You are on page 1of 144

Lasers and Quantum Optics

J.T. Mendonca
Instituto Superior Tecnico
2
Contents
1 Introduction 5
2 Elementary Balance Equations 9
2.1 Planck Law of Thermal Radiation . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Einstein Coecients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.3 Optical pumping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.4 Laser cavity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.5 Simplied Laser model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.6 Relaxation oscillations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.7 Short Laser Pulses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.7.1 Q-switching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.7.2 Mode locking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.8 Amplied spontaneous emission . . . . . . . . . . . . . . . . . . . . . . . . 30
2.9 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3 Quantum theory of Radiation 33
3.1 Field Quantization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.2 Fock States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.3 Phase Operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.4 Field Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.5 Coherent States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.6 Displacement Operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.7 Density Operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
3
4
3.8 Other eld representations . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
3.8.1 P-representation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
3.8.2 Q-representation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
3.9 Generalized representation . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
3.10 Photon angular momentum . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.11 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4 Semi-Classical Theory of Atom-Field Interaction 73
4.1 Time-dependent Perturbations . . . . . . . . . . . . . . . . . . . . . . . . . 73
4.2 Interaction Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.3 Radiative Transitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
4.4 Rabi Oscillations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
4.5 Density matrix description . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
4.6 Selection Roles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
4.7 Susceptibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
4.8 Semi-Classical Laser Theory . . . . . . . . . . . . . . . . . . . . . . . . . . 90
4.9 Laser cooling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
4.10 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
5 Quantum Theory of Atom-Field Interaction 101
5.1 Hamiltonians . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
5.2 Interaction Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
5.3 Interaction with a single eld mode . . . . . . . . . . . . . . . . . . . . . . 106
5.4 Spontaneous emission . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
5.5 Modied refractive index . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
5.6 Quantum theory of the laser . . . . . . . . . . . . . . . . . . . . . . . . . . 118
5.7 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
6 Quantum Coherence 127
6.1 Field Correlations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
6.2 Intensity Correlations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
6.3 Squeezed States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
6.4 Photon entanglement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
Chapter 1
Introduction
These are the lecture notes of a semester course on the elements of Quantum Optics and
Lasers. Teaching stated in March 2008, and this course resulted from an initial proposal
of two distinct courses, one on Quantum Optics, and another one on Lasers. One of
the main diculties of the present course is therefore due to the co-existence of two
distinct (although strongly related) subjects. Such an hybrid purpose makes teaching of
the present course dicult but challenging.
Another kind of diculty results from the existence of two dierent groups of students
attending the same course. Half of the students are more technically oriented, and the
other half more theoretically motivated. So, some parts of this course risk to be seen as
obvious to one group, while being very demanding for the other one. We have tried to
circumvent this diculty by alternating easy chapters with more formal and dicult ones.
As a results, the structure of the course follows a pedagogical sequence, in detriment of a
more logical one. In particular, it would have been more logical to write the semi-classical
theory of chapter 3 before, and not after, the quantum eld theory of chapter 4. But the
present structure seems to work well in the class.
In order to illustrate the contents of these lectures, it is useful to look at the Table 1,
where there dierent levels of conceptual understanding of matter-radiation interaction are
represented. Chapter 2 is devoted to the simplest level of description of electromagnetic
radiation, where photons are point particles with no phase, as considered by geometric
5
6
optics. On the other end, matter is assumed as seen by the old quantum theory of Planck
and Einstein. Simple balance equations can make the account of the number of photons,
as well as the number of atoms in two or three distinct energy levels. In this context we
discuss a simple but useful model for the laser, and use it to illustrate various aspects of
laser operation. In Chapter 4 we take a step forward, and use the full wave description
of electromagnetic radiation, based on Maxwells equations. This Chapter deals with
the so-called Semi-Classical theory of matter-radiation interaction, where radiation is
described in classical terms and the atoms are described by quantum mechanics and, in
particular, by the Shroedinger equation. This semi-classicla approach allows us to derive
laser equations, where the photon phase is now present, thus completing the simple model
of Chapter 2.
In Chapter 3 we discuss the elements of the quantum theory of radiation as described
by the quantum eld theory, which is the basic foundation of modern Quantum Op-
tics. A remarkable result of the electromagnetic eld quantization is the reappearance
of Maxwells equations, in the form of Heisenber equations for the eld operators. The
important concept of a coherent state and the asociated displacement operator, is intro-
duced here. It must be said that Quantum Optics corresponds to the non-relativistic
version of quantum electrodynamics, where the Schroedinger equation (and not Dirac
equation) is used to describe the atoms. In Chapter 5, we then consider matter-radiation
in its full quantum description, where both atoms and photons, or mater and radiation,
are described in purely quantum terms. In particular, this will lead to a third, and more
rened, level of laser equations.
As usual in lecture notes, the present text strongly borrows from many well known
books and reference papers, specially from those included in the Bibliography. A future
version of the present course will eventually give more attention to quantum optical appli-
cations, such as quantum teleportation, quantum computation and atomic manipulation
by lasers. It will also include a discussion of nonlinear optics and a short chapter on
quantum vacuum properties.
IST, 8th of June 2009.
7
Second
quantization
Quantum Optics
QED
Fields
(particles+waves)
Quantum
Mechanics
Maxwells
equations
Waves
Classical
Mechanics
Geometric optics
approximation
Particles
Matter Radiation Type of description
Figure 1.1: Table 1: Description of matter and radiation
8
Chapter 2
Elementary Balance Equations
2.1 Planck Law of Thermal Radiation
Let us consider the electromagnetic radiation conned inside an optical cavity. For sim-
plicity, we assume a cubic cavity with side L, and perfectly conducting walls. The classical
electric eld of the radiation trapped inside the cavity can be described by the wave equa-
tion
_

1
c
2

2
t
2
_

E = 0 (2.1)
where c is the speed of light in vacuum. The tangential components of the eld

E has to
vanish at the cavity walls. If we assume wave solutions of the form

E(r, t) =

E
0
exp(i

k r it) (2.2)
this implies that the three wavevector components have to obey the following conditions
k
i
= 2n
i
/L , (i = x, y, z) (2.3)
where n
i
are zero or integers. Each eld mode in the cavity will then be characterized by
a distinct choice of values of (n
x
, n
y
, n
z
).
Let us consider the number of distinct eld modes n(

k), that can be dened in the


cavity, in the wavevector interval between

k and

k + d

k. By dierentiation of (2.3), we
9
10
get
d

k =
_
2
L
_
3
dn
x
dn
y
dn
z
(2.4)
Noting that we have 2 independent polarization for each mode, we obviously have n(

k) =
2dn
x
dn
y
dn
z
. We can also write it as n(

k) = n

k
d

k, where n

k
is a density of modes. We
obtain
n

k
d

k = 2
_
L
2
_
3
d

k = 2V
d

k
(2)
3
(2.5)
where V = L
3
is the cavity volume. By performing an integration over the solid angle
_
d = 4, we can establish the number of modes in the wavenumber interval between k
and k + dk, as
n
k
dk = V
k
2

2
dk (2.6)
where n
k
is the number of modes per unit wavenumber. If we now want to determine the
corresponding number of modes per unit frequency, knowing that = kc in vacuum, we
obtain
n

d = V

2

2
c
3
d (2.7)
In general conditions, we can extend the volume V to innity, and convert the sum over
the discrete modes by an integration over k, or , using the conversion role

2
_
k
2
dk =
V

2
c
2
_

2
d (2.8)
Usually, we consider the number of modes per unit volume, and therefore take V = 1.
As we will see, each of eld modes will have an energy E
n
= (n + 1/2), where the
integer n is the number of photons, and is the Planck constant divided by 2. If these
eld modes are in thermal equilibrium with the conning medium, at a temperature T,
the probability for a given mode to have a number of photons n is determined by the
Boltzmann energy distribution
P
n
() =
1
Z
exp(E
n
/k
B
T) (2.9)
where k
B
it the Boltzmann constant and Z is the partition function, dened by the
11
normalization condition

n
P
n
= 1, as
Z =

n=0
exp(E
n
/k
B
T) (2.10)
When we replace this in equation (2.9), we notice that the zero point energy /2 cancels
out, and we are left with
P
n
() =
e
nx

n=0
e
nx
= (1 e
x
)e
nx
(2.11)
where we have used x = /k
B
T. We can now calculate the mean number of photons for
the eld mode with frequency at a temperature T, as
< n >=

n=0
nP
n
= (1 e
x
)

n=0
ne
nx
(2.12)
This can also be written as
< n >= (1 e
x
)
d
dx

n=0
e
nx
(2.13)
Using once more the geometric series summation, we get
d
dx

n=0
e
nx
=
d
dx
1
1 e
x
=
e
x
(1 e
x
)
2
(2.14)
This leads us to the Planck function
< n >=
e
x
1 e
x
=
1
e
x
1
(2.15)
In order to obtain the mean energy density in the frequency range between and +d,
we dene it as
W
T
()d = < n > n

d (2.16)
Using equations (2.7) and (2.15), we obtain
W
T
() =

3

2
c
3
1
exp(/k
B
T) 1
(2.17)
This is the famous Planck formula for the radiation energy density in thermal equilibrium
at temperature T.
12
2.2 Einstein Coecients
We now assume that thermal equilibrium between radiation and matter is attained by
elementary mechanisms of emission and absorption of photons by atoms, satisfying the
quantum condition that = E
2
E
1
, where is the frequency of the photons and E
1
and E
2
are the energy values two quantum states of the atoms, assuming that E
2
> E
1
.
Let us consider the number of atoms per unit volume in the lower energy state N
1
,
and the corresponding number in the upper state N
2
. According to Einstein, there are
three distinct processes of photon emission and absorption. First, absorption of photons
by atoms in the lower energy state will increase the number of atoms in the upper level,
as determined by
_
dN
2
dt
_
abs
= B
12
N
1
W() (2.18)
Such an increase is obviously proportional to the number of lower level atoms N
1
, and to
the intensity of radiation W(), with a constant of proportionality B
12
. In contrast with
this absorption process, emission of photons from the atoms in the upper level and the
subsequent increase of the population of this energy level, takes place by two dierent
processes, as described by
_
dN
2
dt
_
em
= A
21
N
2
B
21
N
2
W() (2.19)
The rst term describes spontaneous decay, and the second is related to stimulated emis-
sion. This second term is the exact reverse of absorption, but the rst term has no
analogue in classical physics. Two other coecients of proportionality for the emission
processes, A
21
and B
21
, are introduced. The total rate of change of the atomic population
in the upper energy level, can then be written as
dN
2
dt
= [A
21
+ B
21
W()]N
2
+ B
12
N
1
W() (2.20)
Considering that the total number of atoms N = N
1
+N
2
remains constant, we can also
write
dN
2
dt
=
dN
1
dt
(2.21)
13
The particular case of thermal equilibrium corresponds to the stationary condition
[A
21
+ B
21
W()]N
2
= B
12
N
1
W() (2.22)
which leads to the thermal energy distribution
W
T
() =
A
21
(N
1
/N
2
)B
12
B
21
(2.23)
Using the Boltzmann energy distribution for the atomic populations, and neglecting de-
generacy of energy levels, we can assume that
N
1
N
2
= exp
_

k
B
T
_
(2.24)
which then leads to
W
T
() =
A
21
B
12
1
exp(/k
B
T) (B
21
/B
12
)
(2.25)
This expression reduces to the Planck law (2.17) if we take the values of the emission and
absorption coecients equal to
A
21
B
12
=

3

2
c
3
,
B
21
B
12
= 1 (2.26)
As we will see, such a statement can only be completely justied by using the quantum
theory radiation.
Another interesting statement associated with the three Einstein processes is that we
can used the above balance equations to establish a rate equation for the total number of
photons n exchanged with the medium. This is due to the assumption that every emission
(or absorption) process increases (or decreases) the photon number by one. The resulting
rate equation is then given by
d
dt
n = (N
2
N
1
)BW() + AN
2
(2.27)
where we have used A = A
21
and B = B
12
= B
21
. We can see that
d
dt
n =
dN
1
dt
=
dN
2
dt
(2.28)
14
A
21
B
21
B
12

!
E
2
= E
1
+ h"
!
E
1
N
2
, g
2
N
1
, g
1
| 2 >
| 1 >
Figure 2.1: Einstein coecients: absorption, induced emission and spontaneous emission
of an atom with two quantum levels.
If the energy exchange between the atoms and the medium is made in a small frequency
domain around the atomic transition frequency , it is possible from here to derive the
number of photons for a given eld mode, n(), such that W() = n(). Replacing
this in equation (2.27), we obtain the photon rate equation for a single eld mode, as
d
dt
n() = (N
2
N
1
)w n() + AN
2
(2.29)
where we have introduced the new quantity
w =

2
c
3
A
V
2

(2.30)
and where 1/A, the spontaneous lifetime, and equations (2.26) were also used. The photon
rate equation (2.29) can be used as a simplied model equation for s phenomenological
laser theory, as shown later in this chapter. We should notice, in particular, that the
number of photons at a given eld mode can increase with time, if N
2
> N
1
, or in
other words, if there is inversion of population. Optical methods leading to inversion of
population are discussed next.
15
2.3 Optical pumping
It was Kastler who suggested for the rst time, in 1950, the idea of modifying the pop-
ulations of the energy levels of the atoms by using an intense light beam. This is called
optical pumping. Such an idea was explored later by Basov and Prokhorov for microwaves.
In order to understand this process, let us consider the interaction of an intense electro-
magnetic eld of frequency , with a gas of atoms possessing two energy levels such that
E
2
E
1
=
21
, and let us calculate the resulting changes in the populations N
1
and N
2
due to the action of the eld. Our starting point will be the balance equation
dN
2
dt
= AN
2
+ B(N
1
N
2
)W() (2.31)
where we have used B
12
= B
21
= B, which corresponds to g
1
= g
2
. Noting that the total
number of atoms is conserved, N = N
1
+ N
2
, this can also be written as
dN
2
dt
= BNW() [A + 2BW()] N
2
(2.32)
This can easily be integrated by assuming that the spectral energy density remains con-
stant during the interaction process. This is only valid if the fraction of energy absorbed
by the atoms remains negligible with respect to the total electromagnetic wave energy. As
a plausible initial conditon, we consider the case where the upper energy level is initially
empty, N
2
(t = 0) = 0. Such a condition is valid for a gas in thermal equilibrium, at a
temperature T, such that
21
k
B
T. The solution of equation (2.32) is
N
2
(t) =
BNW()
A + 2BW()
1 exp [(A + 2BW())t] (2.33)
This results shows that, for short times, the population of the upper energy level grows
linearly with time, as N
2
(t) BNW()t, whereas for very long times, such that t
[A + 2BW()]
1
, it tends to a constant
N
2
() =
BNW()
A + 2BW()
(2.34)
This asymptotic value depends on the electromagnetic wave intensity and, for very large
intensities such that W() A/2B, we get N
2
() N/2.
16
0 0.5 1 1.5 2 2.5 3
0
0.1
0.2
0.3
0.4
N2/N
t
BW()t
Figure 2.2: Temporal evolution of the upper level population, N
2
(t), for a two-level
atom pumped by nearly resonant radiation with energy density W(). The value of
A/2BW() = 0.1 was used.
This means that, for very intense radiation interacting with a gas of two-level atoms,
the maximum possible value for the population of its upper energy level is equal to half
of the total number of atoms N
2
= N
1
= N/2. It is therefore not possible by this means
to obtain an inversion of population such that N
2
> N
1
. It should also be noticed that
the saturation of the radiative transition, which corresponds to the equality between the
two populations N
2
= N
1
is more dicult to achieve in the optical domain than in the
radio-frequency domain, because the Einstein coecient of spontaneous emission, A is
much larger for optical transitions, which results in a faster decay of the upper energy
level, due to spontaneous emission (uorescence). This is obvious from equation (2.34),
which for A 2BW(), leads to N
2
() N[BW()/A] N.
Let us now consider the optical pumping with a three level atom system. as proposed
by Basov and Prokhorov em 1953. This will lead to an inversion of population, as shown
next. We will call [1 > the lowest energy level, and [2 > and [3 > the two excited states.
17
|3>
|2>
|1>
B
31
W
p
A
32
A
31 A
21
W()
Figure 2.3: Scheme for optical pumping of a three-level atom.
The electromagnetic wave pump with energy density W
p
() is nearly resonant with the
radiation transition between the two extreme energy states, such that E
3
E
1
=
31

. Given that the saturation state is very dicult to achieve in the optical domain, we
will always have N
3
N, where N = N
1
+ N
2
+ N
3
is the total number of atoms (per
unit volume). On the other end, the atoms at level [3 > can radiatively decay into the
intermediate level [2 >, by assuming dipolar electric transitions, emitting photons with
frequency
32
= (E
3
E
2
)/.
We can now show that, in such conditions, inversion of population between the two
excited states, [2 > and [3 > can be attained for appropriate values of the Einstein
coecients. We start with the balance equations for the three level populations
dN
1
dt
= A
21
N
2
+ A
31
N
3
B
31
W
p
(N
1
N
3
) (2.35)
dN
2
dt
= A
21
N
2
+ A
32
N
3
+ B
32
W(N
3
N
2
)
dN
3
dt
= (A
31
+ A
32
)N
3
+ B
31
W
p
(N
1
N
3
) B
22
W(N
3
N
2
)
18
Given the conservation of the total number of atoms, N, these equations have to verify
the condition
dN
1
dt
+
dN
2
dt
+
dN
3
dt
= 0 (2.36)
We now look at stationary solutions for the system of equations (2.33). From the rst
two equations we get
A
21
N
2
+ A
31
N
3
= B
31
W
p
(N
1
N
2
) (2.37)
A
21
N
2
A
32
N
3
= B
32
W(N
3
N
2
)
For convenience, we introduce the parameter r, such that
r =
1
N
B
31
W
p
(N
1
N
3
) (2.38)
It represents the fraction of atoms transferred by optical pumping onto the upper level
[3 >. Equations (2.38) can then be written as
A
21
N
2
+ A
31
N
3
= Nr (2.39)
(A
21
+ B
32
W)N
2
= (A
32
+ B
32
W)N
3
These equations can easily be solved for N
2
and N
3
, leading to
N
2
=
Nr(A
32
+ B
32
W)
A
21
(A
31
+ A
32
) + B
32
W(A
21
+ A
31
)
(2.40)
and
N
3
=
Nr(A
21
+ B
32
W)
A
21
(A
31
+ A
32
) + B
32
W(A
21
+ A
31
)
(2.41)
These expressions show that it is possible to obtain an inversion of population N
3
> N
2
,
provided that the following condition is fullled
A
21
> A
32
(2.42)
or equivalently, if the atoms at the state [2 > decay more rapidly into the lower level
state [1 > than the atoms at the upper level state [3 > decay into this level intermediate
level [2 >. When this inequality is satised it becomes possible, at least in principle, to
19
amplify radiation at the frequency
32
= (E
3
E
2
)/, which means, increase the energy
density W(), for =
32
. But equations (2.40)-(2.41) also show that, for an increasing
value of W() the dierence (N
3
N
2
) tends to decrease and eventually saturate.
Based on this three-level optical pumping mechanism, or any other alternative mech-
anism for inversion of population, it is then possible to built up and optical amplier, or
a laser (light amplication of stimulated emission of radiation). But, in order to produce
the laser eect, we need to add to the amplifying medium that we have just described, a
feedback process provided by an optical cavity, as described next.
2.4 Laser cavity
The laser is an oscillator or amplier in the optical frequency range, where the ampli-
fying medium (for instance, a gas or a solid, with an inversion of population between
two quantum level) is connected to a feedback system allowing the electromagnetic sig-
nal to repeatedly cross the amplifying region. Shalow and Townes, and independently
Prokhorov, in 1958, were able to demonstrate that a Fabry-Perot cavity could be used to
provide the feedback process. The rst laser (now known as the Rubi laser) was built and
operated by Maiman in 1960.
Let us consider the interferometer dened by two plane and parallel mirrors, separated
by a distance L (see gure). Assuming that the mirrors have a reection coecient close
to unit, we know that dierent standing-wave modes of the electromagnetic radiation
can be excited inside this system. these modes are such that an integer number q of
half-wavelengths can t the cavity length L, or
q
= 2L/q, where
q
is the wavelength of
a given standing-wave mode. The frequency interval between two consecutive modes is
then given by
=
2c

q1
=
c
L
(2.43)
This frequency interval is therefore equal to the inverse of the time spent by light to travel
inside the cavity from one mirror to the other, and than back to the same mirror. For a
20
L
(a)
(b)
R =L
Figure 2.4: Optical cavity with parallel plane mirrors (a), and spherical mirrors (b).
cavity of 1 meter this corresponds to a frequency interval of 150 MHz. The interval is
in general, much shorter than the natural bandwidth of the resonant radiative transition
where the inversion of population takes place. This means that it is possible to excite
simultaneously various cavity modes, which will be discussed later.
Another important question is related with the width of the cavity resonance,
q
,
which is inherent to each of the cavity modes
q
= 2c/
q
. Such a width can be dened
in terms of the quality factor of the cavity
Q
q
=

q

(2.44)
This quantity is determined by 2 times the ration between the energy contents and the
power losses of electromagnetic radiation kept inside the cavity. These power losses result
essentially from: i) reection losses due to the imperfection of the mirrors; ii) diraction
losses at the mirrors leading to energy leaking outside. It should also be noticed that, in a
laser, the reection losses can never be reduced to zero, because one of the cavity mirrors
has to be partially transparent in order to allow the passage of laser radiation. On the
21
other hand, the diraction losses depend strongly on the mirror conguration, and can
be reduced if, instead of plane mirrors we use a confocal cavity conguration, made of
two spherical mirrors of equal radius of curvature R, place at a distance L = R from each
other. In this case the focal points of the two mirrors coincide.
Finally, in order avoid mode competition inside the laser cavity, we have to reduce the
number of modes to a minimum. The ideal case of single mode operation is attained when

trans
, where
trans
determines the natural width of the atomic transition. This
implies a laser cavity length L of the order of c/
trans
.
2.5 Simplied Laser model
Let us assume that inversion of population between two atom levels is obtained by some
unspecied process, and let us study the photon rate equation for a given eld cavity
mode, coupled with the evolution equation for the population dierence (N
2
N
1
). In
the photon rate equation (2.29), we neglect the spontaneous emission term AN
2
, which
is irrelevant for the laser eect. On the other hand, we add to this equation a damping
term, describing the photon losses in the cavity. We can then write
d
dt
n = (N
2
N
1
)w n
c
n (2.45)
where
c
is the cavity damping rate. On the other hand, we can write the evolution
equation for the atom population in the upper energy level as
dN
2
dt
= w
21
N
1
w
12
N
2
(N
2
N
1
)wn (2.46)
Comparing this with equation (2.20), we can see that an extra term w
21
N
1
introduced in
order to account for any possible inversion of population mechanism, pumping the atoms
to the upper energy level. On the other hand, in the decay term w
12
N
2
we can also
include not the spontaneous decay which will not contribute to the specic cavity mode
under discussion here, but other possible de-excitation processes caused, for instance, to
atomic collisions.
22
A similar equation can be established for the population of the lower energy level, and
can be written as
dN
1
dt
= w
12
N
2
w
21
N
1
+ (N
2
N
1
)wn (2.47)
It can be seen that
d
dt
(N
1
+ N
2
) = 0 (2.48)
which is compatible with a two level atom model, where the total number of atoms
N = N
1
+ N
2
is constant. By subtracting equations (2.46) and (2.47) we can also easily
establish the evolution equation for the population dierence D = N
2
N
1
, as given by
dD
dt
= N(w
21
w
12
) D(w
21
+ w
12
) 2wDn (2.49)
Noting that w
21
and w
12
are transition rates, and consequently are proportional to the
inverse of a transition time, we can introduce a time constant , such that
=
1
(w
21
+ w
12
)
(2.50)
On the other hand, we notice that, in the absence of laser radiation (n = 0), equation
(2.49) predicts an equilibrium value for the inversion of population D = D
0
, where
D
0
=
(w
21
w
12
)N
(w
21
+ w
12
)
= (w
21
w
12
)N (2.51)
This quantity is called the unsaturated inversion for reasons that will become apparent
below. Using the denitions of and D, we can write equation (2.49) in the form
dD
dt
=
1

(D
0
D) 2wDn (2.52)
This equation can be coupled with equation (2.45), written in the form
d
dt
n = Dw n
c
n (2.53)
They form the basic equations for phenomenological laser model. These are nonlinear
coupled equations which will be discussed next. Let us rst consider the steady state
solutions. From equation (2.52), we get steady state inversion of population D =

D, with

D =
D
0
1 + 2wn
(2.54)
23
This quantity diers from the value D
0
because of the saturation associated with the
second term in the denominator, which is proportional to the laser intensity. Saturation
therefore leads to a decrease in the steady state inversion. On the other hand, from
equation (2.53), we get two possible steady solutions, one trivial n = 0, and the other
given by

Dw =
D
0
w
1 + 2wn
=
c
(2.55)
which gives n = n
0
with
n
0
=
(D
0
w
c
)
2
c
w
(2.56)
We can see from this result that a suciently high value of the pumping process is needed,
w
21
> w
12
, in order to satisfy the physical condition n
0
> 0. This means that the laser
eect will imply that
D
0
w
c
(2.57)
Below the critical value for the inversion of population D
0crit
=
c
/w, no laser eect can
ever occur. This threshold laser condition is however quite misleading, because it implies
that n
0
= 0, a condition that can only be fullled in the absence of spontaneous emission,
as will be shown later. Only by using the quantum theory of the laser can we establish a
rigorous account of the laser threshold condition.
Let us now study the approximate time-dependent solutions for the above laser equa-
tions. As a simplifying assumption, we assume that the inversion of population D only
evolves on a time scale much longer than the transition time rate . This allows us to
neglect the term dD/dt 0 in equation (2.52), which leads to the approximate solution
D(t)
D
0
1 + 2wn(t)
(2.58)
Replacing this in equation (2.53), we obtain a closed equation for the laser photon number,
as given by
dn
dt
=
D
0
wn
1 + 2wn

c
n (2.59)
By expanding the denominator, we arrive at a simple nonlinear equation of the form
dn
dt
= an bn
2
(2.60)
24
0 2 4 6 8 10
0
20
40
60
80
100
n(t)/n(0)
at
Figure 2.5: Photon number evolution, as given by equation(2.62). The value of bn(0)/a =
0.01 was used.
with
a = D
0
w
c
, b = 2D
0
w
2
(2.61)
This shows that the photon number n(t) initially grows, for a > 0, as n(t) n(0) exp(at).
But, at later times, the nonlinear term slows down the exponential growth and saturates
the photon number at the value n
0
= a/b. A more exact solution of the nonlinear equation
(2.59) and describing such a saturation process, is given by
n(t) =
a n(0) exp(at)
a + b n(0)[exp(at) 1]
(2.62)
2.6 Relaxation oscillations
Let us now assume a small perturbation around the steady state solutions. This can be
done by writing
D =

D + , n = n
0
+ (2.63)
25
where

D and n
0
were determined above, and where we assume that

D and n
0
.
Replacing this in equations (2.53), and noting that n
0
is a constant, we get
d
dt
= w(

D + )(n
0
+ )
c
(n
0
+ ) (2.64)
But, by denition, the constants

D and n
0
are such that, for D =

D and n0n
0
, we have
dn/dt = 0. This leads us to conclude, from the evolution equation (2.53), that

Dw
c
= 0 (2.65)
This allows us to simplify equation (2.64), leading to
d
dt
= w(n
0
+ ) (2.66)
We can now linearize this equation, by noting that and are two very small quantities.
As a result, their product is a second order quantity which can be neglected in a simplied
model. WE are then reduced to
d
dt
= wn
0
(2.67)
We can apply a similar perturbation procedure to equation (2.52), writing it as
d
dt
=
1

(D
0


D ) 2w(

D + )(n
0
+ ) (2.68)
Noting that the equilibrium quantities satisfy the relation
1

(D
0


D) 2w

Dn
0
(2.69)
Neglecting the nonlinear term containing , we can then write equation (2.68) as
d
dt
=

2w

D (2.70)
The linearized coupled equations (2.67) and (2.70) can be solved for the perturbations
and of the equilibrium laser solutions. By elimination of we can easily get an evolution
for of the form
d
2

dt
2
+
2
0
+
d
dt
= 0 (2.71)
26
where we have used the denitions

0
= w
_
2n
0

D , = 1/ (2.72)
This is a damped oscillator, which has well known solutions of the form
(t) = Aexp(t/2) cos(t + ) (2.73)
The constants of integration A and are the initial amplitude and phase, and the oscillator
frequency is dened by
=
_

2
0

2
/4 (2.74)
We can conclude from here that, near the laser steady state solution

D and n
0
, the per-
turbed population dierence D(t) oscillates around its equilibrium value

D with frequency
, and tends to this value on a time scale of order = 1/.
Similarly, the laser intensity, or the corresponding photon number n(t), oscillated
around n
0
with the same frequency, as determined by
n(t) = n
0
_
1 +
w

sin(t + )
_
(2.75)
These are called the relaxation oscillations of the laser system, showing that any small
perturbation around the equilibrium will oscillate and tend to zero on a time scale of
. Such oscillations can be observed for steady state laser operation. For instance, for
a Ruby laser, we typically have 2/ 10 s, and 1 ms. Other modes odf laser
operation will be considered next.
2.7 Short Laser Pulses
2.7.1 Q-switching
One particular way of achieving short laser pulse operation is to rotate one of the mirrors
of the optical cavity. This is called a Q-switching process, because only for a very small
27
fraction of the rotation period will the cavity possess a large Q-factor. Or, in other words,
the cavity losses will only be negligible for a very small fraction of time when the two
cavity mirrors become parallel to each other. This means that the mirror rotation is a
practical way to switch o for most of the time the quality factor Q = /
c
of the laser
cavity. Here Q and
c
are time dependent quantities.
Because the stimulated radiative transitions are inhibited when the cavity mirrors are
not aligned, we can expect a sudden burst of laser radiation at the aligned position, with
much higher intensities than in the case of a steady-state laser. In the present situation,
the term in 1/ in the above inversion equation (2.52) can be neglected with respect to
the term dD/dt, and this equation is reduced to
dD
dt
2wDn (2.76)
On the other hand, in the photon number equation, we can replace the instantaneous
value of D(t) by its initial value D
i
, leading to
dn
dt
(D
i
w
c
) n (2.77)
which has the approximate solution
n(t) n
i
exp(t) , (D
i
w
c
) (2.78)
Replacing this in equation (2.76), and integrating, we obtain
D(t) D
i
_
1 +
2wn
i

_
1 e
t
_
_
(2.79)
These approximate solutions are only valid during the initial stage of the laser pulse,
when the number of photons grows exponentially at a constant rate , and the inversion
of population only suers a small temporal decay. At later times, however, this picture
has to be changed. When exp(t) becomes signicantly larger than 1, equation (2.79)
becomes
D(t) D
i
_
1
2wn
i

e
t
_
(2.80)
28
Laser active medium
Fixed
mirror
Rotating mirror
Figure 2.6: Scheme of an Q-switching laser. A rotating mirror provides the change in the
quality factor of the optical cavity
The inversion of population described by this new expression will decrease down to
zero, D 0, at a time t = t
1
, such that
n
max
= n
i
e
t
1


2w
(2.81)
In order to describe the evolution of the photon number n(t) for later times t > t
1
, we
have to go back to equation (2.53) with D 0. The resulting solution will then be
n(t) n
max
exp[
c
(t t
1
)] (2.82)
According to this new expression, the photon number will decay to zero within a time
scale determined by the cavity losses 1/
c
. These approximate solutions describe a very
strong peak of high laser intensity.
29
2.7.2 Mode locking
Another method of short laser pulse operation is the so-called mode locking, where the
laser oscillates in multi-mode regime. Record values of short pulse duration, of the order
of a few femtosecond have been attained using this method. The output electric eld can
written as
E(t) =
N

n=1
E
n
exp[i(
n
t +
n
)] (2.83)
where we have N laser modes with amplitudes E
n
, oscillating at the frequencies
n
,
with relative phase shifts determined by
n
. In the incoherent case, where the laser eld
modes are independent from each other and the phases
n
can be considered as randomly
distributed in the interval [0, 2], the total laser intensity is given by
I
tot
=

n
[E
n
[
2
N[E
0
[
2
(2.84)
where we assume that the mode amplitudes are similar, E
n
E
0
. However, the active
laser medium has nonlinear properties, which means that for a large enough amplitude
E
0
the various modes become coupled to each other. In this case, the eld phases tend
to lock into the same value
n

0
, and the total eld becomes
E(t) = E
0
e
i
0
N

n=1
exp(i
n
t) (2.85)
The frequency mode separation is determined by = c/L, where L is the cavity
length. We can then write the mode frequencies as
n
=
0
n. This leads to
E(t) = E
0
e
i(
0
t+
0
)
N1

n=0
exp(inct/L) (2.86)
Performing the geometric series summation, we can write it as
E(t) = E
0
e
i(
0
t+
0
)
sin(N/2)
sin(/2)
(2.87)
with = ct/L. The corresponding laser intensity is then given by
I(t) = [E
0
[
2
sin
2
(N/2)
sin
2
(/2)
(2.88)
30
This corresponds to a periodic laser pulse output with maximum intensity
I
max
= [E
0
[
2
N
2
= NI
tot
(2.89)
This means that, for N 1, short laser pulses are periodically emitted with very large
intensities and very short durations. Typical values are individual pulse durations of the
order of a few femtoseconds, emitted with a periodicity of a few picoseconds.
2.8 Amplied spontaneous emission
In the above description of the laser operation, in steady state or in ultra-sort pulses,
we have always neglected the spontaneous emission term because it plays a negligible
role in the laser output parameters. But it is important to notice that in any amplifying
medium, such that D > 0, spontaneous emission is always present and that it can also be
amplied, disturbing sometimes the output of a laser amplier.
In order to describe such a process, we go back to the photon balance equation and
include a spontaneous emission term. Neglecting the photon cavity losses, this equation
can be written as
dn
dt
= Dwn + AN
2
(2.90)
where N
2
is the number population of the upper quantum level of the amplifying medium.
This equation has to be modied in order to account for losses, noting that only a small
fraction of the spontaneously emitted photons will propagate inside the medium and will
be able to contribute to the amplication process. It is therefore useful to rewrite this
equation as
dn
dt
= Dwn + A
_

4
_
N
2
(2.91)
where is the angle solide, out of the total angle solid 4, which corresponds to prop-
agation in the forward direction inside the amplifying medium. This quantity can be
estimated as = S/L
2
, where S is the area of the section of the amplier and L is its
lemgth.
31
Assuming that g = WD, and that N
2
stays independent of n, which is valid for low
intensities, or small photon numbers, we can integrate equation (2.91) and get
n(t) =
_
A
4
_
N
2
g
_
e
gt
1
_
(2.92)
where we have assumed that n(0) = 0. For suciently long times, such that gt 1, the
amplied spontaneous emission grows exponentially as exp(gt).
If the lower level quantum state population stays very low during the amplication
process, such that N
2
N
1
, we can approximately write (in the unit volume V = 1)
N
2
g

N
2
w(N
2
N
1
)

1
w
=

2

2
c
3
A
(2.93)
We can then write the approximate formula
n(t)

2

4
3
c
3
e
gt
(2.94)
Noting that = 2c/
2
, we can rewrite this expression in terms of the wavelength,
as
n(t) =

4

8
4
c
4
e
gt
= 2

4
e
gt
(2.95)
Finally, it is common to refer to the intensity I = cn, which lead to
I(t) = 2
hc
2

5
e
gt
(2.96)
This gives a rough estimate of the intensity associated with the amplied spontaneous
emission, as expected at the output of a laser amplier. It should be noticed that, although
this amplied light is also responsible for depopulation of the upper quantum level of the
amplifying medium, it is random in phase and polarization, and cannot contribute to the
nal level of the output laser signal. Therefore, the temporal coherence of this amplied
spontaneous emission spectrum is essentially dierent from that of laser radiation.
2.9 Problems
P2.1 - Determine the maximum frequency emitted by a blackbody at a given tem-
perature T. Determine this frequency value for three typical cases: i) the microwave
32
background radiation, T = 3
0
K; ii) room temperature, T = 20
0
C; iii) the interior of a
fusion reactor, T = 10 KeV .
P2.2 - Show that the general solution of equations (2.31)-(2.32) is given by the ex-
pression
N
1
(t) =
_
N
1
(0) N
1 + f()
1 + 2f()
_
exp A[1 + 2f()]t + N
1 + f()
1 + 2f()
(2.97)
where f() = BW()/A. Show that, under certain conditions, equation (2.33) can be
obtained from here.
P2.3 - Solve the nonlinear equation for the number of photons (2.60), and show that
it can be written as (2.62). Solve numerically equation (2.59) and compare the results
with the analytical approximate solution.
P2.4 - Establish a numerical model for Q-switching, and compare it with the approx-
imate solutions given by equations given in the text.
P2.5 - Establish a relation between the natural bandwidth of the radiative transition,
and the number of modes that can be excited in a given laser cavity. Determine plausible
conditions for the excitation of 100 laser modes, and determine the expected intensity for
mode locking.
Chapter 3
Quantum theory of Radiation
3.1 Field Quantization
We start from Maxwells equations, for the electric


E =

B
t
,

D = (3.1)
and magnetic elds


H =

J +

D
t
,

B = 0 (3.2)
where, for elds produced by charges and currents in vacuum, we have

D =
0

E ,

B =
0

H (3.3)
and the charge and current densities satisfy the continuity equation

t
+

J = 0 (3.4)
It is also useful to introduce the scalar and vector potentials, V and

A, which are related
to the elds by the equations

B =

A ,

E = V


A
t
(3.5)
33
34
It is known that Maxwells equations are gauge invariant. Here it is convenient to choose
the Coulomb gauge, which is determined by the condition


A = 0 (3.6)
Using this gauge, we can derive, from Maxwells equations the following wave equation
for the vector potential
_

1
c
2

2
t
2
_

A =
0

J +
_
1
c
2
V
t
_
(3.7)
On the other hand, the scalar potential satises, in this gauge, the Poisson equation

2
V =
1

0
(3.8)
At this point, it is useful to introduce the transverse and the longitudinal currents

J =

J
T
+

J
L
(3.9)
such that


J
T
= 0 ,

J
L
= 0 (3.10)
This decomposition is generally valid, because it is known from vector anaysis (Helmoltz
theorem) that any vector eld can be dened as a sum of of two components, one with
zero divergence, and the other with zero rotational. From the continuity equation (3.4),
we can then easily get


J
L
=

t
(3.11)
Taking the time derivative of the Poisson equation (3.8), and comparing it with this
equation, we can then conclude that

J
L
=
0

V
t
(3.12)
Relacing equations (3.9) and (3.12) in the potential equation (3.7), we can see that the
longitudinal current cancels out the scalar potential term, and the resulting equation is
_

1
c
2

2
t
2
_

A =
0

J
T
(3.13)
35
This shows that the vector potential is only coupled to the transverse current. Let us rst
consider the eld solutions in the absence of sources, which means that we take

J
T
= 0 (3.14)
We then consider a cubic cavity with side L and volume V = L
3
. In this case, we can
expand the vector potential in terms of a Fourier series of the form

A(r, t) =

k
_

A
k
(t)e
i

kr
+

A

k
(t)e
i

kr
_
(3.15)
where the components of the wavevector

k, take for each eld mode the only possible
values compatible with the boundary conditions
k
j
=
2
L

j
, (j = x, y, z) (3.16)
with
j
= 0, 1, 2, ... The change from a discrete sum of eld modes to the integral over
a continuum can be made by taking the limit L . The sum in equation (3.15) will
then be replaced by an integral, according to the transformation

k
2
_
L
2
_
3
_
d

k (3.17)
where additional the factor of 2 accounts for the two possible polarization states of the
electromagnetic eld. Using equation (3.15), we can see that the Coulomb gauge condition
(3.6) is satised by

k

A
k
(t) =

k

A

k
(t) = 0 (3.18)
This shows that the elds are purely transverse.. Replacement of equation (3.15) in the
wave equation (3.13)-(3.14) also shows that each eld mode amplitude has to satisfy the
following equation

2
t
2

A
k
(t) +
2
k

A
k
(t) = 0 (3.19)
where the mode frequency is dened by
k
= kc. The solution for the time dependent
amplitude is then

A
k
(t) =

A
k
exp(i
k
t) (3.20)
36
Returning to equation (3.15),we obtain the complete vector potential solution as

A(r, t) =

k
_

A
k
exp(i

k r i
k
t) + c.c.
_
(3.21)
Using equations (3.5), we can then get the corresponding electric and magnetic eld
solutions as

E(r, t) =

k
_
i

A
k
exp(i

k r i
k
t) + c.c.
_
(3.22)
and

B(r, t) =

k
_
i

A
k
exp(i

k r i
k
t) + c.c.
_
(3.23)
Let us now consider the energy of the eld inside the cavity, as determined by
W =
1
2
_
V
(
0
E
2
+
0
H
2
)dr (3.24)
Using the above mode expansion, we can easily arrive at the following time averaged value
for the electromagnetic energy per mode, as

W
k
= 2
0
V
2
k
(

A
k

k
) (3.25)
where V = L
3
is the cavity volume.
It is now useful to consider position-like and momentum-like variables, Q
k
and P
k
,
such that the vector potential in equation (3.21) can be written as

A
k
=
1
_
4
0
V
2
k
(
k
Q
k
+ iP
k
)e
k
(3.26)
where e
k
is the unit polarization vector, and the quantities Q
k
and P
k
are assumed real.
replacing this in the expression of the average energy per mode (3.25), we get

W
k
=
1
2
(P
2
k
+
2
k
Q
2
k
) (3.27)
This is formally identical to the energy of a classical harmonic oscillator. We can then say
that the electromagnetic eld inside an empty cavity is formally identical to an innite
sum of harmonic oscillators.
The electromagnetic eld can then be quantized in the same way as the harmonic
oscillator, by replacing the quantities Q
k
and P
k
by the position and momentum operators,
q
k
and p
k
, of the corresponding quantized harmonic oscillator.
37
3.2 Fock States
Let us rst consider a single mode of the eld with frequency . For simplicity, we drop
the index k in all the quantities. The corresponding Hamiltonian operator is
H =
1
2
(p
2
+
2
q
2
) (3.28)
We know that this a hermitic operator H = H

, and that q and p obey the commutation


relation [q, p] = i. we now dene a pair of new operators, a and a

, such that
a =
1

2
(q + ip) , a

=
1

2
(q ip) (3.29)
This is equivalent to write
q =
_

2
(a + a

) , p = i
_

2
(a a

) (3.30)
The operators a and a

are called the destruction and creation operators of the given


oscillator eld mode. Using their denitions, we can then calculate
a

a =
1
2
(2H + i[q, p]) =
1

_
H
1
2

_
(3.31)
and
aa

=
1
2
(2H i[q, p]) =
1

_
H +
1
2

_
(3.32)
By adding these two expressions, we obtain
H =
1
2
(a

a + aa

) (3.33)
And, by taking the dierence, we get the commutator
[a, a

] = (aa

a) = 1 (3.34)
From this we can extract new expressions for the Hamiltonian operator
H =
_
a

a +
1
2
_
=
_
aa

1
2
_
(3.35)
38
It is also sometimes useful to use the photon number operator
N = a

a (3.36)
Let us now consider the energy eigenstate [n >, corresponding to the eigenvalue energy
E
n
of the electromagnetic eld mode. We can write the eigenvalue equation as
H [n >=
_
a

a +
1
2
_
[n >= E
n
[n > (3.37)
If we apply the operator a to the left of this equation, we get

_
aa

a +
1
2
a
_
[n >= E
n
a[n > (3.38)
Using the commutation relation (3.34), this can also be written as

_
a

a + 1 +
1
2
_
a[n >= E
n
a[n > (3.39)
which can aso be written as
H a[n >= (E
n
)a[n > (3.40)
If we dene
E
n1
= E
n
(3.41)
this shows that the state vector (a[n >) is equal to the eigenvector of the energy eigenstate
E
n1
, apart from a normalization constant. We can then write
a[n >= c
n
[n 1 > (3.42)
which allows us to state that
H [n 1 >= E
n1
[n 1 > (3.43)
Now, let [0 > be the ground state of the oscillator, with energy E
0
: from equation (3.40),
we have
H a[0 >= (E
0
)a[0 > (3.44)
39
But, by denition, E
0
is the lowest possible energy of the oscillator, which means that we
have to assume that
a[0 >= 0 (3.45)
Going back to equation (3.37), we conclude that
H [0 >=
1
2
[0 >= E
0
[0 > (3.46)
And, from the above condition (3.41), we can nally establish the value of the energy
states of the eld mode oscillator
E
n
=
_
n +
1
2
_
(3.47)
Comparing this with equation (3.37), we can also obtain the eigenvalues of the photon
number operator N, as
N[n > a

a[n >= n[n > (3.48)


In order to complete our discussion of the eld mode oscillator, we only need to determine
the normalization constant introduced in equation (3.42). Multiplying this equation by
its Hermitian conjugate, we get
< n[a

a[n >=< n 1[c

n
c
n
[n 1 > (3.49)
The normalization condition for both [n > and [n 1 > then implies that [c
n
[
2
= n. We
can then write equation (3.42) as
a[n >=

n[n 1 > (3.50)


A similar analysis also leads to
a

[n >=

n + 1[n + 1 > (3.51)


This shows that, by applying the operator a

successively n times to the ground state


vector [0 >, we can generate the state [n >, apart from a multiplying normalization
factor. The result is
[n >=
1

n!
_
a

_
n
[0 > (3.52)
40
It can also be easily realized that the oscillator eigenvectors obey the orthogonality relation
< n

[n

>=
n

n
(3.53)
These eigenvectors can therefore be used as a bases for the representation of an arbitrary
state [ > of the oscillator, given that they also obey the completeness condition

n=0
[n >< n[ = I (3.54)
where I is the identity operator. Noting that each photon mode with frequency
k
= kc
and wavevector

k is equivalent to a simple harmonic oscillator, we can call n

k
the number
of photons existing in this eld mode, and N

k
= a

k
a

k
the corresponding mode number
operator. Furthermore, the destruction and creation photon operators can be dened by
the expressions
a

k
[n

k
>=

k
[n

k
1 > , a

k
[n

k
>=
_
n

k
+ 1[n

k
+ 1 > (3.55)
The state of the total electromagnetic eld can be determined by the product of the states
of the individual oscillator modes, as
[n

k
1
, n

k
2
, ....n

k
j
, .... >= [n

k
1
> [n

k
2
> ...[n

k
j
> .... (3.56)
or, in a more compact notation
[n

k
>=

j
[n

k
j
> (3.57)
Finally, the eld operators can be determined by replacing the vector potential amplitudes

k
by the corresponding operators, following equations (3.26) and (3.30), as

k
=
_

2
0
V
k
a

k
e

k
(3.58)
The total vector potential operator for the mode

k will be obtained from equations (3.58)
and (3.21), where the conplexe conjugate is replaced by the Hermitian conjugate, as shown
A

k
=
_

2
0
V
k
_
a
k
exp(i

k r i
k
t) + h.c.
_
(3.59)
41
with the total vector potential operator given by

A =

k
A

k
e

k
(3.60)
where the unit vector is perpendicular to the wavevector, e

k = 0. Similarly, the
corresponding electric and magnetic eld operators

E

k
and

B

k
will be given by

k
=
_

k
2
0
V
_
ia

k
exp(i

k r i
k
t) + h.c.
_
e

k
(3.61)
and

k
=
_

2
0
V
k
(

k e

k
)
_
ia

k
exp(i

k r i
k
t) + h.c.
_
e

k
(3.62)
It is sometimes useful to consider the two possible polarization states of the eld, and
replace the operators a

k
and a

k
appearing in these expressions by a

k,
and a

k,
, where
takes two dierent values corresponding to the two orthogonal polarization states. The
commutation relations, valid for these operators, are
[a

k,
, a

] = (

)
,
(3.63)
and
[a

k,
, a

] = [a

k,
, a

] = 0 (3.64)
3.3 Phase Operator
In the Fock representatiom, or number state representation discussed above, the number
of photons per eld mode is completely specied, but no information is given in what
concerns the eld phase. Here we consider a phase operator exp(i), which is dened by
the expression
a =

N + 1 e
i
(3.65)
42
where N a

a is the number operator. This allows us to split the destruction operator


a into its amplitude and phase components. Taking the Hermitian conjugate, we obtain
an alternative denition of the phase operator
a

= e
i

N + 1 (3.66)
Multiplying this equation by a, and using (3.65), we can easily get
aa

N + 1 e
i
e
i

N + 1 (3.67)
But we know from the commutation relation [a, a

] = 1 that
aa

= N + 1 (3.68)
Comparing this with equation (3.67), we conclude that
e
i
e
i
= I (3.69)
On the other hand, by using equations (3.65) and (3.50), we can state that
a[n >

N + 1 e
i
[n >=

n[n 1 > (3.70)


which leads to
e
i
[n >=

N + 1
[n 1 > (3.71)
Similarly, we can derive
e
i
[n >=
a

N + 1
[n >= [n + 1 > (3.72)
This means that, in the Fock state representation [n >, the phase operators exp(i)
and exp(i) have nonvanishing matrix elements determined by
< n 1[e
i
[n >= 1 , < n + 1[e
i
[n >= 1 (3.73)
Or, in a more compact form
< n[e
i
[n 1 >= 1 (3.74)
43
Instead of these exponential phase operators, we can also use sine and cosine operators,
dened by
cos =
1
2
_
e
i
+ e
i
_
, sin =
1
2i
_
e
i
e
i
_
(3.75)
It can easily be recognized by using equations (3.73) or (3.74) that the non-vanishing
matrix elements of these new operators are determined by
< n[ cos [n + 1 >=
1
2
, < n[ sin [n + 1 >=
1
2i
(3.76)
Notice that these sin and cosine operators are Hermitian. It is now interesting to con-
sider the phase properties of the Fock states. If a given electromagnetic eld mode with
wavevector

k and polarization state is in a Fock state [n >, it means that its number of
photons is exactly determined, and no uncertainty can be associated with this number,
n = 0. On the other hand, equation (3.76) shows that
< n[ cos [n >=< n[ sin [n >= 0 (3.77)
This means that the eld phase is completely undetermined, and can equally take any
value in the interval (0, 2). Let us now consider the electric eld operator (3.62), and
use equations (3.65) and (3.66), we can determine its mean value, as well as its square
mean value. These quantities are determined by
< n[

E[n >= 0 (3.78)


and
< n[E
2
[n >=

0
V
_
n +
1
2
_
(3.79)
This shows that the square mean deviation of the electric eld is determined by
E =
_
< n[E
2
[n > < n[

E[n >
2
_
1/2
=
_

0
V
_
1/2
_
n +
1
2
_
1/2
(3.80)
We conclude that the eld of a given Fock state oscillates in a perfectly sinusoidal way, with
a perfectly dened oscillation amplitude. But that its phase is completely arbitrary, which
explains the existence of this uncertainty on the value of the instantaneous amplitude.
44
3.4 Field Hamiltonian
Let us examine here the various forms of the electromagnetic eld Hamiltonian operator,
in the absence and in the presence of radiation sources. In the absence of sources, the
Hamiltonian operator can be written as
H =
1
2
_
[
0
E
2
(r, t) +
0
H
2
(r, t)]dr =

k
_
a

k
a

k
+
1
2
_

k
(3.81)
Notice that here, the elds

E and

H are operators. If [ > is the state vector of the
electromagnetic eld, then this eld will evolve according to the Schroedinger equation,
as
i

t
[(t) >= H[(t) > (3.82)
This equation can be solved formally to give
[(t) >= U(t)[(0) > , U(t) = e
1
i
H
(3.83)
The operator U(t) introduced here is such that U

= U
1
, and satises the initial condition
U(0) = I. In this Schrodinger picture, the operators are time independent and the
evolution of the eld is described by the state vector [(t) >. In the case of the radiation
eld, it is more appealing to use the Heisenberg picture, where the state vector remains
constant, and the eld operators are time dependent. This picture is closer to our classical
intuition of the elds. Transformation into the Heisenberg picture can easily be performed
using the operator U(t). Let us consider an arbitrary operator A. Using the formal
solution (3.83), and the denition of mean value, we can write
< A >=< (t)[A[(t) >=< (0)[U

(t)AU(t)[(0) > (3.84)


This last expression can be seen as the mean value of A calculated in the Heisenberg pic-
ture, where now the state vector is time invariant [ > (0) >, and the time dependent
representation of the same operator is given by
A

(t) = U

(t)AU(t) (3.85)
45
By deriving this expression with respect to time, and considering the denition of U(t)
given by equation (3.83), we obtain
A

t
=
1
i
[A

, H] (3.86)
By making this generic operator A

equal to the destruction and creation eld mode


operators a

and a, and using the second expression in equation (3.81), we obtain

t
a

k
= i
k
a

k
,

t
a

k
= i
k
a

k
(3.87)
Similarly, by making A

qual to the electric and magnetic eld operators



E and

H, and
using the rst expression in (3.81), we obtain

E
t
= c
2


B (3.88)
and

B
t
=

E (3.89)
We recognize here Maxwelle equations in vacuum. This leads us to the conclusion that the
eld operators in the Heisenberg picture evolve according to equations that are formally
identical to the classical Maxwells equations, but where the classical eld vectors were
replaced by quantum eld operators.
Let us now examine the eld Hamiltonian in the presence of radiation sources. If we
have a source current, we have to add to the previous expression, an interaction term,
given by
H
int
=
_

J(r, t)

A(r, t)dr (3.90)
If we now write the vector potential operator

A(r, t) in terms of the creation and destruc-
tion operators, we get
H
int
=

k,
_

2
0
V
k
_
a

k,
_
e

k,


J(r, t)e
i

kr
dr + a

k,
_
e

k,


J(r, t)e
i

kr
dr
_
(3.91)
The evolution of the state of the eld can now be better described in the interaction
picture. This is an intermediate picture, between Schroedinger and Heisenberg, where the
46
unitary transformation U(t) excludes the interaction Hamiltonian. The time evolution of
the state vector is then simply determined by the interaction Hamiltonian H
int
, and is
described by the Schroedinger equation
i

t
[(t) >= H
int
[(t) > (3.92)
We then have the formal solution
[(t) >= e
1
i
H
int
[(0) > (3.93)
Using equation (3.91), we can write the exponential operator as
e
1
i
H
int
=

k,
e

H(

k,)
(3.94)
where the Hamiltonian operator per eld mode takes the form

H(

k, ) = (

k,
a

k,

k,
a

k,
) (3.95)
and we have dened the auxiliary quantities

k,
=
i

_

2
0
V
k
_
t
dt

_
dr [e

k,


J(r, t

)]e
i(

kri

k
t

)
(3.96)
By choosing the initial quantum state of the electromagnetic eld as the vacuum state,
we can then write the state vector solution (3.93) as
[(t) >=

k,
exp(

k,
a

k,

k,
a

k,
)[0 >

k,
(3.97)
The nal state of the radiation produced by a classical current

J is a superposition of
coherent states, which can be represented as [

k,
>, as dened by
[

k,
>=

k,
[

k,
> (3.98)
where each coherent state mode results from the following transformation of the initial
quantum vacuum
[

k,
>= exp(

k,
a

k,

k,
a

k,
)[0 >

k,
(3.99)
Notice that the quantities (

k,
and (

k,
are not operators, because they have been as
a integral over the classical current

J. In the following section we will discus the basic
properties of such eld states.
47
3.5 Coherent States
Let us consider the coherent state of a single mode, and write it simply as
[ >= exp(a

a)[0 > (3.100)


A coherent state can be dened as the eigenstate of the destruction operator a, with the
complex eigenvalue , as indicated by the eigenvalue equation
a[ >= [ > (3.101)
Let us develop the coherent state vector in the Fock states representation [n >. We
can write
[ >=

n
C
n
()[n > (3.102)
where C
n
() are development coecients to be specied. By applying the destruction
operator a to this expression, we get
a[ >=

n
C
n
()

n[n 1 > (3.103)


Replacing this in equation (3.101), we obtain the following relation between consecutive
coecients
C
n
() =

n
C
n1
() (3.104)
By recurrence, this leads to
C
n
() =

n

n!
C
0
() (3.105)
On the other hand, by normalizing the coherent state vector [ >, we get
< [ >= 1 ,

n
[C
n
()[
2
= 1 (3.106)
Using equation (3.105), we can then write
[C
0
()[
2

n
[[
2n
n!
= [C
0
()[
2
exp([[
2
) = 1 (3.107)
48
Or
C
0
() = exp([[
2
/2) (3.108)
Finally, replacing this in equations (3.105) and (3.102), we obtain
[ >= exp([[
2
/2)

n!
[n > (3.109)
This completes the denition of the coherent state in the Fock representation. Such a
denition guarantees that the coherent state is normalized < [ >= 1. But it also
implies that two dierent coherent states, [ > and [ >, are not orthogonal, as shown
by
< [ >= exp
_

[[
2
2

[[
2
2
_

n
n!
= exp
_

[[
2
2

[[
2
2
+
_
(3.110)
This leads to
[ < [ > [
2
= exp
_
[ [
2
_
(3.111)
This equation clearly shows that the coherent states are not orthogonal. However, the
coherent states form a complete set [ >, allowing us to represent an arbitrary eld
state in terms of coherent states. This will be shown later.
In order to understand the physical relevance of the coherent states, it is now useful to
consider the mean eld energy associated with a coherent state. From equation (3.109),
we can easily see that the probability for nding n photons in a given coherent state [ >
is given by
P
n
() [ < n[ > [
2
= [C
n
()[
2
= exp([[
2
)
[[
2n
n!
(3.112)
This is a Poisson distribution. It has the following probability recurrence relation
P
n
() =
[[
2
n
P
n1
() (3.113)
and it attains the maximum value when n is equal to the integer part of [[
2
. The mean
value of the eld energy can then be determined by
< H > () =

n
P
n
()
_
n +
1
2
_
(3.114)
49
On the other hand, we can calculate its value by using directly the denition of a mean
value, which gives
< H > () < [H[ >= < [
_
a

a +
1
2
_
[ > (3.115)
By taking the Hermitian conjugate of equation (3.101), we have
< [a

< [ (3.116)
which leads to
< [a

a[ >=

(3.117)
which allows us to write
< H > () =
_
< N > +
1
2
_
, < N >= [[
2
(3.118)
By comparing this with equation (3.114), we obtain the additional expression for the
mean photon number in a coherent state
< N >=

n
P
n
()n (3.119)
For [[ 1, equation (3.118) shows that < H > () is nearly equal to the energy
of maximum probability. Let us now consider the square mean value of the energy, as
dened by
< H
2
> () =
2

2
< [
_
a

a +
1
2
_
2
[ > (3.120)
Using equation (3.101), and the commutator [a, a

] = 1, we can easily obtain


< H
2
> () =
2

2
_
[[
4
+ 2[[
2
+
1
4
_
(3.121)
From here, we can deduce the energy mean deviation, as
H()
_
< H
2
> () < H >
2
() =

[[
4
+ 2[[
2
+
1
4

_
[[
2
+
1
2
_
2
= [[
(3.122)
50
We can then write the relative energy mean deviation of a coherent state as
H()
< H > ()

1
[[
(3.123)
Similarly, we could obtain the relative mean deviation of the photon number, as
_
n
n
_

1
[[
(3.124)
This gives us the uncertainty of the energy, and of the number of photons per mode, for a
coherent quantum state. We can see that, for [[ much larger than one, such uncertainties
become vanishingly small. At this point we could wrongly conclude that the coherent state
would approach a Fock state, in the limit of very large eigenvalues [[. This is not true
because, in the same limit, the uncertainty will also vanish, in contrast with the Fock
states, as shown next. It will then become apparent that, in such limit, the coherent state
will approach a classical state of the radiation eld.
Let us then consider the mean value of the phase operator. For the cosine operator,
we have
< [ cos [ >=
1
2
e
||
2

n+1
+

n+1
n!
_
(n + 1)
(3.125)
Introducing = [[ exp(i), we can also write this expression as
< [ cos [ >= [[ cos e
||
2

n
[[
2n
n!
_
(n + 1)
(3.126)
This shows that the expectation value for cos is proportional to cos , the cosine of the
phase of the eigenvalue . For a large mean photon number < N >= [[
2
1, this can
be approximately written as
< [ cos [ >
_
1
1
8[[
2
_
cos (3.127)
In a similar way, we could estimate the mean value of the square of the cosine operator.
In the limit of large mean photon numbers, the result is
< [ cos
2
[ >
_
1 +
1
4[[
2
_
cos
2
+
1
4[[
2
(3.128)
51
This allows us to estimate the phase uncertainty of the coherence state, in the limit of
large photon numbers, or [[
2
1, as
(cos ) =
_
< [ cos
2
[ > < [ cos [ >
2
=
sin
2[[
(3.129)
Using equations (3.124) and (3.129), we can establish an uncertainty principle relating
the photon number and the eld phase for a coherent state, as
n (cos ) =
1
2
sin (3.130)
Equations (3.124) and (3.129) also show us that, when [[ increases, we get a better
knowledge of both the mean photon number and the mean phase. In the limit [[ ,
we will have a complete knowledge of both quantities, which corresponds to the classical
limit.
In order to conclude our discussion on the basic properties of coherent states, we
consider now the mean value of electric eld operator, as determined by
< [E[ >= i
_

2
0
V
_
< [a[ > e
i(

krt)
< [a

[ > e
i(

krt)
_
(3.131)
Using equation (3.101) and its Hermitic conjugate < [a

=< [

, and using the nor-


malization condition < [ >= 1, we can easily reduce this expression to
< [E[ >= 2
_

2
0
V
[[ sin(

k r t + ) (3.132)
where is the phase of the complex eigenvalue . In a similar way, we could nd that
< [E
2
[ >=
_

2
0
V
_
_
1 + 4[[
2
sin
2
(

k r t + )
_
(3.133)
The root square mean deviation of the electric eld amplitude is therefore given by a
constant
E =
_

2
0
V
_
1/2
(3.134)
We conclude that, in the classical limit, the relative uncertainty tends to zero, E/ <
E > 1/[[, which corresponds to the classical eld case, where the electric eld ampli-
tude and phase are exactly determined. We nally notice that, when applied to the
52
position and momentum like operators p and q, it can be shown by similar calculations
that the coherent state minimizes the quantum uncertainty, in such a way that
p q =
_
< N > +
1
2
_
(3.135)
It will be equal to /2 for the ground state.
3.6 Displacement Operator
We show here that a coherent state can be considered as a displaced harmonic oscillator
state. In order to do this, we rst relate the coherent state [ > with the vacuum state
[0 >. We have seen that, any Fock or photon number state [n >, can be dened by a
operator that applies n times the creation operator a

on the vacuum state, as dened by


equation (3.52), repeated here for convenience
[n >=
_
a

_
n

n!
[0 > (3.136)
By replacing this in the denition of the coherent state, as given by equation (3.109), we
obtain
[ >= e
||
2
/2

n=0

n!
[n >= e
||
2
/2

n=0

n
_
a

_
n
n!
[0 >= e
||
2
/2
e
a

[0 > (3.137)
We also notice that
e

a
[0 >=
_
1

a +
(

a)
2
2
...
_
[0 >= [0 > (3.138)
This means that we can write equation (3.137) as
[ >= D()[0 > (3.139)
where we have introduced the operator
D() = e
||
2
/2
e
a

a
(3.140)
53
This is the so-called displacement operator, and can be written in a more compact and
appealing form by using the Baker-Hausdor theorem. It states that, if A and B are two
non-commuting operators satisfying the conditions
[A, [A, B]] = [B, [A, B]] = 0 (3.141)
them, the following equalities are true
e
A+B
= e
A
e
B
e
[A,B]/2
= e
B
e
A
e
+[A,B]/2
(3.142)
In order to demonstrate this important result, we start by introducing an auxiliary oper-
ator of the form
f() = e
A
e
B
(3.143)
where is a c-number and not an operator. Taking the derivative with respect to , we
get
df
d
= Af() + e
A
e
B
B (3.144)
By using the identity operator
I = e
A
e
A
(3.145)
we can rewrite the last term of equation (3.144) as
e
A
Be
B
= e
A
Be
A
e
A
e
B
= e
A
Be
A
f() (3.146)
This allows us to write equation (3.144) as
df
d
=
_
A + e
A
Be
A
_
f() (3.147)
In order to proceed further, let us consider a new operator, dened by
g() = e
A
Be
A
(3.148)
We note that g(0) = B. By expanding g() as a power series in , we have
g() = g(0) +
_
dg
d
_
0
+

2
2
_
d
2
g
d
2
_
0
+ ... (3.149)
54
We have
dg
d
= Ae
A
Be
A
e
A
Be
A
A = [A, g()] (3.150)
And, consequently
_
dg
d
_
0
= [A, B] (3.151)
Similarly, we obtain
_
d
2
g
d
2
_
0
=
_
A,
dg
d
_
0
= [A, [A, g()]]
0
= [A, [A, B]] (3.152)
and so on. Replacing this in the expansion (3.149), we arrive at the equality
e
A
Be
A
= B + [A, B] +

2
2
[A, [A, B]] + ... (3.153)
But, if we assume that the operators satisfy the conditions (3.141), this will reduce to
e
A
Be
A
= B + [A, B] (3.154)
Replacing this in equation (3.147), we get
df
d
= (A + B)f() + [A, B]f() (3.155)
But, from the same starting conditions (3.141), we also notice that (A + B) commutes
with the commutator [A, B]. This means that the dierential equation (3.155) can be
treated as a normal equation envolving commuting variables. Integrating this equation,
we get the obvious solution
f() = f(0) exp
_
(A + B) + [A, B]

2
2
_
(3.156)
Noting that f(0) I is the identity operator, this solution can also be written as
f() = exp
_
(A + B) + [A, B]

2
2
_
(3.157)
If we take = 1, and multiply from the left both sides of this equation by exp([A, B]/2),
we obtain
e
A
e
B
e
[A,B]/2
= e
A+B
(3.158)
55
This demonstrates the rst identity of the theorem (3.142). Obviously, because the oper-
ators A and B commute with their commutator [A, B], this identity can also be written
as
e
A+B
= e
[A,B]/2
e
A
e
B
(3.159)
The second identity in (3.143) could be demonstrated in a similar way. Let us now consider
the particular case where
A = a

, B =

a (3.160)
With this denitions we get [A, B] = [[
2
. Replacing this in the denition of the
displacement operator D(), as given by equation (3.140), we obtain
D() = exp(a

a) (3.161)
We also get
D

() = exp(

a a

) (3.162)
This shows that the displacement operator is unitary
D()D

() = e
a

a
e

aa

= D

()D() = I (3.163)
which also implies that D

() = D
1
(). Another important property of the displacement
operator D() is that, when applied to the destruction operator a, transforms it into a+,
therefore displacing the amplitude of a. This can be shown by starting from
D
1
()aD() = e

aa

ae
a

a
(3.164)
On the other hand, we have seen that, for any operators A and B, we can write the
expansion
e
A
Be
A
= B [A, B] +

2
2
[A, [A, B]] + ... (3.165)
In the particular case where A = a

and B = a, this leads to the commutator [A, B]


[a

, a] = 1, and [A, [A, B]] = 0, or


e
a

ae
a

= a + (3.166)
56
Replacing in equation (3.164), this leads to
D
1
()aD() = e

a
(a + )e

a
= (a + ) (3.167)
This demonstrates the displacement property. Similarly, we could also show that
D
1
()a

D() = (a

) (3.168)
Finally, we remind that, according to equation (3.140), the coherent state [ > is also the
result from the displacement of the vacuum state [0 >, due to the action of the operator
D().
3.7 Density Operator
At this point it is useful to introduce the density operator, and to apply it to the case
of a coherent state. Let us rst consider a pure Fock state [n >, corresponding to a well
dened number of photons in a given eld mode. We can dene the corresponding density
operator as
= [n >< n[ (3.169)
We can see from this denition that the density operator for a pure quantum state is
a projector operator. And, the sum of all the projector operators onto Fock states of a
given eld mode is the identity operator

n=0
[n >< n[ = I (3.170)
Furthermore, we also notice that, from the above dention

2
= [n >< n[n >< n[ = [n >< n[ = (3.171)
This property is only valid for a pure state. We also have < n[[n

>=
nn
, which
indicates that the density matrix in the Fock state representation is diagonal. On the
57
other hand, for any operator A, we can determined the corresponding mean value by
using the trace
< A >= Tr.(A) =< n[A[n > (3.172)
For a pure coherent state [ >, we can dene a similar density operator as
() = [ >< [ (3.173)
Using the normalization condition for the coherent state, < [ >= 1, we can see that
this operator is also normalized
< () >=< [()[ >= 1 (3.174)
When we represent this operator in the basis of Fock states, we have
() =

n,m

nm
[m >< n[ (3.175)
Using equation (3.137), we obtain for the density matrix elements, the following result

nm
=< n[()[m >=< n[ >< [m >= e
||
2
n
(

)
m

n!m!
(3.176)
We can see from here that the o-diagonal terms of the density matrix are not zero. These
o-diagonal terms play an important role in the calculation of the mean value of electric
eld for a coherent mode. Using the mean value denition, we have
< E >= Tr.(E) =

n
< n[()E[n > (3.177)
Using equation (3.176), we realize that only the terms associated with m = n 1 will
survive in this sum. This is due to the existence of the creation and destruction operators
a

and a in the denition of the operator electric eld E. The result takes the form
< E >= i
_

2
0
V

n
_
< n[()[n 1 >

ne
i(

krt)
< n[()[n + 1 >

n + 1e
i(

krt)
_
(3.178)
Replacing the index n by n + 1 in the rst term of this expression, we can rewrite more
simply as
< E >=
_

2
0
V

n + 1
_
i < n + 1[()[n > e
i(

krt)
+ h.c.
_
(3.179)
58
We can see that the possible existence of a non-zero mean value < E >,= 0 implies the
non-vanishing of the o diagonal term < n + 1[()[n >. Using the decomposition into
real amplitude and phase of the complex eigenvalue = [[ exp(i), we can see from
equation (3.176) that these o diagonal terms are
< n + 1[()[n >= e
||
2
+i
[[
2n+1
n!

n + 1
(3.180)
Let us now consider the completeness of the coherent states. We have seen that two
dierent coherent states are not orthogonal, as shown by equation (3.111). However, the
coherent states form a complete set, allowing us to represent an arbitrary quantum eld
state in terms of coherent states. This can be shown by considering the following integral
_
(

)
n

m
e
||
2
d
2
(3.181)
By using the decomposition = [[ exp(i), we can write d
2
= [[d[[d, which allows
us to write this integral as
_

0
d[[[[
n+m+1
e
||
2
_
2
0
de
i(mn)
= n!
nm
(3.182)
Using this result and the Fock representation of the coherent states, as given equation
(3.137), we can then nd that
_
[ >< [d
2
=

m
[n >< m[
_
(

)
n

n!m!
e
||
2
d =

n
[n >< n[ (3.183)
We know that the Fock states form a complete and orthonormal set of state vectors, as
shown by equation (3.170). This allow us to write that
_
[ >< [d = I (3.184)
where I is the identity operator. This is equivalent to say that the set of coherent states
is complete. This set can even be called overcomplete, by noting from equation (3.111)
that two distinct coherent states [ > and [ > are nearly orthogonal when the dierence
[[ is much larger than one. If we represent a given coherent state [ > in the coherent
state basis [ >, we can use equation (3.184) and write
[ >=
1

_
[ >< [ > d (3.185)
59
from which we obtain the expansion
[ >=
_
C

()[ > d (3.186)


where the expansion coecients are
C

() = exp
_

[[
2
2

[[
2
2
+

_
(3.187)
This demonstrates the overcompleness property of the coherent states.
Let us now return to the density operator, and consider the more general case of a
statistical mixture of Fock sates. This corresponds to a situation where the radiation
eld is not in a well dened quantum state but where we can say that there is a nite
probability P
n
for the radiation eld to be in a given Fock state [n >. The simple denition
of equation (3.169) has to be replaced by
=

n
P
n
[n >< n[ (3.188)
Obviously, these probabilities have to obey the conditions

n
P
n
= 1 , P
n
0 (3.189)
The corresponding mean value for an arbitrary operator A is now given by
< A >= Tr.(A) =

n
P
n
< n[A[n > (3.190)
If the radiation eld is in thermal equilibrium with a background environment at a tem-
perature T, we know that the probability for nding a photon state [n > with energy
E
n
= (n + 1/2) is determined by equation (2.11), or
P
n
() =
_
1 e
/k
B
T
_
e
n/k
B
T
(3.191)
Noting that the photon number operator N a

a has the eigenvalue n for the Fock state


[n >, we can write for thermal radiation
=
_
1 e
/k
B
T
_

n
e
a

a/k
B
T
(3.192)
60
In more general situations, the statistical mixture is dened by specifying the probability
P

for the eld mode to be in a given quantum state [ >. The corresponding density
operator will be
=

[ >< [ (3.193)
where the sum is taken over all the possible quantum states for the radiation eld. The
mean value of an arbitrary operator A will become
< A >=

< [A[ >= Tr.(A) =

n
< n[A[n > (3.194)
Expanding in terms of the Fock states, we have, as before
=

n,m

nm
[n >< m[ (3.195)
In alternative to the Fock representation, we can use the expansion in terms of coherent
states, using the closure property (3.184). we then have
=
_ _

[ >< [dd (3.196)


where the matrix elements are determined by

=
1

2
< [[ > (3.197)
At this point, we can introduce the so-called R-representation of the density operator,
dened by
R(

, ) =< [[ > exp


_
[[
2
2
+
[[
2
2
_
(3.198)
Replacing this denition in equation (3.199), we obtain
=
1

2
_ _
R(

, )e
(||
2
+||
2
)/2
[ >< [dd (3.199)
Other possible representation of the quantum eld states will be discussed next
61
3.8 Other eld representations
3.8.1 P-representation
We have seen previously that the coherent states satisfy the orthonomality and closure
properties dened by
< [ >= e
||
2
,
1

_
[ >< [d = I (3.200)
where d d'()d() represent in fact a double integration. This suggests the use of
a dierent representation of the density operator in the basis of the coherent states as
=
_
P(,

)[ >< [d (3.201)
where the quantity P(,

), depending on both and

, represent the probability for


the quantum eld mode to be at a given coherent state [ >. Obviously, it has to satisfy
the normalization condition
_
P(,

)d = 1 (3.202)
Such a condition is necessary to verify the normalization of the density operator itself,
as determined by Tr. = 1. The quantity P(,

) is called the P-representation of the


density operator. It is particularly useful to determine the mean value of operators of the
form a
m
a
n
, as dened by
< a
m
a
n
>= Tr.(a
m
a
n
) (3.203)
Using the expansion (3.201), we have
< a
m
a
n
>=
_
P(,

)Tr.([ >< [a
m
a
n
)d (3.204)
Now, using the eigenvalue equations a[ >= [ >, and < [a

< [, we can write


< a
m
a
n
>=
_
P(,

) < [a
m
a
n
[ > d =
_
P(,

)
m

n
d (3.205)
This formula shows that P(,

) can be seen as the probability density for the possible


values of and

. But we have to take this conclusion carefully, because the same result
62
cannot is not valid for < a
n
a
m
>. We notice however that any operator involving a
and a

can be converted into sums of terms of the type a


m
a
n
, by using the commutator
[a, a

] = 1. As an example, we have a
2
a

= a

a
2
+ 2a. On the other hand, the quantity
P(,

) can eventually take negative values, which is not compatible with the concept
of probability. For this reason it is sometimes called a quasi-probability.
Let us now write the quasi-probability P(,

) in terms of the density operator .


For this purpose, we dene two coherent states [ > and [

>, corresponding to two


distinct eigenvalues of the destruction operator a. Using the P-representation for the
density operator, as given by equation (3.201), we can write
<

[[ >=
_
P(,

) <

[ >< [ > d (3.206)


On the other hand, we have already demonstrated that
< [ >= exp
_

[[
2
2

[[
2
2
+

_
(3.207)
This means that we can transform equation (3.206) into
<

[[ >= e
(||
2
+|

|
2
)/2
_
P(,

) e
||
2
e

d (3.208)
At this point, it is useful to consider the particular case of

= . We can write the


complex eigenvalues and in their explicit complex form as = x+iy, and = q +ip.
The exponential factor inside the interval will become

= 2i(xp yq) (3.209)


Using d = dxdy, we get
< [[ > e
||
2
=
_
f(x, y)e
2i(xpyq)
dxdy (3.210)
with
f(x, y) = P(,

)e
||
2
= P(x, y)e
(x
2
+y
2
)
(3.211)
We can see from here that right hand side of equation (3.210) is the two dimensional
Fourier transformation of the function f(x, y), apart from a factor of 4. We can then take
63
the inverse Fourier transformation, and write
f(x, y) = 4
_
< [[ > e
||
2 dqdp
(2)
2
(3.212)
Using equation (3.211), we can then write the explicit expression for the quasi-probability
P(,

), as
P(,

) = e
||
2
=
_
< [[ > e

d
(2)
2
(3.213)
If we apply this denition to a radiation eld which with probability one at the coherent
state [
0
>, we simply have
P(,

) = (
0
) (3.214)
which is a two dimensional delta function, given the complex nature of the eigenvalue
0
,
such that (
0
) (x x
0
)(y y
0
).
3.8.2 Q-representation
An alternative representation of the quantum eld can also be constructed, by using
another quasi-ditribution, dened by
Q(,

) =
1

< [[ > (3.215)


In order to see how such a representation an be establish we start from the normalization
condition Tr. = 1. Representing the density matrix in the set of Fock states, we have
Tr. =

n
< n[[n >= 1 (3.216)
Using the completeness relation
1

_
[ >< [d = I (3.217)
this can be transformed into
Tr. Tr.(I) =
1

n
_
< n[ >< [[n > d (3.218)
64
or
Tr. =
1

n
_
< [[n >< n[d =
1

_
< [[ > d (3.219)
where we have use

n
[n >< n[ = I. Comparing with equations (3.215) and (3.216), we
conclude that
_
Q(,

)d = 1 (3.220)
Let us now calculate the mean value f an operator of the type a
m
a
n
. We get
< a
m
a
n
>= Tr.(a
m
a
n
) =

n
< n[ a
m
a
n
[n > (3.221)
Introducing the identity (3.217) between the operators a
m
and a
n
, we can transform this
expression into
< a
m
a
k
>=
1

n
< n[ a
m
[ >< [a
k
[n > d =
=
1

_

m

k
(

n
< n[[ >< [n >)d =
1

_

m

k
< [[ > d (3.222)
Using the denition of Q(,

), as given by equation (3.215), we nally conclude that


< a
m
a
n
>=
_
Q(,

)
m

n
d (3.223)
This shows that the new quasi-probability, Q(,

), plays a symmetric role when com-


pared with P(,

). One can allow us to calculate mean values of operators of the type


(a
m
a
n
), and the other of operators (a
m
a
n
).
3.9 Generalized representation
Let us now consider a generalized representation of the density operator , which reduces
to the previous P and Q representations in particular situations. Let us then call F(,

)
this generalized quasi-distribution, dened in such a way that
=
_
F(,

) (, a;

)d (3.224)
65
where the operator is dened by
(, a;

) =
1

2
_
e
(,

)
e
(

)
e

(a)
d (3.225)
As a particular case, let us consider the function (,

) appearing in this denition, as


equal to
(,

) =
[[
2
2
(3.226)
Using the identity
e
a

a
= e
a

a
e
||
2
/2
(3.227)
we can then write the operator (3.225) for this particular choice of (,

), as
(, a;

) =
1

2
_
e
a

a
e

d (3.228)
Using the completeness relation for the coherent states, we can transform this expression
into
(, a;

) =
1

3
_
d

_
d[, e

(a)
[

><

[ e
(

)
(3.229)
This can be rewritten as
(, a;

) =
1

3
_
d

_
d e

)
e
(

)
[

><

[ (3.230)
We can now use the two dimensional Dirac delta function

2
(

) (

)(

) =
1

2
_
e

)
e
(

)
d (3.231)
where we notice that
_
d represents a double integral. From here, we get
(, a;

) =
1

_
(

)(

)[

><

[d

=
1

[ >< [ (3.232)
Replacing this in the denition of the generalized quasi-distribution, we obtain
=
_
F(,

) [ >< [d (3.233)
We can see from here that the choice of (, a;

) as given by equation (3.226) reduces


the generalized quasi distribution to that of the P-representation
P(,

) = F(,

) (3.234)
66
Let us now consider a dierent choice of (, a;

) and slightly change equation (3.226),


replacing it by
(,

) = +
[[
2
2
(3.235)
In this case, we can use the identity
e
a

a
= e

a
e
a

e
||
2
/2
(3.236)
and replace it in equation (3.225). The result is
(, a;

) =
1

2
_
e
(

)
e

(a)
d (3.237)
Let us now calculate the mean value
<

[(, a;

)[

>=
1

2
_
e
(

)
e

)
d = (

) (3.238)
Replacing this in equation (3.224), we obtain
< [[ >=
_
F(

) <

[(, a;

)[

> d

= F(,

) (3.239)
Therefore, the choice of (3.235) leads us to the Q-representation, as stated by
F(,

) = Q(,

) (3.240)
Finally, let us consider the condition
(,

) = 0 (3.241)
In this case, it can be shown that the generalized distribution F(,

) will reduce to the


Wigner distribution, as dened by
W(,

) =
2

2
e
2||
2
_
< [[ > e
2(

)
d (3.242)
67
3.10 Photon angular momentum
An important and sometimes forgotten aspect of the electromagnetic eld theory is the
angular momentum. Let us start by considering the linear momentum of electromagnetic
radiation. As shown by classical theory, this quantity is given by the integral o Poyntings
vector (

E

H) over a given volume V , or

P =
1
c
2
_
V
(

E

H)dr =
0
_
V
(

E

B)dr (3.243)
In quantum theory, we have to replace the electric and magnetic elds by their corre-
sponding operators. Notice here that the property for the external product of the elds
(

E

B) = (

B

E) remains valid for the eld operators, because these eld operators
obey the commutation relation
_

E ,

B
_
= 0. By following a procedure similar to that
used to establish the energy operator, we arrive at the expression

P =

k,

k,
(3.244)
where the operator linear momentum

P

k,
, associated with a given wavevector

k and a
given polarization state , is determined by

k,
= (a

a + aa

)[e

k
(e

k)] (3.245)
Noting that [e

k
(e

k)] = (e

k
e

k
)

k =

k, we can more simply write the linear momentum
operator for each eld mode (dropping, as usual, the reference to

k and )

k,
=

k(a

a + aa

) =

k
_
a

a +
1
2
_
(3.246)
This shows that the photon momentum operator is proportional to the energy operator
H. The expectation value for this operator, when the eld is at some given Fock state n
is simply given by
<

P

k,
>=

k
_
n +
1
2
_
(3.247)
At rst sight this seems almost identical to the expectation value for the energy operator
H. However, there is an important dierence related with the absence of net momentum
68
for the vacuum state. For n = 0, the zero point momentum

k/2 is exactly canceled by


the zero point momentum for the mode with the same frequency and polarization but
pointing in the opposite direction

k. This leads to <



P >= 0 for the vacuum state [0 >.
Let us now turn to the angular momentum. In classical terms, we can dene the eld
angular momentum

J by the expression

J =
0
_
V
_
r (

E

B)
_
dr (3.248)
If we represent the magnetic eld in terms of the vector potential, as

B =

A, we can
write

E (

A) = (

E )

A +

j
E
j
A
j
(3.249)
where j = x, y, z. This expression can be obtained by using the well-known vector identity
for the double external product
a (

b c) = (a c)

b (a

b)c (3.250)
This can be rewritten as
a (

b c) =

j
a
j

bc
j
(a

b)c (3.251)
again with j = x, y, z. Taking the particular case of a =

E,

b = and c =

A we obtain

E (

A) =

j
E
j
A
j
(

E )

A (3.252)
We can then write the eld angular momentum in the form

J =
0
_
V
_

j
E
j
(r )A
j
r (

E )

A
_
dr (3.253)
Furthermore, it can also be noted that
r (

E )

A = (

E )(r

A) (

E

A) (3.254)
This leads to

J =
0
_
V
_

j
E
j
(r )A
j
+ (

E

A) (

E )(r

A)
_
dr (3.255)
69
Integrating the last term in this expression by parts, we get
_
V

E (r

A)dr =
_
V
_


E(r

A)
_
dr
_
V
(

E)(r

A)dr (3.256)
The rst of these two terms can be transformed into a surface integral. Extending the
volume to innity, and assuming that the elds all go to zero at innity, we get a zero
contribution from this surface integration. For the second term, we can use the Poisson
equation associated with a charge distribution in vacuum, and we get

E = /
0
. For
purely transverse radiation eld, we take = 0, and this term can also be take equal to
zero. We are then left with the following expression for the eld angular momentum

J =
0
_
V
_

j
E
j
(r )A
j
+ (

E

A)
_
dr (3.257)
In order to nd the quantum operators corresponding to these two dierent terms of the
feld angular momentum, let us introduce the Fourier transformation, such that for each
Fourier component of the eld, the following replacements can be made
i

k , r = i

k
i

k
(3.258)
After averaging over time, we get for the photon angular momentum of each eld mode

k
=
0
_

j
E

j,

k
(

k
)A
j,

k
+ (

k
)
_
(3.259)
and, for the total angular momentum

J =
_
V

k
d

k
(2)
3
(3.260)
As will be shown next, the two terms of the angular momentum have very distinct physical
meanings. The rst one is associated with the orbital angular momentum

L and the second
to the polarization state or spin,

S. We can then write equation (3.259) in the form

k
=

L

k
+

S

k
(3.261)
Let us now replace the elds

E and

A by their respective operators. We will get

k
= i

j
a

k
(

k
)a
j

k
,

S

k
= ia

k
a

k
(e

k
e

k
) (3.262)
70
where a
j

k
are the components of a

k
= a

k
e

k
, and e

k
are the eld polarization directions.
We obviously have a

k
= a

k
e

k
. Let us now consider three possible polarization states of
the eld with respect to the same direction of propagation alomg the z-axis,

k = ke
z
.
These states are
e

=
1

2
(e
x
ie
y
) , e
z
(3.263)
This leads to the following values for the external product in the spin term
(e

) = ie
z
, (e
z
e
z
) = 0 (3.264)
Replacing this in equation (3.262), we obtain the following values for the spin operator

S = a

a e
z
,

S = 0 (3.265)
For simplicity we have dropped the subscript

k, as usual. Applying this a given Fock
state [n >, we get
< n[

S[n >= n e
z
, 0 (3.266)
Tis shows that the photon polarization determines its spin, and that the spin can be
attributed to a single photon property, as clearly shown when we consider the particular
case of n = 1. The photon is clearly a particle with spin 1, with three possible values for
its projection along the propagation direction, S
z
= 0, . In free space the third state,
corresponding to S
z
= 0 is absent, because the electromagnetic wave eld propagating
in free empty space is purely tramsverse. But in a medium with free charges, such as
a plasma this longitudinal spin state is present, and the corresponding photon state is
called a plasmon.
Such a simple treatment cannot be extended to the orbital angular momentum op-
erator

L. For a generic eld state vector, [psi >, the expectation value for the orbital
angular momentum will be
< [

L[ >= i
_

j
< [a

k
(

k
)a
j

k
[ >
d

k
(2)
3
(3.267)
Here it is useful to consider the momentum representation of the state vector, as dened
by the quantities

j
(

k =<

k[a
j,

k
[ > (3.268)
71
We then have
< [

L[ >= i
_

j

j
(

k)(

k
)
j
(

k)
d

k
(2)
3
(3.269)
Comparing with the above denition of the photon spin, we notice that the total spin
state of the electromagnetic eld can also be determined by
< [

S[ >= i
_

(

k)

(

k)
d

k
(2)
3
(3.270)
where

(

k) is a vectorial wave function with components given by (3.268), or

k =<

k[a

k
[ > (3.271)
We clearly see that the photons are spin 1 particles, with three possible states S
z
as
indicated above, and described by a three-dimensional spinor [

>= a

k
[ >. On the
other hand, the existence of a nite value for the photon orbital angular momentum,
implies the non-orthogonality between the wavevector

k and this spinor [

>, as shown
by equation (3.269).
In general, the eigenfunctions of the eld total angular momentum

J, and its projection
J
z
can be found in analogy with the elementary Quantum Mechanics. The appropriate
eigenvectors are spherical harmonic vectors

Y
jm
, such that
J
2

Y
jm
= j(j + 1)

Y
jm
, J
z

Y
jm
= m

Y
jm
(3.272)
These vector wave functions can take the form of

Y
jm
= vY
jm
(n) (3.273)
where n =

k/k is the unit vector in the propagation direction, and Y
jm
(n) are the usual
scalar spherical harmonic functions
Y
jm
(n) = i
j

(2j + 1)(j [m[)!


4(l +[m[)!
P
m
j
(cos )e
im
(3.274)
Here, P
m
J
are the associated Legendre polynomials, and and the polar angles of the
direction dened by n. The vector v can be shown to take the possible three values
v = n , v =

n
_
j(j + 1)
, v =
(n
n
)
_
j(j + 1)
(3.275)
72
3.11 Problems
P3.1 - Exploring the analogy between an LC circuit and an harmonic oscillator, show
that this circuit can be quantized.
P3.2 - Show that the electric eld operator can be written in the form

E = i
_

2
0
V
_
a(r)e
it
a

(r)e
it

(3.276)
P3.3 - Using [a, a

] = 1, derive the evolution equations (3.87) for the operators a and


a

.
P3.4 - Prove that the operators N, cos and sin are Hermitian. What about a, a

and e
i
?
P3.5 - Derive equation (3.125) for the cosine operator.
P3.6 - Derive equations (3.132) - (3.134) for the electric eld operator, and its root
square mean deviation.
P3.7 - Demonstrate de second form of the Baker-Hausdor theorem as given by the
last term of equation (3.142).
Chapter 4
Semi-Classical Theory of Atom-Field
Interaction
4.1 Time-dependent Perturbations
In order to understand the mean properties of the atom interactions with a radiation eld,
we consider rst a semi-classical approach where we assume the quantum description of
a two-level atom and describe radiation as a classical eld. The full quantum description
of both atom and radiation will be considered later.
This semi-classical description will allow us to understand the emission and absorption
mechanisms of light by atoms, and to establish the basis of the laser theory. The total
Hamiltonian operator relevant to this problem is given by
H = H
a
+ H
int
(4.1)
where H
a
is the unperturbed Hamiltonian of the atom, with the eigenvalue equation
H
a
[j >= E
j
[j > (4.2)
where E
j
are the energy values of the atomic quantum levels, with the corresponding
eigenstates [j >. Here we will consider a simple two-level atom model, where j = 1, 2. In
equation (4.1) we have also used the interaction Hamiltonian H
int
, which described the
73
74
coupling between the atom and the radiation eld, as specied later. In the absence of
such coupling, or when H
int
= 0, the atomic state vectors have solutions of the form
[j, t >= exp(iE
j
t/)[j > (4.3)
In the presence of coupling, H
int
,= 0, the two atomic states become coupled by the
radiation eld, and the state vector solution will have to contain a superposition of the
two atomic states, as described by
[(r, t) >=

j=1,2
C
j
(t) exp(iE
j
t/)[j > (4.4)
where the coecients C
j
(t) have to obey the normalization condition
< (r, t)[(r, t) >=
_
[(r, t)[
2
dr = [C
1
(t)[
2
+[C
2
(t)[
2
= 1 (4.5)
The quantities [C
j
(t)[
2
are equal to the probability for the atom to be in a given state
j. In order to determine the value of these probabilities in the presence of radiation, we
have to consider the time dependent Schroedinger equation
i

t
[ >= (H
a
+ H
int
)[ > (4.6)
Using the superposition of states (4.4), we obtain

j=1,2
_
i
C
j
t
+ E
j
C
j
_
exp(iE
j
t/)[j >=

n=1,2
(H
a
+ H
int
)C
n
exp(iE
n
t/)[n > (4.7)
By using the unperturbed eigenvalue equation (4.2), this is reduced to
i

j=1,2
C
j
t
exp(iE
j
t/)[j >=

n=1,2
H
int
C
n
exp(iE
n
t/)[n > (4.8)
Or, in explicit form
i
C
1
t
[1 > +i
C
2
t
e
i
0
t
[2 >= C
1
H
int
[1 > +C
2
H
int
e
i
0
t
[2 > (4.9)
where we have introduced the transition frequency
0
= (E
2
E
1
)/. Multiplying this
equation on the left, successively by < 1[ and < 2[, and using the orthonormality condition
< j[n >=
jn
, we obtain the following two equations
i
C
1
t
= C
1
H
11
+ C
2
H
12
e
i
0
t
, i
C
2
t
= C
1
H
21
e
i
0
t
+ C
2
H
22
(4.10)
75
where we have introduced the matrix elements of the interaction Hamiltonian as dened
by
H
ij
=< i[H
int
[j >=
_

i
(r)H
int

j
(r)dr (4.11)
Before studying the solutions of equations (4.10) we need to discuss the explicit form of
these matrix elements.
4.2 Interaction Hamiltonian
Let us consider the generic case where the two quantum levels [1 > and [2 > are associated
with the electron eigenstates in the atom. It is known that the total Hamiltonian of an
electron in the atom, in the presence of a radiation eld, takes the form
H =
1
2m
[ p + e

A(r, t)]
2
e[V (r, t) + V
0
(r)] (4.12)
where p is the canonical momentum, V
0
(r) is the internal electric potential of the atom,
and

A and V are the vector and scalar electromagnetic potentials, describing the radiation
eld.
In the Coulomb gauge, the radiation eld satises the conditions V (r, t) = 0, and
A(r, t) = 0: We can then identify the unperturbed atom Hamiltonian as
H
a
=
p
2
2m
eV
0
(r) (4.13)
and the interaction Hamiltonian is
H
int
=
e
m
p

A(r, t) +
e
2
2m
A
2
(r, t) (4.14)
This last expression can be approximated in the following way. First, for low intensity
radiation elds, the quadratic term in A
2
can be neglected. Second, we can assume that
the wavelength of the electromagnetic radiation is much larger than the dimensions of
76
the atom. This allows us to develop in the remaining term of equation (4.14), the vector
potential

A(r, t) around the position of the atom r
0
, as

A(r, t) =

A(t) exp(i

k r) =

A(t) exp(i

k r
0
)
_
1 + i

k (r r
0
) + ...
_


A(t) exp(i

k r
0
)
(4.15)
For simplicity, we will assume that the atom is at the origin of the coordinates, r
0
= 0.
This is called the dipole approximation, for reasons that will become apparent later. The
resulting interaction Hamiltonian is then
H
int
=
e
m
p

A(t) (4.16)
This can be written in an alternative and nearly equivalent way, in terms of the electric
eld

E =

A/t. For a sinusoidal wave, we have

E =

E
0
exp(it), and we can write

A(t) =
i

E(t) =
i

E
0

exp (it) (4.17)


On the other hand, if we use the Heisenberg evolution equation for the position operator,
we can write
p = m
dr
dt
=
m
i
[r, H
a
] (4.18)
The matrix elements of the operator momentum p, are then given by
< 1[ p[2 >=
m
i
< 1[[rH
a
[2 > < 1[H
a
r[2 > = im
0
< 1[r[2 > (4.19)
where we have used the transition frequency
0
= (E
2
E
1
)/.This suggests us to replace
p in equation (4.16) by i
0
r, and the vector potential

A by the electric eld

E. The
result is
H
int
=
e
0

r

E(t) (4.20)
Noting that the radiative transition will only occur in the immediate vicinity of the
transition frequency,
0
, we can nally write the interaction Hamiltonian in the
dipole approximation as
H
int
= er

E(t) (4.21)
77
We can further transform this expression by using the identity operator

n
[n >< n[ = I.
For a two-level atom, this reduces to
[1 >< 1[ +[2 >< 2[ = I (4.22)
We can then obtain
H
int
= eI (r e) IE(t) = (p
12
[1 >< 2[ + p
21
[2 >< 1[)E(t) (4.23)
where we have used the unit polarization vector e =

E/E, and dened the electric dipole
moment of the atomic transition in the direction of e by
p
12
= p

21
= e < 1[(r e)[2 > (4.24)
In the particular case of linear polarization along the x-axis, we simply have (r e) = x.
4.3 Radiative Transitions
Let us now return to the evolution equation of the atomic states. It can easily be realized
that o-diagonal terms of the interaction Hamiltonian are determined by
H
12
= H

21
= p
12
E(t) (4.25)
We also notice that the diagonal terms of this Hamiltonian are equal to zero, because
they contain factors of the form < 1[2 >= 0. We therefore have
H
11
= H
22
= 0 (4.26)
Replacing this in the evolution equation (4.10), we obtain
i
C
1
t
= C
2
H
12
e
i
0
t
, i
C
2
t
= C
1
H

12
e
i
0
t
(4.27)
For convenience, we take the particular case of a real electric eld E(t) = E
0
cos(t).
This then leads to the following coupled equations
C
1
t
= i
R
e
i
0
t
cos(t)C
2
,
C
2
t
= i

R
e
i
0
t
cos(t)C
1
(4.28)
78
where the new quantity
R
is called the complex Rabi frequency, and is dened by

R
=
1

p
12
E
0
=
1

[p
12
[E
0
e
i
(4.29)
where is the phase of the dipole moment matrix element. The physical meaning of this
frequency will be explained later. Let us discuss the solutions of the coupled equations
(4.28). First, we consider the case of an atom which is initially in the lower energy level
[1 >. This means that our initial conditions are
C
1
(0) = 1 , C
2
(0) = 0 (4.30)
Let us also assume that the electric eld amplitude E
0
is so weak that the occupation
probability of the lower energy level is not signicantly changed by the interaction Hamil-
tonian. We can then take C
1
1 as a constant in the second equation (4.28). The
solution is then simply given by
C
2
(t) = i

R
_
t
0
e
i
0
t

cos(t

)dt

(4.31)
This can easily be integrated, leading to
C
2
(t) =
1
2

R
_
1 e
i(
0
+)t
(
0
+ )
+
1 e
i(
0
)t
(
0
)
_
(4.32)
The transition probability from the lower to the upper energy state can then be cal-
culated from [C
2
(t)[
2
. This solution shows that such a transition represents a resonant
process, which is only signicant when the eld frequency approaches the transition
frequency
0
. In this case, the second term in equation (4.32) will be dominant, and we
can approximately write
[C
2
(t)[
2

[
R
[
2
4

1 e
i(
0
)t
(
0
)

2
= [
R
[
2
sin
2
[(
0
)t/2]
(
0
)
2
(4.33)
The neglected of the non-resonant term is known as the rotating wave approximation.
The transition probability attains a maximum at resonance, when =
0
, with a value
that evolves proportional to the square of time
[C
2
(t)[
2
=
1
4
[
R
[
2
t
2
(4.34)
79
This quadratic increase is only valid for a short time interval, after which the approx-
imation of a constant C
1
(t) is no more valid, and the solution (4.33) brakes down.
It is also useful to integrate of a nite spectral range , across which the electric
eld amplitude and the electric dipole matrix element [p
12
[ can be assumed constant. The
resulting transition probability will be given by
[C
2
(t)[
2
[
R
[
2
_

0
+/2

0
/2
sin
2
[(
0
)t/2]
(
0
)
2
d (4.35)
In the limit of short time intervals, such that 1/t, the integral in this expression
is approximately given by t
2
/4. Such a quadratic dependence with time disappears in
the long time interval limit, 1/t, where this is replaced by t/2. Notice that, in
this last limit, we can use the following denition of the Dirac delta function
lim
t
sin
2
[(
0
)t/2]
(
0
)
2

t
2
(
0
) (4.36)
which allows us to write
[C
2
(t)[
2
=

2
t[
R
[
2
(
0
) (4.37)
Dening the transition probability per unit time as w
12
() = [C
2
(t)[
2
/t, and using the well
known property of the Dirac delta function, (x) = (x)/, we can write this probability
as
w
12
() =

2
[
R
[
2
(E
2
E
1
) (4.38)
This is a very famous expression, usually called the Fermi golden rule. It can be used
to calculate the Einstein coecient B
12
, as shown next. For this purpose, we write the
explicit expression of the Rabi frequency, such that
[
R
[
2
=
[p
12
[
2

2
[E
0
[
2
(4.39)
For an electric eld with linear polarization along the z-axis, we have [p
12
[ = e[z
12
[. At
resonance =
0
, we then get
w
12
() =

2
e
2

2
[z
12
[
2
[E
0
[
2
(4.40)
80
Now, using the denition of the electromagnetic energy density W() =
0
[E
0
[
2
/2, we
obtain
w
12
() =
e
2

2
[z
12
[
2
W() (4.41)
The factor multiplying the energy density W() will then determine the absorption co-
ecient, rst introduced by Einstein using phenomenological arguments. We can state it
as
B
12
=

3
e
2

2
[z
12
[
2
(4.42)
The additional factor of 1/3 appearing in this expression results from taking the average
over all the possible orientations of the atomic z-axis, and assuming that all the directions
are equally valid, which is obviously true for thermal radiation.
4.4 Rabi Oscillations
In the previous section we have considered solutions for the two-atom in the presence
of radiation which are only valid in the limit of short interaction times, such that the
population of the lower energy level could be considered constant. Here we consider more
general solutions, for the atom, which stay valid for arbitrarily long interaction times.
Let us go back to the coupled mode equations (4.28), for the coecients C
1
(t) and C
2
(t)
which determine the probability of occupation of the lower and higher energy states of a
two-level atom. In the rotating wave approximation, these equations reduce to
C
1
t
=
i
2

R
e
it
C
2
,
C
2
t
=
i
2

R
e
it
C
1
(4.43)
where =
0
is the frequency detuning. In order to solve these equations, we
introduce an auxiliary quantity
b
2
= C
2
e
it
(4.44)
This leads to
C
1
t
=
i
2

R
b
2
,
b
2
t
+ ib
2
=
i
2

R
C
1
(4.45)
81
After a second time derivative, we get a closed equation for b
2
, in the form

2
b
2
t
2
+ i
b
2
t
=
1
4
[
R
[
2
b
2
(4.46)
By trying solutions of the form b
2
(t) exp(it), it can easily be realized that the frequency
is determined by
=
1
2
( ) , =
_
()
2
+[
R
[
2
(4.47)
which leads to a general solution
b
2
(t) =
_
Be
it/2
+ B

e
it/2
_
e
it/2
(4.48)
where B and B

are two constants of integration. Let us assume that the atom is initially
in its lower energy level, which means that C
1
(0) = 1 and b
2
(0) = 0. In such conditions,
we have B = B

, and this solution is reduced to


b
2
(t) = 2iBe
it/2
sin(t/2) (4.49)
Replacing this in the second of equations (4.45), we get
C
1
(t) =
2B

R
e
it/2
_
cos(t/2) i

sin(t/2)
_
(4.50)
Comparing this with the initial condition C
1
(0) = 1, we conclude that the constant of
integration has to be taken equal to B =

R
/2. The probability for the atom to be in
the excited state [2 > is then determined by
[C
2
(t)[
2
= 4[B[
2
sin
2
(t/2) =
[
R
[
2
[
R
[
2
+ ()
2
sin
2
_
t
2
_
[
R
[
2
+ ()
2
_
(4.51)
This shows that the occupation of the upper level oscillates with a frequency . In the
resonant case of =
0
, or = 0, we are reduced to = [
R
[. This shows that the
Rabi frequency [
R
[ is the natural frequency of oscillation of the energy level occupation
probability, at resonance. In this case, the above expression reduces to
[C
2
(t)[
2
=0
= sin
2
_
1
2
[
R
[t
_
(4.52)
82
This means that the occupation probability attains its maximum value after multiples of
the time interval t = /2[
R
[. Noting that
[C
1
(t)[
2
+[C
2
(t)[
2
= 1 (4.53)
we conclude that the occupation probability oscillated between the two states of the
atomic transition, and that, on the average, its occupation is 1/2 for both states.
4.5 Density matrix description
It is sometimes useful to describe the Rabi oscillations in terms of the density matrix
elements. For a two-level atom in a pure quantum state, the density operator can simply
be dened as
= [ >< [ (4.54)
with
[ >= C
1
(t)[1 > +C
2
(t)[2 > (4.55)
We can write it more explicitly as
= [C
1
[
2
[1 >< 1[ +[C
2
[
2
[2 >< 2[ + C
1
C

2
[1 >< 2[ + C

1
C
2
[2 >< 1[ (4.56)
This can be represented by a 2 by 2 matrix
=

11

12

21

22

(4.57)
with the diagonal matrix elements dened by

11
=< 1[[1 >= [C
1
(t)[
2
,
22
=< 2[[2 >= [C
2
(t)[
2
(4.58)
and the o-diagonal terms by

12
=< 1[[2 >= C
1
(t)C

2
(t) ,
21
=

12
(4.59)
83
We see that the diagonal terms are real quantities and their sum equal to one,
11
+
22
= 1.
We can also see that, for a collection of identical atoms N per unit volume, there will be
a population of atoms in the lower state N
1
(t), and a population in the upper state N
2
(t),
such

11
= N
1
(t)/N ,
22
= N
2
(t)/N (4.60)
The temporal evolution of the density matrix elements can be derived from the above
equations for the coecients C
j
(t), according to
d
dt

ij
= C
i
dC

j
dt
+ C

j
dC
i
dt
(4.61)
using equations (4.28), we obtain
d
11
dt
=
d
22
dt
= i cos(t)
_

R
e
i
0
t

12

R
e
i
0
t

21
_
(4.62)
and
d
12
dt
=
d

21
dt
= i
R
cos(t)e
i
0
t
(
11

22
) (4.63)
In order to solve these equations we notice that the products of the cosine with the
exponential functions lead to two kinds of oscillating terms, with high and low frequencies,
(
0
+ ) and (
0
). After integration, the high frequency terms will give a small
contribution, due to their rapid oscillations, which can be neglected. This is the rotating
wave approximation, already used before. In this approximation, the above equations can
be reduced to
d
11
dt
=
d
22
dt
=
i
2

R
e
it

12

i
2

R
e
it

21
(4.64)
and
d
12
dt
=
d

21
dt
= i
R
e
it
(
11

22
) (4.65)
where, as before, = (
0
). These equations are known as the optical Bloch
equations. The original Bloch equations, which are formally identical, were derived to
describe the spin states in a magnetic eld. In order to solve these equations, we consider
solutions of the form
(
11
,
22
) e
t
, (
12
,

21
) e
tit
(4.66)
84
Replacing this in the above Bloch equations, we get a system of algebraic equations which
can be written in the following matrix form
_

0
i
R
/2
i

R
/2
0

i
R
/2
i
R
/2
i

R
/2
i

R
/2
i
0
i
R
/2
i
R
/2
0
i
_

_
_

11

22

12

21
_

_
= 0 (4.67)
In order to obtain a non-trivial solution for the matrix elements
ij
, we have to make the
determinant of the (4 4) matrix of the coecients equal to zero, This leads to

2
(
2
+
2
+[
R
[
2
) = 0 (4.68)
This has four distinct solutions, given by

1,2
= 0 ,
3,4
= i = i
_

2
+[
R
[
2
(4.69)
We can then write the general solution of the Bloch equations as

ij
= a
ij
+ b
ij
e
it
+ c
ij
e
it
(4.70)
where a
ij
, b
ij
and c
ij
are constants, and we have additional exponential factors for the
o-diagonal terms, as stated in equation (4.66). In particular, for the initial conditions

11
(0) = 1 ,
22
(0) =
12
(0) = 0 (4.71)
we get the solutions

22
(t) =
[
R
[
2

2
sin
2
_

t
2
_
,
12
(t) =
[
R
[
2

2
sin
_

t
2
__
i cos
_

t
2
_
sin
_

t
2
__
e
it
(4.72)
In the absence of detuning, = 0, this reduces to

22
(t) = sin
2
_
[
R
[
t
2
_
,
12
(t) = ie
i
sin
_
[
R
[
t
2
_
cos
_
[
R
[
t
2
_
(4.73)
where is the phase of the complex Rabi frequency,
R
= [
R
[ exp(i).
85
4.6 Selection Roles
We have seen that, in the electric dipole approximation, the probability for photon emis-
sion by an atom is proportional to the matrix element < n[z[m >. We usually say that
an electric dipole transition is forbidden, if such its associated matrix element is equal to
zero. The calculation of these quantities will therefore allow us to determine the selection
rules under which such transitions are possible. In order to do that, we consider the initial
and nal atom states as represented by wave functions of the form

nlm
(r) = R
nl
(r)Y
lm
(, ) (4.74)
As we have seen earlier, this general form is valid when the eective potential acting on
the electrons in the atom have spherical symmetry. Otherwise, it is impossible to separate
the radial part of the wave function (or the orbitals R
nl
), from the angular part Y
lm
. In
order to avoid misunderstanding with the quantum numbers involved, we introduce the
following notation for the initial and nal states of the radiative transition, [m > and n >
respectively.

n
1
l
1
m
1
(r) =< r[m > ,
n
2
l
2
m
2
(r) =< r[n > (4.75)
We can then dene the dipole matrix element as
z
21
< 2[z[1 >=
_

n
2
l
2
m
2
(r) z
n
1
l
1
m
1
(r)dr (4.76)
Noticing that, in spherical coordinates we have z = r cos = r
_
4/3Y
01
(, ), and
dr = r
2
drd, where d is the element of solid angle, we can write this matrix element as
z
12
=
_
4
3
_

0
dr r
3
R

n
2
l
2
(r)R
n
1
l
2
m
1
(r)
_
d Y

l
2
m
2
Y
10
Y
l
1
m
1
(4.77)
where the arguments of the spherical harmonics are not shown, for simplicity. Let us rst
look, inside the solid angle integral and the angular integration on . It is simply given
by
_
2
0
exp(im
2
) exp(im
1
)d (4.78)
86
This integral is only nonzero for m m
2
m
1
= 0, which would give us a rst selection
rule. However, this rule is not general, because it corresponds to the case of a linearly
polarized electric eld along Oz. For a more general polarization case, we will have to
replace in the above integrals, the operator z by the new operator
r e = r (sin cos e
x
+ sin sin e
y
+ cos e
z
) (4.79)
where e, with components (e
x
, e
y
, e
z
), is the unit polarization vector of the electromag-
netic radiation, e =

E
0
/E
0
. If we use the explicit expressions for the rst few spherical
harmonics
Y
10
=
_
3
4
cos , Y
1,1
=
_
3
8
sin e

(4.80)
we can write
sin cos =
1
2
_
8
3
(Y
1,1
Y
1,+1
) , sin sin =
1
2i
_
8
3
(Y
1,1
+ Y
1,+1
) (4.81)
from where we get
r e =
_
4
3
r (e
+
Y
1,1
e

Y
1,+1
+ e
z
) (4.82)
where e

= (e
x
ie
y
)/

2 describe the left and right circular polarization states, and


where the last term correspond to that appearing in the above integrals. We therefore
see that, apart from this term, there appears an additional azimuthal dependence due to
the terms in Y
1,1
, which takes the form
_
2
0
exp(im
2
) exp(i) exp(im
1
)d (4.83)
This integral is only dierent from zero when m m
2
m
1
= 1, thus leading to
a more general azumuthal selection rule, m = 0, 1. These new allowed transitions
correspond to radation electric elds circularly polarized in the plane Oxy, as described
by the unit polarization vectors
e

=
1

2
(e
x
ie
y
) (4.84)
In order to conclude our discussion on the selection rules, we should notice that the
operator position z, or more generally r, are odd operators. Consequently, they can only
87
establish a transition between initial and nal states of a dierent parity. Noting that
the parity of these states is dened by (1)
l
, we conclude that the radiative transition
can only occur for l l
2
l
1
= 1. This result can be directly obtained by solving
the radial integration in equation (4.77). Finally, the perturbation Hamiltonian ( p

A) is
independent of spin, which means that spin cannot change during the radtiative transition,
or s = s
2
s
1
= 0. We can then write the selection rules for dipolar radiation transitions
as
l = 1 , m = 0, 1 , s = 0 (4.85)
Such rules are valid as long as we can use the quantum numbers (n, l, m, s) to describe
the atomic energy states. In the present or spin-orbit coupling, for instance, the energy
states take the form [n, l, J, m
J
>, where

J =

L +

S is the total angular momentum,
and J, and m
J
the corresponding quantum numbers. New selection rules for the dipole
electric transitions can then be nd, as
l = 1 , J = 0, 1 , m
J
= 0, 1 (4.86)
Similar discussions could also be made for magnetic dipole and electric quadrupole tran-
sitions.
4.7 Susceptibility
Let us now consider the behavior of an electromagnetic wave propagating in a material
medium, made up of a large number of atoms or molecules. We have seen before that
the interaction of the wave with a single atom is determined by the transition probability
from one quantum state to another, with emission or absorption of a photon. From a
macroscopic point of view, the wave propagation is characterized by the refractive index
n(), dening the relation between the wave frequency and its wavenumber k, as
n() =
kc

(4.87)
88
where c is the speed of light in vacuum.In general, this quantity is complex, n() = n

+in

,
and its imaginary part determines if the wave is damped (n

< 0), or amplied (n

> 0)
in the medium. On the other hand, the refractive index can be dened in terms of the
susceptibility (), in such a way that
n()
2
= 1 + () (4.88)
This new quantity is the ratio between the electric eld applied to the medium

E and the
resulting polarization

P. Let us consider an electric eld of the form
E(t) = E
0
cos(t) (4.89)
For isotropic media, the polarization is determined by

P(t) =
1
2

E
0
_
()e
it
+ ()e
it

(4.90)
For anisotropic media we would have to replace the function () by a susceptibility
tensor. Noting that, the susceptibility is also complex in the general case, such that
() =

+ i

, we can easily establish the following relations

= n

2
n

2
1 ,

= 2n

(4.91)
Let us now see how the above semi-classical theory can be used to calculate the real and
imaginary parts of both the susceptibility and the refractive index. We know that the
polarization of a single atoms is given by
p(t) = e < (t)[z[(t) > (4.92)
where the direction of the incident electric eld was assumed in the Oz direction. For a
two level atom, such that
[(t) >= C
1
e
iE
1
t/
[
1
> +C
2
e
iE
2
t/
[
2
> (4.93)
with a transition frequency
0
= (E
2
E
1
)/ very close to the eld frequency , we have
p(t) = e [C

1
(t)C
2
(t)z
12
+ c.c.] , z
12
=<
1
[z[
2
> (4.94)
89
Solving for the coecients C
1
(t) and C
2
(t), as we did before, we arrive at
p(t) =
2
0
(
2
0

2
)
e
2

[z
12
[
2
E
0
sin(t) (4.95)
For a medium with N atoms per unit volume, the polarization vector will be given by

P(t) =
N

j=1
p
j
(t) = N p(t) (4.96)
The atoms are assumed oriented randomly with respect to the direction of the electric
eld. This means that the axis Oz chosen to describe the atom quantum state can be
at an arbitrary angle with respect to the eld

E
0
. WE then have to replace in the
above results [z
12
[
2
by [z
12
[
2
cos
2
. Taking the average over all the possible angles, we get
< cos
2
>= 1/3. We can then write the electric susceptibility of the medium as
() =
2
3
e
2
N

0
(
2
0

2
)
[z
12
[
2
(4.97)
This can also be written in a more compact form as
() =
2
p
f
12
(
2
0

2
)
(4.98)
where we have used

2
p
=
e
2
N
3
0
m
, f
12
= 2
m
0

[z
12
[
2
(4.99)
The quantity
p
is called the plasma frequency of the medium, and f
12
is the oscillation
strength of the atomic transition. If, instead of a single two-level atomic model, e consider
the possible existence of several radiative transitions in the atom, with transition frequen-
cies given by
ij
close to the eld frequency , we can easily generalize the expression of
the susceptibility of the medium to
() =
2
p

i,j
f
ij
(
2
ij

2
)
(4.100)
where f
ij
are now the corresponding oscillator strengths.
90
4.8 Semi-Classical Laser Theory
We are now in conditions to describe the semi-classical model of the laser. We start from
Maxwells equations, and write the electric eld

E in terms of the current density

J and
the displacement vector

D, as
(

E) =
0

t
_

J +

D
t
_
(4.101)
We now replace

D by its expression in terms of the polarization vector

P, as

D =
0

E +

P (4.102)
We also use Ohms law,

J =

E, were is the conductivity of the medium. this then


leads to

E (

E)
0

E
t
+
1
c
2

E
t
2
=
0

P
t
2
(4.103)
We now consider the eld inside an optical cavity, with the axis along Oz. Assuming
that the eld is transverse,

E = 0, and neglecting the weak eld dependence along
the perpendicular directions Ox and Oy, we can write

E(r, t) =

E(z, t), and the above
equation can be reduced to
_

2
z
2

1
c
2

2
t
2
_

E
0

E
t
=
0

P
t
2
(4.104)
Next, we assume that the eld

E and

P are linearly polarized along the transverse direc-
tion Ox, and forget for the vectorial character of these elds. We consider eld solutions
characterized by a given frequency , and a wavenumber k, of the form
E(z, t) =
1
2
E
0
(z, t)e
i(kzt)+i(z,t)
+ c.c. (4.105)
where the eld amplitude E
0
(z, t) and phase (z, t) are assumed as slowly varying func-
tions of space and time. We also assume that they are real. Similarly, the polarization
eld response can be described by
P(z, t) =
1
2
P
0
(z, t)e
i(kzt)+i(z,t)
+ c.c. (4.106)
91
Let us consider the rst order derivative of the electric eld with respect to the space
variable
E
z
=
1
2
_
ik + i

z
+
1
E
0
E
0
z
_
E
0
e
i(kzt)+i(z,t)
+ c.c (4.107)
with a similar expression for the time derivative. On the other hand, due to this slow
character of the amplitudes and phase, we can use the envelope approximation, which
allows us to write
_

z

1
c

t
_
E
i
2
_
k +

c
_
E
0
e
i(kzt)+i(z,t)
+ c.c = 2ikE (4.108)
where we have used = kc. Replacing this in the wave equation (4.104), we get
2ik
_

z
+
1
c

t
_
E
0

E
t

0

2
P (4.109)
Using equation (4.107), a similar time derivative expression, and splitting the resulting
equation into real and imaginary parts, we get the evolution equations for the eld am-
plitude E
0
, and for the eld phase , as given by
_

z
+
1
c

t
_
=
_
k

c
_

k
2
0
E
0
'(P
0
) (4.110)
and
_

z
+
1
c

t
_
E
0
=

2
0
c
E
0

k
2
0
(P
0
) (4.111)
If we now apply these equations to a eld mode in a laser cavity, we have to identify the
wavenumber k with one of the possible discrete values for the eld modes in the cavity,
as
k
m
= m

L
=

m
c
(4.112)
This means that we assume propagation along the z-axis, with the eld E(z, t) evolving
in space as exp(ik
m
z). We also have to assume that there is re-injection of the same wave
mode in the cavity, always propagating in the same direction. This is for instance veried
in a ring laser conguration. We can then neglect the space evolution and only focus on
the temporal evolution of the eld amplitude. The phase equation (4.110) is then reduced
to

t
= (
m
)

2
0
E
0
'(P
0
) (4.113)
92
Similarly, for the eld amplitude, we can write

t
E
0
=
1
2

m
E
0


2
0
(P
0
) (4.114)
where we have introduced the damping coecent for the cavity mode
m
, such that

m
=
(
m
)

0
=

m
Q
(4.115)
where the quantity Q is the quality factor of the cavity. For a perfect cavity, we would
have Q and
m
0. But in such a perfect cavity, all the photons would be kept
inside, and no laser emission could be produced. These simplied evolution equations for
the eld amplitude and phase will be the starting point of our quasi-classical laser model.
We now have to couple them to the polarization eld P, which depends on the atomic
populations of the two energy levels of the atoms involved in the radiation transition at
(or near) the eld frequency . We have already seen that the polarization is related with
the electric eld by and expression which depends on the inversion of population, as given
by
P
0
(t) =
[p
12
[
2

( + i)
()
2
+
2
(
22

11
)E
0
(t) (4.116)
The evolution equations for the density matrix elements (4.64) and (4.65), can be gen-
eralized to include nite decay time of the two atomic levels, 1/
i
, with i = 1, 2. This
includes spontaneous decay as well as the eventuall existence of collisional decay. We can
also include nite excitation rates,
i
. This then leads to
d
22
dt
=
2

22
R(
22

11
) (4.117)
and
d
11
dt
=
1

11
+ R(
22

11
) (4.118)
where we have introduced the rate coecient
R =
1
2
[
R
[
2

()
2
+
2
(4.119)
with =
1
+
2
. This coecient determines the rate at which the population dierence
is varying in time, and depends on the intensity of the radiation eld. In steady state, we
93
simply have
(
22

11
) =
(
2
/
2

1
/
1
)
1 + 2R/
1

N
inv
1 + R/R
s
(4.120)
We can see that the quantity N
inv
is the population dierence in the absence of radiation.
We can now assume that the population dierence in the laser cavity remains nearly
constant, which is certainly true in the steady state regime. Replacing equation (4.120)
in equation (4.116), and then in the phase and eld amplitude equations, we obtain

t
= (
m
) +
A
2[1 + (B/A)(
0
V/2)E
2
]
(4.121)
and

t
E
0
=
1
2

m
E
0
+
AE
0
2[1 + (B/A)(
0
V/2)E
2
0
]
(4.122)
where V is the volume of the laser cavity, and where we have used the new quantities
A = [p
12
[
2

N
inv
()
2
+
2
(4.123)
and
B
A
=
[p
12
[
2

2
(2/
0
V )
()
2
+
2
(4.124)
It is now useful to introduce the number of photons in the cavity n() , which can be
dened by
n()
W()

V =

0
2
E
2
0

V (4.125)
where W() is the energy density at the frequency . The laser equations become

t
= (
m
) +
A
2(1 + Bn/A)
(4.126)
and
n
t
=
m
n +
An
2(1 + Bn/A)
(4.127)
These are the quasi-classical laser equations in their nal form. We can see that, for a small
level of photons inside the cavity, we can assume that the term Bn in the denominator
is very small. And, using this approximation, we can study the threshold of the laser
instability. By expanding the denominator in the above equations, we obtain

t
= (
m
) +

2
(A Bn) (4.128)
94
and
n
t
= (A
m
)n Bn
2
(4.129)
The laser threshold corresponds to the condition of increase in the photon number, or
dn/dt > 0, in the regime where n 0. We see clearly from this last equation that such a
threshold condition is dened by
A >
m
(4.130)
This simply states that the photon gain has to be larger than the cavity losses. This is
equivalent to state that the inversion of population N
inv
has to be larger than a certain
critical quantity N
crit
, where
N
crit
=

0

m
[p
12
[
2

[()
2
+
2
] (4.131)
At resonance, this reduces to
N
crit
=

0

m
[p
12
[
2

=

0

[p
12
[
2

Q
(4.132)
From the approximate laser equation (4.129) we can see that the steady state condition
is attained for n = 0, as well as for an equilibrium intensity n = n
0
, such that
n
0
=
A
m
B
(4.133)
By inserting this steady state solution into the phase equation (4.128), we also obtain a
constant phase solution for a eld frequency such that
=

m
+ S
1 + S
(4.134)
where S is a stabilizing factor dened by
S =
A Bn
0
2
=

m
2
(4.135)
We can see that, for this stabilized situation, the laser frequency is shifted away from
the cavity frequency
m
by a factor which depends on the cavity losses
m
. Only for low
cavity losses, such that
m
2, we will have
m
. This eect of line displacement
is usually quite small, and is called the mode pulling eect.
95
4.9 Laser cooling
Due to momentum conservation, when an atom emits one photon with frequency and
wavevector

k, it also receives a momentum

k, which is called the momentum recoil.


Inversely, when an atom absorbs a photon, it will absorb the photon momentum

k =
kn. If such an absorption is followed by spontaneous emission at the same frequency
= k/c, the emitted photon will propagate in an arbitrary direction dened by n

=

k

/k.
As a result, in the emission-absorption cycle the atom translational momentum is not
conserved, and it will receive a net momentum given by
p = (

) =

c
(n n

) (4.136)
If such a cycle is repeated many times by the same atom, the momentum loss by sponta-
neous emission will be averaged to zero and a net momentum gain will be obtained. We
can write the total momentum balance after r successive of absorption-emission as

P = r
_

k
r

j=1

j
_
(4.137)
The corresponding variation along the direction of the incident photons n, will be
P = rk k
r

j=1
cos
j
(4.138)
where
j
are the angles of spontaneous emission with respect to n. Taking the average
for a large number r 1, we get
< P >= rk k <

j
cos
j
> rk (4.139)
If r is the uorescence rate, we can dene the ponderomotive force acting on the atom
(the momentum variation per unit time) as

F
p
= r

k (4.140)
96
It is obvious that this uorescence rate is determined by the product of the spontaneous
decay rate = 1/A with the upper level energy level population of the atom, or
r =
22
(4.141)
Going back to the evolution equation for the density matrix elements describing the
radiation coupling between two energy levels [2 > and [1 >, and including the spontaneous
decay, we have
d
dt

22
=
22
+
i
2

R
(
21

12
) (4.142)
where
R
is the Rabi frequency associated with the laser eld incident on the atoms. This
has to be completed by the evolution equations for the of-diagonal matrix elements, as
determined by
d
dt

21
=
_
i +

2
_

21
+ i
R

22

i
2

R
(4.143)
where we have used detuning =
0
, between the atomic transition frequency

0
= (E
2
E
1
)/ and the laser frequency. Similarly, by taking the complex ocnjugate
of this equation, we get the evolution equation for
12
=

21
. In steady state, d/dt = 0,
these equations reduce to

22
=
i
2

R
(
21

12
) (4.144)
and

12
= i

R
2
2
22
+ 1
(i/2)
=

21
(4.145)
This leads to
(
21

12
) =
R
(2
22
1)

(
2
/4 +
2
)
(4.146)
Replacing in equation (4.144), and using equations (4.139)-(4.140) we can write the pon-
deromotive force of the incident laser beam on the atom as

F
p
=

k
22
=

k

2
R
2
2
R
+ 4
2
+
2
(4.147)
Let us now assume that the atom is moving with velocity v. Due to the Doppler eect,
it will perceive the laser photons as having the shifted frequency

k v. This
97
will introduce an additional detuning with respect to the transition frequency
0
. The
resulting expression for the force acting on the moving atom will then be

F
p
=

k

2
R
4( +

k v)
2
+
2
+ 2
2
R
(4.148)
For low intensities such that the radiative transition is not saturated,
2
R

2
/2, we can
neglect the Rabi frequency in the denominator of the force. Using the development
1
4( +

k v)
2
+
2

1
(4
2
+
2
) + 8(

k v)

1
(4
2
+
2
)
_
1
8
(4
2
+
2
)

k v
_
(4.149)
The force acting on the atom can then approximately determined by

F
p
=

k

2
R
(4
2
+
2
)
_
1
8
(4
2
+
2
)

k v
_
(4.150)
This can be rewritten in a more suggestive form, as

F
p
=

F
0
M

k
2k
2
v (4.151)
where M is the atomic mass, with

F
0
=

k

2
R
(4
2
+
2
)
(4.152)
and
= k
2
2
4

2
R

M(
2
+
2
/4)
(4.153)
Let us consider the particularly interesting case when the atom moves in the direction of
the laser propagation,

k | v. We are then reduced to

F
p
=

F
0

1
2
Mv (4.154)
We can then see that two dierent terms occur in the ponderomotive force

F
p
. The rst
term,

F
0
is a constant radiation pressure term, which is independent of the atom velocity.
In contrast, the second term can be seen as a friction term which depends on the atom
velocity. It is positive for an atom moving in the opposite direction with respect to the
laser beam (the atom slows down), and negative for an atom moving along the beam (the
98
atom is pushed forward). If now the atom is located in a region with two opposite laser
beams, and moves along the laser beams, it will experiment a total force in the

k direction
is given by

F
p
=
_

F
0


2
Mv
_

F
0
+

2
Mv
_
= Mv (4.155)
This shows that, with two counterpropagating beams, the pressure term term is zero and
the friction force is always negative, leading to a deceleration of the motion. The atom
will therefore slow down (or cool down) along that direction. If the same occurs for every
atom in the gas, the gas will then cool down. This is the laser cooling eect, which leads
to extremely low temperatures in the micro-Kelvin domain.
Let us now briey discuss the theoretical limits of this cooling process. These limits
can be discussed by using a statistical approach to the spontaneous emission. For that
purpose, we study the evolution of the atomic momentum distribution in the cold gas,
f(p, t), in one-dimension. This evolution is determined by a Fokker-Planck equation of
the form

t
f(p, t) =

p
_
A(p) + D(p)

p
_
f(p, t) (4.156)
where A(p) and D(p) are the friction and diusion coecients. A steady state solution
of this equation is given by
f(p) = f
0
exp
_

_
p
0
A(p

)
D(p

dp

_
(4.157)
We know that the friction coecient is related to the averaged particle momentum by
d
dt
< p >= A(p) (4.158)
Comparing with the the above results, we can write
A(p) = Mv = p = 8k
2


2
R
M(4
2
+
2
)
2
(4.159)
Noting that, in typical experimental situations we have
R
, we can for order of
magnitude estimates use a much simpler expression
A(p)
k
2
M
p (4.160)
99
Turning now to the diusion coecient D(p), we know that by denition, it is given by
D(p) =
1
2
< p >
2
t

1
4

2
k
2
(4.161)
The integration of the steady state solution (4.157) then leads to
_
p
0
A(p

D(p

dp


_
p
0
k
2
p

M
4

2
k
2

dp

=
2
M
p
2
(4.162)
We can then reduce the steady state distribution f(p) to a Maxwell distribution such that
f(p) = f
0
exp
_

p
2
2MT
eff
_
(4.163)
with the eective temperature (in energy units)
T
eff


2
(4.164)
We can see that this temperature is limited by the natural life-time of the atomic transi-
tion.
4.10 Problems
P4.1 - Show that the Shroedinger equation, with the Haamiltonian dened by equation
(4.12), is invariant under the gauge transformation

A =

A

+ , V = V


t
, [ >= e
ie/
[

> (4.165)
where is an arbitrary function of r and t.
P4.2 - Show that the quadratic term in A
2
of the Hamiltonian (4.12) can be eliminated
by a transformation of the form
[ >= [

> exp
_

ie
2
2m
_
A
2
(t

)dt

_
. (4.166)
P4.3 - Show that, for randomly polarized radiation, we obtain the factor of 3 in the
denominator of equation (4.97).
100
P4.4 - By using the denition of the oscillator strength f
1j
, and assuming the identity
I =

j
[j >< j[, show that

j
f
1j
= 1.
P4.5 - Include a spontaneous decay rate C
2
in equations (4.43), and determine the
corresponding solutions.
P4.6 - Assume three atomic levels, with allowed radiation transitions [3 >[2 > and
[2 > [1 >, and a forbidden transition [3 > [1 >. Establish the evolution equations
for the coecients C
j
(t), and determine the respective solution by using the rotating
wave approximation. Assume that the atom is initially in the lower level [1 >, and that
(E
3
E
2
) = (E
2
E
3
) = /, where is the radiation frequency.
Chapter 5
Quantum Theory of Atom-Field
Interaction
5.1 Hamiltonians
The interaction of the radiation eld with an atom can be described by an Hamiltonian
operator H which contains three distinct terms
H = H
a
+ H
f
+ H
int
(5.1)
where H
a
is the Hamiltonian of the atom, H
f
that of the radiation eld, and H
int
described
the interaction between the atom and radiation. Let us consider the rst term. In
the elementary formulation of Quantum Mechanics, the Hamiltonian H
old
describing the
electron energy levels inside the atom is determined by the time-independent Schroedinger
equation
H
old
(r) = E(r) , H
old
=
p
2
2m
+ V (r) (5.2)
where E is the internal energy of the atom. We can also represent the atomic wave function
as a superposition of eigenfunctions
j
(r), corresponding to dierent energy eigenvalues
E
j
, as
(r) =

j
b
j

j
(r) (5.3)
101
102
We can now look at the wave function (r) as a scalar eld, and follow a quantization
procedure similar to that used for the vector eld

A(r) describing the electromagnetic
radiation. In such a procedure, the quantity (r) will become an operator, and the
coecients b
j
introduced in equation (5.3) will also become operators. We can now say
that such operators obey the relations
=

j
b
j

j
,

j
b

j
(5.4)
while the wave functions
j
(r) will remains ordinary complex functions describing the
spatial structure of the electron probability distribution, for a given energy eigenstate E
j
.
They will satisfy the usual orthonormality condition
_

i
(r)
j
(r)dr =
ij
(5.5)
We can now dene the mean value of the atom Hamiltonian as
H
a
< [H
old
[ >=

i,j
b

i
b
j
_

i
(r)H
old

j
(r)dr (5.6)
Using equation (5.2), and the orthogonality condition (5.5), we then get
H
a
=

j
E
j
b

j
b
j
(5.7)
where E
j
are the internal atomic energy eigenvalues, and the operators b

j
and b
j
are the
creation and destruction operators of these electron internal states. Let us now discuss
some of the properties of these new operators. First, we can dene a vacuum state [0 >
for the electron in the atom, such that
[
j
>= b

j
[0 > , <
j
[ =< 0[b
j
, b
j
[0 >= 0 (5.8)
But, in contrast with the case of photons, only one electron can occupy this new state,
which means that, if we apply twice the creation operator b

j
to the electron vacuum state
[0 > , we cannot nd a new electron state, because of the Pauli exclusion principle. This
means that
b

j
b

j
[0 >= b

j
[
j
>= 0 b

j
b

j
= 0 (5.9)
103
This property of the operator b

j
can be generalized to
b

j
b

k
+ b

k
b

j
b

j
, b

k
= 0 (5.10)
where the operator , is called the anti-commutator, and applies to fermions. Other
anti-commutation relations are
b

j
, b
k
=
jk
, b
j
, b
k
= 0 (5.11)
Th quantum state of an atom with Z electrons. occupying the quantum states j
1
, j
2
, ...j
Z
,
can be described by
[
{j}
>= b

j1
b

j2
...b

jZ
[0 > (5.12)
This state vector satises an eigenvalue equation of the form
H[ >= E
tot
[ > (5.13)
where the total internal energy of the atom is given by the sum of the individual energy
levels E
tot
= E
j1
+ E
j2
+ ... + E
jZ
. The solutions of this equation satisfy orthonormality
conditions similar to those established for single electron states, such that
<
{j}
[
{k}
>=
j1,k1

j2,k2
...
jZ,kZ
(5.14)
An alternative way to represent the single electron Hamiltonian in an atom can be derived
from the eigenvalue equation
H
a
[i >= E
i
[i > (5.15)
By inserting the identity operator

j
[j >< j[ = I in this equation, we obtain for the
Hamiltonian
H
a
=

i
E
i
[i >< i[ =

i
E
i

ii
(5.16)
where we have used the atomic transition operator

ij
= [i >< j[ (5.17)
For the simple but important case of a two-level atom, the atomic Hamiltonian reduces
to
H
a
= E
1

11
+ E
2

22
(5.18)
104
Using the atomic transition frequency = E
2
E
1
, this can be rewritten as
H
a
=
1
2
(
22

11
) +
1
2
(E
1
+ E
2
)(
11
+
22
) (5.19)
Noting that, for the two-level atom, the quantity
11
+
22
=

j=1,2
= [j >< j[ = I is
the identity operator, we can see that the second term in this expression is a constant,
and can be ignored. On the other hand, we can use the notation
z
= (
22

11
), and
the atom Hamiltonian can be simply written as
H
a
=
1
2

z
(5.20)
5.2 Interaction Hamiltonian
If we restrict our discussion to the dipole approximation, we can write the interaction
Hamiltonian as
H
int
= er

E (5.21)
where r is the electron position inside the atom, and

E the radiation electric eld. We
have previously shown that this is equivalent to the alternative expression
H
int
=
e
m
p

A (5.22)
where p is the electron momentum, and

A is the vector potential. If we use twice the
identity operator I =

i
[i >< i[, we can represent the position operator in equation
(5.21) as
er = e

j
[i >< i[r[j >< j[ (5.23)
Using the atomic transition operator
ij
= [i >< j[, we can also write this expression as
er = e

j
< i[r[j >
ij
=

j
p
ij

ij
(5.24)
where we have used the electric dipole matrix elements
p
ij
= < i[er[j > (5.25)
105
Let us now turn to the electric eld

E. In the general case we have, for a plane wave
decomposition, we have for each photon mode

k, the operatior

E
k
= i
_

k
2
0
V
_
a
k
e
i(

kr
k
t)
a

k
e
i(

kr
k
t)
_
e
k
(5.26)
In the dipole approximation, the eld is assumed uniform and independent of the atom
position. We can then assume the atom at the centre of our reference frame, and take
r = 0 in this expression. Furthermore, if we include the remaining exponentials exp(it)
in the expressions of the operators a
k
and a

k
, we arrive at the eld operator

E
k
= iC
k
e
k
(a
k
a

k
) (5.27)
where C
k
=
_

k
/2
0
V , and e
k
is the unit polarization vector. By inserting equations
(5.24) and (5.27) in the expression of the interaction Hamiltonian, we obtain
H
int
= i

i,j

k
g
ij
k

ij
(a
k
a

k
) (5.28)
where we have introduced the new matrix elements
g
ij
k
=
1

( p
ij
e
k
)C
k
(5.29)
For a two level atom, and assuming that the unit polarization vector e
k
is real, we have
p
21
= p
12
, and using g
k
= g
12
k
= g
21
k
, we can write the interaction Hamiltonian in the form
H
int
= i

k
g
k
(
12
+
21
)(a
k
a

k
) (5.30)
It is now useful to introduce the notations

+

21
= [2 >< 1[ ,


12
= [1 >< 2[ (5.31)
and the above expression for H
int
becomes
H
int
= i

k
g
k
(
+
+

)(a
k
a

k
) (5.32)
It can be seen that the operator

= [1 >< 2[ takes an atom from its upper energy level


[2 >, and brings it into the lower level [1 >, wherehas
+
produces the opposite transition.
106
By looking at H
int
we can then identify four distinct processes. One, described by the term

+
a
k
corrsponds to the excitation of the atom and the destruction of one photon. The
term

k
represents the opposite process. The term

a
k
describes the de-excitation of
the atom and the destruction of one photon, resulting in an energy loss (for the total atom
+ eld system) of 2. Similarly, the forth tern,
+
a

k
corresponds to an energy gain
of 2. Obviously, these two last terms cannot satisfy energy conservation, and should
be dropped. The neglect of these two terms is equivalent to assuming the rotating wave
approximation discussed before. In this approximation, we can then write the interaction
Hamiltonian in the form
H
int
= i

k
g
k
(
+
a
k

k
) (5.33)
Finally, we can write the explicit expression for the total Hamiltonian of the atom plus
radiation eld, as given by equation (5.1), where H
a
=
z
/2, H
int
is determined by
(5.33), and the eld Hamiltonian is given by H
f
=

k

k
(a

k
a
k
+ 1/2).
Before proceeding further, it is useful to notice that the three operators

and
z
can
be identied with the Pauli matrices, as dened by

+
=
_
0 1
0 0
_
,

=
_
0 0
1 0
_
,
z
=
_
1 0
0 1
_
(5.34)
From these denitions we can easily calculate the following commutation relations
[

,
+
] =
z
, [

,
z
] = 2

(5.35)
5.3 Interaction with a single eld mode
We can simplify the picture of the atom radiation interaction by assuming that the two-
levels atom interacts with a single eld mode. We can then drop the sum over

k in
equation (5.32) and write the total Hamiltonian in the form
H = H
0
+ H
int
(5.36)
107
where the unperturbed atom plus eld terms are included in
H
0
=
k
_
a

a +
1
2
_
+
1
2

z
(5.37)
and the interaction term is reduced to
H
int
= ig(
+
a

) (5.38)
Let us consider the expicit description of the coupled system. We introduce the state
vector of the atom plus eld system [ >, which can be written as
[ >=

n
(C
1,n
[1, n > +C
2,n
[2, n >) (5.39)
where the state vectors [j, n > represent the system of photons in the eld mode and
the atom in the energy state j = 1, 2. The time evolution of [ > is described by the
Schroedinger equation
i

t
[ >= H
int
[ > (5.40)
The coecients C
j,n
will evolve in time due to the interaction. At this point, it is useful to
introduce the interaction picture. It corresponds to the use of an unitary transformation
described by
H

= U

(t)H
int
U(t) , U(t) = exp
_

H
0
t
_
(5.41)
Before explicitly performing such transformation, we notice that only the rst term of H
0
acts on the operators a and a

of H
int
, and only the second term acts on

. Using the
relation
e
A
Be
A
= B + [A, B] +

2
2
[A, [A, B]] + ... (5.42)
we can write
e
i
k
a

at
ae
i
k
a

at
= ae
i
k
t
(5.43)
Similarly, we have
e
i
k
zt

z
e
i
k
zt
=
+
e
it
(5.44)
From this we can derive the nal expression for the Hamiltonian in the interaction repre-
sentation, as
H

= ig
_

+
ae
it
+ a

e
it
_
(5.45)
108
with =
k
. Going back to the Schroedinger equation, we can replace equation
(5.39) in (5.40), and obtain
i

t
_

n
(C
1,n
[1, n > +C
2,n
[2, n >)
_
= H
int
_

(C
1,n
[1, n

> +C
2,n
[2, n

>)
_
(5.46)
By projecting this evolution equation for [ > in the direction of [2, n >, we get
i

t
C
2,n
=

(C
1,n
< 2, n[H
int
[1, n

> +C
2,n
< 2, n[H
int
[2, n

>) (5.47)
But the interaction Hamiltonian can only couple the mode [2, n > with the mode [1, n +
1 >. This means that we are reduced to
i

t
C
2,n
= C
1,n+1
< 2, n[H
int
[1, n + 1 > (5.48)
In the same way, we can also derive
i

t
C
1,n+1
= C
2,n
< 1, n + 1[H
int
[2, n > (5.49)
Using the interaction representation, which allows us to replace H
int
by its transformed
H

, as given by equation (5.45), we obtain

t
C
2,n
= ig

n + 1e
it
C
1,n+1
(5.50)
and

t
C
1,n+1
= ig

n + 1e
it
C
2,n
(5.51)
These equations are formally very similar to those obtained in the quasi-classical theory.
However, it will be shown below that they contain qualitative dierences, and they will
lead to new physical results.
Assuming that the atom is initially in the excited state [2 >, we can write C
2,n
(0)
C
n
(0) and C
1,n+1
(0) = 0. The solutions for these particular initial conditions are then
given by
C
2,n
(t) = C
n
(0)
_
cos(
n
t/2)
i

n
sin(
n
t/2)
_
e
it/2
(5.52)
109
and
C
1,n+1
(t) = C
n
(0)
2ig

n + 1

n
sin(
n
t/2)e
it/2
(5.53)
where we have used

2
n
= ()
2
+ 4g
2
(n + 1) (5.54)
The inversion of population D(t) can now be calculated as
D(t) =

n
[[C
2,n
(t)[
2
[C
1,n+1
(t)[
2
] (5.55)
Using equations (5.52) and (5.53), we obtain
D(t) =

n
[C
n
(0)[
2
_
()
2

2
n
+
4g
2
(n + 1)

2
n
cos(
n
t)
_
(5.56)
where [C
n
(0)[
2
is the probability to have an initial state with n photons. In the particular
case of absence of radiation at t = 0, we have [C
n
(0)[
2
=
n0
, and the above expression is
reduced to
D(t) =
1

2
0
_
()
2
+ 4g
2
cos(
0
t)

(5.57)
where
0
=
_
()
2
+ 4g
2
. The most interesting thing about this result is that it predicts
the existence of Rabi oscillations between the two energy levels of the atom, at the fre-
quency
0
, even in the absence of any initial radiation. This is a purely quantum vacuum
eect, which was not predicted by the quasi-classical theory. It results from spontaneous
de-excitation of the atom, thus conrming the phenomenological arguments rst advance
by Einstein.
In the case of a nite amount of photons, such that < n >,= 0 at t = 0, equation (5.57)
conrms the quasi-classical Rabi oscillations induced by a nearly resonant radiation eld.
But, apart from that, it also predicts the occurrence of a collapse of the Rabi oscillations,
followed by a much later revival. Again, the collapse-revival features are due to the
quantum character of the radiation eld, and were absent in the quasi-classical theory
where Rabi oscillations were purely sinusoidal. Assuming a coherent state [ >, the
period of the Rabi oscillations is determined by

R
=
1

n
=
1
_
()
2
+ 4g
2
< n >
(5.58)
110
where < n >= [[
2
. In contrast, the collapse time is approximately given by

c
=
1
2g
_
1 +
()
2
4g
2
< n >
_
1/2
(5.59)
Finally, the revival time can be estimated as

rev

2m

< n >
g
_
1 +
()
2
4g
2
< n >
_
1/2
(5.60)
or, simply

rev
4m

< n >
c
(5.61)
where m is an integer. This means that the revival will occur at regular times after the
collapse of the Rabi oscillations. This can be conrmed by a numerical representation of
the exact solution (5.56).
5.4 Spontaneous emission
Here we consider the decay process of an atom from the upper to the lower energy level,
by assuming that such a decay can take place over an innite number of radiation eld
modes, and not just one as considered in the previous section. As a result, the previous
Rabi oscillation at the frequency
0
will disappear and will be replaced by an irreversible
process. The interaction Hamiltonian for one atom in an innity of radiation eld modes
can be written, in the interaction picture, as
H

k
_
ig

k
(r
0
)a

k
e
it
+ h.c.
_
(5.62)
Here we have reintroduced the spatial dependence of the eld, in order to account for the
possible orientations of the wavevectors

k with respect to the atom, corresponding to the
same mode frequency
k
= kc. We have
g

k
(r
0
) = g

k
e
i

kr
0
(5.63)
111
And the state vector for the system of the atom plus radiation can be written as
[(t) >= C
2
(t) [2, n = 0 > +

k
C
1,

k
(t) [1, n = 1 >

k
(5.64)
where only the vacuum eld, corresponding to a number of photons n = 0, and the single
photon state n = 1 in any of the possible eld modes

k was retained. Replacing this in
the Schroedinger equation
i

t
[ >= H

[ > (5.65)
we obtain the evolution equations for the coecients
d
dt
C
2
= i

k
g

k
(r
0
)e
it
C
1,

k
(5.66)
and
d
dt
C
1,

k
= ig

k
(r
0
)e
it
C
2
(5.67)
Integrating this last equation, we obtain
C
1,

k
(t) = ig

k
(r
0
)
_
t
0
e
it

C
2
(t

)dt

(5.68)
And, replacing in the rst one, we get
d
dt
C
2
=

k
[g

k
(r
0
)[
2
_
t
0
e
i(tt

)
C
2
(t

)dt

(5.69)
In order to solve this integro-dierential equation we have to introduce a few simplifying
assumptions. First, we replace the summation over the eld modes

k by an integral, by
extending to innity the volume of integration. We have

k
2V
_
d

k
(2)
3
(5.70)
We can also use spherical variables in the wavevector space, such that
_
d

k =
_
2
0
d
_

0
sin d
_

0
k
2
dk (5.71)
Using the denition of g

k
, we can easily arrive at the expression
[g

k
(r
0
)[
2
=

k
2V
[p
12
[
2
cos
2
(5.72)
112
After integration over the solid angle d = sin
2
dd, we can then obtain
d
dt
C
2
=
[p
12
[
2
6
2

0
_
t
0
dt

_

0
k
2
dk
k
e
i(tt

)
C
2
(t

) (5.73)
We can now replace k =
k
/c, and integrate over
k
, which leads to
1
c
3
_

0

3
k
e
i(tt

)
d
k


3
c
3
_

0
e
i(
k
)(tt

)
d
k
= 2

3
c
3
(t t

) (5.74)
where we have assumed that
k
. Replacing this in equation (5.73) then leads to
d
dt
C
2
(t) =

2
C
2
(t) (5.75)
where the decay constant is given by
=
1

3
[p
12
[
2
3
2
c
3
(5.76)
We can then easily integrate the evolution equation for C
2
(t), and describe the evolution
of the upper energy level population in the absence of radiation, as

22
= [C
2
(t)[
2
= e
t
(5.77)
This result shows that an excited state of the atom, described by the state vector [2 >,
will spontaneously decay in time, in the absence of any surrounding photon, following an
exponential law. This demonstrates Einstein hypothesis on the existence of a spontaneous
decay process. Replacing this result in equation (5.68) we can also determine the nal
state of the radiation eld resulting from spontaneous decay. We get
C
1,

k
(t) = ig

k
(r
0
)
_
t
0
e
it

/2
dt

(5.78)
Integrating this for the same initial conditions, we obtain
C
1,

k
(t) = g

k
(r
0
)
1 e
itt/2
+ i/2
(5.79)
Replacing this in the expression for the state vector of the entire system, as given by
equation (5.64), we have
[(t) >= e
t/2
[2, n = 0 > +[1 > [(t) > (5.80)
113
where we have introduced the eld state vector as dened by
[(t) >=

k
g

k
e
i

kr
0
1 e
itt/2
( + i/2)
[n = 1 >

k
(5.81)
For very long times, we can use the asymptotic expression
[

>=

k
g

k
e
i

kr
0
( + i/2)
[n = 1 >

k
(5.82)
and the asymptotic state of the atom plus radiation eld system will be given by
[

>= [1 > [

> (5.83)
where we have obtained a complete decoupling between the atomic quantum state and
the radiation eld state. The energy initially contained in the excited atom is randomly
distributed over an innity of radiation eld modes. However, by placing the atom in a
very small cavity, where only a very small number of spatial modes exist for frequencies
close to the transition frequency , we can inhibit spontaneous emission and signicantly
reduce the value of the decay constant . This as been observed experimentally.
5.5 Modied refractive index
Resonant atomic transitions can be used to manipulate the refractive index of an optical
medium. This leads to very interesting new phenomena, such as the electromagnetic
induced transparency, or EIT. In order to understand such manipulating processes we
start by considering a three level atom, with two possible radiation transitions. The
corresponding atom Hamiltonian is
H
a
=

j
E
j
[j >< j[ , (j = 1, 2, 3) (5.84)
The two atomic transition frequencies are
1
= (E
3
E
1
)/ and
2
= (E
3
E
2
)/, where
E
3
is assumed as the upper level energy state, and radiation between the states E
2
and
114
E
1
is forbidden. Let us consider the resonant interaction of this atom with radiation. We
only need to retain two relevant eld modes, with a well dened direction of propagation,
and frequencies equal to
1
and
2
. As shown previously, the interaction Hamiltonian
can be written in the form
H
int
=

2
_

1
e
i
1
t
[3 >< 1[ +
2
e
i
2
t
[3 >< 2[
_
+ h.c. (5.85)
where
i
= [
i
[ exp(i
i
), for i = 1, 2, are the complex Rabi frequencies associates with
the two resonant couplings between the atom and radiation. We now have to solve the
Schroedinger equation
i

t
[ >= (H
a
+ H
int
)[ > (5.86)
We use the following type of solution
[ >=

j=1,2,3
C
j
(t)e
iE
j
t/
[j > (5.87)
where the quantities [C
j
(t)[
2
represent the probability for the atom to occupy the energy
eigenstate [j >. Replacing this in equation (5.86) and using the above expressions for the
two Hamiltonian operators H
a
and H
int
, we can easily obtain coupled evolution equations
for the coecients C
j
(t), which take the form
d
dt
C
1
=
i
2

1
C
3
,
d
dt
C
2
=
i
2

2
C
3
,
d
dt
C
3
=
i
2
(
1
C
1
+
2
C
2
) (5.88)
Let us consider the following initial conditions, corresponding to the case where only the
two lower energy levels are occupied, according to
[(0) >= C
1
(0)[1 > +C
2
(0)[2 > (5.89)
and C
3
(0) = 0. In order to solve the above coupled equations for these initial quantum
state, we write
d
2
dt
2
C
3
=
i
2
_

1
dC
1
dt
+
2
dC
2
dt
_
=
1
4
([
1
[
2
+[
2
[
2
)C
3
(5.90)
with similar equations for the other two coecients
d
2
dt
2
C
1
=
1
4
([
1
[
2
C
1
+

2
C
2
) (5.91)
115
and
d
2
dt
2
C
2
=
1
4
(
1

2
C
1
+[
2
[
2
C
2
) (5.92)
The desired solution for the upper level coecient is then
C
3
(t) =
i

sin
_

2
t
_

1
C
1
(0) +
2
C
2
(0) (5.93)
with
=
_
[
1
[
2
+[
2
[
2
(5.94)
For the other two coecients, we get
C
1
(t) =
1

2
__
[
1
[
2
cos
_

2
t
_
+[
2
[
2
_
C
1
(0) 2

2
sin
2
_

4
t
_
C
2
(0)
_
(5.95)
and
C
2
(t) =
1

2
_
2
1

2
sin
2
_

4
t
_
C
1
(0) +
_
[
2
[
2
cos
_

2
t
_
+[
1
[
2
_
C
2
(0)
_
(5.96)
Notice that the oscillating frequency (/2) is the geometric mean of the two distinct
Rabi frequencies [
1
[ and [
2
[. The most interesting aspect of these solutions is that
they predict the existence of a case where the upper energy level of the atom is never
populated, or in other words, where the coecient C
3
(0) is always equal to zero. From
equation (5.93) we see that this corresponds to the case when

1
=
2
, C
1
(0) = C
2
(0) (5.97)
This can be achieved if the two lower levels are very close to each other and are initially
equally populated, but in phase opposition. More specically, in the case where
C
j=1,2
= [C
j=1,2
[ e
i
j=1,2
, (
2

1
) = (5.98)
with
[C
1
(0)[ = [C
2
(0)[ =
1

2
(5.99)
In such special conditions, the solutions (5.93) - (5.96) will reduce to
[C
1
(t)[
2
= [C
2
(t)[
2
=
1
2
, [C
3
(t)[
2
= 0 (5.100)
116
As a result, the atom acts as if the upper energy level did not exist. This means that if
a laser beam is propagating in a medium with atoms prepared in such a particular initial
atomic energy state, the eld will not perceive any atom, because its state will not change.
As a consequence, the medium will appear as a perfectly transparent medium, as if it was
an empty space.
The question however remains on how to prepare the atoms in order to make them
transparent to radiation, or in other words, how to lead the atoms into the appropriate
initial quantum state. This can be achieved by using a pump radiation eld, which pre-
pares the atoms to become transparent to the second radiation eld (also called, the probe
eld). This is the principle of the electromagnetic induced transparence, or EIT, which
will be considered next. EIT is then described by slightly modifying the previous model,
assuming a strong pump eld with a frequency equal to one of the radiation transition
frequencies, let us say
2
= (E
3
E
2
)/, and a probe eld with a frequency nearly equal
to the other transition frequency =
1
+, where [[ is a small detuning. As well
will show, under certain conditions, the medium will become transparent to this probe
eld. Due to the existence of such a detuning, ,= 0, the interaction Hamiltonian of
equation (5.85) will slightly change, with
1
replaced by .
Our nal purpose here is to determine the polarization vector of the medium, and the
resulting refractive index. We have see that this polarization vector can be determined by
the o-diagonal matrix elements of the atoms in the medium. It is then useful to consider
the evolution equations for the density matrix elements
31
,
21
and
32
. They can be
written as
d
dt

31
= i
1

31

i
2

1
(
33

11
)e
it
+
i
2

21
e
i
2
t
(5.101)
d
dt

21
= i
2

21

i
2

23
e
it
+
i
2

31
e
i
2
t
(5.102)
and
d
dt

32
= i
2

32

i
2

2
(
33

22
)e
i
2
t
+
i
2

12
e
it
(5.103)
These equations can be simplied by introducing two new quantities

31
=
31
e
it
,
21
=
21
e
i(+
2
)t
(5.104)
117
The evolution equations for these new quantities take a much simpler form
d
dt

31
= i
31
+
i
2

1
+
i
2

2

21
(5.105)
and
d
dt

21
= i
21
+
i
2

2

31
(5.106)
This system of coupled equations can be written in a matrix form as
d
dt
= M +

(5.107)
where we have dened
M =
_
i
i

2
/2
i
2
/2
i
_
, =
_
i
31

21
_
,

=
_
i
1
/2
0
_
(5.108)
Integration of this equation for the initial conditions (0) = 0, or
31
(0) =
21
(0) = 0,
leads to the formal solution
(t) =
_
t
exp[M(t t

)]

dt

= M
1

(5.109)
Inverting the matrix M, we get
M
1
=
2i
[
2
[
2
+ 4
2
_
2

2
2
_
(5.110)
Replacing this in the formal solution, we obtain

31
(t) =

1
[
2
[
2
+ 4
2
e
it
(5.111)
We should notice that the complex Rabi frequency
1
is proportional to the amplitude
of the probe eld E(), as by denition
1
= p
31
E()/, where p
31
is the corresponding
dipole moment. We can now use the denition of the polarization vector, as

P() =
0
()

E() (5.112)
where () is the susceptibility of the medium, which determines the refractive index
n() =
_
1 + () (5.113)
118
We have previously seen that the susceptibility is related to the o-diagonal density matrix
elements, by the expression
() = 2N
a
p
13

31

0
E()
e
it
(5.114)
which leads to our nal expression
() =
N
a

0
[p
13
[
2

([
2
[
2
+
2
/4)
(5.115)
This result shows that, when 0, the susceptibility tends to zero, and the refractive
index tends to 1. This means that exact optical transparency is achieved in the presence
of a pump eld such that [
2
[
2
,= 0. This can be called an induced transparency eect,
because the presence of the pump eld is necessary to prepare the atomic quantum states,
which were assumed initially in the ground state,
11
(0) = 1, and
22
(0) =
33
(0) = 0. In
this way, a material medium is made to behave as if it was absent.
In our analysis we have neglected spontaneous decay processes, which would introduce
an imaginary part in the susceptibility, such that () =

() +i

(). In this case,

would be proportional to the spontaneous decay rate from [2 > to [1 >, which is negligible
for a dipole forbidden transition.
5.6 Quantum theory of the laser
We have shown in previous chapters that the main properties of the laser can be described
by a semi-classical theory, where light is described as a classical eld and only the atom
states are quantized. Such a theoretical approach is able to account for the laser threshold
condition, the laser growth rate instability and the steady state of the laser system. Such
a semi-classical model can also be extended to pulsed laser regimes. However, the semi-
classical laser theory ignores spontaneous emission and quantum correlation eects that
can be important near the laser threshold. It is then useful to consider a full quantum
description of the laser.
119
Let us rst consider the Hamiltonian operator pertinent for the laser system. As
noticed before, it has three terms corresponding to the radiation eld H
field
, the atom
H
a
and the interaction H
int
. In this model we have also to consider the atom coupling
with the non-lasing modes of the radiation, which are important for the spontaneous
emission. The atomic states can also be coupled with the environment, due to collisions
with electrons or with other atoms, an eventually with acoustic oscillations of the medium.
All these additional eects can be described by a thermal bath, and included in three
additional Hamiltonian terms, the bath Hamiltonian H
B
, and the atom-bath and eld-
bath interaction Hamiltonians, H
aB
and H
fieldB
. The total Hamiltonian in its nal form
will that be
H = H
field
+ H
a
+ H
int
+ (H
B
+ H
aB
+ H
fieldB
) (5.116)
If we are just considering one cavity mode, the eld Hamiltonian reduces to
H
field
= (a

a + 1/2) (5.117)
with the bosonic commutation relations
[a, a

] aa

a = 1 (5.118)
The Hamiltonian of a single two-atom level can take the form
H
a
= E
1
b

1
b
1
+ E
2
b

2
b
2
=

j=1,2
E
j
b

j
b
j
(5.119)
We can simplify this operator by taking the zero of the energy scale as coincident with
the energy of the lower level, which allows us to write E
1
= 0 and E
2
=
0
, where
0
is
the atomic transition frequency. This reduces the above operator to
H
a
=
0
b

2
b
2
(5.120)
where the fermionic anti-commutation relations hold
b
2
, b

2
b
2
b

2
+ b

2
b
2
= 1 (5.121)
Furthermore, we have previously shown that the atom-edl interaction can be written as
H
int
= g(b

1
b
2
+ b

2
b
1
)(a + a

) (5.122)
120
As we have seen, the term b

1
b
2
a

represents creation of a photon and destruction of an


electron state of the atom in the upper level [2 >, with creation of an electron state in
the lower level [1 >. The other terms were already discussed and have similar meanings.
Introduction of the rotating wave approximation then leads to the neglect of the two
terms that dont conserve energy. As a result, we have
H
int
= g(b

1
b
2
a

+ b

2
b
1
a) (5.123)
In a laser cavity, we have not just one atom, but a large number of atoms N, which implies
a generalization of the atoms and the interaction Hamiltonians, which become
H
a
=
0
N

i=1
b

2i
b
2i
(5.124)
and
H
int
= g
N

i=1
(b

1i
b
2i
a

+ b

2i
b
1i
a) (5.125)
Let us now consider the evolution equations for the eld operators. We know that, in the
Heisenberg picture, an arbitrary operator A evolves in time according to the equations
dA
dt
=
A
t
+
i

[H, A] (5.126)
If we take A as equal to the destruction eld operator a, and assuming that it is not
explicitly dependent on time, we have, for H = H
field
the evolution equation
da
dt
=
i

(H
field
a aH
field
) = i(a

aa aa

a) (5.127)
Using the commutation relation (5.118), this is reduced to
da
dt
= i
_
a

aa (1 + a

a)a
_
= ia (5.128)
Let us consider now the interaction of this radiation eld mode with a thermal bath. This
can be done by replacing in the above equation (5.128) H
field
by (H
field
+H
B
+H
fieldB
),
or
da
dt
=
i

[(H
field
+ H
B
+ H
fieldB
), a] (5.129)
121
As shown the previous section, this can be written as
da
dt
= ia
c
a + F(t) (5.130)
where
c
represents the damping rate of the eld mode in the cavity, and F(t) is an
operator representing a uctuating force described by the following statistical properties
< F(t) >=< F

(t) >= 0 , < F

(t)F(t

) >= 2
c
n
0
(tt

) , < F(t)F

(t

) >= 2
c
(n
0
+1)(tt

)
(5.131)
where n
0
is then mean photon number of the laser eld mode for a thermal bath with
temperature T. For lasers in the optical domain, and at room temperature, we can assume
that n
0
0.
Now, let us consider a similar coupling between the atoms and the thermal bath. For
this purpose, it is useful to introduce the atom dierence operator

d = b

2
b
2
b

1
b
1
(5.132)
The corresponding evolution equation is
d
dt

d =
i

[H
a
,

d] (5.133)
Using equation (5.124) it can easily be shown that
d
dt

d = 0 (5.134)
We also consider two other auxiliary operators
= b
1
+ b
2
,

= b

2
b
1
(5.135)
which evolve according to
d
dt
= i
0
,
d
dt

= i
0

(5.136)
We are now in position to consider our atom in a thermal bath. As we have seen, in the
semi-classical theory of the atom, that the atom occupation numers N
1
and N
2
could be
described by the following balance equations
dN
2
dt
= w
21
N
1
w
12
N
2
,
dN
1
dt
= w
21
N
1
+ w
12
N
2
(5.137)
122
where w
21
was the pumping transition rate, and w
12
the spontaneous decay rate, and
by denition, we had N
j
=< b

j
b
j
>. In the full quantum mechanical treatment, these
equations have to be replaced by the following equations for the atom population operator

N
j
b

j
b
j
, as determined by
d

N
2
dt
= w
21

N
1
w
12

N
2
+
2
(t) ,
d

N
1
dt
= w
21

N
1
+ w
12

N
2
+
1
(t) (5.138)
where
j
are new uctuation force operators, similar to F(t), and having similar statistical
properties
<
j
(t) >= 0 , <
j
(t)
k
(t

) >= G
jk
(t t

) (5.139)
where G
jk
is some correlation operator, to be specied. By substraction of equations
(5.138), we obtain the evolution equation for the dierence operator dened in equation
(5.132), as given by
d
dt

d = 2(w
21

N
1
w
12

N
2
) + (5.140)
where =
2

1
. This can also be written as
d
dt

d =

d
0


d) + (5.141)
where we have used

= w
12
+ w
21
,

d
0
=
w
21
w
12
w
12
+ w
21
(5.142)
Comparing this with the semi-classical laser model, we can see that

d
0
corresponds to the
unsaturated inversion operator, and that

= 1/ is the inverse of the relaxation time .


Similarly, we can rewrite the evolution equations for the operators and

, (5.136) in a
more complete form, as
d
dt
= i
0
+

(t) ,
d
dt

= i
0

+
+
(t) (5.143)
Finally, we consider the interaction Hamiltonian contribution to the evolution of the eld
mode operator a. Such a contribution can be written as
_
da
dt
_
int
=
i

[H
int
, a] = g[( a

a +

a), a] = ig (5.144)
123
Generalizing to N atoms, we obtain
_
da
dt
_
int
= ig
N

i=1

i
(5.145)
By adding this result to that of equation (5.130), we obtain the contribution of all the
terms of the total Hamiltonian H dened in (5.116), leading to the eld mode operator
equation
da
dt
= (i +
c
)a ig
N

i=1

i
+ F(t) (5.146)
Similarly, we can calculate the contribution of H
int
to the evolution of the atom operator
, by using
_
d
dt
_
int
=
i

[H
int
, ] = iga

d (5.147)
Generalizing this to any one of the N atoms, a adding the above contributions coming
from the other Hamiltonian terms, we obtain
d
j
dt
= (i
0
+ )
j
+ iga

d
j
+
j
(t) (5.148)
for j = 1, 2, ...N. We also can write
d

j
dt
= +(i
0
)

j
iga

d
j
+
j+
(t) (5.149)
It also follows, for the dierence operator
d

d
j
dt
=

d
0


d
j
) + 2ig(
j
a

j
a) +
jd
(t) (5.150)
By following a procedure similar to that used in the semi-classical laser theory, we can
eliminate the operators
j
from equation (5.148), and obtain a closed equation for the
eld operator a, in the form
da
dt
= (i G)a Ca

aa +

F(t) (5.151)
where we have dened the coupling C and gain G operators, as
C =
4g
4

2
N

d
0
, G =
c
+
g
2

N

d
0
(5.152)
124
and used the total uctuation force operator

F(t) = F(t) i
g

j=1

j
(t) (5.153)
Formally, equation (5.152) is very similar to that of the semi-classical theory, with two
main qualitative dierences. First, the eld mode amplitude is replaced here by the eld
mode operator a. Second, an additional uctuation term

F(t) is included, which has no
equivalent in the semi-classical approach. This new term results from the eld and atoms
interactions with the thermal bath. Therefore, the quantum laser equation takes the form
of an operator Langevin equation.
It is useful to discuss the qualitative features of the laser eld below the threshold.
In this case, we have G < 0. We also have a very small mean value for the eld number
operator a

a, which means that the nonlinear term Ca

aa can be neglected. The resulting


approximate equation, valid below the laser threshold, is
da
dt
= a +

F , = i[G[ (5.154)
This can be formally integrated to give
a(t) = a(0)e
it
+
_
t
0

F(t

)e
i(tt

)
dt

(5.155)
The rst term in this solution is the ballistic term, depending on the initial eld conditions,
and the second one is the force term due to the existence of uctuations. It is appropriate
to represent

F(t) as the sum of a succession of instantaneous perturbations, occurring at
random times t
k
, with random amplitudes and phases, as

F(t) =

k
f
k
e
i
k
(t t
k
) (5.156)
Inserting this in the above formal solution, and neglecting the ballistic term, which cor-
responds to choose a(0) = 0, we obtain
a(t) =

k
f
k
exp (i(t t
k
) [G[t + i
k
) (5.157)
125
This shows the existence of successive damped oscillations, starting at random intervals,
showing a kind of intermitent behavior. This solution also shows that the mean value
of the eld operator is equal to zero, < a(t) >= 0. Let us now look at its coherence
properties, by considering
< a

(t)a(t

) >= e
i(tt

)
< a

(t

)a(t

) >= e
i(tt

)
< n(t

) > (5.158)
where < n(t) > is the average photon number. Its explicit calculation is not performed
here, but can be written in the form
< n >=

c
[G[
(n
th
+ n
sp
) (5.159)
where n
th
represents the number of photons of the laser eld mode with frequacy , as it
would be at thermal equilibrium at temperature T, and n
sp
is the number of spontaneously
emitted photons.
When inversion of population is exactly equal to zero, we have Nd
0
= 0, and we
have [G[ =
c
. In this case, the eect of stimulated emission and absorption exactly
compensate, and we have < n >= n
th
+ n
sp
. For positive inversion, we have [G[ <
c
,
and the number of photons increases, < n >> n
th
+ n
sp
, even below the laser threshold.
Finally, at threshold, [G[ 0 as the amplication factor
c
/[G[ tends to innity, and the
number of photons becomes much larger than n
th
+n
sp
. However, divergence of < n > is
prevented by the nonlinear term Ca

aa, which can no longer be neglected. As a result, the


mean number of photons will remain nite at the threshold, as it should. Finally, above
threshold, the eld coherence properties will dramatically change, with the uncorrelated
damped signal replaced by a strongly correlated eld with only a negligible uctuation
level.
5.7 Problems
P5.1 - Establish the matrices (5.34), and demonstrate the commutation relations
(5.35). Determine the matrices
x
and
y
, such that

=
1
2
(
x
i
y
).
126
P5.2 - Rewrite the interaction Hamiltonian (5.38), by using the creation and destruc-
tion operatores b

and b, introduced in equation (5.7).


P5.3 - Derive the evolution equations (5.51), and establish the corresponding solutions
(5.52)-(5.53).
P5.4 - Derive the coupled system of equations (5.88) which describe a three level atom
model with two allowed radiative transitions.
P5.5 - Demonstrate equations (5.128) and (5.147).
Chapter 6
Quantum Coherence
6.1 Field Correlations
In order to understand the mean qualitative dierences between the classical and the
quantum theory of radiation, we discuss here the problems of eld correlations. We start
with the simplest case of amplitude correlations, using as a typical example, the famous
Young experiment. In such an experiment, radiation emitted from a single source is made
to pass through two distinct pinholes, and is made to interfere at a plane screen. The
total electric eld measured at a given position r on the screen is the sum of the elds
coming from the two pinholes, placed at positions r
1
and r
2
. Assuming that the pinholes
are at distances d
1
and d
2
from the observation point, we can write for the observed eld
on the screen

E(r, t) = u
1

E(r
1
, t
1
) + u
2

E(r
2
, t
2
) (6.1)
where the coecients u
j
depend on the geometric conguration and typically are inversely
proportional to the distances d
j
, and t
j
are retardation times with respect to the sources,
as dened by
t
j
= t
d
j
c
, d
j
= [r r
j
[ (6.2)
127
128
with j = 1, 2. The observed radiation intensity, averaged over one cycle of the eld
oscillation, is given by
I(r, t) =

0
2
c[

E(r, t)[
2
= I
1
(r, t) + I
2
(r, t) +
0
c(u

1
u
2
)'
_

(r
1
, t
1
)

E(r
2
, t
2
)
_
(6.3)
where I
j
is the value of the intensity received at the observation point at
I
j
(r, t) =

0
2
c[u
j
[
2
[

E(r
j
, t
j
)[
2
(6.4)
In a steady state experiment, we integrate the intensity over a large period of time T,
much larger than the wave period. The time integrated mean intensity, will be determined
by
I(r) =< I
1
(r, t) > + < I
2
(

(r), t) > +
0
c(u

1
u
2
)'
_
<

E

(r
1
, t
1
)

E(r
2
, t
2
) >
_
(6.5)
We see here the appearance of the eld correlation function, which can be generally dened
as
<

E

(r
1
, t
1
)

E(r
2
, t
2
) >=
1
T
_
T
0

(r
1
, t
1
)

E(r
2
, t
1
+ )dt
1
(6.6)
with = t
2
t
1
. For the simple case of a purely monochromatic radiation source at
frequency , we can use

E(r, t) =

E
0
cos(

k r t) (6.7)
This allows us to write
<

E

(r
1
, t
1
)

E(r
2
, t
2
) >=
[E
0
[
2
4
exp
_
i

k (r r
1
) i

k (r r
2
)
_
+c.c. =
[E
0
[
2
2
cos[k(d
1
d
2
)]
(6.8)
from where we get, assuming equal intensities I
1
= I
2
= I
0
, the well known expression
I(r) = 2I
0
1 + cos[k(d
1
d
2
)] = 2I
0
1 + cos (6.9)
where = t
1
t
2
= (d
1
d
2
)/c is the retardation time. In a more general case, each eld
will have a phase
j
(t), and the eld correlation function will be given by
<

E

(r
1
, t
1
)

E(r
2
, t
2
) >=
[E
0
[
2
2
< e
i(t
1
)
e
i(t
1
)
> cos() (6.10)
129
The phase correlation function appearing in this new expression can be characterized by
a correlation time
c
, such that
< e
i(t
1
)
e
i(t
1
)
>= exp
_

[[

c
_
(6.11)
For complete a incoherent phase we have
c
0, and the interference fringes in the Young
experiment will disappear
<

E

(r
1
, t
1
)

E(r
2
, t
2
) >= 0 (6.12)
At this point, it is convenient to dene the rst-order coherence degree, as
g
(1)
(r
1
, t
1
; r
2
, t
2
) =
<

E

(r
1
, t
1
)

E(r
2
, t
2
) >
[< [E(r
1
, t
1
)[
2
>< [E(r
2
, t
2
)[
2
>]
1/2
(6.13)
We can see that, for the generic case considered above, we have
g
(1)
(r
1
, t
1
; r
2
, t
2
) = cos() exp
_

[[

c
_
(6.14)
where we have
c
= 0 for completely incoherent light,
c
for completely coherent
light, and
c
nite for partially coherent light.
Until now we have discussed the eld correlations in purely classical terms. Let us
see how we can describe the Young experiment and eld amplitude correlations from the
point of view of the quantum theory of radiation. For this purpose, we write the electric
eld operator as

E(r, t) =

E
(+)
(r, t) +

E
()
(r, t) (6.15)
where we have

E
(+)
(r, t) =

k
E
k
e

k
a

k
e
i

kri
k
t
(6.16)
and

E
()
(r, t) =

k
E
k
e

k
a

k
e
i

kr+i
k
t
(6.17)
We can then dene, in analogy with the above classical discussion of the eld intensity, a
quantum expression of the mean intensity
I(r) < I(r, t) >=

0
2
c <

E
()
(r, t)

E
(+)
(r, t) > (6.18)
130
The corresponding quantum rst-order degree of coherence will be
g
(1)
(r, t; ) =
<

E
()
(r, t)

E
(+)
(r, t + >
_
<

E
()
(r, t)

E
(+)
(r, t) ><

E
()
(r, t) + )

E
(+)
(r, t + ) >
_
1/2
(6.19)
For the case of a single eld mode, this reduces to
g
(1)
() =
< a

(t)a(t + ) >
< a

a >
(6.20)
Let us now consider the specic case of the Young experiment. The measured mean
intensity at a given position on the screen, is given by
I(r) = I
1
+ I
1
+
0
c '
_
<

E
()
(r
1
, t
1
)

E
(+)
(r
2
, t
2
) >
_
(6.21)
with
I
j
=

0
2
c <

E
()
(r
j
, t
j
)

E
(+)
(r
j
, t
j
) > (6.22)
for j = 1, 2. Let us assume that we detect on the screen only one single eld mode from
each of the two pinholes. We can dene the following operators
a =
1

2
(a
1
+ a
2
) , a

=
1

2
(a

1
+ a

2
) (6.23)
where the normalization factor is used to guarantee that [a, a

] = 1. We also notice that


[a
1
, a
2
] = [a
1
, a

2
] = [a

1
, a
2
] = 0, because these operators correspond to two dierent eld
modes. In particular, they can have the same frequency , but two dierent wavevectors,

k
1
and

k
2
. The quantum state corresponding to n photons at the detector can then be
dened by the operator a

dened above, acting on the total vacuum state


[n >=
(a

)
n

n!
[0 >
1
[0 >
2
=
(a

)
n

n!
[0 > (6.24)
If the pinhole number 2 is absent (or obstructed) the mean number of photons detected
on the screen associated with this eld state n is
< n
1
>< n[a

1
a
1
[n >=
1
n!
< 0[(a)
n
a

1
a
1
(a

)
n
[0 > (6.25)
Using the denition (6.23), we can then show that
< n
1
>=
n
2
(6.26)
131
Similarly, if the pinhole number 1 is absent, we also get
< n
2
>< n[a

2
a
2
[n >=
n
2
(6.27)
Finally, in the presence of both pinholes, we have to use the total eld operators, or the
form
E
(+)
(r, t) a
1
e
ikd
1
+ a
2
e
ikd
2
, E
()
(r, t) a

1
e
ikd
1
+ a

2
e
ikd
2
(6.28)
The resulting mean intensity is then
I(r) =

0
2
c
_
< n[a

1
a
1
[n > + < n[a

2
a
2
[n > +2 cos[k(d
1
d
2
)] < n[a

1
a
2
[n >
_
(6.29)
where we obtain, for the interference term, the following value
< n[a

1
a
2
[n >=
1
2n!
< 0[(a
1
+ a
2
)
n
a

1
a
2
(a

1
+ a

2
)
n
[0 >=
n
2
(6.30)
This shows that the classical result of equation (6.9) for the interference fringes of the
intensity on the screen is exactly recovered here using the quantum eld description of
the Young experiment.
6.2 Intensity Correlations
In order to introduce the intensity correlations, let us consider another famous experiment,
rst performed by Hanbury-Brown and Twiss, where a photon beam originating from a
single source is divided by a semi-transparebt plate in two secondary beams, which are
detected by two photo-detectors. The radiation intensity measured at each detector is
I
1
(t) and I
2
(t), and we can dene intensity correlations as
< I
1
(t)I
2
(t) >=

0
2
c < E

1
(t)E

2
(t + )E
1
(t)E
2
(t + ) > (6.31)
This leads us to the denition of the second-order degree of coherence, as
g
(2)
(r
1
, t
1
; r
2
, t
2
) =
< E

(r
1
, t
1
)E

(r
2
, t
2
)E(r
1
, t
1
)E(r
2
, t
2
) >
< [E(r
1
, t
1
)[
2
>< [E(r
2
, t
2
)[
2
>
(6.32)
132
Let us consider the special case of an intensity measured at a single position r. For steady
state of radiation g
(2)
will only depend on the diference between the two instants t
1
and
t
2
, and not on their value. Its expression can then be reduced to
g
(2)
() =
< I(t)I(t + ) >
< I >
2
=
< E

(t)E

(t + )E(t)E(t + ) >
< E

(t)E(t) >
2
(6.33)
It is also well known, from the denition of intensity that the following inequality holds
I
2
(t
1
) + I
2
(t
2
) 2I(t
1
)I(t
2
) (6.34)
This means that
< I
2
(t) > < I(t) >
2
(6.35)
which implies the following property for the second-order degree of coherence
g
(2)
(0) 1 (6.36)
Let us now turn to the quantum theory. The quantum equivalent to the denition (6.32)
can be written as
g
(2)
(r
1
, t
1
; r
2
, t
2
) =
< E
()
(r
1
, t
1
)E
()
(r
2
, t
2
)E
(+)
(r
1
, t
1
)E
(+)
(r
2
, t
2
) >
< E
()
(r
1
, t
1
)E
(+)
(r
1
, t
1
) >< E
()
(r
2
, t
2
)E
(+)
(r
2
, t
2
) >
(6.37)
In order to illustrate this quantum denition, let us consider the case of intensity corre-
lations involving a single eld mode. We have
g
(2)
() =
< a

(t)a

(t + )a(t)a(t + ) >
< a

(t)a(t) >
2
(6.38)
Using the commutator [a, a

] = 1, we can rewrite this expression as


g
(2)
(0) =
< a

(aa

1)a >
< a

a >
2
=
< a

aa

a a

a >
< a

a >
2
(6.39)
Using the photon number operator N = a

a, this can be replaced by a more suggestive


result
g
(2)
(0) =
< N(N 1) >
< N >
2
=
< N
2
> < N >
< N >
2
(6.40)
Let us apply this to a simple non-classicak eld state, corresponding to the Fock state
[n >: The result is
g
(2)
(0) =
n 1
n
(6.41)
133
We see that, in this case, we have g
(2)
(0) = 0, for the special cases of n = 0 and n = 1. On
the other hand, for a large number of photons, n , it will tend to one. This shows that,
in contrast with the classical denition, the quantum second-order degree of coherence
can take values in the classically forbidden range, between zero and one, 0 g
(2)
(0) 1.
This also shows, once more, that the Fock state has no classical equivalent.
Let us now go back to the Hanbury-Brown Twiss experiment, where we have two
photon detectors. The associated second-order coherence denition will be
g
(2)
() =
< a

1
(t)a

2
(t + )a
1
(t)a
2
(t + ) >
< a

1
(t)a
(
t) >< a

2
(t)a
2
(t)
=
N
1
(0)N
2
() >
< N
1
>< N
2
>
(6.42)
If the source of radiation in this experiment is in a two-mode state [n >, as dened
by equation (6.24), and if the two arms in the experiment are symmetric, we will have
< N
1
>=< N
2
>= n/2, as shown before. We also have
< N
1
N
2
>=
1
4
n(n 1) (6.43)
which leads to the following second-order coherence result
g
(2)
(0) =
n(n 1)
n
= 1
1
n
(6.44)
which again can take values in the non-classical range between zero and one. The condition
g
(2)
(0) < 1, leads to photon anti-bunching eects, which are peculiar of quantum states of
light. In the range of g
(2)
(0) > 1, we will have photon bunching, which as characteristic
of classical eld states.
Quantum anti-bunching is therefore a specic property. It can be more clearly un-
derstood if we consider the existence of very small values of g
(2)
(0), in general, implies
that
g
(2)
() g
(2)
(0) (6.45)
This means that the probability for the arrival of two photons at the detector separated
by some nite interval > 0, is larger than the probability for their simultaneous arrival.
As a result, the successive photons tend to be detected separately, as if their repelled each
134
other. They tend to avoid bunching. In contrast, as we have seen, in a classical eld
state, we tend to have
g
(2)
() g
(2)
(0) (6.46)
which means that we have the opposite situation of bunching at the detector. Classical
sources of radiation (such as a lamp) tend to show variations of intensity, resulting from
this bunching eect. In the case of a laser, the photons tend to anti-bunch, therefore
leading a reduction of intensity uctuations.
It can be shown that for a thermal source of radiation, we have g
(2)
(0) = 2, which
is described by a sub-Poissonian distribution. A coherent state, [ > leads to a value of
g
(2)
(0) = 1, and is described by a Poissonian distribution, as we have shown. Finally, for
a Fock state, we have g
(2)
(0) < 1, which corresponds to a super-Poissonian distribution.
6.3 Squeezed States
It is well known from elementary Quantum Mechanics that, given two Hermitian opera-
tors, A and B, having a commutation relation of the form
[A, B] = iC (6.47)
their mean values cannot be determined exactly at the same time, because they will
necessarily obey the uncertainty principle
A B
1
2
[ < C > [ (6.48)
We dene squeezed state any quantum state where the uncertainty on one of these oper-
ators is less then half of the expected minimum uncertainty, or
(A)
2
<
1
2
[ < C > [ (6.49)
This means that, in a squeezed state, one of the observables has a reduced uncertainty,
and the other one as an increased uncertainty, with respect to the expect minimum value
(A)
2
(B)
2

1
2
[ < C > [ (6.50)
135
As an example, let us consider a given quantum state of the electromagnetic eld, corre-
sponding to a single mode with frequency . The electric eld operator is

E(t) = Ee
_
iae
it
ia

e
it
_
(6.51)
where the operators a and a

satisfy the commutation relation [a, a

] = 1. We know that
they are related with generalized position and momentum operators, q and p, as shown
by equations (????). Here it is useful to consider a simplied version of these operators,
such that
q =
1
2
(a + a

) , p =
1
2i
(a a

) (6.52)
Conversely, we can write
a = q + ip , a

= q ip (6.53)
The commutation relation for the position and momentum operators is
[q, p] =
i
2
(6.54)
The electric eld (6.51) can also be written in terms of these operators as

E(t) = 2Ee [q sin(t) p cos(t)] (6.55)


They are usually called the quadratures of the eld, as they represent the sine and cosine
parts of the eld operator. Their phase dierence is then equal to /2. From the commu-
tator (6.54) we can then derive an uncertainty relation for the mean values of these two
operators, as
q p
1
4
(6.56)
We will therefore have a squeezed state of the radiation eld if we verify any one of these
conditions
(q)
2
<
1
4
, (p)
2
<
1
4
(6.57)
We know from equation (6.56) that they cannot be veried simultaneously. If, in addition
to one of these two conditions, we have
q p =
1
4
(6.58)
136
the squeezed state will be called ideal, because it will also minimize the uncertainty
relation. In order to illustrate these concepts, let us apply then to the Fock states [n >
of the radiation eld mode. We can write
< n[q[n >=
1
2
< n[a[n > +
1
2
< n[a

[n >= 0 (6.59)
and
< n[q
2
[n >=
1
2
< n[(a
2
+ aa

+ a

a + a
2
)[n >=
1
4
(2n + 1) (6.60)
From here we can calculate the mean square deviation, as
(q)
2
=< n[q
2
[n > (< n[q[n >)
2
=
1
4
(2n + 1) (6.61)
The same result could be derived for the other eld quadrature, (p)
2
. We conclude from
here that, even for vacuum, n = 0, the Fock states will never be squeezed, and will never
satisfy any of the two conditions (??). In a sinilar way, we can consider a coherent state
of the same eld mode, [ >. Explicit calculation of the quantities
< [q[ >=
1
2
< [(a + a

)[ > , < [q
2
[ >=
1
4
< [(a + a

)
2
[ > (6.62)
will lead to the following result
(q)
2
=< [q
2
[ > (< [q[ >)
2
=
1
4
(6.63)
with the same value for the other quadrature. This shows that the coherent state is not
a squeezed state, although it reduces the eld uncertainties to their minimum possible
values. Let us now introduce the squeezed operator, as dened by the expression
S() = exp
_
1
2

a
2

1
2
a
2
_
(6.64)
where is a complex number, = r exp(i), where its amplitude r is usually called the
sqeezing parameter. It ca easily be show that this is an unitary operator, satisfying
S
1
() S() = I (6.65)
where I is the identity operator. We can also verify that
S
1
() = S

() = S() (6.66)
137
Let us now apply this unitary squeezing operator to the creation and destruction operators
a and a

. We will get
S

()aS() = a cosh r a

e
i
sinh r (6.67)
and
S

()a

S() = a

cosh r ae
i
sinh r (6.68)
Now, if we rotate the real and imaginary axis of the complex (q, ip) plane by an angle ,
we will have a new destruction operator, dened the new quadrature operators
a

(q

+ ip

) = (q + ip)e
i
= ae
i
(6.69)
By applying the squeezed transformation to this new operator, we obtain
S

()a

S() S

()(q

+ ip

)S() = qe
r
+ ipe
r
(6.70)
We can see that the resulting new eld quadratures are
Q = q

e
r
, P = p

e
r
(6.71)
For r > 0, this shows a reduction of the eld amplitude along the new real axis , and an
increase along the imaginary axis. We are now ready to construct a squeezed coherent
state, starting from vacuum, by successively applying the displacement and the squeezing
operators, D() and S() to the vacuum state [0 >. The resulting state is
[, >= S()D()[0 > (6.72)
It can be shown that, for such a state we have
< a >=< , [a[, >= cosh r

e
i
sinh r (6.73)
After some calculations, we can also get the uncertainty associated with the two new eld
quadratures, as
(Q)
2
=< Q
2
> (< Q >)
2
=
1
4
e
2r
, (p)
2
=
1
4
e
2r
(6.74)
We also have
Q P =
1
4
(6.75)
This shows that the squeezed coherent state [, > dened above is an ideal squeezed
state, where the degree of squeezing is determined by the squeezing parameter r = [.
138
6.4 Photon entanglement
One of the most important and intriguing features of a quantum system is its non-locality.
In order to discuss this problem, and its relevance to the quantum theory of radiation,
let us consider a two-component system, of two spin 1/2 particles. If they were two
independent particles, the total quantum state of the system would simply be [1, 2 >=
[1 > [2 >, which would mean that the state of each particle could me measured without
perturbing the state of the other one. However, in Quantum Mechanics, the total state
vector of a system is usually not given by such a simple expression. For instance, if the
system has total spin equal to zero, S = 0, ts state vector can be represented by the
anti-symmetric form
[ > [1, 2 >=
1

2
([1, + > [2, > [1, > [2, + >) (6.76)
where the signs represent states with spin 1/2 along some given direction. In such
case, the spin state of each of the individual particles is not independent from the state
of the other. This means their quantum states are entangled. Such an entanglement is
associated with non-locality. Even if the two particles in the system move far apart from
each other, their quantum states will stay interconnected, and any measurement or action
made on one of them will instantaneously inuence the other one.
In order to understand this problem of entanglement and non-locality of the quantum
state of two particles, let us dene S
a
(1) as the spin of particle 1 along the direction of
some arbitrary unit vector a. Formally, it can be dened as
S
a
(1) =

S(1) a =

2
(1) a (6.77)
where the components of are the Pauli matrices

x
=
_
0 1
1 0
_
,
y
=
_
0 i
i 0
_
,
z
=
_
1 0
0 1
_
(6.78)
It can be seen from here that the value of S
a
(1) can only be /2, or /2. On the other
hand, by denition, we can say that its mean value is determined by
< S
a
(1) >=< 1, 2[S
a
(1)[1, 2 >= 0 (6.79)
139
Such a result is obvious, because the positive and negative values of the spin are equally
probable. If we now measure simultaneously the spin of particle 1 along the direction
a and the spin of particle 2 along another direction

b, the product of the two measured
spins will be
S
a
(1)S
b
(2) =

2
4
((1) a)
_
(2)

b
_
(6.80)
The mean value of this new operator is given by
P(a,

b) < S
a
(1)S
b
(2) >=

2
4
< 1, 2[ ((1) a)
_
(2)

b
_
[1, 2 > (6.81)
Using equation (6.76), this leads to
P(a,

b) =

2
4
_
a

b
_
=

2
4
cos (6.82)
where is the angle between the vectors a and

b. In the particular case of the two spin
measurements being made along the same direction, a =

b, this probability reduces to
P(a,

b) =
2
/4, which corresponds to the existence of two opposite spin values, one
equal to +/2 and the other to /2, as expected.
Let us now consider a third direction of observation c. IIt is useful o compare the values
of the three probabilities P(a,

b), P(a, c) and P(

b, c). In the particular case where the


three vectors are in the same plane, with angles
a,b
,
a,c
and
b,c
, such that
a,b
=
b,c
=
and
a,c
= 2, we establish the following relations
P(a,

b) = P(

b, c) =

2
4
cos , P(a, c) =

2
4
cos 2 (6.83)
from where we obtain
[P(a,

b) P(a, c) =

2
4
[ cos cos 2[ (6.84)
This quantity can become larger than
2
/4, and can even verify the condition
[P(a,

b) P(a, c) >
_

2
4
+ P(

b, c
_
(6.85)
Such a condition is equivalent to
[ cos cos 2[ > 1 cos (6.86)
140
Take the particular case of = /3. We have cos = 1/2 and the above expression
becomes 1 > 1/2. This means that the inequality (6.85) can be eventually be satised in
the frame of Quantum Mechanics.
Let us now consider the possible existence of hidden variables in the system, as a
consequence of some hypothetical incompleteness of the standard quantum description.
In this case, the spin component S
a
(1) along the axis a will be replaced by a function
(/2)A(a, ), where the variable represents the hidden parameter. Similarly, S
b
(2) will
be replaced by (/2)B(

b, ). But, because the total spin of the system is assumed to be


zero, we will always have
A(a, ) = B(

b, ) (6.87)
On the other hand, we can assign a certain probability p() for the hidden variable to
take a precise value . Such a probability necessarily veries the normalization and non-
negative conditions
_
p()d = 1 , p() 0 (6.88)
By taking the average of the product of the two quantities which replaced S
a
(1) and S
b
(2)
over all the possible values of the hidden variable , we will obtain the joint probability
P
hidd
(a,

b) =

2
4
_
p()A(a, )B(

b, )d (6.89)
In the particular case of a =

b, this would reduce to


P
hidd
(a, a) =

2
4
(6.90)
as it should. Now, let us consider, as previously, the dierence between joint probabilities
associated with dierent choices of the axis over which we measure the spin of the two
particles in the system. Using this new denition, we have
P
hidd
(a,

b) P
hidd
(a, c) =

2
4
_
p()A(a, )
_
B(

b, ) B(c, )
_
d (6.91)
Using the property (6.87), we can also write
P
hidd
(a,

b) P
hidd
(a, c) =

2
4
_
p()A(a, )A(

b, )
_
1 + A(

b, )B(c, )
_
d (6.92)
141
Noting that p() 0 and that the quantities A can only take the values of 1, we can
nally write

P
hidd
(a,

b) P
hidd
(a, c)

<
_

2
4
+ P
hidd
(

b, c
_
(6.93)
This is called the Bell inequality. Comparing this results with the quantum mechanical
result (6.85), we conclude that the Bell inequality is violated by Quantum Mechanics.
Or in other words, standard quantum theory is not compatible with the existence of any
hidden variable.
Violation of Bells inequality was experimentally demonstrated by aspect in 1982, using
photons where spin states are replaced by photon polarization states. In his experiments,
a two photon radiative transition between two levels of the Calcium atom, 4p
2 1
S
0
and
4s
2 1
S
0
, were measured with two rotating polarizers. These atomic quantum levels have
a zero total angular momentum, which implies that the total angular momentum of the
two photons is also zero. These and other similar experiments agree with the quantum
mechanical expression (6.85) for joint probability measurements, and demonstrated that
there are no hidden variables. As a consequence, quantum entanglement at a distance
can persist between two particles of the same system, even if their quantum states are
measured very far apart.
142
Bibliography
[1] L. Allen and J.H. Eberly, Optical resonance and Two-level Atoms, John Wiley and
Sons, New York (1975).
[2] B.H. Bransden e C.J. Joachain, Quantum Mechanics, Prentice Hall, 2 ed. (2000).
[3] C. Cohen-Tannoudji, J. Dupont-Roc and G. Grynberg, Photons and Atoms: Intro-
duction to Quantum Electrodynamics, John Wiley and Sons, New York (1989).
[4] C. Cohen-Tannoudji, J. Dupont-Roc and G. Grynberg, Atom-Photon Interactions,
John Wiley and Sons, New York (1992).
[5] H. Haken, Light, vols I and II, North-Holland, Amsterdam (1981).
[6] U. Leonhardt, Measuring the Quantum State of Light, Cambridge University Press
(1995).
[7] R. Loudon, The Quantum Theory of Radiation, Noth-Holland, Amsterdam (1958).
[8] W.H. Louisell, Quantum Statistical Properties of Radiation, John Wiley and Sons,
New York (1973).
[9] L. Mandel and E. Wolf, Optical Coherence and Quantum Optics, Cambridge Uni-
versity Press (1995).
[10] P. Meystre and M. Sargent III, Elements of Quantum Optics, Springer-Verlag, Berlin
(1990).
[11] P.W. Milonni and J.H. Eberly, Lasers, John Wiley and Sons, New York (1988).
143
144
[12] B.E.A. Saleh and M.C. Teich, Fundamentals of Photonics, Wiley-Interscience, New
York (1991).
[13] M. Sargent III, M. Scully and W.E. Lamb Jr., Laser Physics, Addison-Wesley, Read-
ing MA (1974).
[14] M.O. Scully and S. Zubairy, Quantum Optics, Cambridge University Press (1997).
[15] A.E. Siegman, Lasers, Univ. Sci. Books, Mill Valley CA (1986).
[16] D.F. Walls and G.J. Milburn, Quantum Optics, Springer-Verlag, Berlin (1994).
[17] J. Wilson e J.F.B. Hawkes, Lasers: Principles and Applications, Prentice Hall (1987).

You might also like