation in the integral form is
Z +
0
m
E (x) = 2
e|xx | U (x0 )E (x0 )dx0 .
~

(3.137)

(3.138)

Subscript E indicates that we used the boundary condition for the bound states with E < 0.
Example. For a particle moving in the delta-potential well U = (x) we have
E (x) =

m
E (0)e|x| ,
~2

(3.139)

which is the eigenfunction of the only bound state. This equation must be consistent at x = 0 implying that
1=

m
,
~2

(3.140)

which coincides with the result of III 1B.


B.

Continuous spectrum

To obtain the Greens function at E > 0 we notice that


r
2mE
= 2 = ik ,
~

k=

2mE
.
~2

(3.141)

Therefore, instead of (3.137) we obtain


0
G
E (x, x ) =

im ik|xx0 |
e
.
~2 k

(3.142)

Schr
odinger equation for scattering problem where particles with momentum p = ~k are incident from the left is
equivalent to the following integral equation
Z
+ 0
0
0
0
k+ (x) = eikx G+
(3.143)
E (x, x )U (x )k (x )dx .
The first term in the right-hand-side corresponds to the incident particles, while the second term describes the reflected
and transmitted waves. Indeed, U (x0 ) in the integrand is non-vanishing only at x . a, where a is the range of the

(7

The boundary condition (3.5) has to be slightly modified: 0 (x0 ) = 2m/~2 because of the difference between (3.2) and (3.131).

Kirill Tuchin

Phys. 591 Lecture Notes

45

potential. At |x|  a we can approximate |x x0 | |x| so that asymptotic form of the solution to the Schr
odinger
equation becomes:

( ikx  im R
x + ,
e ~2 k U (x0 )k+ (x0 )dx0 eikx = A1 eikx ,
(3.144)
k+ (x)


R
U (x0 )k+ (x0 )dx0 eikx = eikx + B2 eikx , x .
eikx ~im
2k

Compare this with (3.15),(3.16). Schr


odinger equation in the integral form (3.143) is the starting point of the scattering
theory.
Example. In potential U = (x) (3.143) reads
k+ (x) = eikx

im ik|x| +
e
k (0) .
~2 k

(3.145)

At x = 0 this implies
k (0) =

~2 k

~2 k
.
+ im

This coincides with the wave functions given by (3.15),(3.16), (3.18),(3.19).

(3.146)

Kirill Tuchin

Phys. 591 Lecture Notes


IV.

46

REPRESENTATION THEORY
1.

Representations of states
A.

Hilbert space

Consider a quantum system described by a set of quantum numbers a. Let the corresponding wave function be
a (q). Position q is an eigenvalue of a Hermitian position operator q. For a single particle q is the same as r. From
a mathematical standpoint there is nothing special about the operator q and its eigenvalues q. Therefore, one can
describe the same quantum system by different functions a (p), a (En ) etc., which depend on particle momenta,
energy, etc. These are called representations and the mathematical framework that describes the relationships between
different representations is called the representation theory.
In any representation, expectation values of physical quantities must be the same. It is therefore convenient to
think about an abstract state vector |ai, called ket, that incorporates the physical information about the state of the
system but does not dependent on a particular representation. All possible state vector form an abstract vector space.
Superposition principle dictates that |ai + |bi is also a possible vector state. In other words, the vector space is
linear.
For every ket |ai we introduce a dual state vector bra ha| such that |ai = ha|. We also define the scalar product of
two state vectors |ai and |bi as hb|ai. Clearly, ha|bi = hb|ai. Linear vector space with defined scalar product is called
the Hilbert space.
Similarly to the more familiar Euclidean space, we can project a state vector |ai onto any complete set of linearly
independent basis vectors. In Euclidean space such projection are called coordinates; any vector A is completely
described by a set of ordered coordinates. It is convenient to choose the orthonormal basis vectors. By the same
token, the infinite dimensional Hilbert space can be span by a set of orthonormal basis state vectors. Since spectrum
of any Hermitian operator is complete and orthonormal, it can be chosen as a set of basis state vectors. Unlike the
Euclidean space where projections are real numbers, projections in Hilbert space are functions.
For example, lets take spectrum of the position operator r as a basis set in the Hilbert space. Denote the basis
vectors by |ri. Projections of a state vector |ai onto this basis states are hr|ai. They specify the wave functions of a
system in state with quantum numbers a at any point r. In other words, hr|ai = a (r).
B.

Discrete (energy) representation

Choose as a basis set of functions the eigenfunctions of the Hamiltonian. I will assume that the energy spectrum is
discrete; each eigenstate can be labeled by a set of numbers En . Denote by n (q) = hq|En i the stationary state in the
coordinate representation and by n (q) = hEn |qi = hq|En i its complex conjugate. Orthonormality of {n (q)}n , viz.
Z
m (q)n (q)dq = mn
(4.1)
implies
Z

dqhEm |qihq|En i hEm |En i = mn .

(4.2)

To change from the coordinate representation a (q) = hq|ai of a state vector |ai to energy representation we expand
in a set of functions {n (q)}n (in two equivalent notations)
X
a (q) =
n (q)a (En ) ,
(4.3)
n

hq|ai =

X
n

hq|En ihEn |ai .

(4.4)

Here a (En ) = hEn |ai is the wave function of state q in the energy representation. It has a simple physical meaning:
the probability to find the system with energy En is Pn = |a (En )|2 = |hEn |ai|2 .
In q-representation the normalization condition reads
Z
dqha|qihq|ai = 1 .
(4.5)

Kirill Tuchin

Phys. 591 Lecture Notes

47

Using
ha|qi =
hq|ai =

ha|En ihEn |qi ,

(4.6)

hq|En ihEn |ai ,

(4.7)

we obtain
X
n

ha|En ihEn |ai

X
n

|a (En )|2 = 1 ,

i.e. wave functions in the energy representation are also normalized.


Transformation inverse to (4.3) is (in two notations)
Z
a (En ) = dqn (q)a (q) ,
Z
hEn |ai = dqhEn |qihq|ai
C.

(4.8)

(4.9)
(4.10)

Continuous (momentum) representation

The basis functions in momentum representation are the eigenfunctions of the momentum operator p (q) = hq|pi.
We assume that the momentum spectrum is continuous. Orthonormality condition:
Z
dqp0 (q)p (q) = (p0 p) ,
(4.11)
Z
dqhp0 |qihq|pi = hp0 |pi = (p0 p) .
(4.12)
Expanding the wave function a (a) of a state a in in a complete set of momentum eigenstates we get:
Z
a (q) = dpp (q)a (p) ,
Z
hq|ai = dphq|pihp|ai ,

(4.13)
(4.14)

where we denoted a (p) = hp|ai the state vector |ai in the p-representation. Probability density reads
(p) = |a (p)|2 = |hp|ai|2 .

(4.15)

Inverse transformation:
hp|ai =
D.

dqhp|qihq|ai .

(4.16)

General case

In summary, given state vector |ai in m-representation hm|ai we can transform it to -representation as
X
h|ai =
h|mihm|ai ,

(4.17)

where hq|mi are eigenfunctions of m


in -representation. Conversely,
X
hm|ai =
hm|ih|ai ,

(4.18)

Kirill Tuchin

Phys. 591 Lecture Notes

where hm|i is an eigenfunction of in m-representation.


Notice a convenient identity (in discrete and continuous forms):
Z
X
|mihm| = 1 ,
dp|pihp| = 1 .

48

(4.19)

Example. Consider a set of eigenfunctions of momentum operator in coordinate representation:


p (r) = hr|pi =

1
eipr/~ .
(2~)3/2

(4.20)

Then a set of eigenfunctions of position operator r in momentum representation is


hp|ri =

1
eipr/~ .
(2~)3/2

(4.21)

Additional reading: Merzbacher, Quantum Mechanics 3rd edition, 9.1,9.2,3.2.


2.

Representations of operators

Action of a Hermitian operator F on the state vector |ai is defined as


F |ai = |bi ,

(4.22)

where |bi is another state vector from the same Hilbert space.
hb| = (F |ai) = ha|F = ha|F

(4.23)

Thus, we define F to act on bra from the right. The set of eigenstates |Fm i of F is complete, which means that we
can expand
X
|ai =
|Fm ihFm |ai .
(4.24)
m

Similarly, one can expand an operator A as follows


A =

X
m,n

Amn |Fm ihFn | .

(4.25)

By orthogonality of |Fm i the coefficients Amn read


ni .
Amn = hFm |A|F

(4.26)

In particular, the unity operator is given by


1 =

X
m

|Fm ihFm |

(4.27)

in agreement with (4.19).


 In coordinate representation F is a function of qs and /qs.
b (q) = F a (q) ,

or hq|bi = F hq|ai .

Lets determine the form of F in the energy representation. Noting that


X
hq|ai =
hq|En ihEn |ai ,

(4.28)

(4.29)

hq|bi =

X
n

hq|En ihEn |bi ,

(4.30)

Kirill Tuchin

49

Phys. 591 Lecture Notes

and substituting into (4.28) we get


X
n

hq|En ihEn |bi = F

X
n

hq|En ihEn |ai

(4.31)

Now, multiply both sides by hEm |qi and integrate over all q using (4.2):
X
XZ
XZ
dqhEm |qiF hq|En i hEn |ai =
Fmn hEn |ai ,
dqhEm |qihq|En ihEm |bi =

(4.32)

where
Fmn = hEm |F |En i =

dqhEm |qiF hq|En i .

Finally, action of operator F in energy representation reads


X
hEm |bi =
hEm |F |En ihEn |ai .

(4.33)

(4.34)

Thus, if energy levels are not degenerate, then operator F is represented by an infinite dimensional square matrix
with elements Fmn . Indexes m and n label rows and columns correspondingly. The state vector hEm |bi is represented
by a column, so that the action of operator becomes equivalent to matrix multiplication. R
R
If operator F is Hermitian the corresponding matrix Fmn is Hermitian as well. Indeed if F dq = F dq
then,
Z
Z

Fmn = m F n dq = n F m
dq = Fnm
.
(4.35)

In particular, diagonal matrix elements are real Fnn = Fnn


. Any operator is represented by a diagonal matrix in its
n i = En mn .
own representation. For example, in the energy representation hEn |H|E

 The form of operator F in momentum representation can be obtained in a similar way. We start with the
expansions
Z
hq|ai = dphq|pihp|ai ,
(4.36)
Z
hq|bi = dphq|pihp|bi .
(4.37)

which we substitute into the definition


hq|bi = F hq|ai .
Multiplying both sides with hp0 |qi and integrating over q using (4.12) yields
Z
0
hp |bi = dphp0 |F |pihp|ai ,
with
hp0 |F |pi =

dqhp0 |qiF hq|pi .

(4.38)

(4.39)

(4.40)

Eqs. (4.39),(4.40) describe action of F on state vectors in momentum representation.


Examples.
1. In coordinate representation momentum operator is p = i~x . In momentum representation the same operator
reads using (4.40)
Z
0
hp |
p|pi = dxhp0 |xi
phx|pi .
(4.41)

Kirill Tuchin

Phys. 591 Lecture Notes

By definition, phx|pi = phx|pi. Thus,


hp0 |
p|pi = p

dxhp0 |xihx|pi = p(p0 p)

which means that the matrix elements are diagonal as expected. Now,
Z
hp0 |bi = dp p (p0 p)hp|ai = p0 hp0 |ai ,

50

(4.42)

(4.43)

which indicates that action of p in momentum space is simple multiplication.


2. In momentum representation the form of the position operator x
is found from
Z
hp0 |
x|pi = dxhp0 |xi
xhx|pi .

(4.44)

Recalling (4.20) and using its one-dimensional analogue


hx|pi =

1
eipx/~
2~

we derive x
hx|pi = i~p hx|pi. Thus, the matrix element of momentum operator is
Z
hp0 |
x|pi = i~p dxhp0 |xihx|pi = i~p (p0 p) .
Its action on a state vector reads
hp0 |bi = i~

dphp|ai

(p0 p) = i~ 0 hp0 |ai .


p
p

(4.45)

(4.46)

(4.47)

Therefore,

.
p

(4.48)

[
x, px ] = i~

(4.49)

2
= p + U (i~p ) .
H
2m

(4.50)

p2

hp0 |H|pi
=
(p0 p) + U (i~p )(p0 p) .
2m

(4.51)

x
= i~
It is easy to check that

holds in momentum representation.


3. Hamiltonian in momentum representation reads

The corresponding matrix element

4. Consider a wave function on a unit sphere (, ) = h, |i. We can change to the angular momentum
representation in which L2 and Lz have definite values. Lets define the state vector |l, mi to be an eigenstate
2 |l, mi = ~2 l(l + 1)|l, mi and L
z |l, mi = ~m|l, mi. Then,
of the corresponding operators, i.e. L
X
h, |i =
h, |l, mihl, m|i
(4.52)
l,m

We have seen functions h, |l, mi = Ylm (, ) before in I 6C. Functions hl, m|ai are the wave functions in the
angular momentum representation. They are given by
Z
hl, m|i = dhl, m|, ih, |i .
(4.53)

Kirill Tuchin

51

Phys. 591 Lecture Notes

As a more specific example, let


(, ) =

3
sin sin .
4

(4.54)

We can expand it in spherical harmonics as follows


i
(, ) = (Y1,1 + Y1,1 ) .
2

(4.55)

Hence, function in the angular momentum representation is


i
h1, 1|i = h1, 1|i = ,
2

(4.56)

and all other hl, m|is vanish. A compact way to represent this state vector is to depict it as a column in an
abstract the three-dimensional space span by the orthonormal state vectors |1, 1i, |1, 0i, |1, 1i:

i 1
0
h1, 1|i =
.
(4.57)
2 1
Additional reading: Merzbacher, Quantum Mechanics 3rd edition, 9.39.4.
3.

Eigenvalue problem in matrix form

In continuous representations eigenvalue problem for operator F can be written down as a differential equation.
For example, in the q-representation
F F (q) = F F (q) .

(4.58)

In a discrete representation the eigenvalue problem can be formulated as a matrix equation. To derive it expand an
eigenfunction F (q) in a discrete, say, energy representation:
X
F (q) = hq|F i =
hq|En ihEn |F i .
(4.59)
n

Substitute into (4.58):


F

X
X
hq|En ihEn |F i = F
hq|En ihEn |F i .
n

(4.60)

Now multiply by the function hEm |qi and integrate over all q:
Z
Z
X
X
dqhEm |qiF
hq|En ihEn |F i = F dqhEm |qi
hq|En ihEn |F i .

(4.61)

Using the definition (4.33) and (4.2) we can write


X
X
Fmn hEn |F i = F
mn hEn |F i ,

(4.62)

or, equivalently
X
n

{Fmn F mn } hEn |F i = 0 .

(4.63)

Here F (En ) = hEn |F i is the wave function in the E-representation. There are non-trivial solutions to (4.63) if
det {Fmn F mn } = 0 .

(4.64)

Kirill Tuchin

Phys. 591 Lecture Notes

52

Since this determinant is infinite-dimensional, it has infinitely many solutions F1 , F2 , . . . which are eigenvalues of F .
Once Fn s are known, substitute them into (4.63) to determine the wave functions hEn |F i.
It is convenient to represent functions hEn |F i as a column

hE1 |F i
hE2 |F i .
(4.65)
...
Then (4.64) can be written as


F11 F
F12
F13

F22 F
F23
F21

F32
F33 F
F31
...
...
...


...

...
=0
...
...

(4.66)

Thus, the eigenvalue problem in a discrete representation is equivalent to the problem of matrix diagonalization.
As has been mentioned above, any operator is diagonal in its own representation. Indeed,
Z
Z
Fmn = dqhFm |qiF hq|Fn i = dqhFm |qiFn hq|Fn i = Fn mn .
(4.67)
Example: Harmonic oscillator in the energy representation. Matrix elements of the position operator x
in the
energy representation read according to (4.33)
Z +

xmn =
m
(x)xn (x)dx ,
(4.68)

where n s are given by (3.95). Using the identity (3.99) we obtain that the matrix elements are
r

~
xm,n =
( nm,n1 + n + 1m,n+1 ) .
2m
Similarly,
Z +

(x2 )mn =
m
(x)x2 n (x)dx

(4.69)

(4.70)

i
p
~ hp
n(n 1)m,n2 + (2n + 1)m,n + (n + 1)(n + 2)m,n+2 .
2m
Z +
Z +

2
2

=
m (x)(i~x ) n (x)dx = ~
m
(x)n00 (x)dx
=

(p2 )mn

(4.71)
(4.72)

i
p
~m hp
n(n 1)m,n2 (2n + 1)m,n + (n + 1)(n + 2)m,n+2 .
2


(p2 )mn
m 2 (x2 )mn
1
Hmn =
+
= ~ n +
m,n = En mn .
2m
2
2

(4.73)
(4.74)

Eq. (4.74) is a particular case of (4.67): the Hamiltonian is diagonal in the energy representation.
Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, 11,23.
4.

Unitary operators

Transformation of the wave function between the representations can be thought of as an action of an certain oper (8 . Consider, for example a transformation from a discrete F -representation to the discrete E-representation:
ator U
X
X
(E) .
(F ) = hFm |i =
hFm |En ihEn |i =
Umn hEn |i = U
(4.75)
n

(8

Dont confuse it with the operator of potential energy.

Kirill Tuchin

53

Phys. 591 Lecture Notes

The second equations is a definition of Umn s. In particular we can take |i = |Fk i in which case (4.75) becomes
X
X
X

hFm |Fk i = mk =
Umn hEn |Fk i =
Umn Ukn
=
Umn Unk
,
(4.76)
n

In matrix notation this reads as


U
= 1 ,
U

(4.77)

= U
1 .
U

(4.78)

or

Such operators are called unitary, see II 4. This definition can be generalized to continuous representations as well,
in which case matrix elements Umn become kernel of integral transformation.
To find how an arbitrary operator P transforms from E to F -representation consider its action of some state:
0
(E) = P (E)(E). Substituting (4.75) we get
1 0 (F ) = P (E)U
1 (F )
U

P (E)U
1 (F ) .
0 (F ) = U

(4.79)

So, if we define operator P (F ) in F -representation as 0 (F ) = P (F )(F ), then


P (E)U
1 .
P (F ) = U

(4.80)

This guarantees that the expectation values of physical quantities are the same in different representations:
P (F )U
|(E)i = h(E)|P (E)|(E)i = hP (E)i .
hP (F )i = h(F )|P (F )|(F )i = h(E)|U

(4.81)

We see, that every physical quantity can be represented by an infinite number of operators, differing from one
another by unitary transformations.
 Transformation between different representations in the Hilbert space is of course not the only example of unitary
operators. For example, transformation between difference reference frames in Euclidean space and time evolution
are also described by the unitary operators as we saw in II 5 and II 4. The main property of the unitary operator is
that it preserves the lengths of vectors and the angles between them. In the Euclidean space these are literally
lengths and angles, in the Hilbert space these are normalization and scalar products of the state vectors.
Here are some properties of the unitary transformations.
1. Invariance of the scalar product:
h(E)| 0 (E)i = h(F )| 0 (F )i .

(4.82)

hm (F )|P (F )|n (F )i = hm (E)|P (E)|n (E)i

(4.83)

P (E)U
1 } = tr{P (E)U
1 U
} = tr{P (E)} .
tr{P (F )} = tr{U

(4.84)

2. Invariance of matrix elements

can be proved as (4.81).


3. Invariance of trace:

4. Invariance of the eigenvalue problem P (E)(E) = P (E)


1 P (F )U
U
1 (F ) = P U
1 (F )
U

P (F )(F ) = P (F ) .

(4.85)

5. If operator P (E) is Hermitian, then operator P (F ) is also Hermitian. Indeed,


P (E)U
1
P (F ) = U

1 ) P (E)U
= U
P (E)U
= P (F ) .
P (F ) = (U

(4.86)

Kirill Tuchin

54

Phys. 591 Lecture Notes

6. Transformation of commutation relations:


1
1 1
1
1

1
[P (E), Q(E)]
=U
| {z U} Q(E) U
| {z U} |U {z U} Q(E) |U {z U} P (E) |U {z U}
| {z U} P (E) U

1
1

=U

)U
U
P (F )Q(F

)P (F )U
=U
Q(F

)]U
.
[P (F ), Q(F

(4.87)

7. Any unitary operator can be represented as

S = eiF ,

with

F = F ,

(4.88)

Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, 12, Merzbacher,
Quantum Mechanics 3rd edition, 9.4-9.6, 10.1-10.4.
5.

Schr
odinger equation in momentum representation

Eigenvalue problem for the Hamiltonian can be formulated in the momentum representation. Formally, we can use
(4.50). However, it is not very convenient, for although the kinetic term p2 /2m is a constant, U may be a complicated
function of the operator p . We need an explicit form of the U operator in the momentum space. I will show how to
derive it in one dimension. Generalization to three dimensions is straightforward.
We start with the Schr
odinger equation in the form that does not refer to any particular representation:

H|i
= E|i .
Expand |i =

dp0 |p0 ihp0 |i:

0 ihp0 |i = E
dp H|p
0

(4.89)

dp0 |p0 ihp0 |i .

(4.90)

Now multiply on the left by the bra hp|:


Z
Z
0
0
0

dp hp|H|p ihp |i = E dp0 hp|p0 ihp0 |i .

(4.91)

The matrix element of the Hamiltonian reads


0i =
hp|H|p

p2
|p0 i .
(p0 p) + hp|U
2m

(4.92)

where I used hp|p0 i = (p p0 ). We calculate the matrix element of U as follows


Z
Z
Z
Z
|p0 i = dx0 dxhp|xihx|U
|x0 ihx0 |p0 i = dx0 dxhp|xiU (x)(x0 x)hx0 |p0 i
hp|U
Z
0
1
=
U (x)ei(p p)x/~ dx .
2~

(4.93)

Assembling everything together and recalling that (p) = hp|i is the wave function in the momentum representation
we arrive at the Schr
odinger equation in the p-representation
Z +
p2
|p0 i(p0 )dp0 = E(p) .
(p) +
hp|U
(4.94)
2m

Examples.
1. Bound state in the delta-well U = (x). Note that the potential is an even function of coordinate, so we
expect that the eigenstates can have definite parity.
|p0 i = 1
hp|U
2~

eix(p p)/~ ()(x) =

.
2~

(4.95)

Kirill Tuchin

55

Phys. 591 Lecture Notes

Substituting into (4.94) we get

p2
(p)
C = E(p) ,
2m
2~

(4.96)

where I denoted
C=

(p)dp .

(4.97)

Solution to (4.96) is
(p) =

C
m
.
2
~ p + 2m|E|

(4.98)

I used that E = |E| < 0. Now, inserting this into (4.97) and taking the integral yields
r

m
C=
C.
~ 2|E|

(4.99)

Hence, the only bound state is

Normalization is fixed from

m2
.
2~2

(4.100)

2m
.
~

(4.101)

E=
2 (p)dp = 1 as
C=

2. Bound states in
U = [(x a) + (x + a)] ,

> 0.

First we compute the Fourier image of the potential


Z
i
0
0
h ia(p0 p)/~
|p0 i = 1
hp|U
eix(pp )/~ ()[(x a) + (x + a)] =
e
+ eia(p p)/~ .
2~
2~

(4.102)

(4.103)

Then write down the Schr


odinger equation

where

i
h iap/~
p2
(p)
e
C+ + eiap/~ C = E(p) ,
2m
2~
C =

eiap /~ (p0 )dp0 .

(4.104)

(4.105)

Denote
2 =

2mE
> 0,
~2

m
.
~2

(4.106)

Then
(p) =

~  iap/~
1
e
C+ + eiap/~ C 2
.

p + ~2 2

(4.107)

Substituting in (4.105) we derive a system of two linear equations with respect to C :




C+ + e2a C ,

2a
C =
e
C+ + C .

C+ =

(4.108)
(4.109)

Kirill Tuchin

56

Phys. 591 Lecture Notes

There is a non-trivial solution iff



1
e2a


e2a 1

This is equivalent to the requirement that


1

= e2a ,




= 0.

(4.110)


=
1 e2a .

(4.111)

Solution to this equation gives the energy levels of the system.


Consider first the 0 +0 solution:


=
1 + e2a .

We have
C+ =


1
C+ + e2a C
2a
1+e

C+ = C

(4.112)

(p) = (p) .

(4.113)

We thus found an even stationary state, which is the only even state of this system. It is also the ground state,
because the ground state cannot be odd.
In the limit a  1 the two -wells are very close:
2

E+ =

2m2
.
~2

(4.114)

In the opposite limit a  1 we have e2a  1, so that in the first approximation


. In the second

approximation =
(1 e2a
). Therefore,
E+ =


m2

1 + e2a
.
2
2~

(4.115)

The 0 0 solution is a problem in one of the home assignments.

Additional reading: Merzbacher, Quantum Mechanics 3rd edition, 2.2, 2.5.


6.

Occupation number representation of harmonic oscillator

In III 4B we represented the harmonic oscillator in terms of the ladder operators. The raising operator a
moves
the oscillator up to the next excited state, while the lowering operator a
moves it down. We can give a different
interpretation to these operator. Instead of counting the quantum states (ground state, first excited state, etc.) we
will count the number of quantum excitations, or simply quanta. Each quantum excitation of harmonic oscillator
has energy E = ~, see (3.96). For example, harmonic quantum excitation of the lattice structure of a solid body is
phonon, harmonic excitation of the electromagnetic field is photon. We will say that the nth state contains n quanta
and denote it by the ket |ni. The physical meaning of the ladder operators is adding or removing a single quantum
to the hormonic oscillator:

a
|ni = n|n 1i ,
(4.116)

a
|ni = n + 1|n + 1i .
(4.117)
For this reason a
and a
are also called the annihilation and creation operators.
We can write the excitation number operator n
in terms of the ladder operators:
n
=a
a
.

(4.118)

n
|ni = n|ni .

(4.119)

Indeed, it is easy to check that

Kirill Tuchin

57

Phys. 591 Lecture Notes

Hamiltonian of the harmonic oscillator is


= ~(
H
a a
+ 1/2) = ~(
n + 1/2) .

(4.120)

Thus, states |ni are eigenstates of the Hamiltonian and the excitation number operator. State |0i is the ground state
which we normalize as usual h0|0i = 1. The excited states can be obtain by a successive application of the creation
operator
1
a )n |0i .
|ni = (
n!

(4.121)

You can verify that hn|ni = 1.


State vectors |ni form a complete set of states because they are eigenstates of the Hamiltonian. Therefore, instead
of the coordinate representation that we discussed in III 4 one can introduce the occupation number representation.
For any state
X
|i =
|nihn|i ,
(4.122)
n

where |hn|i|2 is interpreted as a probability to have n excitations. If a wave function is known in the coordinate
representation, then (4.122) implies that
X
hq|i =
hq|nihn|i ,
(4.123)
n

In particular, a wave function describing a single particle in one dimension can be expanded as
(x) = hx|i =

hx|nihn|i =

n=0

Cn n (x) ,

(4.124)

n=0

where n (x) are eigenfunctions of the harmonic oscillator given by (3.95). The probability to have n excitations in
the state |i is |Cn |2 .
In matrix notation the state vectors |ni can be represented as follows

1
0

|0i = 0
0
...

0
1

|1i = 0
0
...

0
0

|2i = 1
0
...

etc

Ladder operators become matrices amn = hm|


a|ni and amn = hm|
a |ni:

0 0 0 ...

0
1 0 0 . . .
1 0 0 ...
0 0

2 0 . . .

,
0
a
=
a

=
2 0 . . .

0 0 0

3 ...
0 0
3 ...
... ... ... ... ...
... ... ... ...

(4.125)

(4.126)

By the way, we see that (


a ) = a
as required. The number operator is a diagonal matrix

0
0
n
=
0
...

0
1
0
...

0
0
2
...

...
...
...
...

(4.127)

 Although the ladder operators are not Hermitian and therefore do not correspond to any physical quantity, their
eigenstates have clear physical interpretation. The eigenvalue problem for harmonic oscillator reads
a
|i = |i .

(4.128)

Kirill Tuchin

58

Phys. 591 Lecture Notes

In the coordinate representation:


a
(x) = (x) .

(4.129)

Employing (3.106) we can write it as a differential equation

r
~ 0
2~
+ (x x0 ) = 0 , with x0 =
.
(4.130)
m
m
Notice, that is a complex number. Replacing x by x0 = x x0 we obtain the same equation as in (3.112). Its
solution is given by (3.113):
 m 1/4

2
em(x0 x0 )x0 /4~ em(xx0 ) /2~ .
(4.131)
(x) =
~


R
It satisfies the normalization condition | |2 dx = 1.(9 We can write it in terms of the expectation value (x)2 =
~/(2m) as follows:

2
(x) = e(x0 x0 )x0 /8h(x) i

(2 h(x)2 i)

1/4

e(xx0 )

/4h(x)2 i

(4.132)

This is precisely the wave function of a coherent state generalized to complex x0 , see (1.133) with p0 = 0. We observe
that the eigenstates of the annihilation operator are the coherent states.
What is the probability that a coherent state contains n excitations of the harmonic oscillator? To answer this
question we can use (4.124) and expand (4.132) into the harmonic oscillator eigenstate basis:
(x) =

Cn n (x) ,

(4.133)

n=0

where n are given by (3.95). Coefficients Cn can be computed using the orthonormality of the n s:
Z +
Cn =
n (x) (x)dx


 r
1/2 1 Z +

2
2
m
m(x0 x
)x0 /4~ m
0

emx /(2~) em(xx0 ) /2~ dx


Hn x
=e
n
~
~
2 n!
Z +

2
2
1
= e( )/2
Hn ()e /2( 2) /2 d
n
2 n!
Z +
n 2

1
( )/2
n 2 d e
2 + 22

=e
(1) e
e
d ,
d n
2n n!

(4.134)

(4.135)
(4.136)

where I used (3.90). Integrating by parts n times we get


Z +
2

2
e

( 2)n
e e 2 d
n
2 n!

2
2
e
n
= e( )/2
( 2)n e /2 = e|| /2 .
n
2 n!
n!

Cn = e(

)/2

(4.137)
(4.138)

Thus, the probability to find a coherent state |i with n harmonic oscillator excitations is
n

||2n ||2
hni hni
e
=
e
,
(4.139)
n!
n!
where hni = h|
n|i = ||2 is the average number of excitations in the given coherent state. Distribution (4.139)
is the Poisson distribution, which gives the probability to have n independent excitations in a coherent state with
average excitation number hni. Independence of excitations is closely related to the fact that the coherent state is as
close to the classical limit as a quantum system can be.
Pn = |Cn |2 =

Additional reading: Merzbacher, Quantum Mechanics 3rd edition, 10.6.

(9

Function (4.131) can be multiplied by any phase. I used this freedom to fix the phase so that the coefficients Cn in (4.133) have simple
form (4.138).

Kirill Tuchin

Phys. 591 Lecture Notes


V.
1.

59

MOTION IN CENTRAL POTENTIAL

General properties of motion in central potential

Motion in central potential U (r) is described by the Hamiltonian:


2

= ~ + U (r) ,
H
2M
where r is a distance form the center, M is mass of the particle. Laplacian in spherical coordinates






1
1
1

1 2
2
2
= 2
r
+ 2
sin
+
r r
r
r
sin

sin2 2

(5.1)

(5.2)

Using (1.106) and (5.2) in (5.1) we have


2

= ~
H
2
2M r r

r2

2
L
+ U (r) .
2M r2

(5.3)

L
2 ] = [H,
L
z ] = 0. Thus, systems in central potential can be in stationary
Since U (r) does not depend on angles [H,
2 and L
z in coordinate representation are spherical harmonics
states with given values of L2 and Lz . Eigenfunctions of L
Ylm (, ). The corresponding eigenvalues are L2 = ~2 l(l+1), l = 0, 1, 2 . . . and Lz = ~m, m = 0, 1, . . . , l. Therefore,
= E in the form
we can look for a solution to the Schr
odinger equation equation H
Elm (r, , ) = REl (r)Ylm (, ) .

(5.4)

In place of the radial function R it is convenient to introduce a different function


(r) = rR(r) .

(5.5)



~2 l(l + 1)
~2 d 2
+ U (r) +

= E .
2M dr2
2M r2

(5.6)

Schr
odinger equation becomes

Since R(r) must be finite at r 0, we conclude that


lim (r) = 0 .

r0

Normalization condition for the wave function of the discrete spectrum:


Z
Z
Z

3
2 2
0
0
0
Elm (r, , )E lm (r, , )d r = EE
REl r dr = EE
2El (r)dr = EE 0 ,
while for the continuous spectrum:
Z
Z

klm
(r, , )k0 lm (r, , )d3 r =

k0 (r)k (r)dr = (k k 0 ) ,

(5.7)

(5.8)

(5.9)

where k 2 = 2M E/~2 0 as usual.


Some properties of the Schr
odinger equation with the central potential:
1. For each l there is (2l + 1)-fold degeneracy corresponding to the values of m. States with l = 0, 1, 2, 3 . . . are
denoted by letters s, p, d, f, . . ..
P ] = 0, all stationary states can be chosen to have definite parity. In fact the spherical harmonics
2. Since [H,
do have definite parity. To see this, note that in spherical coordinates spatial inversion corresponds to the
transformation 0 = and 0 = + . This is easy to verify by writing down the spherical
coordinates of r and observing that r r 0 = r. Under such transformation
P Ylm (, ) = Ylm ( , + ) = (1)l Ylm (, ) .
Thus, spherical harmonics are eigenfunctions of P .

(5.10)

Kirill Tuchin

Phys. 591 Lecture Notes

60

3. Eq. (5.6) looks like the Schr


odinger equation for one dimensional motion in the effective potential
Ul = U (r) +

~2 l(l + 1)
,
2M r2

(5.11)

with boundary condition (5.7). Hence, we can use the results of III 5 for one dimensional motion. In particular,
the energy spectrum is non-degenerate.
4. Suppose that the potential is such that U 0 as r and consider particle with negative energy. Such
particle is constrained to move in a volume around the origin and its energy spectrum is discrete. Multiplying
(5.6) by and integrating
Z
Z
Z
~2

00 dr +
Ul 2 dr =
E2 dr .
(5.12)
2M 0
0
0
Since at 0 as r we can integrate the first term by parts
Z
Z

00 dr =
(0 )2 dr ,
0

(5.13)

which yields for the average energy


E=

~2
2M

(0 )2 +



l(l + 1) 2M
2
+
U
(r)

(r)
dr .
r2
~2

(5.14)

Let us estimate this integral in the case of the potential that is non-vanishing only at r a and has the form
U (r) = A/rn with positive A and n. By uncertainty relation 0 /a, implying that


~2 1 + l(l + 1) 2M A
E
2 n .
(5.15)
2M
a2
~ a
If n > 2, then E is at a = 0, which describes falling onto the center. However, if n < 2, then the
minimum of E is at a finite a, implying that no falling into the center occurs. In this case the energy spectrum
starts with the finite negative value. Recall that in the classical mechanics falling onto the center happens at
any n > 0.

FIG. 7: falling onto the center happens at n < 2.

5. According to the oscillation theorem, if for a given l we arrange the energy eigenvalues in order of increasing
magnitude El0 < El1 < El2 < . . . and assign the corresponding states the radial quantum number nr = 0, 1, 2, . . .,
then the number of zeros of (r) and of R(r) in the interval 0 < r < is nr . (r = 0 is not counted).
Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, 32,18, Merzbacher,
Quantum Mechanics 3rd edition, 12.1.

Kirill Tuchin

61

Phys. 591 Lecture Notes


2.

Spherical waves

Free motion of a particle with given l and m is called the spherical wave. The corresponding solutions of Schr
odinger
equation can be found form (5.6) with U = 0:
~2 l(l + 1)
d2 l

l + k 2 l = 0 .
dr2
r2

(5.16)

Consider first s-states l = 0:


000 + k 2 0 = 0

0 (r) = A sin(kr) + B cos(kr) .

In view of the boundary condition (5.7), B = 0. Substituting into (5.9) we get


Z
Z
0
0
A2 + eik r eik r eikr eikr
0
2
sin(k r) sin(kr)dr =
A
dr
2
2i
2i
0
A2
A2
{2(k 0 k) + 2(k 0 + k) }2 =
=1
=
2
| {z }
2(2i)
2
=0 (k>0, k0 >0)

Therefore,

k0 (r)
=
Rk0 (r) =
r

A=

(5.17)

2
.

2 sin(kr)
.

(5.18)

(5.19)

In general, for l 6= 0 (5.16) has a solution in terms of the spherical Bessel and Neumann functions jl and l :
Rkl (r) = Ajl (kr) + Bl (kr) .
Spherical Bessel functions can be expressed through the Bessel functions of half-integer order:
r
 l
d

sin z
jl (z) =
Jl+1/2 (z) = (1)l
,
2z
dz
z
r

l+1
l (z) = (1)
Jl1/2 (z) ,
2z

(5.20)

(5.21)
(5.22)

Asymptotic behavior of these functions:


jl (z)

zl
(2l+1)!!

,
zl


1 cos z (l + 1) , z  l
z
2

,
zl
(2l1)!!
z l+1
l (z)


1 sin z (l + 1) , z  l
z
2

(5.23)

(5.24)

Evidently, only jl satisfies the boundary condition (5.7). Therefore, the general solution to the Schrodinger equation
for a free particle with given l and m reads
klm (r) = Ajl (kr)Ylm (, ) .

(5.25)

Eq. (5.16) describes one-dimensional motion of a particle in the effective potential Ul = ~2 l(l +1)/2M r2p
. Classically
particle moving with energy E is allowed to move only in the region E U , which corresponds to r rl = l(l + 1)/k.
Thus, at r < rl the wave function l (r) exponentially decays toward r 0. In the classically accessible region r  rl ,
we can neglect the effective potential:
00l (r) + k 2 l (r) = 0 ,

kr  1

l (r) A sin(kr + l ) ,

(5.26)

which is of course seen also in (5.23).


Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, 33, Merzbacher,
Quantum Mechanics 3rd edition, 12.2.

Kirill Tuchin

62

Phys. 591 Lecture Notes


3.

Spherical potential well

As an application of the result we derived in the previous section, consider motion of a particle in the infinitely
deep spherically symmetric potential well:
(
0, r a
U (r) =
(5.27)
, r > a
Solution inside the well is given by (5.25). At r a the wave function vanishes. Thus l (ka) = 0. Let me denote
zeros of the spherical Bessel function of the lth order by Xln . Index n = 1, 2, . . . enumerates zeros starting from the
lowest one. For example, the lowest four zeros are
Xs1 X01 = ,

Xp1 X11 = 4.493 ,

Xd1 X21 = 5.763 ,

Xs2 X02 = 6.283 .

(5.28)

The energy levels are follow from ka = Xln :


Enl =

2
~2 Xln
.
2M a2

(5.29)

The radial quantum number that corresponds to the number of zeros in a given state is nr = n 1.

FIG. 8: Spherical well of finite depth.

Consider now a well of finite depth, see Fig. 8:


U (r) =

U0 , r a
0,

(5.30)

r>a

Bounds states have energies in the interval U0 E 0. Schrodinger equation equation for the s-state reads (l 6= 0
states can be discussed along the same lines but involve a more tedious algebra):
2M
(U0 |E|) = 0 , r a ,
~2
2M
00 2 |E| = 0 , r a
~
00 +

(5.31)
(5.32)

Denote
2M
k2 = 2 (U0 |E|) ,
~

2 =

2M
|E| .
~2

(5.33)

Solution to (5.31) and (5.32) read:


,
= A sin(kr)
= Be

r < a,
r > a.

(5.34)
(5.35)

The logarithmic derivative 0 / must be continuous across r = a implying that


= .
k cot(ka)

(5.36)

Kirill Tuchin

63

Phys. 591 Lecture Notes

5
4

3
2
1
0

ka

3
2

5
2

FIG. 9: Solution to (5.36) is shown in blue. Orange (no bound states), green (single bound state) and red (two bound states)
lines correspond to solution of (5.37) with 2M U0 /~2 = 1, 5, 25.

Solutions to this transcendental equation are a set of discrete numbers kn corresponding to the energy levels in the
well. Notice that k and are related by (5.33) as follows:
2M U0
.
k2 + 2 =
~2

(5.37)

So, geometrically, kn0 s lie at the intersections of curves given by (5.36) and (5.37), see Fig. 9.
It is seen that there exist a solution only if a /2, which translates into a2 U0 2 ~2 /8M . The first excited
state exists if a 3/2 etc.
Additional reading: Merzbacher, Quantum Mechanics 3rd edition, 12.3.
4.

Spherical harmonic oscillator

Schr
odinger equation for a particle moving in the three-dimensional harmonic oscillator potential
U (r) =

M 2 r2
,
2

(5.38)

can be solved by separation of variables in Cartesian and in spherical coordinates. The former is one of the home
assignment. Clearly, physical observables such as the energy spectrum should not depend on the choice of the reference
frame, while the eigenfunctions of operators do depend on it. Let us confirm this by solving the Schrodinger equation
for 3D harmonic oscillator in spherical coordinates. We have


M 2 r2
~2 l(l + 1)
~2 d2
+
+

E
(5.39)

nl nl (r) = 0 .
2M dr2
2
2M r2
Introduce a length parameter
a=

~
M

and dimensional quantities = r/a and = E/~. Then (5.39) can be cast in form
 2

d
l(l + 1)
2

+ 2 nl () = 0 .
d 2
2

(5.40)

(5.41)

Kirill Tuchin

Phys. 591 Lecture Notes

64

Its solution can be expressed in terms of the confluent hypergeometric function F as follows
nl () = Nnl e

where
1

2 2

nr =

F (nr , l + 3/2, 2 ) ,

(5.42)

(5.43)

/2 l+1

l+

3
2

The only property of the confluent hypergeometric function that we need to know is that 0 when only if
nr = 0, 1, 2, . . .. This gives energy spectrum as


3
, nr , l = 0, 1, 2, . . .
(5.44)
Enr l = ~ 2nr + l +
2
Wave functions of the corresponding eigenstates read
1
n l (r)Ylm (, ) .
r r

nr lm =

(5.45)

Note, that energy levels (5.44) depend only on a combination of quantum numbers 2nr + l = n, n is called the
principal quantum number. States are usually referred to by specifying its quantum numbers in form (n + 1)l. For
example, 1s state has n = 0, l = 0, 1p state has n = 0, l = 1, 2s state has n = 1, l = 0 etc., while energy levels are
labeled by n as En .
Levels with n 2 are degenerate. For instance, E2 = (7/2)~ is six-fold degenerate. One s state with l = 0, nr = 1
and five d-states l = 2, nr = 0, m = 2, 1, 0 have this energy. While degeneration of states with different m and the
same l is due to spherical symmetry, degeneration with different ls is accidental. To degeneracy g(n) reads (home
assignment):
1
(1 + n)(2 + n) .
2

g(n) =

(5.46)

For odd n the result is the same.


Schr
odinger equation for 3D harmonic oscillator can be solved also in parabolic coordinates, which is helpful when
discussing atom in external electric field.
Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, , Merzbacher, Quantum Mechanics 3rd edition,
5.

Coulomb potential

Schr
odinger equation with Coulomb potential
U (r) =

e2 Z
r

(5.47)

has special physical significance as it describes motion of electrons in the field of atomic nucleus. In such case e is
electron charge and Z is the nucleus electric charge in units of e.
A.

Discrete spectrum

It is conventional to introduce the atomic units: atomic units of length


a=

~2
,
M e2

(5.48)

also called the Bohr radius and atomic unit of energy


Ea =

e2
M e4
= 2 .
a
~

(5.49)

Kirill Tuchin

65

Phys. 591 Lecture Notes

For electron, these units are a = 5.29 109 cm and Ea = 27.21 eV.
Introduce dimensionless quantities = r/a and = E/Ea . Then Schrodinger equation for the radial component
takes form
 2

2Z
l(l + 1)
d
+
2
+
() = 0 .
(5.50)

d2

2
Bound states have E < 0 so we can denote 2 = 2 > 0:
 2

d
2Z
l(l + 1)
2
+
() = 0 .

d2

(5.51)

At the terms containing the inverse powers of can be neglected yielding the asymptotic solution
() Ae + Be ,

B = 0.

(5.52)

Now we are looking for a solution in the form () = e w(), where w is unknown function that we expand in
power series as follows
w() =

(5.53)

=0

We also need to check the boundary condition at 0. In this case () 0 which must satisfy (5.51). This
gives
:0

2Z 
( 1)2 2

+ l(l + 1)2 = 0 ,


(5.54)

where I dropped terms that vanish faster than others. Equation ( 1) = l(l + 1) has two solutions
= l + 1, l .

(5.55)

Only the first of them is finite at 0. Thus, after applying the boundary conditions we have
X
() = e l+1
.

(5.56)

Loading this in (5.51) produces the following recurrence relation


+1 =

2[( + l + 1) Z]
.
( + l + 2)( + l + 1) l(l + 1)

(5.57)

In order that be finite at large the power series in (5.56) must terminate at some = nr . This happens if
(nr + l + 1) Z = 0 ,

Z
Z
= ,
nr + l + 1
n

(5.58)

where the principal quantum number is given by


n = nr + l + 1 .

(5.59)

Energy levels read


En =

2
M (e2 Z)2 1
Ea =
,
2
~2
2n2

n = 0, 1, 2, . . .

(5.60)

It is evident from (5.60) that for a given n states with l = 0, 1, . . . , n 1 and m = 0, 1, 2, . . . l are degenerate.
The degeneracy g(n) can be computed by summing the arithmetic series
g(n) =

n1
X

(2l + 1) =

l=0

1 + 2(n 1) + 1
n = n2 .
2

(5.61)

Kirill Tuchin

Phys. 591 Lecture Notes

66

The normalized wave functions can be expressed through the associated Laguerre polynomials L as follows
" 
#1/2

l


3
2Z
(n l 1)!
2Zr
2Zr
nl (r)
2l+1
Zr/na
e
Lnl1
=
.
(5.62)
Rnl (r) =
r
na
2n(n + l)!
an
na
A few lowest order radial functions:
1/2
4Z 3
eZr/a ,
a3
 3 1/2 

Z
Zr
R20 (r) =
2
eZr/2a ,
8a3
a
 3 1/2
Zr Zr/2a
Z
R21 =
e
.
24a3
a
R10 (r) =

The normalization condition is

2 2
Rnl
r dr = 1 .

(5.63)
(5.64)
(5.65)

(5.66)

For the future reference let me list here several useful formulas involving expectations values in states (5.62)
1
[3n2 l(l + 1)]
2Z

2
n2
=
[5n2 + 1 3l(l + 1)] ,
2Z 2

1
Z

= 2,
n

2
Z2

= 2
,
n (l + 1/2)

3
Z3

= 3
.
n (l + 1)(l + 1/2)l
hi =

B.

(5.67)
(5.68)
(5.69)
(5.70)
(5.71)

Continuous spectrum

Continuous spectrum lies at E > 0. Introducing k 2 = 2 0 we rewrite the Schrodinger equation (5.50) as follows

 2
2Z
l(l + 1)
d
2
+
k
+

= 0.
(5.72)
d2

2
At the asymptotic behavior of the solution is

() Aeik + Beik .

(5.73)

Thus we are looking for a solution in the form


() = eik l+1

(5.74)

Plugging into (5.73) we obtain


+1 =

2[i( + l + 1)k Z
.
( + l + 2)( + l + 1) l(l + 1)

(5.75)

This power series describes a confluent hypergeometric function. The final solution to the Schrodinger equation reads


2
kl () = eik l+1 F l + 1 ; 2l + 2, 2ik .
(5.76)
ik
Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, 36.

Kirill Tuchin

67

Phys. 591 Lecture Notes


6.

Effective electric potential of the hydrogen atom

Hydrogen atom consists of positively charged proton and negatively charged electron orbiting around its common
center of mass. Since mass of proton is about two thousand times larger than that of electron the reduced mass
of hydrogen atom equals electron mass M with very good accuracy, so that the quantum mechanical problem is to
quantize motion of electron in potential U (r). This potential obeys the Poisson equation
2 U = 4(e) ,

(5.77)

where is the charge density of the hydrogen atom. It can be computed as


(r) = e(r) e|(r)|2 .

(5.78)

where is the wave function on electron in the hydrogen atom. The first term in the right-hand-side of (5.78)
described the proton charge, while the second term describes the electron cloud. Now, satisfies the Schr
odinger
equation
p2
+ U (r) = E ,
2M

(5.79)

where U is a solution to (5.77). We thus arrived at the system of coupled Schrodinger equation and Poisson equations
that must be solved simultaneously.
We can proceed by the method of consecutive approximations. At first, we will neglect the electron cloud contribution in (5.78). Then, U = e2 /r is a Coulomb potential. Solution to (5.79) with such potential has been discussed
in the previous section. The bound state solutions are a system of wave functions nlm (r). The second step is to
substitute these wave functions into (5.77) and calculate the potential U . Clearly, the result depends on electron
state. For illustration, suppose that electron is in the ground state, viz.
100 (r) = 2a3/2 er/a Y00 (, ) .

(5.80)

Poisson equation
1
U= 2
r r
2



2
U = 4(e)
r
r

is convenient to write for a auxiliary function f defined as U = ef /r:




00
3 2r/a 1
f = 4r e(r) 4ea e
.
4

(5.81)

(5.82)

At r > 0, the delta-function does not contribute, so that


f 00 =

4re 2r/a
e
.
a3

(5.83)

R
At large distances U must decrease at least as fast as 1/r2 . This is because atom is neutral ( d3 r = 0) and thus the
monopole contribution to the multipole expansion vanishes. Thus, the boundary conditions at r are f 0 and
f 0 0. Integrating (5.83) twice we get
Z
Z
Z
0
4e 0 00 2r00 /a 00
4e 0 1
e
f (r) = 3
dr
r e
dr = 3
dr a(a + 2r0 )e2r /a = (a + r)e2r/a .
(5.84)
a r
a r
4
a
r0
Finally,
U (r) = e2

1 1
+
r
a

e2r/a .

(5.85)

We see that the Coulomb potential is modified at distances comparable to or larger than the Bohr radius a. Electron
cloud screens the proton potential.
One can continue the iteration process and solve the Schrodinger equation with potential (5.85) to find the modified
wave functions etc.

Kirill Tuchin

Phys. 591 Lecture Notes


7.

68

Origin of degeneracy of the energy spectrum

We have seen many times that there is a connection between the degeneracy of energy levels of a quantum system and
its symmetry. To make a general statement connecting degeneracy and symmetry consider two Hermitian operators
Let [F , K]
6= 0, but [F , H]
= [K,
H]
= 0, i.e. F and K
correspond to conserved quantities, but are not
F and K.
E,F = EE,F and an eigenfunction
simultaneously measurable. Let E,F be a wave function of a stationary state: H

of F : F E,F = F E,F . This is possible because F commutes with H.


because [K,
F ] 6= 0, i.e. K
E,F
Now, E,F is not an eigenfunction of K
/ E,F . On the other hand,
K
E,F ) = K(
H
E,F ) = E K
E,F .
H(

(5.86)

E,F is an eigenfunction of H.
Thus, we come to conclusion that there are at least two different
This indicates that K

E,F .
eigenfunctions of H corresponding to the same level E: E,F and K
but not with each other.
In summary, energy level is degenerated if there are operators that commute with H,
Note, that while symmetry implies degeneracy, degeneracy does not necessarily imply symmetry.
Examples.

= p2 /2m. [H,
p] = 0 and [H,
P ] = 0, where P is the inversion (parity) operator.
1. Free one dimensional motion: H

However [
p, P ] 6= 0. Therefore, levels are two-fold degenerate. Indeed, there are two independent solution to the
Schr
odinger equation : = eikx .

2. For a spherically symmetric potential U (r) the Hamiltonian is invariant under rotations:
L
x ] = [H,
L
y ] = [H,
L
z] = 0 .
[H,

(5.87)

i, L
j ] = i~ijk L
k 6= 0 .
[L

(5.88)

However,

This is the source of the 2l + 1 degeneracy of lth level.


Examples of accidental degeneracy, i.e. degeneracy which is not related to a symmetry.
1. Spherical harmonic oscillator.
2
2
= p + M r .
H
2M
2

(5.89)

pi pk
Tik =
+ M
ri rk
M

(5.90)

Operator

2.
commutes with the Hamiltonian, but does not commute with L
2. Coulomb potential.
2
2
= p e .
H
2M
r

(5.91)

Operator
=
K


1 
+ r ,
L
L P p
2
2M e
r

(5.92)

i, K
j ] 6= 0 and [K
i, K
j ] 6= 0 if i 6= j.
commutes with the Hamiltonian, but [L
Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, 36,10, Merzbacher,
Quantum Mechanics 3rd edition, 12.5.

Kirill Tuchin

69

Phys. 591 Lecture Notes


VI.
1.

ANGULAR MOMENTUM
Angular momentum operator

and discussed some of its properties in the


In I 4 and I 6 I introduced the orbital angular momentum operator L
coordinate representation. In particular, we have seen that its components satisfy the commutation relations (1.57).
In fact, a vector operator satisfying such commutation relations has much reacher mathematical structure than is
reflected by the coordinate representation, which can be used to describe only those physical quantities that have
classical analogues. This is why in this section we are going to study a generic operator satisfying the commutation
relations (1.57) without a reference to any particular representation (apart from a few examples at the end of this
section).
Define the angular momentum operator J whose components satisfy the commutation relation
[Ji , Jk ] = i~ikl Jl .

(6.1)

is a particular example of such an operator. In mathematical physics relations


Orbital angular momentum operator L
(6.1) are referred to as the the Lie algebra. Introduce the Casimir operator
X
J2 =
Jk2 = Jx2 + Jy2 + Jz2 .
(6.2)
k

This operator commutes with all components of J (summation over the repeated indexes is implied):
[J2 , Ji ] = [Jk , Jk , Ji ] = Jk [Jk , Ji ] + [Jk , Ji ]Jk = i~(kil Jk Jl + kil Jl Jk ) = i~ kil (Jk Jl + Jk Jk ) = 0 .
|{z} |
{z
}
sym

(6.3)

anti-sym

Since [J2 , Jz ] = 0 physical quantities J 2 and Jz can simultaneously have definite values. Therefore, there exist a
complete set of states |jmi such that
J2 |jmi = Jj2 |jmi ,
Jz |jmi = ~m|jmi .

(6.4)
(6.5)

In other words, we denote the eigenvalues of J2 and Jz operators as Jj2 and ~m respectively; j denotes the largest value
of m. To determine the possible values of these eigenvalues it is convenient to introduce the following non-Hermitian
operators, which are similar to the ladder operators of harmonic oscillator:
1
J = (Jx iJy ) .
2

(6.6)

Their properties:
[J2 , J ] = 0 ,
1
[Jz , J+ ] = (i~Jy + i(i~)Jx ) = ~J+ ,
2
[Jz , J ] = ~J ,
[J+ , J ] = ~Jz ,

(6.7)
(6.8)
(6.9)
(6.10)

We can express J2 via J and Jz . To this end write


2J+ J = (Jx + iJy )(Jx iJy ) = Jx2 + Jy2 + iJy Jx iJx Jy = Jx2 + Jy2 + ~Jz = J2 Jz2 + ~Jz .

(6.11)

J2 = 2J+ J + Jz2 ~Jz .

(6.12)

J2 = 2J J+ + Jz2 + ~Jz .

(6.13)

Thus,

Simiarly,

Kirill Tuchin

70

Phys. 591 Lecture Notes

Consider
Jz J+ |jmi = (~J+ + J+ Jz )|jmi = ~J+ |jmi + ~mJ+ |jmi = ~(m + 1)J+ |jmi

J+ |jmi |j, m + 1i .

(6.14)

Note, that |j, m + 1i is an eigenstate of Jz corresponding to the eigenvalue ~(m + 1). By the same token,
Jz J |jmi = ~(m 1)J |jmi

J |jmi |j, m 1i .

(6.15)

We see that operators J+ and J are diagonal with respect to j and increase/decrease m by one. Since j m j it
follows that m and 2j must be integers. Therefore, j can be either a non-negative integer j = 0, 1, 2, . . . or a positive
half-integer j = 1/2, 3/2, 5/2, . . ..
We are ready to compute Jj2 . Start from |j, m + 1i J+ |jmi and set m = j:
J+ |jji = 0 ,

(6.16)

1
1
J J+ |jji = (J2 Jz2 ~Jz )|jji = (Jj2 ~2 j 2 ~2 j)|jji = 0 .
2
2

(6.17)

Jj2 = ~2 j(j + 1)

(6.18)

because m j. Now, using (6.13)

Thus,

Examples.

1. In applications it is often convenient to know the matrix elements of operators Jx and Jy . To compute them we
start with the operator equation (6.12) and take its matric element:
hjm|J2 |jmi = ~2 j(j + 1) = 2hjm|J+ J ||jmi + hjm|Jz2 ~Jz |jmi
X
=2
hjm|J+ |j 0 m0 ihj 0 m0 |J |jmi + ~2 (m2 m)
m0 j 0

= 2hjm|J+ |j, m 1ihj, m 1|J |jmi + ~2 m(m 1)

(6.19)

Since Jx and Jy are Hermitian operators


1
1
hj, m 1|J |jmi = hj, m 1| (Jx iJy )|jmi = hj, m| (Jx + iJy )|j, m 1i
2
2

(6.20)

Substituting this into (6.19) we get


2|hj, m|J+ |j, m 1i|2 = ~2 j(j + 1) ~2 m(m 1) = ~2 (j 2 + j m2 + m)

(6.21)

Thus, the only non-vanishing matrix elements are

Consequently,

~ p
hj, m|J+ |j, m 1i = hj, m 1|J |j, mi =
(j + m)(j m + 1) .
2
~p
(j m)(j m + 1) ,
2
i~ p
hj, m 1|Jy |j, mi =
(j m)(j m + 1) .
2
hj, m 1|Jx |j, mi =

(6.22)

(6.23)
(6.24)

2. The choice of z-axis as a quantization direction is a matter of convention. We can instead choose, say, x-axis.
For example, a state with the orbital angular momentum l = 1 and its projection on z-axis lz = 1 is described
by the wave function
r
r
3
3 x iy
i
l=1,lz =1 = i
sin e
= i
.
(6.25)
8
8 r

Kirill Tuchin

71

Phys. 591 Lecture Notes

After a cycle permutation of x, y, z: x y, y z, z x, we get a wave function


r
r
3 y iz
3
= i
(sin sin i cos ) ,
l=1,lx =1 = i
8 r
8

(6.26)

describing a state with orbital angular momentum l = 1 and its projection on x-axis lx = 1.

In general, transformation from one complete set |jmi to another complete set |jmi0 can be done by the rotation
operator

n = eiJn/~
R
,

(6.27)

n |jmi .
|jmi0 = R

(6.28)

which is a generalization of (2.87), as follows:

Acting on the left-hand-side with a unit operator 1 =


|jmi0 =

j 0 m0

j 0 m0

|j 0 m0 ihj 0 m0 | we can write this in a matrix form

n |jmi =
|j 0 m0 ihj 0 m0 |R

X
m0

(j)

Dm0 m |jm0 i ,

(6.29)

I introduced matrix elements


(j)

n |jmi .
Dm0 m = hjm0 |R

(6.30)

(j)

Dm0 m is called the Wigner matrix. This is the matrix element of the rotation operator (6.27) about the vector
through the angle . In the case of the orbital angular momentum, projecting (6.30) onto the coordinate
n
representation we have (see e.g. IV 2):
Ylm (0 , 0 ) =

l
X

(l)

Dm0 m Ylm0 (, ) .

(6.31)

m0 =l

3. In the case j = 1 (6.23) can be represented as a 3 3 matrix:

2 0
0
~
Jx = 2 0
2.
2
0
2 0

Suppose that a particle is in a state with Jx = 0. Denote this state as



a
0 = b .
c

Because Jx is definite, 0 must be an eigenstate of Jx , i.e.


2 0
a
b

0
Jx 0 = 0 2 0
2 b = 2 a + c = 0.
c
b
0
2 0

Thus, b = 0 and a = c. Normalization requires that a = 1/ 2, so finally

1 1
0
0 =
.
2 1

(6.32)

(6.33)

(6.34)

(6.35)

Kirill Tuchin

Phys. 591 Lecture Notes

72

4. Consider a rotator (such as a diatomic molecule) with the moment of inertia I rotating in plane around its
symmetry axis. The Hamiltonian is
2
= Lz .
H
2I

(6.36)

L
z ] = 0, the wave functions of stationary states can be chosen to be also eigenfunctions of L
z = ~ .
Since [H,
The corresponding complete set is
1
m = eim ,
2

Em =

~2 m 2
,,
2I

m = 0, 1, 2, . . . .

(6.37)

All states except the ground state (m = 0) are two-fold degenerate. This is because the Hamiltonian commutes
with the operator of inversion (reflection) with respect to the x-axis defined as Px (x, y) = (x, y) (which is
z ] 6= 0.
equivalent to ), whereas [Px , L
Instead of the states (6.37) that have definite Lz , one can choose states with definite Px = 1:
1
+
m
= cos(m) ,

m
= sin(m) .

Suppose that the rotator is in a state described by the wave function


r
4
2
.
= A cos , A =
3

(6.38)

(6.39)

We can expand
=

X
m


1
A 2i
Cm eim =
e + 2 + e2i .
4
2

It follows that the only three non-vanishing coefficients are


r
r
2
1
C0 =
, C2 =
3
6

(6.40)

(6.41)

Expectation values of physical quantities


hLz i =

X
m

|Cm |2 ~m = 0 ,

X
4~2
,
L2z =
|Cm |2 ~2 m2 =
3
m
hEi =

X
m

|Cm |2 Em =

2~2
.
3I

(6.42)
(6.43)
(6.44)

Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, 26-28,58, Merzbacher,
Quantum Mechanics 3rd edition, Ch. 11,17.
2.

Spin

Thus far we considered only one example of the angular momentum operator, viz. the orbital angular momentum
The corresponding quantum number l can be only integer, which as we will see, is intimately related to the fact
L.
has a classical analogue. In fact, we constructed L
as a generalization of its classical expression.
that L
I pointed out in VI 1 that j can be either integer or half-integer. The first evidence of half-integer angular
momentum can from the Stern-Gerlach experiment. In this experiment a beam of atoms passes through a magnetic
field with fixed direction but varying absolute value in the direction x perpendicular to the beam velocity. After

Kirill Tuchin

Phys. 591 Lecture Notes

73

passing through the magnetic field the beam is projected onto a screen. Atom in magnetic field is subject to a force
of magnitude
Fx = B

B
x

along the x-direction. Coefficient B is a projection of the magnetic moment of a particle in the bean onto the
magnetic field direction. If we compute the magnetic moment using the classical formula
=

e
L,
2M c

where M is electron mass, and then quantize the orbital angular momentum along the magnetic field direction, then
the possible values of the magnetic moment are
B =

e~
m,
2M c

m = 0, 1, 2, . . . , l .

Since l is integer, we expect to see 2l + 1 traces on the screen, which is an odd number. However, the experiment
indicated that the incident beam of Ag and Na atoms produces an even number of traces. This can be explained only
by existence of an additional half-integer contribution to the total angular momentum.
All empirical evidence points to the existence of an intrinsic angular momentum, or spin, of particles. Unlike the
orbital angular momentum of a particle, which describes its motion in space, spin is an intrinsic property of a particle.
Spin has no classical analogy and cannot be visualized as a rotation of a particle around its axis. Spin characterizes
transformational property of the wave function under rotations.
Since spin S is an angular momentum operator, we can apply to it the formalism developed in VI 1. In particular,
it satisfies the commutation relations
[Si , Sj ] = i~ijk Sk .

(6.45)

Let |si be a complete set of eigenstates of operators S2 and Sz ; the corresponding eigenvalues are S 2 = ~2 s(s + 1)
and Sz = ~, = s, s + 1, . . . , s 1, s. It serves as a basis of the spin representation. Spin s can be either integer or
half-integer: s = 0, 21 , 1, 32 , . . .. In the spin representation, a wave function of a particle with spin s can be represented
by a column of 2s + 1 components called the spinor :

s (q)
s1 (q)
(6.46)
... .
s (q)
The spin operator is represented by a square matrix that can obtained from (6.23),(6.24) after replacements J S,
j s and m :
~p
(s )(s + 1) ,
2
~p
hs, 1|Sy |s, i = i
(s )(s + 1) ,
2
hs, |Sz |s, i = ~ .

hs, 1|Sx |s, i =

(6.47)
(6.48)
(6.49)

Of particular importance is the case of s = 1/2. The corresponding representation consists of two basis states with
= 1. In this representation


D1 1 1 1E D1 1 1 1E ~
~ 0 1

, Sx ,
=
, Sx ,
=
Sx =
,
(6.50)
2 2
2 2
2 2
2 2
2
2 1 0

such that(10
and similarly for the other spin components. Introduce the Pauli matrices
~
.
S =
2

(10

Do not confuse the Pauli matrices with the eigenvalues of Sz operator, viz. ~.

(6.51)

Kirill Tuchin

74

Phys. 591 Lecture Notes

From (6.50) we see that in the spinor representations they have the following form:






0 1
0 i
1 0

x =
,
y =
,
z =
.
1 0
i 0
0 1

(6.52)

Properties of the Pauli matrices:


1.

x2 =
y2 =
z2 = 1 =

1 0
0 1

(6.53)

x
y = i
z .

(6.54)

2.

y
z = i
x ,

z
x = i
y ,

Eq. (6.53),(6.54) can be written together in a compact form

i
j = iijk
k + ij 1 .

(6.55)

{
i ,
j } =
i
j +
j
i = 2ij .

(6.56)

3. Using (6.55) we derive


This means that the Pauli matrices anti-commute.
4. Home assignment:

2 = 3 1 ,

(6.57)

a)(
b) = a b 1 + i
(a b) ,
(

(6.58)

Tr
i = 0 ,

(6.59)

Tr(
i
k ) = 2ik ,
(
an 1
n even
n
=
(a )
n1
, n odd
a
(a )

(6.60)
(6.61)

Spin operator is a generator of rotations in the spinor space (see (2.87)):

n = eiSn/~
R

(6.62)

z = e2iSz /~ . The eigenvalues of this operator are


Consider a rotation through 2 about the z-axis R
2

1 for integer
2i
e
= cos(2) + i sin(2) =
= (1)2s
1 for half-integer
Example. In the case s =

(6.63)

1
2

we can write (6.62) using (6.61) as






1
1

n
i 12 n

n + i sin i
n
= cos i
R = e
2
2

2k

2k+1

X
X
1
1
1
1
k
k
n
n
=
(1) i
+i
(1) i
(2k)!
2
(2k + 1)!
2
k=0
k=0
 2k
 2k+1

X
X
1

k
k

=
(1)
1+i
(1)
(n )
(2k)!
2
(2k + 1)!
2
k=0

k=0

.
= cos(/2)
1 + i sin(/2)(n )

In particular, rotation about z-axis is given by

z = cos(/2) 1 + i sin(/2)
R
z =

ei/2
0
0 ei/2

(6.64)

(6.65)

Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, 54-56, Merzbacher,
Quantum Mechanics 3rd edition, 16.1-16.4.

Kirill Tuchin

Phys. 591 Lecture Notes


3.

75

Addition of angular momenta

Consider a system consisting of two parts each having angular momentum J1,2 . Assume that the parts are noninteracting so that [J1i , J2k ] = 0. Then the total system can be in states with definite values of
J12 = ~2 j1 (j1 + 1) ,

and J22 = ~2 j2 (j2 + 1)

(6.66)

and J2z = ~m2 .

(6.67)

and their z-projections


J1z = ~m1 ,

Denote the corresponding state vectors by |j1 m1 j2 m2 i = |j1 m1 i|j2 m2 i. At fixed j1 and j2 there are (2j1 + 1)(2j2 + 1)
eigenfunctions with different m1 and m2 .
Define the operator of the total angular momentum as J = J1 + J2 . It satisfies the same commutation relations as
J1 and J2 :
[Ji , Jk ] = i~ikl Jl .

(6.68)

J2 = J12 + J22 + 2J1 J2 .

(6.69)

The Casimir operator can be written as

Since J2 is the Casimir operator of the algebra (6.68), it commutes with all components of J, in particular, [J2 , Jz ] = 0.
It also commutes with the Casimir operators of subsystems:
[J2 , J12 ] = [J2 , J22 ] = 0 ,

(6.70)

because [J12 , J1k ] = 0. Note, however, that [J2 , J1z ] 6= 0 since [J1 J2 , J1z ] 6= 0.
Since Jz = J1z + J2z commutes with operators J12 , J22 , J1z , J2z , states |j1 m1 j2 m2 i are also its eigenstates corresponding to the eigenvalues Jz = ~m = ~(m1 + m2 ). However, they are not eigenstates of the operator J2 , because
J2 does not commute with J1z and J2z . Operator J1 J2 mixes states differing by one in m1 and m2 . In practice, it
is advantageous to have states with definite values of J2 and Jz . Since J12 and J22 commute with them we can choose
as a basis set of states, eigenstates of these four operators: J2 , Jz , J12 and J22 . I will denote such states by |j1 j2 jmi.
In view of completeness of both sets we can write the following unitary transformation
X
|j1 j2 jmi =
hj1 m1 j2 m2 |j1 j2 jmi|j1 m1 i|j2 m2 i .
(6.71)
m1 ,m2

The matrix elements hj1 m1 j2 m2 |j1 j2 jmi are called the Clebsch-Gordan coefficients. Often, when j1 and j2 are fixed
they are denoted simply by hm1 m2 |jmi to simplify the notation.
Properties of the Clebsch-Gordan coefficients:

1. Clebsch-Gordan coefficients are non-vanishing only if m = m1 + m2 . Indeed, on the one hand


X
Jz |j1 j2 jmi =
hm1 m2 |jmi~(m1 + m2 )|j1 m1 i|j2 m2 i = ~(m1 + m2 )|j1 j2 jmi .

(6.72)

m1 ,m2

On the other hand, Jz |j1 j2 jmi = ~m|j1 j2 jmi implying that m1 + m2 = m.


2. At fixed j1 and j2 number j can take only the following values:
|j1 j2 | j j1 + j2 .

(6.73)

A derivation of this property can be found in the book by Merzbacher, 17.5. Below I will illustrate it with an
example (Example 1).
To each value of j, at fixed j1 and j2 , there correspond 2j + 1 values of m: m = j, j + 1, . . . , j 1, j. The
total number of states with all possible values of j:
jX
1 +j2

(2j + 1) = (2j1 + 1)(2j2 + 1) .

j=|j1 j2 |

(6.74)

Kirill Tuchin

76

Phys. 591 Lecture Notes

According to (6.71) instead of the (2j1 + 1)(2j2 + 1) basis states


{|j1 m1 j2 m2 i}m1 =j1 ,...,j2 ; m2 =j2 ,...,j2 ,

(6.75)

{|j1 j2 jmi}j=|j1 j2 |,...,j1 +j2 ; m=j,...,j

(6.76)

which form a complete set in the Hilbert space, we can use another complete set
spanning the same Hilbert space (and containing the same number (2j1 + 1)(2j2 + 1) of basis states).
3. By convention, Clebsch-Gordan coefficients are real. This implies that the inverse transformation to (6.71) is
X
|j1 m1 i|j2 m2 i =
hj1 j2 m1 m2 |j1 j2 jmi|j1 j2 jmi .
(6.77)
jm

4. Orthogonality and normalization


X
hj1 j2 m1 m2 |j1 j2 jmihj1 j2 m01 m02 |j1 j2 jmi = m1 m01 m2 m02 ,

(6.78)

jm

m1 ,m2

hj1 j2 m1 m2 |j1 j2 j 0 m0 ihj1 j2 m1 m2 |j1 j2 jmi = jj 0 mm0 ,

(6.79)

5. Symmetry under particle permutations:


which implies that

hj1 j2 m1 m2 |j1 j2 jmi = (1)j1 +j2 j hj2 j1 m2 m1 |j2 j1 jmi ,

(6.80)

|j1 j2 jmi = (1)j1 +j2 j |j2 j1 jmi .

(6.81)

Examples.

1. Consider two systems with j1 = 3 and j2 = 3/2. The total number of basis states is (3 2 + 1)( 23 2 + 1) = 28.
Let us list and count all these states in two different representations: |j1 j2 m1 m2 i and |j1 j2 jmi. I will organize
the list starting from the largest m and decreasing it by one. Keep in mind that for a given m, m1 + m2 = m
and j |m|.
|j1 j2 m1 m2 i
m1
3

|j1 j2 jmi

m2 # of states m j # of states
3/2
1
9/2 9/2
1

3
2

1/2
3/2

7/2

9/2
7/2

3
2
1

1/2
1/2
3/2

9/2
5/2 7/2
5/2

3
2
1
0

3/2
1/2
1/2
3/2

9/2
3/2 7/2
5/2
3/2

9/2
7/2
1/2 5/2
3/2

2 3/2
1 1/2
0 1/2
1 3/2

(6.82)

no state with j=1/2

>

1/2


Kirill Tuchin

77

Phys. 591 Lecture Notes

The table continues symmetrically to the values m = 1/2, 3/2, 5/2, 7/2, 9/2. Going down the table I
included states with j = 9/2, . . . , 3/2 in order that the number of states in both sets be the same. It is seen
that for m = 1/2, the state with j = 1/2 does not exist. This agrees with the rule that 3 3/2 j 3 + 3/2.
2. Consider the expectation value of J = J1 + J2 in a state |j1 m1 j2 m2 i. Since J2 = J12 + J22 + 2J1 J2 and
hJx i = hJy i = 0 (home assignment) we get

2
J = ~2 j1 (j1 + 1) + ~2 j2 (j2 + 1) + 2~2 m1 m2 .
(6.83)
3. Using the table of the Clebsch-Gordan coefficients.
(a) j1 = 3/2, j2 = 1/2.

1
3
|j = 1, m = 1i .
|m1 = 3/2, m2 = 1/2i = |j = 2, m = 1i +
2
2

(6.84)

(b) j1 = 2, j2 = 1.
|m1 = 0, m2 = 0i =

3
|j = 3, m = 0i
5

2
|j = 1, m = 0i .
5

(6.85)

(c) j1 = 2, j2 = 1/2.
2
1
|j = 5/2, m = 3/2i = |m1 = 2, m2 = 1/2i + |m1 = 1, m2 = 1/2i .
5
5

(6.86)

Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, , Merzbacher, Quantum Mechanics 3rd edition,
4.

Matrix elements of vector operators

We have noticed in I 4 that commutation relations of vector operators with the operator of orbital angular momen is proportional to the generator of
tum are similar, see (1.59) and (1.60). In II 5 we attributed this to the fact that L
rotations: all vector quantities transform under the rotations in the same way. (This is true for any tensor quantity,
but we will consider vectors for simplicity). This is why we defined a vector operator by (2.94). We can generalize
this definition as follows:
[Ji , Vj ] = i~ijk Vk .

(6.87)

Eq. (6.87) allows us to relate the matrix element of the operator V in the basis |jmi to the matrix elements of
operators J and J V in the same basis. To derive the corresponding expression, start with the following commutator
[J2 , J V ] = 2i~{J2 V (J V )J} ,

(6.88)

which can be proved using (6.87),(6.1) and (C7). Let us compute the matrix elements hjm0 | . . . |jmi on the both sides
of (6.88). Firstly,
hjm0 |[J2 , J V ]|jmi = hjm0 |J2 (J V ) (J V )J2 |jmi = ~2 hjm0 |(J V ){j(j + 1) j(j + 1)}|jmi = 0 . (6.89)
Secondly, consider the following matrix element
hjm0 |(J V )J|jmi = hjm0 |(J V )

j 0 ,m00

|j 0 m00 ihj 0 m00 |J|jmi

(6.90)

It is easy to check that [J2 , J V ] = 0 and [Jz , (J V )] = 0 (because J V is a scalar). Therefore, state vectors |jmi
are eigenstates of operator J V implying that
hjm0 |J V |j 0 m00 i m0 m00 j 0 j 0 .

(6.91)

Kirill Tuchin

Phys. 591 Lecture Notes

78

It follows from (6.90) and (6.91) that


hjm0 |(J V )J|jmi = hjm0 |J V |jm0 ihjm0 |J|jmi .

(6.92)

hjm0 |J2 V |jmi = ~2 j(j + 1)hjm0 |V |jmi .

(6.93)

Lastly,

Combining (6.88),(6.89),(6.92) and (6.93) we finally obtain


hjm0 |V |jmi =

1
hjm0 |J V |jm0 ihjm0 |J|jmi .
~2 j(j + 1)

(6.94)

Equivalently, replace m m0 , take the complex conjugation on both sides and use the fact that all involved operators
are Hermitian to write (6.94) in a slightly different way:
hjm0 |V |jmi =

1
hjm|J V |jmihjm0 |J|jmi ,
~2 j(j + 1)

(6.95)

which indicates that the matrix element hjm|J V |jmi does not depend on m at all.
This is a particular case of the Wigner-Eckart theorem. The Wigner-Eckart theorem separates the geometric and
symmetry-related properties of the matrix elements (in our example hjm0 |J|jmi) from the other physical properties
that are contained in the reduced matrix element (in our example hjm|J V |jmi).
Examples.
1. Suppose that J = J1 + J2 and let us compute the matrix elements hj1 j2 jm|J1,2 |j1 j2 jmi.
According to (6.94)

hjm0 |J1,2 |jmi =

1
~2 j(j

+ 1)

hj, m|J J1,2 |jmihjm0 |J|jmi .

(6.96)

Decompose
1
1
J = (ex iey )J+ + (ex + iey )J + ez Jz .
2
2

(6.97)

and calculate the matrix element


p
1
hjm0 |J|jmi = ~(ex iey ) j(j + 1) m(m + 1)m0 ,m+1
2
p
1
+ ~(ex + iey ) j(j + 1) m(m + 1)m0 ,m1 + ez ~mmm0 .
2

(6.98)

Next, using the identities

1
J J1 = (J2 + J12 J22 ) ,
2

1
J J1 = (J2 + J22 J22 ) .
2

(6.99)

we have
1
~2
hjm|J J1 |jmi = hjm| (J2 + J12 J22 )|jmi =
[j(j + 1) + j1 (j1 + 1) j2 (j2 + 1)] .
2
2

(6.100)

Notice that the reduced matrix element (6.100) does not depend on m and m0 as expected. Plugging (6.100)
and (6.98) into (6.96) we derive


~m
j1 (j1 + 1) j2 (j2 + 1)

hjm|J1 |jmi = ez
1+

,
(6.101)
2
j(j + 1)
j(j + 1)


~m
j1 (j1 + 1) j2 (j2 + 1)
1
+
.
(6.102)
hjm|J2 |jmi = ez
2
j(j + 1)
j(j + 1)

Kirill Tuchin

Phys. 591 Lecture Notes

79

2. Consider electron in a state with given angular orbital momentum l, and z-component m of its total angular
momentum j. The possible values of the total angular momentum are j = l + 1/2 and j = l 1/2. Using (6.101)
and J2 S we have the following diagonal matrix elements
and (6.102) with replacements J1 L


D
1
~m
3
~m
1
1
1E
l(l + 1)
= ez
+
= ez
l + , m; l S l + , m; l
1
,
(6.103)
1
3
1
3
2
2
2
2
2
2l + 1
(l + 2 )(l + 2 ) 4(l + 2 )(l + 2 )


D
1
~m
3
~m
1
1
1E
l(l + 1)
= ez
+
= ez
l , m; l S l , m; l
1
,
(6.104)
2
2
2
2
2
2l + 1
(l 12 )(l + 21 ) 4(l 21 )(l + 12 )


D
1
~m
3
~m(l + 1)
1
1
1E
l(l + 1)
= ez
= ez
l + , m; l L l + , m; l
1+
,
(6.105)
1
3
1
3
2
2
2
2
2
2l + 1
(l + 2 )(l + 2 ) 4(l + 2 )(l + 2 )


D
1
~m
1
1E
l(l + 1)
3
~m(l + 1)
1
= ez
1+

= ez
l , m; l L
.
(6.106)
l , m; l
1
1
1
1
2
2
2
2
2
(l 2 )(l + 2 ) 4(l 2 )(l + 2 )
l + 12
We will see later that the magnetic dipole moment of electron is given by
=

e
.
(L + 2S)
2M c

(6.107)

Its expectation value in two states with different j are


e D
l+
2M c
e D
=
l
2M c

hij=l+ 1 =
2

hij=l 1


1
1
l +
, m; l (L
+ 2S)
2
2

1
1
l
, m; l (L
+ 2S)
2
2

1
1E
e ~m(l + 1)
, m; l
= ez
,
2
2
2M c l + 12
1
1E
e ~ml
, m; l
,
= ez
2
2
2M c l + 12

(6.108)
(6.109)

Additional reading: Landau and Lifshitz, Quantum Mechanics: Non-relativistic theory, 106,107, Merzbacher,
Quantum Mechanics 3rd edition, Ch. 17.

Kirill Tuchin

Phys. 591 Lecture Notes


VII.

80

Kirill Tuchin

Phys. 591 Lecture Notes

81

Appendix A: Fourier analysis

Any integrable function f (x) can be expanded in a complete set of functions {eikn x }n in the interval L/2 x
L/2, known as Fourier series, as follows

f (x) =

fn eikn x ,

kn =

n=

2n
.
L

(A1)

The Fourier coefficients fn are given by


fn =

1
L

L/2

f (x)eikn x dx .

(A2)

L/2

In the limit L (A1),(A2) become continuous Fourier integral expansion


Z +
dk
f (x) =
fk eikx
,
2

Z +
fk =
f (x)eikx dx ,

(A3)
(A4)

Three-dimensional generalization of (A3),(A4) is


f (r) =
fk =

fk eikr

Z +

d3 k
,
(2)3

f (r)eikr d3 r ,

(A5)
(A6)

Additional reading: Arfken at. al. , Mathematical methods for physicists, 7th edition, Ch.19, 20.1-20.4.
Appendix B: Dirac delta function

Dirac delta function (x) is a singular function such that


(
0 , if x 6= 0 ,
(x) =
, if x = 0 ,

(B1)

and
Z

(x)dx = 1 .

(B2)

This definition implies the main property of the delta function: if f (x) is a smooth function and a < 0, b > 0, then
Z

f (x)(x)dx = f (0) .

(B3)

In particular,
Z

f (x)(x)dx = f (0) .

(B4)

Shifting the origin x x a we get in place of (B4)


Z +
f (x)(x a)dx = f (a) .

(B5)

Kirill Tuchin

Phys. 591 Lecture Notes

82

Often one needs do deal with a delta function of the form [(x)], where is a smooth function. Suppose that
(x) has one and only zero at x = x0 . Then integrand in
Z +
f (x)[(x)]dx
(B6)

is non-vanishing only at x = x0 . Thus, we can replace with the first non-vanishing term of its Taylor expansion
(x) = 0 (x0 )(x x0 ). We have
Z +
Z +
f (x)[(x)]dx =
f (x)[0 (x0 )(x x0 )]dx.
(B7)

Making change of variables y = 0 (x0 )(x x0 ) we get using (B4)


Z +
Z +
1
f (x0 )
f (x)[(x)]dx = 0
f [y/0 (x0 ) + x0 ](y)dy = 0
.
|
(x
)|
|
(x0 )|
0

Generally, if (x) has zeros at x = xn , were n is any integer number bigger than zero, then (B8) becomes
Z +
X f (xn )
f (x)[(x)]dx =
.
|0 (xn )|

n
In a particular case (x) = cx, with constant c:
Z +

f (x)(cx)dx =

1
f (0) .
|c|

(B8)

(B9)

(B10)

Delta function can be represented in different forms that are often very useful in applications. As a Fourier integral:
Z +
dk
(x) =
eikx
.
(B11)
2

As a limit of elementary functions:



1
lim
0 x2 + 2
1
sin(x/)
= lim
0
x
2
2
1
1
= lim 2 ex /4
2 0 

(x) =

Note that (B11) and (B13) are mutually consistent as follows:


Z 1/
1
1
sin(x/)
lim
eikx dk = lim
2 0 1/
0
x

(B12)
(B13)
(B14)

(B15)

Three-dimensional delta function is defines as a product of three one-dimensional delta-functions


(r r 0 ) = (x x0 )(y y 0 )(z z 0 ) .

(B16)

Additional reading: Arfken et. al. Mathematical methods for physicists, 7th edition, 1.11.
Appendix C: Levi-Civita symbol

Levi-Civita symbol ijk is defined as

+1 if (i, j, k) is (1, 2, 3), (3, 1, 2), (2, 3, 1) ,


ijk = 1 if (i, j, k) is (1, 3, 2), (3, 2, 1), (2, 1, 3) ,

0 if i = j or j = k or k = i

(C1)

Kirill Tuchin

Phys. 591 Lecture Notes

83

Levi-Civita symbol is completely antisymmetric: ikj = ijk , jik = ikj etc. That is, permutation of any two
nearby indices produces the minus sign. The nearby pairs of indices of ikj are i and k, k and j, j and i. With this
notation the ith component of vector A = B C can be written as Ai = ijk Bj Ck , where summation over the
repeated indices is implied(11 . The scalar product of any two vectors A and B reads A B = Ai Bi .
We can use this, so called index notation to easily prove vector identities. For example,
A (B C) = ei ijk Aj klm Bl Cm = ei (il jm im jl )Aj Bl Cm = B(A C) C(A B) ,
where ei is a unit vector in ith direction.
It is helpful to remember that if matrix [S]ij is symmetric, i.e. [S]ij = [S]ji and [A]ij is antisymmetric, i.e.
[A]ij = [A]ji , then
[S]ij [A]ij = [S]ji [A]ji = [S]ij [A]ij = 0

(C2)

where we first permuted the indices i j, j i and than relabeled i as j and j as i.

The differential operator nabla is defined as = {x , y , z }, where i x


. Some operations with nabla:
i
A ij i Aj = i Ai

(C3)

= ei i

(C4)

A = ijk ei j Ak

(C5)

Examples of differential identities:


(A B) = i (ijk Aj Bk ) = ijk (i Aj )Bk + ijk (i Bk )Aj
= ijk (i Aj )Bk ikj (i Bk )Aj = ( A) B ( B) A
( A) = i ijk j Ak = 0,

(C6)

since i j = j i is symmetric, whereas ijk is asymmetric in any pair of its indices.


A useful property of the Levi-Civita symbol:
ijk ilm = jl km jm kl .

(C7)

It helps simplify multiple cross products.

(11

In other words, we do not write explicitly the summation sign


indices i, j are called dummy indices.

j,k .

This is called the Einstein summation convention. The repeated

Kirill Tuchin

84

Phys. 591 Lecture Notes

34. Clebsch-Gordan coefficients

010001-1

34. CLEBSCH-GORDAN COEFFICIENTS, SPHERICAL HARMONICS,


AND d FUNCTIONS
p
Note: A square-root sign is to be understood over every coefficient, e.g., for 8/15 read 8/15.

1/2 1/2

1
0
+1
1
0
0
+ 1/2 + 1/2 1
+ 1/2 1/2 1/2 1/2 1
1/2 + 1/2 1/2 1/2 1

1 1/2
+ 1 + 1/2

3/2
+ 3/2 3/2 1/2
1 + 1/2 + 1/2

+ 1 1/2
0 + 1/2

1/3 2/3 3/2 1/2


2/3 1/3 1/2 1/2

0 1/2
1 + 1/2
3
+3 3
2
+2 +1 1 +2 +2

21

2/3 1/3 3/2


1/3 2/3 3/2

1 1/2

+ 1/2 + 1/2 3/4 1/4

5/2
+ 5/2
+ 3/2 + 1
1
+ 3/2 0
+ 1/2 + 1

3/21

2
+ 2 0 1/3 2/3
1
3
+ 1 + 1 2/3 1/3
+1
+1
+1
+ 2 1 1/15 1/3 3/5
2
1
1/6 3/10
3
1 1 + 22 2 1 + 10 + 01 8/15
2/5 1/2 1/10
0
0
0
+1 +1 1 +1 +1
+ 1 1 1/5 1/2 3/10
0
+ 1 0 1/2 1/2
2
1
0 2/5
0 0 3/5
0 + 1 1/2 1/2
0
0
0
1 + 1 1/5 1/2 3/10

+ 1 1 1/6 1/2 1/3


0 1/3 2
0 0 2/3
1 + 1 1/6 1/2 1/3 1

4
+4
+2 +2 1
+2 +1
+1 +2

3/2

d 3/2,3/2 =

cos
2
2
1 + cos

= 3
sin
2
2
1 cos

3/2
d 3/2,1/2 = 3
cos
2
2

1 cos
3/2
d 3/2,3/2 =
sin
2
2
3 cos 1

3/2
d 1/2,1/2 =
cos
2
2
3 cos + 1

3/2
d 1/2,1/2 =
sin
2
2

3/2
d 3/2,1/2

...
...

3
1

d `m,0 =

1
1

2
1

1/2 1/2 3/4 1/4 2


3/2 1/2
3/2 + 1/2 1/4 3/4 2
+ 1/2 + 1/2
3/2 1/2 1
2/5 1/2
+ 3/2 1 1/10
+ 1/2 0 3/5 1/15 1/3 5/2
3/2 1/2
1/2 + 1 3/10 8/15 1/6 1/2 1/2 1/2
+ 1/2 1 3/10 8/15 1/6
2
1
1/2 0 3/5 1/15 1/3 5/2 3/2
1
1
3/2 + 1 1/10 2/5 1/2 3/2 3/2
1/2 1 3/5 2/5 5/2
1/2 1/10
3/2 0 2/5 3/5 5/2
2
1/6 3/10
3
1/3 3/5 2 2
3/2 1
1
1 1 2/3 1/3 3
2 0 1/3 2/3 3
2 1 1

4
Y m eim
2` + 1 `

hj1 j2 m1 m2 |j1 j2 JM i

= (1)Jj1 j2 hj2 j1 m2 m1 |j2 j1 JM i

3/2 3/2

1 + cos
1/2
+3 3
2
d 10,0 = cos
d 1/2,1/2 = cos
d 11,1 =
2
2
+ 3/2 + 3/2
1 +2
+2

2 3/2 + 7/2
3
2
1
+ 3/2 + 1/2 1/2 1/2
1/2
1 = sin
7/2 7/2 5/2

d
d
=

sin
1,0
+ 1/2 + 3/2 1/2 1/2 + 1 + 1
+1
1/2,1/2
2
2
+ 2 + 3/2
1 + 5/2 + 5/2
+ 3/2 1/2 1/5 1/2 3/10
5/2 3/2
+ 2 + 1/2 3/7 4/7 7/2
1 cos
0
3
2
1
+ 1/2 + 1/2 3/5
0 2/5
d 11,1 =
+ 1 + 3/2 4/7 3/7 + 3/2 + 3/2 + 3/2
0
0
0
1/2 + 3/2 1/5 1/2 3/10
0
2
+ 2 1/2 1/7 16/35 2/5
+ 3/2 3/2 1/20 1/4 9/20 1/4
5/2 3/2
1/2
+ 1 +1/2 4/7 1/35 2/5
7/2
1/4

1/20

1/4
9/20
+ 1/2 1/2
0 +3/2 2/7 18/35 1/5 + 1/2 + 1/2 + 1/2 + 1/2
1
3
2
1/2 + 1/2 9/20 1/4 1/20 1/4
3
4
1
3/2 + 3/2 1/20 1/4 9/20 1/4 1 1
+ 2 3/2 1/35 6/35 2/5
2/5
+3 +3
0 3/10
+ 1 1/2 12/35 5/14
+
1/2

3/2
1/5
1/2
3/10
3
2
4
1/2 1/2
0 +1/2 18/35 3/35 1/5
5/2 3/2 1/2
1/5
7/2
0 2/5
1/2 1/2 3/5
2
3
1/2 1/2 + 2
+2 +2
1 +3/2 4/35 27/70 2/5 1/10 1/2 1/2 1/2 1/2
3/2 + 1/2 1/5 1/2 3/10 2 2
+ 2 0 3/14 1/2 2/7
+ 1 3/2 4/35 27/70 2/5 1/10
1/2
1/2

1/2

3/2
3
4
0 3/7
3
2
1
+ 1 +1 4/7
0 1/2 18/35 3/35 1/5 1/5
3/2 1/2 1/2 1/2 3
+1
+1
+1
+1
0 +2 3/14 1/2 2/7
1 +1/2 12/35 5/14
5/2 3/2
0 3/10 7/2
3/2 3/2 1
2 +3/2 1/35 6/35 2/5 2/5 3/2 3/2 3/2
+ 2 1 1/14 3/10 3/7 1/5
+ 1 0 3/7 1/5 1/14 3/10
0 3/2 2/7 18/35 1/5
0 +1 3/7 1/5 1/14 3/10
1
0
4
3
2
1 1/2 4/7 1/35 2/5 7/2
5/2
0
0
0
1 +2 1/14 3/10 3/7 1/5
0
0
2 + 1/2 1/7 16/35 2/5 5/2 5/2
+ 2 2 1/70 1/10 2/7 2/5 1/5
1 3/2 4/7
3/7 7/2
+ 1 1 8/35 2/5 1/14 1/10 1/5
2 1/2 3/7 4/7 7/2
0 2/7
0 1/5
0 0 18/35
2 3/2
1
8/35 2/5 1/14 1/10 1/5
4
1 +1
2
1
3
2 +2
1/70 1/10 2/7 2/5 1/5
1
1
1
1
1 + cos

d m0 ,m = (1)mm d m,m0 = d m,m0

22

J
M

m2

+ 1/2 1/2 1/2 1/2


1/2 + 1/2 1/2 1/2

5/2 3/2
+ 3/2 + 3/2
2/5 3/5 5/2
3/5 2/5 + 1/2

0 1 2/5
1 0 8/15
2 + 1 1/15

1
1

0 1 1/2 1/2 2
1 0 1/2 1/2 2
1 1 1

Y`m = (1)m Y`m

m1

J
M

3
cos
m 1 m 2 Coefficients
2 1/2 + 5/2
5/2 5/2 3/2
4
.
.
r
+ 2 +1/2
1 + 3/2 + 3/2
.
.
3
.
.
Y11 =
sin ei
+ 2 1/2 1/5 4/5 5/2 3/2
8
+ 1 + 1/2 4/5 1/5 + 1/2 + 1/2
r 

+ 1 1/2 2/5 3/5 5/2 3/2
5 3
1
Y20 =
cos2
0 + 1/2 3/5 2/5 1/2 1/2
4 2
2
r
0 1/2 3/5 2/5 5/2 3/2
15
1 + 1/2 2/5 3/5 3/2 3/2
Y21 =
sin cos ei
2
1 1/2 4/5 1/5 5/2
8
3/2 1/2 + 2 2 1
r
2 + 1/2 1/5 4/5 5/2
1 15
+ 3/2 +1/2 1 + 1 + 1
2
2
2i
2 1/2
1
Y2 =
sin e
4 2
1
+ 3/2 1/2 1/4 3/4 2

Y10 =

1/2 1/2 1

Notation:

d 22,2 =

2
1 + cos
=
sin
2

6
d 22,0 =
sin2
4
1 cos
d 22,1 =
sin
2
 1 cos 2
d 22,2 =
2

d 22,1

3/7 1/5
+ 1 2 1/14 3/10
1/5 1/14 3/10
0 1 3/7
1 0 3/7 1/5 1/14 3/10
2 +1 1/14 3/10
3/7 1/5

 1 + cos 2

1 + cos
(2 cos 1)
2
r
3
d 21,0 =
sin cos
2
1

cos

d 21,1 =
(2 cos + 1)
2

d 21,1 =

4
2

3
2

2
2

0 2 3/14 1/2 2/7


0 3/7
1 1 4/7
2 0 3/14 1/2 2/7

1
2

4
3
3 3
2 1/2 1/2 4
1 1/2 1/2 4

d 20,0 =

3

cos2

1

Figure 34.1: The sign convention is that of Wigner (Group Theory, Academic Press, New York, 1959), also used by Condon and Shortley (The
Theory of Atomic Spectra, Cambridge Univ. Press, New York, 1953), Rose (Elementary Theory of Angular Momentum, Wiley, New York, 1957),
and Cohen (Tables of the Clebsch-Gordan Coefficients, North American Rockwell Science Center, Thousand Oaks, Calif., 1974). The coefficients
here have been calculated using computer programs written independently by Cohen and at LBNL.

FIG. 10: Clebsch-Gordan coefficients

You might also like