Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Chapter 1

Introduction
The field of quantum optics deals with light and its interaction with matter in a funda-
mentally ”quantum” way. Light is treated as photons, and here matter consists of atoms
whose electrons absorb and emit photons to transition between discrete energy levels. These
relatively simple interactions lead to rich and diverse phenomena that can give insight into
a wider variety of physical systems and have applications on small and large scales. For
example, the principles of quantum optics can be used to create extremely precise force and
position sensors [1][2] and study the crossover into quantum behavior of macroscopic objects
[3][4], and they have applications in quantum computing [5].
We will look at several system configurations that are all built around a cavity created
between two reflective mirrors. Photons reflect off the mirrors and can be contained in the
cavity, although they occasionally ”leak” out. From here, the system can be modified in
many ways to elicit new behavior. For example, a two-level atom can be added to the cav-
ity. The interaction between an atom and photons in a cavity is known as cavity quantum
electrodynamics (cQED). Alternatively, one of the mirrors could be allowed to oscillate as
if on a spring. The study of photons in a cavity interacting with this mirror, a macroscopic
mechanical object, is called cavity optomechanics. Of course, we can include an atom to-
gether with the moving mirror, which has become known as hybrid optomechanics [6]. A
schematic of this hybrid system is shown in Fig. 1.1 [4][7]. These relatively simple hybrid
systems can be used to demonstrate seemingly endless interesting behavior including many
non-classical effects, such as photon anti-bunching and squeezed light [8][9][10].
Over the last three years, we have taken a somewhat exploratory approach to various
cavity systems. In this thesis, we look at some of the outcomes of this approach. In the rest
of this chapter, we describe the cavity, cavity-atom, cavity-mirror, and cavity-atom-mirror
systems in more detail while also explaining the important characteristics and parameters
that are used throughout the paper. In Chapter 2, we explain some of the methods in
quantum theory that are used to calculate the quantities of interest, and describe how these
are implemented in the Quantum Toolbox in Python (QuTiP), which we used for numerical

1
simulations. In Chapter 3, we explain a method for reducing the computational scope of
these simulations for systems with large numbers of photons and phonons. In Chapter 4,
we explain a filter-cavity method which can be used to calculate time-dependent spectra
of cavity systems. Finally, in Chapter 5, we use a cavity system to model the dynamical
Casimir effect, which results in the creation of photon pairs out of a quantum vacuum state.

Figure 1.1: A hybrid optomechanical system with a cavity, two-level atom, and movable
mirror. The labeled parameters are explained in subsequent sections.

1.1 Cavity
A cavity is used to keep photons localized to a region in space. Since photons reflecting off
the mirrors typically make many round trips in the cavity, this strengthens the interactions
with the atom and movable mirror when they are added later.
The use of a cavity also limits the frequencies of photons found inside. Classically, a
cavity of length L between two mirrors can contain an electric field subject to the standing
wave condition for allowed frequencies,

cm
fm = , (1.1)
2L

2
or
πcm
ωm = , (1.2)
L
where c is the speed of light and m is an integer index starting at 1. We assume that only one
mode of the cavity electric field is filled, so that the photons present are only of frequency
ω1 ≡ ωc , each with energy

E = ~ωc . (1.3)

The cavity is modeled as a quantum harmonic oscillator with discrete energy levels given by
 
1
En = ~ωc n + . (1.4)
2
The ladder operators a and a† , which obey the commutation relation a, a† = 1, are
 

used on number states of the oscillator to add and remove quanta (photons); the annihilation
operator a removes one photon from the state |ni and the creation operator a† adds one,
and together a† a is a number operator which gives the number n of photons in state |ni as
follows:


a|ni = n|n − 1i, (1.5a)

a† |ni = n + 1|n + 1i, (1.5b)


a a = hn|a† a|ni = hn|n|ni = n. (1.5c)

These are also related to the electric field; an electric field can be quantized and written in
terms of a and a† . A single mode of the field in the cavity of volume V can be written as
follows if the field does not vary much over the length of the cavity [11]:
r

ae−iωt − a† eiωt .

E(t) = i (1.6)
20 V
The Hamiltonian which represents the energy of the cavity mode is

H = ~ωc a† a. (1.7)

The cavity can be driven by an external laser to add photons to it. This setup results in

3
adding a term to (1.7):

H = ~ωc a† a + ~Y aeiωL t + a† e−iωL t ,



(1.8)

where ωL is the laser photon frequency and Y is the amplitude of the laser driving.
(1.8) is in the Schrödinger picture, in which states evolve with time due to the Hamilto-
nian and operators are constant. We usually transform to the interaction picture, in which
states evolve due to interaction terms of the Hamiltonian and operators evolve due to non-
interaction terms. This results in a time-independent Hamiltonian which is easier to use in
simulations:

H = ~∆a† a + ~Y a + a† ,

(1.9)

where ∆ is the detuning between the cavity resonance frequency and the laser frequency,
ωc − ωL .
We must also include some form of decay, since realistically the cavity is “leaky,” and
photons occasionally pass through a mirror and leave the system. We explain more about
the system’s interaction with the environment in Section 2.1. For now, we just add −iκc a† a
to the Hamiltonian, where κc is the cavity damping rate, and finally we have

H = ~∆a† a + ~Y a + a† − i~κa† a.

(1.10)

In Fig. 1.2, we show the expected photon number a† a for a driven cavity as a function

of detuning ∆. a† a is a Lorentzian with a peak at δ = 0 of linewidth 2κc .



4
Figure 1.2: a† a as a function of ∆; a Lorentzian centered at ∆ = 0 with linewidth 2κc .

Y = 1, κc = 2.

1.2 Cavity and Atom


Now a two-level atom is added to the cavity. The atom has a ground state |gi and an excited
state |ei. It can absorb a photon from the cavity to transition from |gi to |ei, and it can
emit a photon to fall from |ei to |gi. The resonance frequency for this transition is ωa . The
operators we use to act on the atom’s state are σ− and σ+ , which are combinations of Pauli
operators,

1
σ− = (σx − iσy ) ; (1.11a)
2
1
σ+ = (σx + iσy ) . (1.11b)
2

5
σ− and σ+ work as lowering and raising operators for the atom:

σ+ |gi = |ei; (1.12a)


σ− |ei = |gi. (1.12b)

The energy of the atom together with the energy of the cavity give the non-interaction
part of the Hamiltonian,
~ωa
HO = ~ωc a† a + σz . (1.13)
2
σz is a Pauli operator which acts on the state of the atom to give ±1 with corresponding
energies ±~ωa /2, with the point of zero energy in the middle. The zero point can be shifted
to correspond to the ground state; then σz /2 can be replaced with σ+ σ− in the Hamiltonian,
and the possible energies of the atom are 0 and ~ωa .
The energy of the interaction between the atom’s dipole moment p and the cavity mode
is

HI = −p · E, (1.14)

which can be written as [11]

HI = ~ga (σ+ + σ− ) a + a†

(1.15)
= ~ga aσ+ + a† σ− + a† σ+ + aσ− ,


where ga is the cavity-atom coupling strength. The first term in parentheses, aσ+ , corre-
sponds to the atom absorbing a photon from the cavity mode and transitioning to the excited
state; a† σ− corresponds to the atom falling to the ground state and emitting a photon into
the cavity; a† σ+ corresponds to the atom emitting a photon and becoming excited; and σ− a
corresponds to the atom falling to the ground state and absorbing a photon. The last two
terms do not conserve energy.
The total Hamiltonian for this system reads

H = HO + HI
(1.16)
= ~ωc a† a + ~ωa σ+ σ− + ~ga (σ+ + σ− ) a + a†


This is known as the Jaynes-Cummings model in cQED.


We again go to the interaction picture, in which the state of the system evolves due to
the interaction terms of the Hamiltonian (HI ), and the operators evolve due to the non-

6
interaction terms (HO ). The Heisenberg equation of motion for a gives

i
ȧ = [HO , a]
~
= i ωc a† a + ωa σ+ σ− , a
 
(1.17)
= iωc a† a, a
 

= −iωc a,

which shows that in the interaction picture, a picks up time dependence and becomes ae−iωc t .
This can be done similarly for a† , σ− , and σ+ , so that in the interaction picture the operators
become

a → ae−iωc t , (1.18a)
a† → a† eiωc t , (1.18b)
σ− → σ− e−iωa t , (1.18c)
σ+ → σ+ eiωa t , (1.18d)

and the interaction Hamiltonian becomes

HI = ~ga aσ+ ei(ωa −ωc )t + a† σ− e−i(ωa −ωc )t + aσ− e−i(ωa +ωc )t + a† σ+ ei(ωa +ωc )t .

(1.19)

For the sake of simplicity here and throughout the rest of the thesis, we assume that ωc = ωa ,
so that

HI = ~ga aσ+ + a† σ− + aσ− e−2iωc t + a† σ+ e2iωc t .



(1.20)

The last two terms are rapidly oscillating and average to zero over the time scales in which
we are interested, so they may be dropped; this is called the rotating wave approximation
(RWA).
Finally, the Jaynes-Cummings Hamiltonian for a cavity coupled to a two-level atom under
the RWA is

H = ~ωc a† a + ~ωa σ+ σ− + ~ga aσ+ + a† σ− .



(1.21)

You might also like