Вы находитесь на странице: 1из 259

PHOTON WAVE MECHANICS AND EXPERIMENTAL QUANTUM STATE

DETERMINATION

by
BRIAN JOHN SMITH

A DISSERTATION
Presented to the Department of Physics
and the Graduate School of the University of Oregon
in partial fulfillment of the requirements
for the degree of
Doctor of Philosophy
March 2007

ii

Photon Wave Mechanics and Experimental Quantum State Determination, a dissertation


prepared by Brian John Smith in partial fulfillment of the requirements for the
Doctor of Philosophy degree in the Department of Physics. This dissertation has
been approved and accepted by:

Dr. Hailin Wang, Chair of the Examining Committee

Date

Committee in charge:

Accepted by:

Dean of the Graduate School

Dr.
Dr.
Dr.
Dr.
Dr.

Hailin Wang, Chair


Michael G. Raymer, Research Advisor
Jens Nockel
Stephen Hsu
Andrew H. Marcus

iii

c
March
2007
Brian John Smith

iv

An Abstract of the Dissertation of

Brian John Smith


in the Department of Physics
Title:

for the degree of

Doctor of Philosophy

to be taken

March 2007

PHOTON WAVE MECHANICS AND EXPERIMENTAL


QUANTUM STATE DETERMINATION

Approved:
Dr. Michael G. Raymer

In this dissertation, a new method of quantum state tomography (QST) for light
is presented and demonstrated. This QST approach characterizes the transversespatial state of an ensemble of single photons by measuring the transverse-spatial
Wigner function of the ensemble. The first experimental measurements of the full
transverse-spatial state at the single-photon level for light are presented. To perform
these measurements, we developed a novel photon-counting, parity-inverting Sagnac
interferometer.
We also show how this method may be generalized to determine the transversespatial state of an ensemble of photon pairs, which may be entangled. This allows
characterization of the continuous-variable entanglement properties that can arise

v
in photon-pair states. The method introduced measures the two-photon, transversespatial Wigner function, which may be used to demonstrate a Bell-inequality violation.
In treating photons as particle-like entities, as we do in the interpretation of these
experiments, the question of the most appropriate theoretical description comes to
the fore. In order to describe these experiments, we extend a quantum theory of light
called photon wave mechanics, based on a single-particle viewpoint, and we show it
to be equivalent to the standard quantum field theory of light. We show that the
wave mechanics for multi-photon states is identical to the evolution of the coherence
matrices that appear in classical, vector coherence theory. The connection between
classical coherence theory (CCT) and photon wave mechanics allows us to utilize the
well-developed tools of CCT to describe the propagation of multi-photon states. We
present two example calculations to show the utility of the photon wave mechanics
treatment.

vi

CURRICULUM VITAE
NAME OF AUTHOR: Brian John Smith
PLACE OF BIRTH: Two Harbors, Minnesota
DATE OF BIRTH: October 7, 1977

GRADUATE AND UNDERGRADUATE SCHOOLS ATTENDED:


University of Oregon, Eugene, Oregon
Gustavus Adolphus College, Saint Peter, Minnesota
DEGREES AWARDED:
Doctor of Philosophy in Physics, 2007, University of Oregon
Bachelor of Arts in Physics, 2000, Gustavus Adolphus College
Bachelor of Arts in Mathematics, 2000, Gustavus Adolphus College
AREAS OF SPECIAL INTEREST:
Quantum Mechanics, Optics, and Information
Science Policy
PROFESSIONAL EXPERIENCE:
Graduate Research Assistant,
University of Oregon, 2003 - 2007
Graduate Teaching Fellow,
University of Oregon, 2000 - 2003
Teaching Assistant,
Gustavus Adolphus College, 1997 - 2000
GRANTS, AWARDS AND HONORS:

vii
John Borneman Prize (to an outstanding student in the fields of physics
and mathematics), Gustavus Adolphus College, St. Peter, Minnesota,
1999 - 2000
John Chindvall Scholarship in Physics (to an outstanding physics
student), Gustavus Adolphus College, St. Peter, Minnesota, 1998 1999

PUBLICATIONS:

B. J. Smith, M. G. Raymer, Two-photon wave mechanics, Phys. Rev.


A, 74, 062104, (2006).
B. J. Smith and M. G. Raymer, Photon Wave Mechanics, in
CLEO/QELS and PhAST, Technical Digest (CD) (Optical Society of
America, 2006), paper QThD3.
B. J. Smith, B. Killett, M. G. Raymer, I. A. Walmsley, and K. Banaszek,
Measurement of the transverse spatial quantum state of light at the
single-photon level, Opt. Letters 30, 3365-3367 (2005).
M. G. Raymer, B. J. Smith, The Maxwell wave function of the photon,
Proc. SPIE 5866, 293 (2005).
B. J. Smith, B. Killett, A. Nahlik, M. G. Raymer, K. Banaszek, and I. A.
Walmsley, The One- and Two-Photon Transverse Wave Functions:
Theory and Experiment, in CLEO/QELS and PhAST, Technical
Digest (CD) (Optical Society of America, 2005), paper QTuA3.
B. J. Smith, M. G. Raymer, B. Killett, K. Banaszek, and I. A. Walmsley,
The photon transverse wave function and its measurement, in FiO,
OSA Technical Digest Series (Optical Society of America, 2004), paper
FMO1.
B. J. Smith, M. G. Raymer, B. Killett, K. Banaszek, and I. A. Walmsley,
The photon transverse wave function and its measurement, in
CLEO/IQEC and PhAST, Technical Digest (CD) (Optical Society of
America, 2004), paper ITuM4.

viii

ACKNOWLEDGEMENTS

I would like to thank my advisor Professor Michael Raymer, who provided me with
numerous opportunities to learn and grow as an individual and a scientist. Thank
you for your support, guidance, leadership and teaching I will always carry these
with me. I would also like to acknowledge and thank Professor Ian Walmsley at the
University of Oxford, United Kingdom, and Professor Peter Smith at the University
of Southampton, United Kingdom for helpful discussions, and their hospitality when
I visited for collaborative research. I thank Professor Jaewoo Noh at Inha University,
Korea, for helpful hints with down conversion.
I have greatly benefited from the many discussions with, helpful hints from, and
camaraderie of my peers in the lab. To all the members of the Raymer lab during
my tenure I say, Thank you. Dr. Ethan Blansett, Dr. Andy Funk, Guoqiang Cui,
Justin Hannigan, Wenhai Ji, PengFei Nie, Chunbai Wu, Cody Leary, and Hayden
McGuinness I wish you well in all that you do.
I am fortunate to have a wonderful, caring, supportive family. To my mom, Jackie
Gerard, dads Al Gerard and Tom Smith, brother and sister, Rick Smith and Dorothy
Gerard, in-laws John, Marcy and Charlie Colvin, and grandparents, aunts, uncles,
and cousins thank you for all the love, support, and encouragement you have given
me over the years. You have always believed in me, and I am forever grateful.

ix
And last but not in the least, the most important person in my life, my exquisite,
loving, darling, sweet wife, Kelly. You are my cheerleader, first line of support, partner
in life, best friend, and so much more. For your patience, understanding, and all that
you do for me I thank and love you from the bottom of my heart.

To my grandparents John M. and Dorothy E. Anderson, Kenneth R. and Donna L.


Smith you inspire me.

xi

TABLE OF CONTENTS

Chapter

Page

1. INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
Light . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
Quantum State Determination/Tomography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
Photon as a Particle (Photon Wave Function) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
Key Issues Addressed. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2. QUANTUM OPTICS AND PHOTON WAVE MECHANICS . . . . . . . . . . . . . . 23
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
Quantum Optics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
From Einstein to Maxwell Deriving the Single-photon Wave Function . . . . . 31
Quantization of the Single-photon Wave Function . . . . . . . . . . . . . . . . . . . . . . . . . . 43
Photon Wave Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
Connections to Classical Coherence and Photo-detection Theories . . . . . . . . . . 56
Modes Versus States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
Measurement-induced Photon Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
3. MEASURING THE TRANSVERSE SPATIAL STATE OF LIGHT AT
THE SINGLE-PHOTON LEVEL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
Electromagnetism in the Paraxial Approximation . . . . . . . . . . . . . . . . . . . . . . . . . . 71
The Photon Wave Function in the Paraxial Approximation . . . . . . . . . . . . . . . . . 77
The Wigner Distribution Function and Its Properties. . . . . . . . . . . . . . . . . . . . . . . 79
The Transverse Spatial Wigner Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
Parity-inverting Sagnac Interferometer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
Sagnac Interferometer Diffraction Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
Experimental Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
Experimental Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
4. TWO-PHOTON TRANSVERSE SPATIAL-STATE
CHARACTERIZATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119

xii
Chapter

Page

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
Two-photon Transverse Wave Function and Wigner Function . . . . . . . . . . . . . . . 121
Transverse Spatial Disentanglement of a Photon Pair . . . . . . . . . . . . . . . . . . . . . . . 123
Spontaneous Parametric Down Conversion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
Experimental Down-conversion Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
Ghost Imaging and Ghost Wigner Distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
Two-photon Transverse Spatial Wigner Function Measurement and Bell
Inequalities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
5. CONCLUSIONS. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
Future Directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
APPENDICES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
A. SINGLE-PHOTON WAVE FUNCTION LORENTZ TRANSFORMATION
PROPERTIES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178

B. SPONTANEOUS PARAMETRIC DOWN CONVERSION . . . . . . . . . . . . . . . . 197

C. GHOST IMAGING . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214

D. FOURIER OPTICS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226

BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231

xiii

LIST OF FIGURES
Figure
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.

Page

Schematic set up for measurement-induced interaction . . . . . . . . . . . . . . . . . . . . . 65


Longitudinal and transverse wave functions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
Converging and diverging Gaussian beams and associated transverse
spatial Wigner functions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
Parity-inverting Sagnac interferometer with a Dove prism. . . . . . . . . . . . . . . . . . 92
Parity-inverting Sagnac interferometer with the top mirror. . . . . . . . . . . . . . . . . 94
Rotation of the transverse spatial state caused by the top-mirror. . . . . . . . . . . 95
Parity-inverting Sagnac interferometer. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
Linear-optical evolution of the transmitted and reflected fields that pass
through the parity-inverting Sagnac interferometer. . . . . . . . . . . . . . . . . . . . . . . . . 99
Experimental setup to measure the transverse spatial Wigner function. . . . . . 104
Steering mirror motion control setup. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
Detection electronics for single-photon Wigner measurements. . . . . . . . . . . . . . . 106
Raster scan of phase space for a one-dimensional field. . . . . . . . . . . . . . . . . . . . . . 108
Experimental and theoretical plots of the transverse spatial Wigner
function of a diverging Gaussian beam. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
Intensity and field amplitude of the HG10 mode. . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
Experimental and theoretical plots of the transverse spatial Wigner
function of a HG10 mode. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
Top-hat field amplitude and experimental arrangement for its
construction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
Experimental and theoretical plots of the transverse spatial Wigner
function for a propagated top-hat field. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
Displaced double-top-hat field amplitude and experimental arrangement
for its construction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
Experimental and theoretical transverse spatial Wigner functions for the
displaced top-hat field. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
Paraxial two-photon source. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
Entangled photon pair transmission through a turbulent atmosphere. . . . . . . 125
Concurrence and transmission fidelity of an OAM-entangled photon pair
after propagation through a turbulent atmosphere.. . . . . . . . . . . . . . . . . . . . . . . . . 127
Measuring the OAM density matrix using Laguerre-Gauss holograms. . . . . . . 128
Parametric amplification process. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
Spontaneous parametric down-conversion process. . . . . . . . . . . . . . . . . . . . . . . . . . 132
Four possible spontaneous parametric down-conversion configurations.. . . . . . 133

xiv
Figure
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.
48.

Page

Basic components for SPDC setup. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136


Contrast between continuous-wave and pulsed pumps. . . . . . . . . . . . . . . . . . . . . . 137
Second-harmonic generation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
Walk-off effects in second-harmonic generation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
Optimal focusing for second-harmonic generation. . . . . . . . . . . . . . . . . . . . . . . . . . 144
Phase-matching in second-harmonic generation by angle tuning.. . . . . . . . . . . . 146
Experimental second-harmonic generation setup.. . . . . . . . . . . . . . . . . . . . . . . . . . . 147
Experimental down-conversion setup. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
Detection electronics used with the APD and PMT detectors. . . . . . . . . . . . . . . 150
Hong-Ou-Mandel-interference experiment setup. . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
Hong-Ou-Mandel interference experiment results. . . . . . . . . . . . . . . . . . . . . . . . . . . 154
Quantum-imaging experimental setups. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
Advance-wave description of quantum imaging. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160
Experimental ghost-imaging results. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
Conditional single-photon state preparation setup. . . . . . . . . . . . . . . . . . . . . . . . . . 163
Incoherent and coherent ghost-Wigner experimental setups. . . . . . . . . . . . . . . . . 165
Experimental setup to measure the two-photon transverse spatial
Wigner function. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
Diagram of degenerate, type-I, non-collinear SPDC.. . . . . . . . . . . . . . . . . . . . . . . . 209
Schematic diagram of a standard ghost-imaging experiment. . . . . . . . . . . . . . . . 215
Ghost imaging experimental setup. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
Unfolded diagram of ghost-imaging experiment.. . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
Linear optical system in the paraxial approximation. . . . . . . . . . . . . . . . . . . . . . . . 227

CHAPTER 1

INTRODUCTION

Light

Light has played an instrumental role in many developments of our understanding


of nature, and continues to be at the forefront of modern-day, fundamental physical
research. From the time of Euclid (c. 300 B.C.E.), who thought that light traveled
in straight lines, but originated from the eye and strikes the objects seen by the
observer, light has been a continuing theme in scientific discovery. The microscope,
thought to be first developed by the Dutch lens maker, Zacharias Janssen (c. 1590),
opened the way for many discoveries in the medical and biological sciences. The
telescopes of Galileo Galilei (1609) led to the discovery of moons circling Jupiter,
which reinforced the heliocentric model of the universe, leading to his persecution by
the church. Fermats principle of least time, (1657) in which light travels from one
point to another along the path taking the least transit time, may be taken as a precursor to the principle of least action, which is at the foundation of Hamiltonian
and Lagrangian dynamics, as well as modern quantum field theory. Isaac Newton
(1672) and Christian Huygens (1678) put forth the competing corpuscle and wave
theories of light respectively. In 1801, Thomas Young presented his famous double-

2
slit experiment results, and laid to rest (for a time) the particle-wave debate of light,
solidifying Huygens wave theory.
One of the pinnacles of 19th century science, Maxwells theory of electromagnetism
(1864) was the first unified field theory, uniting electricity, magnetism and optical
phenomena in a single theory. Unbeknownst to him, Maxwell had also discovered
the first relativistic quantum theory of photons (light corpuscles). In order for the
equations that now bear his name to take the same form in any non-accelerating
(inertial) frame of reference, the Galilean transformations assumed to transform
position and time coordinates from one inertial frame to another had to be modified.
These required modifications of Galilean relativity led Einstein (1905) to formulate
the theory of special relativity.
Quantum mechanics (QM), one of the most successful physical theories we have
produced, was developed at the beginning of the 20th century, and brought the
particle-wave debate back to center stage. The origins of quantum theory can be
traced to Max Plancks theory of the emission of light by heated material bodies
(1900). In order to correctly predict the observed spectrum, he needed to assume
that light was only emitted with certain discrete energies. Building on the idea
that light was emitted and absorbed with discrete energies, Einstein carried the idea
further, introducing the light quanta, to describe the photoelectric effect, work for
which he later won his only Nobel Prize. Arthur Comptons experiments on the
scattering of X-rays by free electrons followed the same law as the collision between

3
two elastic spheres. Thus the novel idea of light quantization was established both
experimentally and theoretically in the early 1900s, and the wave-particle debate was
upon us once again.
The advent of the laser in 1960, which allowed the creation of light in highly
coherent states, ushered in the field of quantum optics and quantum coherence theory,
and brought with it many technological advances in society from fiber optics
communications to compact-disc data storage. It enabled many experiments to be
carried out such as the ultra-fast probing of molecular bonds with femtosecond laser
pulses and opened the door to completely new fields of research, such as nonlinear
optics. Additionally, the laser enabled the first conclusive experiments testing the
Bell-inequality [1], which highlights the strange (non-local, non-classical) behavior of
correlated (entangled) quantum systems spatially separated from one another, to be
performed using polarization-entangled photons from an atomic-ion cascade emission
[2, 3].
Even today, light is at the fore of modern research. The fast-growing field of
quantum information has relied on many of its proof-of-principle experiments to
be carried out with entangled photons produced with spontaneous parametric down
conversion. In addition, the now well-established field of quantum state determination,
also called quantum state tomography, which was first experimentally carried out in
1993 to characterize the quantum state of light [4], is central to the emerging fields
of experimental quantum information and quantum technologies. Moreover, light is

4
typically the main tool used to investigate or change properties of matter. Therefore,
by gaining as much information as possible about light, we can better understand the
behavior and properties of materials.
In this dissertation, we focus on the characterization of single-photon and twophoton quantum states of light, which play an important role in quantum information
science, quantum metrology, and communications. In the process, we find that it is
useful to discuss the state of the photon, or photons, in terms of coordinate-space
wave functions which obey certain wave equations. This formalism we call photon
wave mechanics. However, before proceeding we first present some background on
the subjects that we will discuss.

Quantum State Determination/Tomography

Physics is the study of nature at its most fundamental level. Owing to the fact
that mathematics seems to be the language in which physical concepts are the easiest
to express, physics is often mathematically very detailed. However, one should not
think that simply because a mathematical concept is beautiful, or intriguing,
that it necessarily corresponds to a physical theory. This is a common mistake made
by non-physicists associating mathematics with physics. Physics is a scientific
study of nature and, as such, is deeply rooted in its observational and experimental
details. Without experimental data to support and drive physical theories, they
become nothing more than mathematical objects, not reflecting any actual physical

5
system. Today there seems to be a lag in the amount and development of experimental
data and discoveries that push our envelope of understanding. Experiments at the
edges of scientific knowledge are becoming increasingly difficult to perform (all of the
easy and obvious experiments have already been done). This lack of experimental data
leads to theoreticians developing pseudo-theories of everything with little real-world
support to back them up. The difficulty of experiments also drives the experimentalist
to devise more creative ways in which to probe the fundamental workings of nature.

The state of a physical system represents our knowledge of the system and provides
information about it in the past and future. In standard quantum mechanics the state
of a system is represented mathematically by a wave function in either the momentumspace representation (p, t), or the coordinate-space representation (x, t) (or a
statistical ensemble of wave functions when the state is mixed). The representations
are typically related by a three-dimensional Fourier-transform relationship.

The

modulus squared of the momentum-space (coordinate-space) wave function is equal


to the probability per unit momentum-space (coordinate-space) volume to find the
system with a particular value of momentum (position). Quantum state determination,
or quantum state tomography, refers to a method by which one may experimentally
gain all possible information about the state of a system.

This enables one to

make the best possible prediction (in a probabilistic sense) about the results of
any measurement or experiment that may be performed on the system. In classical
mechanics, the state of a system is represented by a set of numbers that label the

6
positions and velocities of particles at a given time (or the amplitude and phase of a
field at all points in space). From this set of numbers, one can predict all outcomes
of experiments using Newtonian, Hamiltonian, or Lagrangian mechanics. Here the
state of the classical system is a physical property of the system itself, and could, in
principle, be observed without introducing any disturbance to the system. This is
not the case for quantum-mechanical systems.
It is well known that the state of a single quantum system cannot be determined
[5, 6]. The linearity of quantum mechanics implies that a quantum state cannot
be cloned, that is replicated on another system, allowing for multiple copies to be
made [7]. If cloning were possible, one could perform all possible measurements on the
system and its clones, and thereby determine the state. In addition, the uncertainty
principle [8], which implies that by observing a quantum system, we inherently change
the state of the system through a quantum back action, also prohibits full quantum
state determination of a single quantum system.
One of the fundamental questions in quantum theory, whether or not one can infer
the quantum state of a collection of identically-prepared quantum systems, i.e. an
ensemble, remained unanswered for much of the 20th century due, in part, to these
difficulties. The question was first posed by Pauli in 1933 [9], but was not answered.
It was not until the late 1950s that the first systematic approach to reconstructing
the state of a quantum system from a series of measurements was given by Fano
[10]. Quantum state estimation techniques began with the proposal by Vogel and

7
Risken [11], and the first quantum state determination experiments carried out by
Smithey, et. al. at the University of Oregon in 1993, in which the quantum state of
a single mode of the electromagnetic field was measured [4]. In these experiments,
the amplitude and phase structure of a single electromagnetic field mode, specified
by the polarization, and spatio-temporal mode, were determined by a procedure
called optical homodyne tomography (OHT). The technique of OHT was used to
tomographically reconstruct the Wigner distribution function of the quantum state
from several measured probability distributions. The Wigner function, introduced
in 1932 by Wigner to simplify quantum statistical mechanics problems [12], is a
quasi-probability phase-space distribution that is directly related to the wave function
(density matrix) for a pure- (mixed-) state quantum system. For pure (mixed) states,
this direct relationship between the Wigner function and the wave function (density
matrix) allows for us to obtain one from the other. Thus complete knowledge of the
Wigner function implies complete knowledge of the state of the system [13]. Since
these pioneering experiments were performed over a dozen years ago, the field of
quantum state determination has become widespread, and the various techniques
have become a standard tool in laboratories around the world [14].

A closely related subject, optical phase retrieval aims to characterize the amplitude
and phase structure of a fully coherent, time-stationary, quasi-monochromatic electromagnetic wave field as a function of position, E (x), assuming the polarization and
frequency are known. Here E (x) refers to the complex field amplitude of a scalar

8
electro-magnetic wave, as a function of the spatial coordinate x = (x, y, z). In the case
of partially-coherent light, the field may be characterized by the Wigner distribution
function [15]
1
W (x, k) = 3

d3 x0 E (x x0 ) E (x + x0 ) ei2kx ,

(1.1)

where k = (kx , ky , kz ) is the wave vector, and brackets h i, imply an ensemble average
over all statistical realizations of the field. The task of spatially mapping out the field
amplitude is quite easy, and can be accomplished by using photographic paper, or
charge-coupled-device (CCD) cameras for example. However the ability to determine
the phase as a function of spatial position for an optical field is quite challenging.
The ability to characterize the amplitude and phase information of an optical
field, or amplitude and coherence information in the case of partially-coherent light,
is critical to several areas of study [16]. Optical coherence tomography, which has
found many applications in the biological sciences, relies on the known coherence of
the incident radiation to determine the structure of objects from which it scatters.
There are several other areas of practical importance that the coherence of an optical
field plays a critical role, such as the testing of optical equipment, the study of fluid
dynamics, and photolithography, to name a few. Indeed, one can go as far as to say
that coherence properties of light are the most important aspect, due to their role in
determining interference and other optical correlation effects. Various light sources,
both man-made and naturally occurring, have varying degrees of coherence, and thus
a study of such sources requires the ability to characterize coherence. This makes

9
techniques to measure the coherence of optical fields not only of practical use, but
also of great interest in studying the fundamental nature of light.
There have been several proposed techniques to measure the coherence, (we use the
term coherence for spatial coherence), of optical fields [1719]. The first experimental
methods suffered from an inability to measure fields of arbitrary coherence. Then
in 1995, at the University of Oregon, McAlister et. al. introduced a tomographic
method to measure fields with any state of coherence [20]. This method characterized
the transverse field, assuming the paraxial approximation, and required tomographic
reconstruction of the Wigner function, which led to the possibility of errors in the
inverse transform. After McAlisters development, there were several other transversecoherence measurement methods introduced [2123]. However, none of the proposed
techniques worked at the level of single-photon fields.
In this dissertation, we present and demonstrate the first experimental technique
to fully characterize the transverse spatial state of light (i.e.

transverse spatial

coherence) at the single-photon level. The method measures the transverse spatial
Wigner function W (r, k ), for a single photon, in the two-dimensional plane (x, y)
located at z = 0, of a wave field with arbitrary state of coherence. Here r = (x, y), and
k = (kx , ky ) are the transverse position and wave vectors in the plane perpendicular
to the propagation axis (z). Our method utilizes a parity-inverting Sagnac (commonpath) interferometer to scan the phase space (r, k ). This is done by taking advantage
of the fact that the Wigner function is proportional to the expectation value of the

10
(r, k ) [24]. The interferometer creates two copies of the
displaced parity operator
field at the input beam splitter, which travel through the interferometer in opposite
directions. The fields are directed in and out of the plane of the interferometer, picking
up a geometric phase that causes a rotation of the wave fronts by 180 relative to
one another. This gives a two-dimensional parity inversion of the wave fronts. The
Wigner function is proportional to the integrated intensity detected at the output
using photon-counting detectors. The phase-space point (r, k ) at which the Wigner
function is measured, is controlled by a mirror external to the interferometer.
The first experimental demonstrations characterizing the single-photon transverse
spatial state are presented in this dissertation. We have constructed four different
fields to demonstrate the ability of our technique to measure not only at the singlephoton level, but also highly-divergent fields, which aid in the study of decoherence
effects in scattered light.
Even more interesting than single-photon coherence, the study of two-photon
coherence represents an arena in which classical and quantum theories of light can
truly be distinguished [25]. The most common experimental demonstration of the
non-classical nature of light is the Hong-Ou-Mandel (HOM) interference experiment.
In this experiment, two identical photons are incident at opposite input ports of a
perfect 50:50 beam splitter. The signals of two photon-counting detectors, located at
the output ports of the beam splitter, are sent to a coincidence counter, registering
when the photons arrived at both detectors. If the photons arrive at the beam splitter

11
at the same time, one observes no coincidences between the detectors. This effect is
known as photon bunching, and stems from the boson nature of light. Indeed, if
the same experiment were performed with electrons, one would observe electron antibunching, reflecting the fermion nature of electrons. Now, as one of the photons
arrival time to the beam splitter is varied with respect to the others, coincidence
counts between the two detectors begin to emerge. When the arrival time difference
is greater than the coherence time of the photons, the coincidence counts reach the
classical level when 50 percent of the time one will observe coincidences. Thus, a plot
of the probability of coincidence as a function of arrival time delay shows a drop, or
dip, at zero time delay, known as the Hong-Ou-Mandel dip.
The HOM interference experiment probes the temporal correlations of photon
states by assuming that the photon spatial states are identical. Non-identical photon
states lead to degradation of the visibility in the HOM dip. To address the spatial
state of the photon pair, which generally includes any entanglement between them,
one must devise a new characterization method. The ability to characterize the spatial
state of two-photon states would not only be of interest at a fundamental level, but
it would also find use to characterize photon sources used for quantum information,
metrology and communications schemes.
Another often cited set of experiments that highlight the non-classical nature of
spatially-entangled two-photon states are the quantum-imaging experiments of the
mid 1990s [2629]. These quantum-imaging experiments utilize entangled photon

12
pairs derived from a spontaneous parametric down-conversion (SPDC) source. The
SPDC source is of particular interest and use in quantum-optical implementations of
quantum-information schemes, and tests of fundamental quantum theory [3032]. In
the process of SPDC, an intense laser beam, called the pump, is incident on a nonlinear
optical crystal. Through the nonlinear, optical interaction, a pump photon has a
small, but non-zero probability to split into a pair of daughter photons, traditionally
called the signal and idler photons. Energy and momentum are typically conserved
in the interaction, resulting in correlation, or entanglement, between the daughter
photons.

In the quantum-imaging experiments, the entangled photons from the SPDC


source are directed along two different, spatially-separated paths. In one path, say
the idler-photon path, an aperture (amplitude mask) followed by a large-area photoncounting detector, is placed. In the other path, the signal-photon path, a lens, followed
by a small, point-like photon-counting detector, is placed. As the point-like detector is
scanned in the plane perpendicular to the beam axis, the coincidence rate between the
two detectors is recorded. The coincidence rate as a function of the signal detectors
position in the transverse plane maps out the aperture in the idler-photon path.
Each individual detector count rate remains relatively constant, but the coincidence
rate reflects the aperture transmission function. The distance between the SPDC
source and aperture, and the SPDC source, imaging lens, and point-like detector
are determined by a thin-lens-like equation [26]. These experiments depend only

13
on amplitude correlations and not phase correlations, and thus do not give complete
information about the two-photon state. It has been shown that these experiments do
not require entangled photons to observe the coincidence image, classically-correlated
photons suffice. However, there are some benefits, such as increased spatial resolution
and visibility, when using the entangled-photon source.
In this dissertation, we present an experimental technique to fully characterize
the two-photon transverse spatial state of light. The method measures the twophoton transverse spatial Wigner function W (r1 , k1 ; r2 , k2 ), a generalization of
the single-photon transverse spatial Wigner function, in a pair of planes perpendicular
to the propagation axis. This allows characterization of not only amplitude correlations,
but also phase correlations between photon pairs.
Our approach is based on the single-photon Wigner function method described
above. In the two-photon case, two parity-inverting Sagnac interferometers are used
to measure the two-photon Wigner function of spatially separated photons traveling
in different directions, such as the entangled photons encountered in a SPDC source.
The signals of the detectors placed at the outputs of the interferometers are sent to
a coincidence counter, whose count rate is proportional to the two-photon transverse
spatial Wigner function. This method enables one to violate a Bell inequality based
on the Wigner function [33].
The question of interpretation arises as to what is measured for single-photon and
two-photon states in these experiments. We advocate that it is not the electromagnetic

14
field that is characterized; we go so far as to claim the electromagnetic field does not
exist for a single photon. Rather we explain our results in terms of Wigner functions
derived from single-photon wave functions, and that the electromagnetic field is an
emergent quantity when considering many photons.

Photon as a Particle (Photon Wave Function)

The concept of the light quantum was first introduced by Einstein in 1905 [34]
to describe the then-recently-observed photo-electric effect [35]. In his paper [34],
Einstein writes, According to the assumption considered here, when a light ray
starting from a point is propagated, the energy is not continuously distributed over
an ever increasing volume, but it consists of a finite number of energy quanta, localized
in space, which move without being divided and which can be absorbed or emitted
only as a whole. This statement captures the essence of the view of the photon
as a quantum particle that many physicists hold in one form or another. More
evidence of the particle-like nature of the light quantum was provided by the results of
Arthur Comptons electron-X-ray scattering experiments [36]. In these experiments,
an electron and X-ray scatter from one another. The resulting change in energy and
momentum of the two objects can be easily described from a billiard-like collision, in
which a point-like electron scatters from a similarly point-like light quantum.
In spite of these early developments, an acceptable quantum theory of electromagnetism based upon the standard particle-wave-function viewpoint did not develop.

15
The first satisfactory quantum treatment of electromagnetism was not given until 1927
by Dirac [37], and later clarified by Fermi [38]. Diracs quantum treatment of light
was not given in terms of wave mechanics, in which particles are the fundamental
quantum objects described by wave functions. Rather, he presented a quantized field
theory (QFT), where the field is the fundamental physical entity. The term photon,
which we typically use instead of light quantum, was not used until 1926 by Gilbert
N. Lewis to describe the interaction of neutral-atom valence bonds [39], and not to
describe the light quanta of Planck and Einstein. Indeed, the term photon has been
met with opposition for its catch-all nature [40]. However, photon is a very convenient
word to describe what we now mean as a fundamental excitation of the quantized
electromagnetic field.

There is good reason that the particle view of the photon did not lead to a
quantum theory of light, as opposed to the quantum theory of the electron, where
particles abound. Much of the difficulty in developing a quantum theory of the
photon as a particle stems from its inherent relativistic nature, due to its zero rest
mass. The absence of rest mass, along with its internal, spin-1 degree of freedom,
led Newton and Wigner to the conclusion that the photon is, strictly speaking, nonlocalizable [41, 42]. To what extent the photon may be localized has been carefully
examined [4348]. Bialynicki-Birula has found that photons can be localized in space
with an exponential falloff in the energy spatial density and photo-detection rate
[48]. Nevertheless, faster than exponential falloff cannot be achieved, as far as is

16
known. The strict non-localizability of the photon implies that there is no position
operator, and thus no position eigenstates for the photon. This leads us to the
conclusion that there is, in the usual sense, no probability density for the position
of the photon, and thus a coordinate-space (position-representation) wave function
cannot be consistently introduced. This has led to the course grained, photo-detection
model of the photon probability amplitude [4953], that has been used to explain
the majority of experimental results. Non-canonical position operators have been
introduced [5457] to try avoiding these difficulties, but still leave something to be
desired.
Nevertheless, several candidates have been proposed for the photon wave function
in coordinate-space [52, 53, 5863]. The first attempt to introduce a coordinatespace photon wave function, viewed as a description of a single particle, was given by
Landau and Peierls [58]. However, it was quickly noted that the Landau-Peierls (LP)
wave function is a highly non-local object [9]. This function was also independently
rediscovered in the 1980s [6467].
By extending what one means by wave function, to a complex vector-function
of space and time coordinates (x, t), that adequately describes the quantum state
of a single photon, it is possible to define a wave function for the photon. The
mathematical object that is now accepted by many as the photon wave function is
closely related to the Riemann-Silberstein (RS) vector
B (x, t)
D (x, t)
+i
,
F (x, t) =
20
20

(1.2)

17
where D (B) is a field analogous to the electric- (magnetic-) displacement field, and
0 (0 ) is the permittivity (permeability) of the vacuum [68]. The RS vector obeys
the complex form of the Maxwell equations, which in free space may be written as
it F (x, t) = c F (x, t) ,
(1.3)
F (x, t) = 0.
The use of the RS vector as the photon wave function has been advocated by many
over the past 75 years [59, 60, 6885]. There are several reasons for choosing the
RS vector, and other equivalent formulations, as the single-photon wave function in
coordinate space. For example, one can arrive at the RS vector from a particle view,
by starting with Einstein kinematics, and derive the photon wave function in much
the same way that Dirac did for the electron (see chapter II).
The subtlety of using the RS vector as the photon wave function lies in the fact that
|F (x, t)|2 is not the position probability density of standard non-relativistic quantum
mechanics, but rather is the local spatial energy density. Thus, the standard scalar
product between two RS vectors F and F0
0

hF | F i =

F (x, t) F0 (x, t) d3 x,

(1.4)

cannot be interpreted as the probability amplitude for finding a photon in state F,


when it is known to be in state F0 . If one tried to push this interpretation, there
are several problems, the least of which is the fact that the integral in Eq. (1.4)
is not Lorentz invariant [68, 73]. Lorentz invariance is expected since a probability
amplitude is a number. The proper, Lorentz-invariant scalar product is found to take

18
a non-local form in coordinate space [68, 73] (see Chapter II and Appendix A)
1
(F| |F ) = 2
2 ~c
0

dx

F (x, t) F0 (x0 , t)
,
dx
|x x0 |2
3 0

(1.5)

where we have introduced the non-standard notation ( F| |F0 ), for the Lorentz-invariant
scalar product to emphasize that it is not just the standard local integral. With this
scalar product, the expectation value of any observable such as energy or momentum,
takes on a simple, bilinear form that is formally analogous in form to the expectation
value of standard quantum mechanics [59, 68, 70, 71, 73]. Thus, the single-photon
wave function, its equation of motion and scalar product comprise the workings of
single-photon wave mechanics.
The problem of treating multiple photons, and not just single photons, in a
consistent wave-mechanics format, in which the state of the photons is described
by a wave function that obeys a wave equation (for pure states), has not yet been
accomplished. This is quite disappointing since many proof-of-principle experiments
in quantum information, and those probing the fundamentals of quantum theory
(such as measurements of Bell inequalities), are performed with entangled photon
pairs [3032], and could benefit from such a description. For example, one would like
to be able to explain few-photon interference experiments, which are often discussed
loosely in terms of overlapping wave packets or interference of Feynman paths, in
terms of photon wave functions. This is particularly true when the experiments
involve localized photon wave packets.

19
It is often useful, and sometimes easier to use the wave mechanics approach to
solving a problem than to use the full QFT. For example, one typically does not use
QFT to treat the helium atom, but rather one uses Schrodinger wave mechanics. The
ability to attack a problem from different approaches allows for deeper insight into the
issues at hand, and can lead to a better understanding of the problem. The quantumfield theoretic approach gives the correct answer, however, it is useful to be able to
solve problems using several different methods. Indeed this opinion was emphasized
in Feynmans 1965 Nobel Lecture [86], in which he notes, I, therefore, think that a
good theoretical physicist today might find it useful to have a wide range of physical
viewpoints and mathematical expressions of the same theory (for example, of quantum
electrodynamics) available to him. Thus the photon-wave-mechanics approach can
lead to a more intuitive understanding of experiments, and give different insights into
the physics occurring. It is also satisfying to note that photons can be treated in the
same way as electrons, at the level of a quantum particle.

In this dissertation, we show, for the first time, how to treat multi-photon states
in a consistent photon-wave-mechanics approach. We develop the wave mechanics for
multi-photon states by expanding on the single-photon wave-mechanics formalism.
We define the multi-photon wave functions, and determine their equations of motion.
In treating mixed states, we point out how to obtain the reduced density matrix from
a given multi-photon state. In the process, we find useful connections between photon
wave mechanics (PWM) and several other well-known theories, such as classical

20
coherence theory. As an example calculation using PWM, we show how to treat
the problem of multi-photon interference on a beam splitter.
We find that classical vector coherence theory [87] is closely related to photon
wave mechanics. We show that the two-photon wave function and its equation of
motion are equivalent in form to the second-order, classical coherence matrices [87].
This implies that the evolution of a two-photon state can be described using the welldeveloped tools of classical coherence theory. As a demonstration of the utility of this
close relationship, we show how a pair of photons, entangled in their orbital-angularmomentum degrees of freedom, disentangle as they propagate through a turbulent
atmosphere.

Key Issues Addressed

We begin the dissertation with the theoretical descriptions of light used to describe
our experiments. A review of standard quantum optics is presented to show the
connection between QFT and PWM. The monochromatic mode expansion of the
electric field operator, which is widely known, and the non-monochromatic mode
expansion [88], which is not as well known, are discussed. We present a derivation
of the single-photon wave function in coordinate space [59, 60]. When the canonicalquantization procedure is carried out on the single-photon wave-function theory,
the non-orthogonality of the non-monochromatic wave-packet modes of Titulaer and
Glauber [88] naturally arise. We note that the scalar product for the wave function

21
in the coordinate representation, which is a non-local integral Eq. (1.5), is tied to the
non-orthogonality of the non-monochromatic wave packet modes of quantum optics.

We extend the single-photon theory to multi-photon states by constructing multiphoton wave functions and determining their proper wave equations. It is shown
that the PWM for two photons is closely related to second-order vector, classical
coherence theory (CCT) [87], and the two-photon detection amplitude of quantum
optics [52, 87], which is also called the biphoton amplitude [89]. These connections
are generalized to n-photon wave mechanics, n-th order CCT, and n-photon detection
amplitudes, highlighting the connections between the theories. We perform a calculation
of measurement-induced photon interaction as an example of the utility of the PWM
approach and how it can be implemented.

After developing the theory of photon wave mechanics, we turn to characterization


of a single-photon state. In particular, we focus on the transverse spatial state of the
photon in the paraxial approximation. Here the single-photon wave function may be
represented by the transverse spatial wave function, which we introduce in analogy
to the paraxial treatment of classical radiation. From this transverse spatial wave
function we construct the transverse spatial Wigner function for such a state.

We present an experimental technique to directly measure the transverse spatial


Wigner function of an ensemble of single photons. This is done through use of a
novel, parity-inverting Sagnac interferometer. The first complete measurements of

22
the transverse spatial state of light at the single-photon level are given for a series of
different field distributions in order to demonstrate the technique.
We then move from characterization of single-photon states, to characterization
of two-photon states. Again, we examine the transverse spatial state of the photons.
The multi-photon generalizations of the single-photon transverse spatial wave and
Wigner functions are given. In particular, we discuss the two-photon transverse
spatial wave function and Wigner function. We show how one can utilize the close
relationship between PWM and CCT to calculate the disentanglement of a pair of
spatially-entangled photons traversing a realistic turbulent atmosphere.
An experimental technique to measure the two-photon transverse spatial Wigner
function is presented. We show how one can use this method to violate a Bell
inequality. Our experimental progress towards realization of this experiment is given
by presenting HOM interference and quantum-imaging results from our SPDC source.
The idea of quantum-imaging can be generalized to a non-local Wigner function,
which we introduce.
The dissertation concludes with comments on current and future work. There are
several appendices that cover the Lorentz-transformation properties of the photon
wave function, the Lorentz-invariance of the photon-wave-function scalar product,
the theory of spontaneous parametric down conversion, and quantum-imaging.

23

CHAPTER 2

QUANTUM OPTICS AND PHOTON WAVE MECHANICS

Introduction

Quantum field theory (QFT) has its origin in Diracs exposition on the quantization
of the electromagnetic field [37]. In this treatment, Dirac noted that the Hamiltonian
of the electromagnetic field may be expressed as a sum of harmonic-oscillator modes
with different resonant frequencies. He then proceeded to quantize the electromagnetic field by imposing the now-famous commutation relations on the canonicallyconjugate variables (field amplitudes, or field quadratures) that arise from the Poisson
brackets in the classical discussion. The fields were raised to the status of operators,
and expanded in terms of creation and annihilation operators of the fictitious harmonic
oscillator. Photons were interpreted as excitations of the field that arose from the
application of the creation operator acting on the vacuum state of the electromagnetic
field. This is the basis of QFT and quantum optics (QO), in which the details have
since been developed and fleshed out. Yet there is still debate as to the nature of
the photon. There is a discrete click in a photo-detector, which signals the arrival
of a photon. The recent development of controlled single-, pair-, and few-photon

24
states, which may be manipulated, and measured, begs the question, Is a photon
just a monochromatic field excitation, as in the canonical Dirac theory?
To contrast this development, consider the story of the electron. It was first
discovered experimentally in cathode-ray-tube experiments, and viewed as a click
in an electron detector. The model of atomic physics, in which negatively-charged
electrons orbit a positively-charged nucleus, relies on a particle view of the electron,
and is well seated in its predictive power. Indeed, the Dirac equation of the electron,
which was the first quasi-successful attempt to unite special relativity and the quantum
theory of electrons, resulted in a relativistic wave equation for the electron (viewed
as a particle), and predicted the existence of a new particle, the positron, or antielectron. The electron was still treated as a particle in this theory. It is only when one
quantizes the Dirac wave function of the electron, by elevating it to the status of
an operator, that the true QFT of light and matter arises. In this theory, localized
electrons can be described in terms of non-relativistic wave-packet modes.
In this chapter, we begin with a review standard quantum optics the Dirac,
monochromatic theory of electromagnetism in free space [37], and its wave-packet
(non-monochromatic) counterpart, developed by Titulaer and Glauber in 1967 [88].
This is followed by a derivation and review of the single-photon wave function in
coordinate space, which has developed over the past decade [59, 60, 68, 73]. We
then show that both the monochromatic and non-monochromatic quantized theories
can be derived directly from quantization of the single-photon wave function. In

25
doing so, we also show that the scalar product for the single-photon wave function,
which is a non-local integral in coordinate space [68, 73], gives the appropriate
overlap of the Titulaer-Glauber (TG) wave-packet states. This is closely connected to
the well-known non-orthogonality of the TG wave-packet modes under the standard
scalar product. The single-photon wave function theory is then extended to multiphotons states, culminating in a complete wave mechanics theory of photons. Close
connections between classical coherence theory, photo-detection theory, and photon
wave mechanics are given explicitly. We discuss how to treat entanglement in the state
of two photons, and how to correctly reduce a two-photon state to a single-photon
density matrix. The distinction between the modes of a quantum field and states
of a particle are then discussed. We end the chapter with an example calculation of
multi-photon interference.

Quantum Optics

The quantized electromagnetic theory first developed by Dirac starts from the
Maxwell theory of classical electromagnetism. The positive-frequency part of the
classical electric and magnetic-induction fields may be expanded in monochromatic
modes in free space as [25, 52, 87, 90]
r
d3 k
~k
E (x, t) = i
k, uk, (x) eik t ,
3
2
(2)
0

r
X Z d3 k
~k
k
(+)
B (x, t) = i
k,
uk, (x) eik t ,
3
2
c
|k|
(2)
0

(+)

XZ

(2.1)

(2.2)

26
where ~ is the Planck constant divided by 2. Here 0 is the permittivity of the
vacuum, the sum is taken over two orthogonal polarization indices , and k = c |k|
is the frequency associated with a given wave vector k. Here E(+) (x, t) is implicitly
assumed to be the transverse part of the electric field. The monochromatic, planewave modes are
uk, (x) = ek, eikx ,

(2.3)

where the ek, are unit polarization vectors. The Hamiltonian for the transverseelectromagnetic field expressed in terms of the normal-mode expansion coefficients
k, , is
H=

XZ


d3 k ~k

k,
k,
k,
k, .
(2)3 2

(2.4)

One may express these quantities in terms of the real-valued, canonically-conjugate


variables qk, and pk, , known as the quadrature amplitudes, given by
r
qk, =


k

k, + k,
,
2~

r
pk, = i


1

k, k,
,
2~

(2.5)

which arise in the classical Hamiltonian and Lagrangian treatments of electrodynamics.


With the help of the inverse relations,
r
k, =

k
2~

pk,
qk, + i
k


,

k,

r
=

k
2~

pk,
qk, i
k


,

(2.6)

this leads to a Hamiltonian of the form


1X
H=
2


d3 k
2 2
2
3 k qk, + pk, ik {qk, , pk, } ,
(2)

(2.7)

27
where {qk, , pk, } = 0 is the Poisson bracket of the two canonically-conjugate variables.
This is formally equivalent to a sum of Hamiltonians for classical harmonic oscillators
with resonant frequencies k . In the canonical-quantization scheme of fields, the
fields become operators, in which canonically-conjugate amplitudes, now operators,
obey the following commutation relations (note that for Fermion fields, the conjugate
operators obey anti-commutation relations)
[
qk, , pk0 ,0 ] = i~,0 (2)2 (3) (k k0 ),

(2.8)

where the caret notation emphasizes that these are now operators. Here ,0 is a
Kronecker delta, which is non-zero only when the subscripts are equal, and (3) (k k0 )
is a three-dimensional Dirac delta function. Upon quantization, the normal-mode

amplitudes k, and k,
become annihilation and creation operators a
k, and a
k, ,

respectively [25, 52, 87, 90]. These operators obey the inherited commutation relations
h

a
k, , a
k0 ,0

= ,0 (2)2 (3) (k k0 ) .

(2.9)

The state space of the free electromagnetic field HR , (R implies radiation field)
on which the field operators act, consists of a tensor product of the state spaces of
an infinite number of harmonic oscillator states,

HR = Hj ,

(2.10)

j=1

where Hj , j = (k, ), is the harmonic-oscillator state space associated with the wavevector and polarization pair j = (k, ). One possible orthonormal basis of Hj is

28
{|nj i}, where nj = 0, 1, 2, 3, . . . labels the energy levels of the harmonic oscillator j.
Thus HR may be expanded in the orthonormal number basis {|n1 i |n2 i |nl i } =
{|n1 , n2 , . . . , nl , . . .i}, which is often referred to as the Fock basis. The vacuum state
of the electromagnetic field is the state with all harmonic oscillators in the ground
state, nj = 0 for all j
|vaci = |01 i |02 i |0l i = |01 , 02 , . . . , 0l , . . .i .

(2.11)

The mode-j creation operator a


j , acting on the vacuum state of the electromagnetic
field creates a single field excitation, quanta, or photon, with wave vector k, and
polarization ,
a
j |vaci = a
j |01 , 02 , . . . , 1j , . . . , 0l , . . .i .

(2.12)

The positive-frequency parts of the electric and magnetic-induction field operators


may be expressed in terms of the annihilation and creation operators as
(+) (x, t) = i
E

XZ

(+) (x, t) = i
B

XZ

d3 k
(2)3

d3 k
(2)3

~k
a
k, uk, (x) eik t ,
20

(2.13)

~k
k
uk, (x) eik t ,
a
k,
20
c |k|

(2.14)

with the negative-frequency parts given by the respective Hermitian conjugates,


h
i
h
i
()
(+)
()
(+)

E (x, t) = E (x, t) and B (x, t) = B (x, t) . The entire electromagneticfield Hamiltonian operator may be expressed in terms of the annihilation and creation
operators as
=
H

XZ



d3 k
1

~k a
k, a
k, +
.
2
(2)3

(2.15)

29
k, as
This may be written in terms of the number operators n
k, = a
k, a
=
H

XZ



d3 k
1
.
~k n
k, +
2
(2)3

(2.16)

This is the standard approach taken in the quantum-mechanical treatment of the


electromagnetic field. The interaction of the quantized electromagnetic field with
atomic systems is typically introduced through an interaction term (usually the dipole
interaction in non-relativistic treatments) in the full atom-field Hamiltonian. We will
not go into the details of this, as it has been treated elsewhere [25, 52, 87, 90], and it
is not imperative to the current discussion.
The free-space field operators in Eqs. (2.13) and (2.14) are expanded in terms of
infinite plane waves, which work fine for simple models. However, when interactions
with localized objects are considered, such as atoms or molecules, the plane-wave
description of the electromagnetic field operators becomes insufficient. In such a
case, one may expand the electric and magnetic-induction field operators for a given
polarization = 1, in terms of non-orthogonal, polychromatic, spatio-temporal
wave-packet modes vl, (x, t) [88]. The non-orthogonal, polychromatic modes are
related to the monochromatic, orthonormal, plane-wave modes uk, (x), through the
unitary transformation Ul, (k) by
r Z
~
d3 k
k Ul (k) uk, (x).
vl, (x, t) = i
2
(2)3

(2.17)

This may be inverted using the unitary relation


X
l

Ul (k) Ul (k0 ) = (2)3 (3) (k k0 ) ,

(2.18)

30
to give
uk, (x) = i

X
l

2 ik t
e Ul (k) vl, (x, t).
~k

(2.19)

The annihilation and creation operators a


k, and a
k, are also changed by this unitary
transformation, leading to new annihilation and creation operators bl, and bl, given
by
bl, =

d3 k
k, ,
Ul (k) a
(2)3

(2.20)

which obey boson commutation relations


h
i
bl, , b
m, = l,m , .

(2.21)

Assuming circular polarization, as we do throughout this chapter, the positivefrequency parts of the electric and magnetic-induction field operators may then be
expressed in terms of the non-monochromatic modes for each polarization , as
1 X
(+)
E
vl, (x, t) bl, ,
(x, t) =
0 l

(2.22)

i X
(+)
B
vl, (x, t) bl, .
(x, t) =
c 0 l

(2.23)

and

Here we have made use of the following relationship between the unit polarization
vectors (for circular polarization) to simplify the expression for the magnetic-induction
field [68, 73]
k
ek, = iek, .
|k|

(2.24)

The full positive-frequency parts of the electric and magnetic-induction field operators
are given by adding together the two polarization parts.

31
One particular advantage to using the plane-wave expansion, or any other complete,
orthonormal expansion of monochromatic mode functions, denoted generically by
uk, (x), is that they are orthonormal under the standard definition of the scalar
product given by
Z
huk, | uk0 ,0 i =

uk, (x) uk0 ,0 (x) d3 x = (2) (3) (k k0 ) ,0 ,

(2.25)

where the last equality holds only for the monochromatic modes. This is not the case
for the non-monochromatic modes, in which the non-orthogonality under the scalar
product in Eq. (2.25) can be seen to arise from different weightings given to different
frequency components due to the

k factor in Eq. (2.17). The wave-packet modes

vl, (x, t), are not orthogonal under a scalar product of the form in Eq. (2.25). This
is one major disadvantage to using such an expansion. However, as we will show, the
weighting of the different monochromatic modes by the

k factor may be canceled

out by defining a new scalar product for the wave-packet modes, which leads to a
different interpretation of these mode functions.

From Einstein to Maxwell Deriving the Single-photon Wave Function

In the previous section we reviewed the standard Dirac quantum theory of electromagnetism in vacuum [37]. The classical Maxwell fields were raised to the status
of operators acting on a Hilbert space, or state space, of the electromagnetic field.
These operators obey the Maxwell equations. This is in contrast to the historical

32
and contemporary quantum-mechanical treatments of the electron, in which case
the single-particle wave function is first introduced to describe the evolution of one
electron (typically in the context of non-relativistic Schrodinger evolution). Then the
relativistic Dirac equation of the electron field is introduced and the electron field is
quantized.
Here we aim to show that this particle-like approach to the electron may also be
applied to the case of the photon. We begin by following closely Diracs approach
to finding the equation of motion for the single electron, a spin-1/2 particle, from
Einstein kinematics. After arriving at the equations of motion for a single photon,
taken as a particle-like object with zero rest mass and spin-1, we discuss the scalar
product and normalization of the wave function, showing that the scalar product
must be a non-local integral in coordinate space.
We begin by reviewing the approach taken by Dirac to arrive at the relativistic
equation of motion for the electron, now called the Dirac equation, which led to
the prediction of its anti-particle, the positron. Starting from the Einstein energymomentum-mass relationship

E 2 = c2 |p|2 + mc2

2

(2.26)

one can easily arrive at a relativistic theory for scalar fields, the well-known KleinGordon equation. This is done by multiplying both sides of Eq. (2.26) by the
scalar wave function (x, t), and replacing the energy and momentum with their

33
corresponding quantum operators
E i~t ,

(2.27)

p i~,

(2.28)

which leads to
t2

(x, t) = c (x, t) +

mc2
~

2
(x, t) .

(2.29)

Note that the Schrodinger equation may be derived from Eq. (2.26) by taking
the positive square root of both sides, and expanding the right-hand side (RHS) for
non-relativistic particles, i.e., c |p|  mc2 , to give
E=

c2 |p|2 + (mc2 )2 mc2 +

c2 |p|2
.
2mc2

(2.30)

Dropping the constant rest energy of the particle mc2 , making the canonical operator
substitutions (2.27) and (2.28), and multiplying by the wave function (x, t), we
arrive at the free-space Schrodinger equation
i~t (x, t) =

~2 2
(x, t) .
2m

(2.31)

There are difficulties that arise when one tries to treat (x, t) in the Klein-Gordon
equation, Eq. (2.29), as a wave function for a particle [9193]. The modulus squared
of the function (x, t) is not positive definite, and therefore cannot be interpreted
as a probability. This negative probability arises from the square energy term
in Eq. (2.26), or the second derivative in time in Eq. (2.29), which do not enter
into the Schrodinger equation. To remedy this negative-probability issue, Dirac tried

34
to find an equation linear in both the time derivative and spatial derivatives. He
reasoned that in order to satisfy special relativity, both space and time had to be
treated on equal footing, and in order to avoid the negative-probability issues, the
time derivative had to be linear, and thus so did the spatial derivatives. The next step
in Diracs approach was to take the positive square root on each side of Eq. (2.26),
without making any approximations. This was accomplished by changing the number
of components of the wave function on which the equation was to operate, going from
a one-dimensional scalar, to a four-component wave function. The energy term on
the left-hand side (LHS) of Eq. (2.26) was multiplied by the 4 4 identity matrix,
I44 . On the RHS of Eq. (2.26), Dirac looked for a set of matrices = (x , y , z ),
and 0 , now known as Dirac matrices, so that

0 mc + c p

2

= c |p| + mc


2 2

I44 ,

(2.32)

which would allow for one to take the positive square root of both sides of Eq. (2.26).
For Eq. (2.32) to be satisfied, the matrices = (x , y , z ) and 0 must obey the
Dirac anti-commutation relations

{l , m } = I44 l,m ,

in order that the cross terms that arise when squaring the LHS of Eq.

(2.33)

(2.32)

cancel each other. The anti-commutation relation works only when the number of
components of the wave function is four. This leads to a wave equation of the following

35
form


c2
i (t + c ) 0 m
I44 (x, t) = 0.
~

(2.34)

The Dirac matrices are also related to the generators of rotation for spin-1/2 particles,
usually taken to be the Pauli matrices. Equation (2.34) is one form of the Dirac
equation, which may be recast into a more explicit Lorentz-covariant form, by changing
the representation of the Dirac matrices so that
(i m) (x, t) = 0.

(2.35)

Here we have used the God given units [92] in which ~ = 1, c = 1, and the 4 4
identity is implicitly assumed present with the mass term. From here one typically
identifies two components of the four-component wave function with the two-spinor
wave function for the electron, and the other two components are identified with the
two-spinor wave function of the positron. The wave function is then quantized in the
usual way [92, 93]. The term spinor, short for spin-tensor, arises from the treatment
of general internal degrees of freedom (the number of internal degrees of freedom is
related to the spin) of particles under rotations. In particular, a spin-1/2 particle,
such as the electron, has two internal degrees of freedom, leading to two-independent
geometric components.
This treatment of the electron may be replicated for the photon, treated as a
spin-1 particle with zero rest mass. Note that the number of components for a spin-j
particle, j = 0, 1/2, 1, 3/2, 2 . . ., is given by n = (2j + 1), and comes from the general
treatment of rotations for n-component wave functions in three dimensions [94]. This

36
leads us to a three-component wave function for the photon. Beginning with Eq.
(2.26) and setting m = 0, we take the square root of both sides, leading to

E = c p p.

(2.36)

In the energy-momentum representation, where a momentum-space wave function


can be well defined, we focus on the transverse, three-component momentum-space
wave function (p). Transversality implies that
p (p) = 0,

(2.37)

so that we may make use of the vector identity




p p (p) = p p (p) + p p (p)
(2.38)
= p p (p) ,
to linearize Eq. (2.36). This leads us to the conclusion that the proper choice for the
Hermitian Hamiltonian operator on the RHS of Eq. (2.36) is
= icp,
H

(2.39)

where we have introduced the label = 1, which we will see corresponds to the
helicity of the photon. When this Hamiltonian is substituted into Eq. (2.36) it gives
the following momentum-space wave equation
E (p) = icp (p) = c |p| (p) .

(2.40)

This equation can be put into a form that more closely resembles the Dirac equation,
with explicit spin dependence, by noting the following feature of the spin-1 matrices

37
and the vector cross product
a b = i (a s) b,

(2.41)

where a and b are ordinary three-component vectors, and s = (sx , sy , sz ) is a threecomponent vector composed of the three spin-1 matrices (generators of rotations for
spin-1 particles)

0 0 0

sx = 0 0 i
,

0 i 0

0 0 i

sy = 0 0 0
,

i 0 0

0 i 0

sz = i 0 0
.

0 0 0

(2.42)

This leads to the following form of the photon Hamiltonian


= icp = c (s p) ,
H

(2.43)

and the corresponding momentum-space wave equation


E (p) = c (s p) (p) = c |p| (p) .

(2.44)

The helicity dependence is now explicitly present, as can be seen by noting that the
helicity operator, i.e., the projection of the spin onto the direction of propagation, is
h=s

p
.
|p|

(2.45)

For completeness, one must treat both helicities on equal footing, which can be
done by creating a six-component, spinor wave function [68, 73]. However, since the
helicities do not mix in free space, we treat each helicity independently.

38
Interpretation of the momentum-space photon wave function (p), must now be
addressed prior to transformation into coordinate space. The momentum-space wave
function (p), is typically interpreted as the probability amplitude in momentum
2




space [60]. This means that (p) d3 p (2~)3 gives the probability of finding a
photon with helicity , and momentum in a momentum-space volume d3 p about p. In
the standard non-relativistic quantum mechanics of massive particles, the momentumspace wave function and the coordinate-space wave function, which is interpreted as
the probability amplitude in coordinate space, are related by a Fourier transform
relationship. However, it is well-known that photons, being inherently relativistic
particles, are non-localizable, and thus have no well-defined coordinate-space wave
function in this usual sense [41]. Thus one may not interpret the Fourier transform
of (p), given by
(LP )

Z
(x, t) =

d3 p i(pxc|p|t)/~
e
(p) ,
(2~)3

(2.46)

as a coordinate-space wave function. Nonetheless, this has been done in the past, and
interpreted, albeit mistakenly, as the photon wave function [58, 64, 66]. Here we have
denoted this pseudo wave function with the superscript LP for Landau-Peierls, the
first to propose this form of the wave function in coordinate space [58]. There are
several reasons for not choosing this function as the true single-photon wave function,
including the fact that the wave function is non-locally connected to the classical
electromagnetic field [59, 60, 68, 73].

39
To obtain the coordinate-space representation of the wave function and wave
equation (2.44), instead of the standard Fourier transformation of the momentumspace wave function, we opt to weight the transformation with a function of the
magnitude of the momentum (equivalently, the energy) [59, 60, 68, 73], and make the
standard operator substitutions Eqs. (2.27) and (2.28). This leads to the following
form of the coordinate-space wave function
Z
(x, t) =

d3 p i(pxc|p|t)/~
f (|p|) (p) ,
3e
(2~)

(2.47)

where the weighting function f (|p|), is yet to be determined. One way of obtaining
the weighting function f (|p|), is to note that the only localizable, scalar quantity
that can be associated with a photon is its energy [60, 68, 73, 95]. Indeed, it has
been shown that for massless particles with spin greater than one, even the energy is
non-localizable [95]. This leads us to a weighting function of the form
f (|p|) =

p
c |p|,

(2.48)

so that the coordinate-space wave function defined in Eq. (2.47) is equal to the
energy-density amplitude. For this reason we refer to the wave function defined in
Eq. (2.47), with f (|p|) =

c |p|, as the energy-density amplitude or energy-density

wave function or simply the Bialynicki-Birula-Sipe (BB-S) wave function, after the
people who proposed this as the wave function [59, 60, 68, 73]. The wave equation
obeyed by any function of the form given in Eq. (2.47) is given by
i~t (x, t) = ~c (x, t) = i~c (s ) (x, t) .

(2.49)

40
Note that Eq. (2.49), with the zero-divergence condition, Eq. (2.37), which in
coordinate space is given by
(x, t) = 0,

(2.50)

are equivalent to the complex form of the Maxwell equations given in Chapter I, Eq.
(1.3). Also note, the non-local Landau-Periels wave function in Eq. (2.46) obeys the
same wave equation as the BB-S wave function. However, they have very different
interpretations, Lorentz-transformation properties, and normalizations [59, 60, 68,
73].
The probabilistic interpretation of the photon wave function, and scalar product
of two different single-photon wave functions are most clearly defined in momentum
space. The probabilistic interpretation of quantum mechanics requires a definition of
the scalar product between two different states and , that is used in calculating
transition probabilities. We denote the scalar product of photon wave functions with
the non-standard notation ( k ), to emphasize that this is not the usual scalar
product. The Born rule states that the modulus squared of the scalar product of
two normalized wave functions |(k )|2 , is to be interpreted as probability of finding
a photon in state , when it is known to be in state . The probability is a real,
dimensionless number, and, being a true observable, must be invariant under all
Lorentz transformations. A choice as to the dimensions and interpretation of the
photon wave functions in momentum-space must be made prior to the determination
of the form of the Lorentz-invariant scalar product. As stated above, the single-photon

41
wave function in momentum space (p), is equal to the probability amplitude
for finding the photon with polarization (helicity) , and momentum p, that is,

2


(p) d3 p is interpreted as the probability of finding the photon with polarization ,
and momentum in a momentum-space volume d3 p about p [60]. The normalization for
single-photon momentum-space wave functions, taking both helicities into account,
is given by
(k ) =

XZ

d3 p
(p) (p) = 1.
(2~)3

(2.51)

This form of normalization imposes constraints on how the momentum-space wave


functions behave under Lorentz transformations. Lorentz-transformation properties
of the momentum-space wave functions may also be determined from the Lorentztransformation properties of the energy and momentum expectation values, which
take on the following forms
  Z
=
H

d3 p

3 c |p| (p) (p) ,


(2~)

(2.52)

d3 p

3 p (p) (p) .
(2~)

(2.53)

and
  Z
=
P

Here the notation ( ), means the expectation value, which comes from taking the
ensemble average over all possible photon wave functions. To find the form of the
Lorentz-invariant scalar product, we first invert the transformation in Eq. (2.47),
Z
1

(p) =
d3 x (x, t) ei(pxc|p|t)/~
f (|p|)
Z
(2.54)
1
3
i(pxc|p|t)/~
=p
d x (x, t) e
.
c |p|

42
Then upon substitution into the momentum-space scalar product in Eq. (2.51), this
gives the scalar product in coordinate space
1
(k ) = 2
2 ~c

dx

d3 x

(x, t) (x0 , t)
.
|x x0 |2

(2.55)

For a more detailed discussion of the Lorentz-transformation properties of the photon


wave function see Appendix A, and references therein.
Note that this is a non-local integral, which is not surprising in light of the fact
that the photon number is not a local quantity in coordinate space [51]. It is also
interesting to note the following forms for the energy and momentum expectation
values in the coordinate-space representation [59, 68, 73]
Z
Z


1
1

0
3
3 0
| =
k H
d
x
d
x
2 (x , t) H (x, t)
0
2 2 ~c
|x x |
Z
Z
1
1
0

0
= 2
d3 x d3 x
2 (x , t) (i~cs ) (x, t)
0
2 ~c
|x x |
Z
= d3 x (x, t) (x, t) ,

(2.56)

and
Z
Z
1
1

3
3 0
0
| =
d
x
d
x
k P
2 (x , t) P (x, t)
2
0
2 ~c
|x x |
Z
Z
1
1
0

0
(2.57)
= 2
d3 x d3 x
2 (x , t) (i~) (x, t)
0
2 ~c
|x x |
Z
1
=
d3 x (x, t) (x, t) .
2ic


| implies that the operator O
act on the ket to its
Here the notation k O


right, | ), and the double lined bra ( k, indicates the double integral, and nonlocal weighting occur in front of the operator. The last line in Eq. (2.56) indicates
that it is the energy (equivalently, the momentum), and not photon number, that is

43
a localizable scalar quantity in the coordinate-space representation. Thus we see that
the wave functions in coordinate space (x, t), are interpreted as energy amplitudes.

Quantization of the Single-photon Wave Function

We may now proceed to quantize the single-photon theory in the same manner
that one does for the Dirac equation [92, 93, 96]. We raise the photon wave function
(x, t), to the status of a field operator. We expand the wave function in modes
{j, (x, t)} that are orthonormal with respect to the non-local norm defined in Eq.
(2.55). The subscripts j, represent spatial and spin (helicity) degrees of freedom.
The expansion amplitudes become annihilation and creation operators, bj, and bj,
respectively. The photon field operator may then be expressed as

(x, t) =

j, (x, t) bj, + H.c.,

(2.58)

j,

where H.c. stands for Hermitian conjugate. The canonical boson commutation relations
for the bj, and bj, operators are
h
i
bj, , b = jl .
l,

(2.59)

The mode functions are orthonormal with respect to the non-local scalar product
defined for the single-photon wave function
1
(j, k l, ) = 2
2 ~c

dx

d3 x

j, (x, t) l, (x0 , t)
= jl .
|x x0 |2

(2.60)

44
The mode functions {j, (x, t)}, are in general non-monochromatic, and thus we may
consider the energy matrix, with non-zero off-diagonal elements Hj,l


|l,
Hj,l = j, k H
Z
Z
1
1
0

0
3
= 2
d x d3 x
2 j, (x , t) Hl, (x, t)
0
2 ~c
|x x |
Z
= d3 x j, (x, t) l, (x, t)

(2.61)

= hj, |l, i ,
where, in the last line we have defined notation for what is typically, (but not in the
in the present theory), taken as the scalar product
Z
hj, |l, i =

d3 x j, (x, t) l, (x, t) .

(2.62)

Note that this quantized field theory, which was arrived at from the single-photon
wave function, is nearly equivalent to the Titulaer-Glauber theory of electromagnetism
discussed earlier, with the exception of the non-local scalar product in Eq. (2.60).
Because the Hamiltonian is Hermitian, we may diagonalize this matrix using a unitary
operator Uj,l . This transforms the modes {j, (x, t)} to another set of modes
n
o
l, (x, t) given by
l, (x, t) =

Uj,l j, (x, t) ,

(2.63)

j,

which are also orthonormal with respect to the scalar product defined in Eq. (2.60).
The operators are also transformed, leading to new creation and annihilation operators,

a
l, =

X
j,

bj, ,
Uj,l

(2.64)

45
and
a
l, =

Uj,lbj, ,

(2.65)

j,

which also obey the standard, bosonic commutation relations


h

a
j, , a
l,

= jl .

(2.66)

In this new basis, the energy matrix is diagonal, giving the following matrix elements

Hj,l


 Z

= j, l, = d3 x j, (x, t) l, (x, t) = ~j jl ,


(2.67)

where we have introduced the eigen-frequencies j . Note that the energy eigenstates
for each polarization (helicity) obey the following eigenvalue equation
j, (x, t) = i~c (s ) j, (x, t)
H
= ~c j, (x, t)

(2.68)

= ~j j, (x, t) .
One can easily verify that this eigenvalue equation is solved by the following mode
functions
r
j, (x, t) =



~j
kj
uj, (x, t) + i uj, (x, t) ,
2
kj

(2.69)

where the spatial vector functions uj, (x, t), are given by
1
uj (x, t) = ekj , ei(kj xj t) .
V

(2.70)

Here the kj are wave vectors orthogonal to the transverse unit polarization vectors
ekj , , and V is the quantization volume (taken to infinity at the end of a calculation)

46
[25, 52, 87, 90]. The photon field operator may thus be expressed as
(x, t) =

j, (x, t) a
j, + H.c..

(2.71)

j,

Note that the energy basis is orthogonal with respect to both the non-local scalar
product of Eq. (2.60), and the standard form for the scalar product in Eq. (2.62),
but is not normalized in the latter case. However, we may use the energy basis to
create a set of orthonormal modes under the standard form of the scalar product
Eq. (2.62). These orthonormal mode functions may be written as
p
j, (x, t) = j, (x, t) / ~j
r 

1
kj
uj, (x, t) .
=
uj, (x, t) + i
2
|kj |

(2.72)

This leads to the following form for the field operator


Xp
(x, t) =

~j j, (x, t) a
j, + H.c.
j,
r

X ~j 
kj
=
uj, (x, t) + i
uj, (x, t) a
j, + H.c.
2
|kj |
j, r

X ~j 
kj
=
ekj , ei(kj xj t) a
ekj , + i
j, + H.c..
2V
|k
|
j
j,

(2.73)

can be written in an identical


Notice that the quantized Reimann-Silberstein vector F,
form as the quantized single-photon field operator Eq. (2.73)
=F
(+) (x, t) + F
() (x, t).
F

(2.74)

(+) , is given by the complex sum of


Here the positive-frequency RS vector operator F
positive-frequency electric and magnetic-induction field operators, Eqs. (2.13) and

47
(2.14),
r
0 (+)
(+)

(+) (x, t)]


[E (x, t) + ic B
F (x, t) =
2
r
X Z d3 k
~k
k
=i
a
k, [uk, (x) + ic
uk, (x)]eik t ,
3
4
c|k|
(2)

(2.75)

where the mode functions uk, (x), are plane wave amplitudes, Eq. (2.1). When the
electromagnetic field operators are expanded in a discrete sum of plane waves instead
of the continuous spectrum given above, the equivalence of the two theories is clear
r

X ~k 
k
(+)
(x, t) =
F
uk, (x, t) + i
uk, (x, t) a
k, + H.c.
4
|k|
k,
r
(2.76)

X ~k 
k
i(kxk t)
ek, e
ek, + i
a
k, + H.c..
=
4V
|k|
k,
Comparing Eqs. (2.73) and (2.76), we see that they are equivalent (up to

2). We

also note that they both obey the complex form of the Maxwell equations, Eqs. (2.49)
and (2.50). This may be summarized by stating that in quantum field theory, the
photon wave functions are the mode functions of the quantized Reimann-Silberstein
vector. Conversely, the quantum field theory of light is constructed by canonicallyquantizing the single-photon wave function.
Indeed, if we examine the real and imaginary parts of the single-photon wave
function and their equations of motion derived from Eq. (2.49), we find

t R (x, t) = c I (x, t) ,

(2.77)

t I (x, t) = c R (x, t) .

(2.78)

and

48
Along with the zero divergence conditions
R (x, t) = I (x, t) = 0,

(2.79)

these equations are identical to the Maxwell equations, with the electric field identified
with the real part of the photon wave function, and the magnetic induction field
identified with the imaginary part of the photon wave function (up to a couple
constants)
R =

0 E,

I =

0 cB.

(2.80)

One should be cautious interpreting this as proof that the electric and magnetic fields
for a single photon actually exist, however. We suggest that the photon wave function
be taken as the fundamental physical object, and that the macroscopic electric and
magnetic fields appear as emergent properties of a collection of many photons. This
is similar to the macroscopic spin associated with a collection of several atoms, which
is determined by weak measurements on the entire ensemble [9799].
In this section, we have shown that the monochromatic Dirac, and polychromatic
Titulaer-Glauber quantized field theories of electromagnetism can be derived from the
photon energy-density amplitude wave function and its equations of motion, in much
the same way that one arrives at the quantum field theory for electrons. The photon
wave function and its equations of motion are found by linearizing the Einstein energymomentum-mass relation for massless, spin-1 particles, and then the single-particle
theory is canonically quantized. We presented the Lorentz-invariant scalar product
of the photon wave function, which is non-local in the coordinate representation.

49
Photon Wave Mechanics

In the previous sections we reviewed the theory of single-photon wave functions.


The main results are the form of the single-photon wave function, its interpretation
as the energy-density probability amplitude, whose modulus squared is related to the
probability of localizing the photon energy at one point in space-time, and the form
of the normalization integral and scalar product, which are non-local in coordinate
space. We also noted a useful connection of this non-local scalar product with the
wave-packet theory of quantum optics. In doing so, we showed how such wave-packet
modes can be made orthogonal with respect to a new inner product.
In this section, we will build upon the single-photon wave mechanics theory,
developing a two-photon wave mechanics, and show explicitly the relationship of this
theory to other well-known theories, such as photo-detection theory, classical and
quantum optical coherence theory [49, 50, 87], and the biphoton amplitude [89, 100
102] that is used in most discussions of spontaneous parametric down conversion
experiments. The theory is then extended to multi-photon states with known photon
number.
The two-photon wave function (2) (x1 , x2 , t), which is related to the probability
amplitude for localizing the energies of the two photons at two different spatial points
x1 and x2 at the same time t, can be expressed in free space as a sum over tensor

50
products of single-photon wave functions [83]
(2) (x1 , x2 , t) =

(1)

(1)
Clm l (x1 , t) m
(x2 , t),

(2.81)

l,m

where the coefficients Clm , symmetrize the wave function, and is the tensor product.
The modulus squared of the expansion coefficients |Clm |2 , gives the probability of the
 (1)
, are solutions
photons being in the states labeled by l and m. The basis states m
of the single-photon wave equations, and the subscript implicitly includes the spin
dependence.
The equation of motion in vacuum of the two-photon wave function, referred to
as the Maxwell-Dirac equation for the two-photon state [83], is given by the sum of
the Hamiltonians for the individual photons
(2)

(2)

i~t (2) = ~c 1 1 (2) + ~c 2 2 (2) ,

(2.82)

where the differential operators are understood to act on the appropriate components
of the tensor product,
(2)

1 = 3 I,

(2)

2 = I 3 ,

(2.83)

and

1 0
,
I=

0 1

1 0
,
3 =

0 1

1 0 0

1=
0 1 0 .

0 0 1

(2.84)

The two-photon wave function also obeys the zero-divergence conditions


j (2) = 0,

j = 1, 2,

(2.85)

51
in which the differential operator acts on the appropriate tensor component.
The spatially-varying refractive index of a linear medium may be treated in a
phenomenological manner, resulting in modified single-photon Hamiltonians [68, 73].
The two-photon wave function correspondingly changes to
(2)

(2)

i~t (2) = ~v1 1 (1 + 1 L1 ) (2) + ~v2 2 (2 + 2 L2 ) (2) ,

(2.86)

where the material dependent quantities are evaluated at the local coordinates L1(2) =


L x1(2) and v1(2) = v x1(2) . The divergence condition is also modified and becomes
(j + j Lj ) (2) = 0,

j = 1, 2.

(2.87)


Here v1(2) = v x1(2) is the local value of the speed of light in the medium given by
p
v (x) = 1/  (x) (x),

(2.88)

where  (x) and (x) are the local values of permittivity and permeability of the
medium. The matrix Lj , (j = 1, 2), is also given in terms of the local values of the
permittivity and permeability of the medium, and is defined as

.
p
p
L (x) = I ln  (x) (x) + 1 ln  (x) / (x) 2
where

(2.89)

0 1
.
1 =

1 0

(2.90)

Tracing over the tensor product of the Hermitian conjugate of the two-photon wave
function with itself, and integrating over all space gives the expectation value of the

52
product of the two photons energies
Z Z



T r (2) (2) d3 x1 d3 x2 = hE1 E2 i .

(2.91)

If the state of the two photons is not entangled, this simplifies to give hE1 E2 i =
hE1 i hE2 i. One may define a related joint probability density for finding the energy
of one photon at the space-time point x1 and the other at x2 , (xj = (xj , t) , j = 1, 2)
(2)
E

(2)

(x1 , x2 , t) = T r

(2)

(x1 , x2 , t)

i.
(x1 , x2 , t) hE1 E2 i,

(2.92)

and a corresponding current density


(2)
jE

n
o.

(2)
(2)
(x1 , x2 , t) = T r (x1 , x2 , t) [I 3 s + 3 s I] (x1 , x2 , t) hE1 E2 i,
(2.93)

where the trace is taken over the identity component of the tensor product in square
brackets. This probability density and current density obey a continuity equation
(2)

(2)

t E (x1 , x2 , t) + (1 + 2 ) jE (x1 , x2 , t) = 0.

(2.94)

The state of a two-photon system may also be described by a related two-time


wave function, which we denote by (2) to distinguish it from the one-time wave
function (2) . This two-time wave function is given by the tensor product of two
single-photon wave functions evaluated at different space-time coordinates
(2) (x1 , t1 ; x2 , t2 ) =

(1)

(1)

Cjk j (x1 , t1 ) k (x2 , t2 ).

(2.95)

j,k

This two-time wave function obeys a pair of wave equations


i~

(2)
(2)
(x1 , t1 ; x2 , t2 ) = ~cj j (2) (x1 , t1 ; x2 , t2 ) ,
tj

j = 1, 2,

(2.96)

53
and the zero-divergence conditions
j (2) (x1 , t1 ; x2 , t2 ) = 0,

j = 1, 2.

(2.97)

In a linear medium, the equations of motion for the two-time, two-photon wave
function are modified in essentially the same manner as the one-time equations. One
simply makes the following substitutions
j j + j L j ,

j = 1, 2,

(2.98)

and
c vj ,

j = 1, 2.

(2.99)

These two descriptions of the two-photon state are equivalent under the standard
measurement-collapse hypothesis of quantum mechanics. Consider a two-photon state
described by the wave function in Eq. (2.81). Suppose a measurement of the position
of a photon at time T1 reveals its position to be R1 . The one-time state in Eq. (2.81),
is collapsed into the two-time state
(2) (R1 , T1 ; x2 , t2 ) =

(1)

(1)

Cjk j (R1 , T1 ) k (x2 , t2 ),

(2.100)

j,k

which obeys the same wave equations (2.96) and (2.97), that the two-time wave
function does. Thus we see that the two descriptions are equivalent.
The two-photon wave function contains all obtainable knowledge about the state
of the two-photon system. Of course this only describes a pure-state photon pair.
In general one may construct two-photon density matrices from two-photon wave

54
functions, just as in standard quantum mechanics [103]. However, when one wishes
to calculate the reduced density matrix for a system of photons, care must be taken
when choosing the proper tracing operation. As an example, consider a two-photon
pure state describe by Eq. (2.81). The density matrix for this state is

(2) (x1 , x01 , x2 , x02 , t) = (2) (x1 , x2 , t) (2) (x01 , x02 , t) .

(2.101)

Suppose that we discard all information about the photon labeled by number 2. In
standard quantum mechanics we would simply set x02 = x2 , and integrate Eq. (2.101)
over x2 . However, this is not the proper way to eliminate all information about
photon 2, since this does not eliminate the dimensions of energy for photon 2. In
fact, following this procedure gives you information about photon 2 pertaining to
its average energy. To properly extinguish all information regarding photon 2, one
must use the proper scalar product, which for the case of photons, is the non-local
expression in coordinate space given by Eq. (2.55). Thus the proper reduced density
matrix for photon 1 is given by

(1)

(x1 , x01 , t)

1
= 2
2 ~c

d x2

d3 x02

(2) (x1 , x01 , x2 , x02 , t)


.
|x2 x02 |2

(2.102)

It is important to know when to use the two different normalizations, that


is, scalar products, associated with the photon wave functions. When one would
like to determine the probability of detecting or localizing photon energy, then the
local, modulus-squared formulation, Eq. (2.25), is correct. However, when one is

55
determining overlaps, transition amplitudes, and reduced density matrices, the nonlocal scalar product of Eq. (2.55), should be implemented.
The two-photon wave mechanics just described can be extended to any photon
number. An n-photon pure state-system may be characterized by an n-photon wave
function given by

(n) (x1 , x2 , , xn ; t) =

X
{a}

(1)

C{a} {a}j (xj , t),

(2.103)

j=1

n
o
(1)
l
are a set of single-photon basis states, the C{a} are expansion

where the

coefficients that symmetrize the state, the sum is taken over the set of basis labels
{a} = {a1 , a2 , , an }, and {a}j is the j-th entry of the basis label {a}. The tensor
product is taken over the n different single-photon states evaluated at different spatial
coordinates. The equation of motion for this n-photon wave function is found by
adding the Hamiltonians associated with the different photon coordinates

(n)

i~t

(x1 , x2 , , xn ; t) =

n
X

(n)

~c j j (n) (x1 , x2 , , xn ; t) ,

(2.104)

j=1
(n)

where the j

are a generalization of the matrices defined in Eq. (2.83). As in

the two-photon case, there is a generalized n-time, n-photon wave function which is
related to the one-time, n-photon wave function defined above, through the standard
measurement-collapse hypothesis of quantum mechanics. It is straightforward to see
this relationship, and I shall not go though the argument, but note that it follows
directly from the two-photon case. The reduced density matrix for subsets of photons

56
is obtained in the same manner as for the two-photon case, with a double, non-local
integration for each photon state to be eliminated.

Connections to Classical Coherence and Photo-detection Theories

We now turn to the correspondence between classical coherence theory (CCT) of


vector fields and photon wave mechanics. First, one can see the close connection of
the two-photon wave functions (both one-time and two-time) and their equations of
motion, with second-order optical coherence theory, by first writing out the secondorder correlation functions and their equations of motion. The most general secondorder coherence description of the electromagnetic field is given by the second-order
coherence matrices (tensors) [87]
Ei

hD ()
(+)
E (x1 , x2 ) = E() (x1 ) E(+) (x2 ) = Ej (x1 ) Ek (x2 )
Ei

hD ()
(+)
H (x1 , x2 ) = H() (x1 ) H(+) (x2 ) = Hj (x1 ) Hk (x2 )
Ei

hD ()
(+)
M (x1 , x2 ) = E() (x1 ) H(+) (x2 ) = Ej (x1 ) Hk (x2 )
Ei

()
hD ()
(+)
(+)
N (x1 , x2 ) = H (x1 ) E (x2 ) = Hj (x1 ) Ek (x2 ) ,

(2.105)

where E() , H() are the positive- and negative-frequency components of the electric

and magnetic field vectors, x1(2) = x1(2) , t1(2) are space-time coordinates, and j, k
{x, y, z} label the Cartesian components of the electric and magnetic fields. The
notation hi implies an ensemble average over all statistical realizations of the fields,
and [ ] indicates a 3 3 matrix of correlations functions. These matrices completely
describe the correlations between various components of the electric and magnetic

57
fields at two different space-time coordinates. In particular, they give a complete
description of second-order partial coherence of an optical field (including spatial,
temporal and polarization coherence).
The evolution of these matrices are governed by a set of first-order linear differential
equations, that we call the first-order Wolf equations, which can be derived from the
Maxwell equations
j A +

1
B = 0,
c tj

j F = 0,

j = 1, 2
(2.106)

j = 1, 2.

Here the curl and divergence are understood to act on the appropriate vector in the
tensor-product, the matrix pairs (A, B) {(E, N ) , (M, H) , (N, E) , (H, M )}, and
the matrix F is any of the four coherence matrices. Equations (2.106) completely
describe the evolution of the second-order coherence properties of an optical field
as it propagates. Using these equations, one can show that each component of
the coherence matrices obeys the Wolf equations [87], a well-known set of secondorder differential equations, recently highlighted for their relation to the two-photon
detection amplitude [104]. The Wolf equations can be written as


2j

1 2
2 2
c tj


Flm (x1 , x2 ) = 0,

j = 1, 2,

(2.107)

where Flm is the lm component of any one of the second order coherence matrices.

58
The second-order coherence matrices may be combined into a single complex
matrix that is identical in form to the two-photon wave function

(2) (x1 , x2 ) =

(2)
+1+1

(x1 , x2 )

(2)
+11

(x1 , x2 )
,

(2)
(2)
1+1 (x1 , x2 ) 11 (x1 , x2 )

(2.108)

, labeled by the helicity 1 , 2 (or handedness


where the block matrix elements (2)
1 2
of the polarization) of the fields at positions x1 and x2 respectively, are given by
(2)
(x1 , x2 ) =
1 2

i
0
{E (x1 , x2 ) + [1 N (x1 , x2 ) + 2 M (x1 , x2 )] 0 1 2 H (x1 , x2 )} .
2
c
(2.109)

We call (2) (x1 , x2 ) the second-order, Reimann-Silberstein (RS) coherence matrix,


since it can be easily derived from the complex RS vector Eq. (1.2). In vacuum,
the equations of motion for the second-order RS coherence matrix follow from the
Maxwell equations, and are identical to the equations describing the two-time, twophoton wave function behavior Eqs. (2.96) and (2.97)

i~

(2)
(2)
(x1 , x2 ) = ~cj j (2) (x1 , x2 ) ,
tj

j = 1, 2,

(2.110)

and
j (2) (x1 , x2 ) = 0,

j = 1, 2.

(2.111)

In a linear medium, these equations are modified in the same way as those for
the two-photon wave function as described above. From this result, we see that
the evolution of two-photon states is identical to that of the second-order optical

59
correlation functions. Thus one can make use of the well-developed theories of secondorder optical coherence propagation to describe the behavior of two-photon states.
The n-photon wave function can also be identified with the more general n-th order
coherence tensors [87] in the same way the two-photon wave function is related to the
second-order coherence tensors.
The two-time description has another relationship with CCT. Starting from the
two-time equations of motion (2.96), we divide by i~c and take the time derivative of
both sides, to get
1 2 (2)
(x1 , t1 ; x2 , t2 ) = ij
c2 t2j


1 (2)
(x1 , t1 ; x2 , t2 ) ,
c tj

= j j (2) (x1 , t1 ; x2 , t2 )

j = 1, 2
(2.112)

= 2j (2) (x1 , t1 ; x2 , t2 ) .

Here we have made use of the vector identity F = 2 F + ( F),


and the fact that the wave function has zero divergence. These are nothing more
than the Wolf equations (2.107), which have been previously discussed in relation to
two-photon detection amplitudes [104]. Note that the Wolf equations can be derived
from the two-photon Maxwell-Dirac equations, but the converse is not true. This is
essentially the same relationship between the Dirac equation and the Klein-Gordon
equation for electrons. The two-photon Maxwell equations contain more information
about the evolution of the two-photon field than the Wolf equations. This can be
easily seen from the fact that in order to derive the Wolf equations from the two-

60
photon Maxwell equations, we had to take a derivative, which removes information
about the dynamics.
The relationship between two-photon wave mechanics and quantum-optical photodetection theory is found in the two-photon detection amplitudes [49, 50, 87]. The
two-photon detection amplitude of standard quantum optics is given by



(2)
(+) (x1 ) E
(+) (x2 ) (2) .
AD (x1 , x2 ) = hvac| E

(2.113)

(+) (xj ) is the positive-frequency part of the electric-field operator evaluated at


Here E
the space-time coordinate xj , of detector Dj ,

(j = 1, 2). The two-photon detection

amplitude is proportional to the joint probability amplitude for detecting one photon
at the space-time coordinate x1 , and the second photon at x2 . This has also been
called the biphoton amplitude for the case of the highly-entangled photon pairs
derived from the spontaneous parametric down-conversion source [89]. The twophoton detection amplitude is directly proportional to the real part of the twophoton wave function defined in Eq. (2.81) above. To see this more explicitly, note
that the single-photon wave functions, from which the two-photon wave function is
constructed, can be linked to the classical Maxwell fields through the RS vector.
There the real part of the single-photon wave function is identified with the electric
field and the imaginary part with the magnetic field (up to two constants). Following
this identification in the two-photon case, we see that the two-photon wave function

61
may be written as
(2)
1 2 (x1 , x2 , t) =


i
0 h (+) (+)
(+) (+)
(+) (+)
(+) (+)
E1 E2 1 2 c2 B1 B2 + ic 2 E1 B2 + 1 B1 E2
,
2
(2.114)
(+)

where we have included explicit polarization dependence with the s. Here Ej




(+)
Bj
is the positive-frequency part of the electric (magnetic-induction) field, which
is evaluated at the space-time coordinate xj , and a tensor product is implicit between
each E and B field. Thus if our detector is insensitive to magnetic fields, the real
part of the two-photon wave function is identical in form to the two-photon detection
amplitude. Of course this is not surprising since the photon-wave functions are
based upon energy localization, which triggers photon-counting detectors. The energy
from photons is absorbed by such detectors, and transformed into a photo-current of
electrons. It is clear that the electric-field operators in Eq. (2.113) evolve according to
Maxwells equations, as do the components of the two-photon wave function. Indeed,
one may carry this full circle and note that one may define a two-photon detection
amplitude based on the full electromagnetic field, which is identical in form to the
two-photon wave function. This idea would prove useful if the detector absorbs light
by exciting magnetic dipoles, as in some atomic transitions.

Modes Versus States

There is a direct correspondence between the modes of a quantum-optical (QO)


field and the states of photon wave mechanics (PWM). This relationship can be seen

62
by writing out the states for a few examples commonly encountered in quantum
optics. The simplest example is that of a single-mode, one-photon state, which in the
standard quantum-optical treatment may be denoted by the Hilbert-space ket,
|1 i = a |vaci ,

(2.115)

where a is the creation operator for a single photon in the spatio-temporal wavepacket mode (x, t). In photon wave mechanics, the single-photon state, which
contains all of our information about the photon, is represented by the single-photon
wave function, which can be identified with the mode function of quantum optics
(1) (x, t) = (x, t) .

(2.116)

We see that the state of a field is given by Eq. (2.115), whereas the state of the
particle is given by Eq. (2.116). Both are completely described by specifying the
function (x, t), which plays the role of a mode function in quantum optics and a
wave function of a particle in photon wave mechanics. In either case, the function
(x, t) obeys the complex form of the Maxwell equations (2.49).
To see the equivalence between modes of QO fields and states of PWM more
generally, consider a single-photon pure state in the plane-wave basis. In quantum
optics, the single photon state is given by
(1) X

=
Ck, ak, |vaci,

(2.117)

k,

where, as before, labels the two polarization states associated with a particular
wave vector k. The same state may be expressed in PWM by the single-photon wave

63
function for a given helicity
r
(1)

(x, t) =

Ck, i

~
[ek, + ick ek, ] ei(kxt) .
2V

(2.118)

To see the formal equivalence of the two formulations, we look at the generalized
photo-detection amplitude, defined in terms of the electric and magnetic-induction
field operators


(1)
D (x, t) = hvac| E(+) (x, t) + icB(+) (x, t) (1) ,

(2.119)

where denotes the helicity of the photon state. Combining Eqs. (2.117) and (2.119),
we get the generalized photo-detection amplitude for the state defined in Eq. (2.117)
(1)
D

(x, t) =

r
Ck, i

~
[ek, + ick ek, ] ei(kxt) ,
20 V

(2.120)

where the electric and magnetic-induction field operators are


(+)

(x, t) = i

~
ek, ei(kxt) a
k, ,
20 V

(2.121)

~
k ek, ei(kxt) a
k, .
20 V

(2.122)

k,

and
(+)

(x, t) = i

X
k,

Comparison of equations (2.118) and (2.120) lead to the conclusion that the
single photon wave function of PWM can be given in terms of the generalized photodetection amplitudes in the following form for a given helicity
(1)
(x, t) =

(1)

0 D (x, t) .

(2.123)

64
To state the result in terms of quantum field theory (QFT), the photon wave functions
are the mode functions of the quantized Reimann-Silberstein (RS) field, which equals
(up to constants) E + icB. Conversely, QFT for light is constructed by quantizing
the single-photon wave function.

Measurement-induced Photon Interactions

To demonstrate the use of the photon wave mechanics developed above, we consider
the description of measurement-induced photon interactions that arise when photons
are incident on an optical beam splitter [105108]. Such interactions are the basis of
linear-optics quantum computing (LOQC) [109]. Consider the optical set up shown in
Fig. 1. A pure n-photon state is incident on port 1, and an ancilla photon is incident
on port 2 of an optical beam splitter. The input wave function can be written as the
product of the n-photon state with the ancilla photon state
in (x1 , x2 , t) = (n) (x1 , x1 , , x1 ; t) (1)
a (x2 , t)

(2.124)


n
= (1) (x1 , t)
(1)
a (x2 , t) .
Here the notation n implies a tensor product of n identical states (assuming all
n photons at port 4 are in the same spatial state (1) , and (1)
a is the ancilla-photon
wave function, which could, in general, differ from (1) . The beam splitter performs
a linear transformation of the single-photon wave functions [110112]
(1) (x1 , t) (1) (x3 , t) + r(1) (x4 , t) ,
(2.125)
(1)

(1)

(x2 , t) r

(1)

(x3 , t)

(x4 , t) ,

65
where r ( ) are real reflection (transmission) coefficients, and x3 (x4 ) label the position
of the output ports 3 (4).

FIGURE 1. Schematic set up for measurement-induced interaction. An n-photon


state at input port 1 and a single-photon ancilla state at input port 2 are incident
on a beam splitter characterized by reflectivity, R. The probability of finding one
and only one photon at output port 3, and thus n photons at output port 4 is
PM I = Rn1 [R n (1 R)]2 . When R = n/ (n + 1), one never finds a single photon
at the output port 3. Through the linearity of quantum mechanics, if there is a
superposition of photon number states at input port 1, and a single photon was
found in output port 3, the n-photon component at output port 4 would be completely
removed.

If the photons are indistinguishable, that is, if they have the same coordinate-space
wave functions (1) = (1)
a , the output photon wave function is given by

n  (1)

out (x3 , x4 , t) = (1) (x3 , t) + r(1) (x4 , t)
r (x3 , t) (1) (x4 , t)

n
X
n m nm n  (1)
(nm)
(m+1)  (1)

r
=
r (x3 , t)
(x4 , t)

m=0
m

m  (1)
(nm+1) o
(1) (x3 , t)
(x4 , t)
,

(2.126)

66
where we have used the binomial theorem to get the second line, and

n
n!

m! (n m)! ,
m

(2.127)

is the binomial coefficient. The first (second) terms in the second equality of Eq.
(2.126) are interpreted as having m + 1 (m) photons at output port 3, and n m
(n m + 1) photons at output port 4. The amplitude to find a single (one and
only one) photon in output port 3, and thus n photons in output port 4, is found
from Eq. (2.126) by the projective measurement hypothesis, that is by looking at

 
n
the coefficient of the term (1) (x3 , t) (1) (x4 , t) . This amplitude is given

by AM I (n) =

n1

[R n (1 R)], where R = r2 is the reflectivity of the beam

splitter, and the subscript M I implies the measurement-induced nonlinearity. This


amplitude arises from the interference between the two possible ways in which a single
photon can emerge from the beam splitter at output port 3 the ancilla and all nphotons can be reflected, which can only happen one way, or the ancilla photon and
one photon from the n-photon state can be transmitted, which can occur n different
ways. The associated probability for finding a single photon at output port 3 and
n-photons at output port 4 is PM I = Rn1 [R n (1 R)]2 , given by the modulus
squared of AM I (n). We can see that this can be adjusted between zero and one,
meaning that one can determine with certainty, conditioned on a single-photon output
at output port 3, when an n-photon input state at port 1 will yield an n-photon state
at output port 4. In particular, if there is a superposition of number states incident at

67
input port 1 along with the ancilla photon at input 2 (assuming AM I (n) is adjusted to
be zero), a detection of a single photon at output port 3 guarantees that the n-photon
component of the superposition is not present at the output port 4.
This is a demonstration of the measurement-induced interaction between photons
on a beam splitter. The simplest example of such a nonlinear interaction is the HongOu-Mandel interference (HOMI) experiment [113], in which two identical photons are
incident on a 50:50 beam splitter and always emerge at the same output port. This
can be easily seen as a result of the measurement-induced probability for n = 1 and
R = 1/2, PM I (1) = (1/2 (1 1/2))2 = 0.
If the photons are not indistinguishable in this example, the visibility of the
interference that gives rise to the measurement-induced interaction amplitude will
diminish, reducing the overall effectiveness of the interaction. This can be easily
accounted for in multi-photon wave mechanics, in which the photons are described
by different spatial states. For the case in which the input n-photon state has all
photons in the single photon state (x, t), and the ancilla photon in the state (x, t),
the probability of finding a single photon at output port 3 and n photons at output
port 4, assuming that we do not filter out any other states, is


PM I (n) = Rn1 R2 + n2 (1 R)2 2nR (1 R) ||2 ,

(2.128)

where is the overlap between the two single-photon states

= h| i ,

(2.129)

68
as defined in Eq. (2.25) above. The overlap ranges from zero to one, and the


corresponding probability goes from the classical value of Rn1 R2 + n2 (1 R)2
to quantum limited value of zero. This analysis is particularly nice for treating the
outcomes of HOMI experiments. In the case of only one incoming photon at each
input port, this leads to the coincidence probability

PC.C. = 1 2RT 2RT ||2 ,

(2.130)

where T = 1 R is the transmissivity of the beam splitter, and is again the overlap
of the two, single-photon wave functions. It is clear that for identical input wave
functions, ||2 = 1, and we obtain the usual HOM dip. As the overlap diminishes,
the coincidence probability increases to its classical value of 1 2RT , in which each
photon may be treated as a single particle, having a probability of R (T ) to reflect
(transmit) from (through) the beam splitter, and there are two equally probable ways
to do so (hence the factor of two).

69

CHAPTER 3

MEASURING THE TRANSVERSE SPATIAL STATE OF LIGHT AT THE


SINGLE-PHOTON LEVEL

Introduction

Measurement of physical systems is what makes physics a science as opposed


to simply mathematics. The theories that we concoct are used to describe actual
experiments carried out in laboratories, and with these, we hope that the good
theories actually reflect what is occurring in our experiments. In the previous chapter
we presented two equivalent theories of electromagnetic phenomena, quantum field
theory, and photon wave mechanics. They have been experimentally tested in several
ways, however, until only recently has anyone examined the relationship between
the two. The reasons for this seem to be twofold. First, due to their inherent
relativistic nature, photons behave quite differently from electrons or other massive
particles. This difference between light and matter led to the quantization of the
classical electromagnetic field prior to anyone proposing a photon wave function [37].
Early attempts to define a photon wave function failed for several reasons, not the
least of which was the non-locality of the earliest proposed wave function [58]. Then
in 1949, Newton and Wigner [41] showed that photons were, strictly speaking, non-

70
localizable. This leads to the conclusion that a photon position operator cannot exist,
and thus, there are no localized eigenstates of such an operator. Therefore, positionrepresentation wave functions for photons, describing the probability for localizing
the photon as a particle-like entity, cannot be well defined. Nevertheless, photon
wave functions based on energy density amplitudes are well defined, as reviewed in
Chapter II.
In the field of quantum optics, measurements have played important roles in
shaping the development of theoretical models. In the 1960s, the advent of the laser
and the rise of nonlinear optics led to a plethora of new experimental results that
were in need of good theoretical description. Glauber, Mandel, Wolf, Sudarshan,
and others put forth the foundations of what is now known as quantum optics
and quantum optical coherence theory, in order to describe many of these new
experimental results. Of particular use for our purposes here is the theory of quantum
optical detection theory used to describe single-photon and few-photon detection
events [4951, 114118].
A long-standing question in physics was whether or not a quantum state could
be inferred by measurements on an ensemble, or collection of identically-prepared,
quantum systems. Wolfgang Pauli posed the question of whether the Schrodinger
wave function (x), can be inferred from probability distributions of position and


2
2
momentum, i.e., from | (x)| and (p) [9]. The general problem of quantumstate reconstruction was articulated by U. Fano in his article on density matrices in

71
1957 [10]. The first experimental demonstration of quantum state reconstruction
optical homodyne tomography of the electromagnetic field quadruture amplitudes
was performed by D. T. Smithey, M. Beck, M.G. Raymer, and A. Faridani in 1992,
at the University of Oregon [4]. This experimental demonstration set off the field of
experimental quantum state reconstruction, which is now a commonly used technique
in many laboratories across the globe.
In order to characterize the state of an electromagnetic field, there are several
degrees of freedom that must be examined. In the first quantum state tomography
experiments, the photon number and phase of the electromagnetic field were examined
and characterized. From the quantum-field-theoretic viewpoint, a single field mode
was detected, and the photon number and phase characteristics of that mode were
inferred from measurements made on a collection of identically prepared systems. In
the present case, we consider fields with only a single photon, with fixed polarization
and spectrum, and seek to gain information about the coordinate-space wave function
as defined in Chapter II.

Electromagnetism in the Paraxial Approximation

In typical optical experiments one most often encounters light confined to a beam
geometry with a varying spatial profile in the plane perpendicular to the beam axis
for a given polarization, as depicted in Fig. 2 In this geometry a mode of the field has
four degrees of freedom (DOFs): one polarization DOF, two transverse spatial DOFs,

72
and a spectral DOF. The complete characterization of a spatial mode of light in this
geometry thus requires measurement of all degrees of freedom, except polarization
(where we have assumed a particular polarization).

FIGURE 2.
Longitudinal and transverse wave functions.
(a) Schematic
representation of longitudinal and transverse wave functions. (b) Transverse wave
function probability |T (x, y; z)|2 at the plane z = 0 for a Hermite-Gaussian beam.
(c) Transverse wave function T (x, y; z), evaluated in the plane z = 0, and y = 0, for
a Hermite-Gaussian beam. (d) Longitudinal wave function probability |L (z, t)|2 , as
a function of the constant phase-front coordinate z ct.

Characterization of the polarization state of light has been theoretically and


experimentally discussed for both classical and quantum states of the electromagnetic
field [119]. This essentially equates to measuring the Stokes parameters of the electro-

73
magnetic field. The characterization of the spectral degree of freedom has been
examined in the case of intense electromagnetic fields, utilizing second-order nonlinear
optical interactions. Techniques such as spectral-shearing interferometry for direct
electric-field reconstruction (SPIDER) [120] and frequency-resolved optical gating
(FROG) [121] have been developed to reconstruct the electric-field spectral amplitude
E(), and spectral phase (). The first experiments to fully characterize the the
state of an electromagnetic field were carried out by Smithey et. al. in the early 1990s,
at the University of Oregon [4]. These experiments relied on tomographic techniques
to reconstruct the quadrature-amplitude Wigner distribution function W (q, p), of
small-amplitude electromagnetic fields, such as the non-classical squeezed-vacuum
state. The Wigner distribution function is a quasi-probability distribution function
in phase space that will be discussed more thoroughly later in this chapter. More
recently, Lvovsky et. al. [122] performed similar reconstructions of the quadrature
Wigner distribution for single-photon states.

Characterization of the transverse spatial state of optical beams has been of


interest for quite some time (see Refs. [18, 19] and references therein). A breakthrough
in the mid-1990s occurred with the introduction of a practical tomographic technique
[20] designed to reconstruct the transverse spatial Wigner function W (x1 , x2 , z), in
a particular transverse plane (defined by the z-axis coordinate). The transverse
spatial Wigner function will be detailed in the next section. Other techniques have
since been developed to measure the two-point correlation function (x1 , x2 ) of an

74
electromagnetic field [22, 23, 123], which is directly related to the transverse spatial
Wigner function as shown in the following section. One major drawback with these
techniques is that they are unable to measure extremely low light levels. Overcoming
this limit is a significant advantage of the technique we will discuss in this chapter.
The Maxwell equations in free space can be written in terms of the RiemannSilberstein (RS) vector F (x, t), as (see Eqs. (1.3) and (1.4))
c F (x, t) = it F (x, t) ,
(3.1)
F (x, t) = 0,
where the free-space RS vector F (x, t), is defined as
r
F (x, t) =

0
[E (x, t) + icB (x, t)] .
2

(3.2)

Each component of the RS vector obeys the Helmholtz equation





1 2
2 t F (x, t) = 0,
c
2

(3.3)

which can be found by taking the curl of Eq.(3.1), using the vector identity
F = 2 F + ( F), and the zero divergence of the RS vector, Eq.(3.1). In the
paraxial approximation, we assume that the electromagnetic field, which propagates
along the z-axis, can be described by a RS vector written as a slowly-varying (in z
and t) envelope function f (x, t) multiplied by a carrier wave ei(Kzt) ,
F (x, t) = f (x, t) ei(Kzt) + c.c.,

(3.4)

where K is the magnitude of the wave vector, is the central frequency of the field,

75
and c.c. stands for complex conjugate. The Helmholtz equation can now be written





1 2
1 2
i(Kzt)
2
2
2 t f (x, t) e
= + z 2 t f (x, t) ei(Kzt)
c
c


1
1 2
i(Kzt)
2
2
=e
+ 2iKz + z 2i 2 t 2 t f (x, t) = 0,
c
c
2

(3.5)

where 2 = x2 + y2 is the transverse Laplacian. This may be simplified further


by assuming that the transverse and longitudinal dependence of the RS vector can
be factored as well. This implies the slowly-varying envelope may be expressed (for
constant linear polarization) as
1
f (x, t) = e f (z, t) u (x, y; z) .
2

(3.6)

Here e is a unit polarization vector in the plane perpendicular to the propagation


axis, and the functions f (z, t) and u (x, y; z), are known as the pulse envelope and
the transverse spatial mode functions respectively. Separating out the transverse and
longitudinal dependence of the field by combining Eqs. (3.5) and (3.6) leads to


z2

1 2
+ 2iKz 2i 2 t 2 t f (z, t) = 0,
c
c

(3.7)

for the longitudinal dependence, which we will neglect in the rest of our treatment,
and

2 + 2iKz + z2 u (x, y; z) = 0,

(3.8)

for the transverse dependence. The paraxial approximation implies the transverse
beam dimensions remain small compared to the distance of propagation. This statment
translates into neglecting the second derivative with respect to z in Eq.(3.8) above,

76
giving the paraxial propagation equation

2 + 2iKz u (x, y; z) = 0.

(3.9)

Solutions of this paraxial propagation equation have been studied in great detail, and
lead to three basic families of orthonormal modes known as the Hermite-Gaussian
(HG), Laguerre-Gaussian (LG) and Ince-Gaussian (IG) modes [124130]. In the work
presented in this dissertation, we focus on the transverse spatial mode functions
u (x, y; z) and their characterization, and neglect the pulse envelope f (z, t).
It is worth noting that the differential equation of motion for paraxial propagation
given in Eq. (3.9) may also be expressed in integral form. If the transverse spatial
mode u (x, y; z) is known in one plane z = za , then the transverse spatial mode in
another plane a distance d away is given by the Fresnel integral [125127]
u (x, y; z = za + d) =

(3.10)

In discussing the characterization of electromagnetic fields used in the laboratory,


the traditional approach is to neglect the magnetic field, and examine only the electric
field E (x, t). The motivation for this approach is that most materials, and typical
photo-detectors, respond to the electric-dipole field in the first approximation, and
not the magnetic field. Also, in free space, the electric and magnetic fields are
closely related to one another through the Maxwell equations, and knowledge of one
gives complete information of the total field. Techniques to measure the longitudinal
envelope of a pulse of light have been developed to gain both amplitude and phase

77
information (SPIDER and FROG) [120, 121]. Similarly, there have been several
demonstrations of measurements of the transverse envelopes of optical beams [20
23]. However, to date, there have not been ample experimental techniques developed
to completely characterize these aspects of the electromagnetic field at the singlephoton level. In this chapter we present a new experimental method of characterizing
the transverse spatial state of light at the single-photon level.

The Photon Wave Function in the Paraxial Approximation

Considering an electromagnetic field with only a single photon present, one may
completely describe the state of the field (or photon) by a single-photon wave function,
as discussed in Chapter II. In free space the real and imaginary parts of the singlephoton wave function are closely related, and thus characterization of one ensures
knowledge of the other. This is equivalent to the relationship between the electric
and magnetic-induction fields of classical electromagnetism in free space. As pointed
out in Chapters I and II, the photon wave function for a given polarization (helicity)
is equivalent to the RS vector

(x, t) = F (x, t) .

(3.11)

The single-photon wave function for a given helicity obeys the same first order wave
equations as the RS vector. In particular, each component of the wave function obeys
a Helmholtz equation of the same form as Eq. (3.3) above. One may then proceed

78
along the same lines as done in the previous section, choosing a particular polarization,
assuming the photon to propagate along the z-axis, and splitting the wave function
into transverse and longitudinal dependence. One can thus define longitudinal and
transverse envelopes
1
(x, t) = L (z, t) T (x, y; z) exp [i (Kz t)] + c.c.,
2

(3.12)

where we have dropped the polarization vector. Here L (z, t) is called the longitudinal
(L) wave function, which describes the wave packet along the direction of propagation
as sketched in Fig. 2. The transverse (T ) wave function T (x, y; z), characterizes the
wave packet in the plane perpendicular to the direction of propagation, also sketched
in Fig. 2. The evolution of the longitudinal and transverse wave functions is identical
to the evolution pulse envelope and transverse spatial mode function of the classical
field respectively. Concentrating on the transverse wave function T (x, y; z), we note
that in the paraxial approximation, it obeys the paraxial Helmholtz equation

2 + 2iKz T (x, y; z) = 0,

(3.13)

of the same form as Eq. (3.9) discussed above for the RS vector.
In this way, we can use the techniques developed to measure transverse and
longitudinal envelopes of classical electromagnetic fields, in the characterization of
single-photon states. However, we must be careful to note that simply because
these techniques work in the classical domain, does not imply that they will work
effectively at the single-photon level. For example, the techniques used to characterize

79
the longitudinal envelopes of classical pulses rely on nonlinear interactions between
the field to be characterized and a known auxiliary field. At the level of a single
photon, such nonlinearities are extremely small, and thus lead to practical problems
implementing techniques such as SPIDER and FROG.

The Wigner Distribution Function and Its Properties

A central concept in classical statistical physics is the phase-space probability


distribution function P (x, p; t), which gives the joint probability of a massive particle
at position x having momentum p. Knowledge of the joint probability distribution
P (x, p; t) for a single classical system allows one to predict the probabilities for all
possible outcomes of experiments performed on the system. This concept can be
carried over to the quantum-mechanical regime, where probabilities abound. The
Wigner distribution function [12, 131] W (x, p; t), was the first such distribution to
be introduced. However, the Wigner function cannot be interpreted as a true joint
probability distribution. For example, the Wigner function may be negative, and thus
may not be interpreted as a probability. Also, the Heisenberg uncertainty principle
prevents one from determining position and momentum simultaneously, so does it
even make sense to talk about a phase-space probability distribution in quantum
mechanics? Such issues may be safely swept aside by reinterpreting the Wigner
function as a quasi-probability distribution. The Wigner function has many of the
attributes that one would look for in a probability distribution, as discussed below.

80
The Wigner function for a quantum state is defined in terms of the density operator
as
1
W (x, p; t) =
(~)3

 p 
,
d3 hx | (t) |x + i exp 2i
~

(3.14)

in the position representation. It can also be expressed in the following momentumrepresentation form
1
W (x, p; t) =
(~)3


q x
.
d3 q hp + q| (t) |p qi exp 2i
~

(3.15)

For a pure state, these simplify to the following forms


Z
 p 
1

3
W (x, p; t) =
d

(x

,
t)

(x
+
,
t)
exp
2i
,
~
(~)3
Z

q x
1

3
d
q

(p
+
q,
t)

(p

q,
t)
exp
2i
.
W (x, p; t) =
~
(~)3

(3.16)

The Wigner function has several useful properties that we now list.
(1) The Wigner function is real valued, indicating the information contained in the
complex-valued density matrix, or wave function, is incorporated in a single real
function.
(2) The marginal integrals of the Wigner function yield the probability distributions
in position x, and momentum p
Z
Z

W (x, p; t) d3 x = P (p, t) = hp| (t) |pi ,


(3.17)
3

W (x, p; t) d p = P (x, t) = hx| (t) |xi ,


where is the density operator for the quantum system in question, and |xi
and |pi are the position and momentum eigenstates of the system.

81
(3) The Wigner function is normalized to unity
Z

d3 p W (x, p; t) = 1.

dx

(3.18)

(4) The standard overlap probability between quantum states, |h | i|2 in Dirac
notation, can be calculated from the Wigner functions associated with each
state
2

|h | i| =

dx

d3 p W (x, p; t) W (x, p; t) .

(3.19)

This is a special case of the formula [13]


n
o Z

|xi =
T r AB = d3 x hx| AB

(3.20)

where
1
WO (x, p; t) =
(~)3



(t) |x + i exp 2i p ,
d3 hx | O
~

(3.21)

(= A or B)
is an arbitrary operator.
and O

(5) To treat reduced density matrices, in which one traces over some degrees of
freedom in the density matrix, one must integrate the Wigner function over the
same degrees of freedom [132].

(6) For a normalized pure state, the probability-current vector density j (x, t),
satisfying the continuity equation
j (x, t) + t ( (x, t) (x, t)) = 0,

(3.22)

82
can be obtained from the first moment of the Wigner function in momentum
[133]
Z
j (x, t) =

p
d3 p W (x, p; t)
~

1
=
[ (x, t) (x, t) (x, t) (x, t)] .
2i~

(3.23)

(7) It was pointed out by Royer in 1977 [24] that the Wigner function may be
expressed as the expectation value of a displaced parity operator
E
1 D
1

W (x, p) =
D (x, p) D (x, p)
,
(~)3

(3.24)

is the parity operator which takes x x, (that is


(x, t) =
where
(x, p) is the phase-space displacement operator
(x, t), and p p). Here D


(p

)
(x, p) = exp i
D
,
~

(3.25)

introduced by Glauber in discussion of coherent states [114]. Here


r and p
are
the position and momentum operators (not coordinates). Thus we see that the
Wigner function W (x, p), for a pure state can be thought of as the overlap
of the wave function with its mirror image reflected about x, p, giving a
measure of how much the state is centered about x, p. This interpretation of
the Wigner function as the expectation value of a displaced parity operator has
strong implications for experimental determination of the Wigner function.
The Wigner function for a quantum system completely characterizes the system in
the sense that one may calculate probabilities for all possible outcomes of experiments
from the Wigner function.

Just as one may calculate all possible experimental

83
outcomes from the density matrix (for an arbitrary quantum state) or the wave
function for a pure quantum state, the Wigner function can be put on the same
footing as these mathematical objects. Indeed, on may invert the integrals in the
definition of the Wigner function Eqs.(3.14), (3.15), and (3.16) to give the density
matrix, or in the case of a pure quantum state, the quantum wave function,
0

(x, x ) = hx| |x i =




(x x0 ) p
x + x0
, p exp 2i
,
d pW
2
2~

(3.26)




x + x0
(x x0 ) p
d pW
, p exp 2i
.
2
2~

(3.27)

and
0

(x ) (x) =

The latter can be used to extract the wave function by setting x0 = 0, and then
renormalizing the resulting function. Thus the Wigner function does indeed completely
characterize the quantum state of the system.

The Transverse Spatial Wigner Function

In the case of classical optics, in which the electric field for fully-coherent light, or
the first-order coherence function for partially-coherent light, are used to characterize
the optical field, the Wigner function has found several uses [15, 134]. The first-order
coherence function, also called the mutual coherence function, (and sometimes the
second-order coherence function) is defined as [87]


(1)
jk (x, t; x0 , t0 ) = Ej (x, t) Ek (x0 , t0 ) ,

(3.28)

84
where Ej (x, t) is the j-component of the electric field evaluated at (x, t), and the
angle brackets imply the ensemble average over all possible realizations of the fields.
Partially-coherent light is typically described by temporally-stationary fields, which
implies the field statistics are independent of the exact time, and only time differences
(1)

matter. In this case, the mutual coherence function jk (x, t; x0 , t0 ) is dependent only
on the time difference, = t t0 , so that it may be written as

(1)
jk (x, x0 , ) = Ej (x, 0) Ek (x0 , ) .

(3.29)

The Wigner distribution function used in classical optics is defined in a similar


manner to the quantum case. In the case of partially-coherent light, the first-order
coherence function takes the place of the density matrix, and for fully-coherent light,
the complex electric field (in the sense of analytic signal treatment) takes the place of
the wave function. Thus the Wigner distribution function is a matrix (each component
describing the different polarization coherence properties), with entries
1
Wjk (x, k; t, t ) =
()3
0

(1)
jk


k
,
(x , t; x + , t ) exp 2i
~
0

(3.30)

in the case of partially-coherent light, and


1
Wjk (x, k; t, t ) =
()3
0

Ej


k
(x , t) Ek (x + , t ) exp 2i
.
~
0

(3.31)

Here we have switched from the momentum vector, p to the wave vector k, which
are proptional to one another, p = ~k. In the paraxial approximation, the electric
field may be separated out into longitudinal and transverse dependence as discussed

85
above. In this regime it is often helpful to introduce a transverse field amplitude, and
a transverse spatial Wigner distribution function. Consider a quasi-monochromatic,
linearly-polarized light field propagating in the z-direction. As in Eq. (3.6), the
field is a product of a longitudinal part f (z, t), and a transverse part u (r; z), where
r = (x, y) is the two-component position vector in the transverse plane, and we have
dropped the polarization vector. Focusing on the transverse part, we construct a
four-parameter transverse spatial Wigner function
1
Wu (r, q; z) = 2

d2 u (r ; z) u (r + ; z) exp (2iq ),

(3.32)

where q = (qx , qy ) is the two-component transverse wave vector. We may thus borrow
this form from classical optics to construct a transverse spatial Wigner function for a
paraxial single-photon by simply replacing the transverse spatial amplitude u (r; z),
with the transverse spatial wave function T (r; z)
1
W (r, q; z) = 2

d2 (r ; z) (r + ; z) exp (2iq ),

(3.33)

where we now drop the T subscript. The experiments that we discuss in the next
sections are designed to measure these two quantities.
To illustrate the typical behavior of the transverse spatial Wigner function, we
sketch three examples in Fig. 3. We consider only one transverse spatial dimension,
characterized by the transverse position and wave-vector component (x, kx ). The
darkened circle in Fig. 3a(b) represents a contour of the transverse spatial Wigner
function of a collimated beam. The elongated ovals schematically depict the transverse

86
spatial Wigner function of a diverging (converging) beam. The direction in which the
contours are tilted reflect the fact that for a diverging (converging) beam a positive
position x correlates to a positive (negative) wave-vector component kx . Figure
3.c(d) show pictorial representations of diverging (converging) paraxial beams. The
dotted lines represent the phase-front of the beam, that is, the surface of constant
phase, and the solid curves represent the transverse-beam intensity profile at different
longitudinal coordinates z.

FIGURE 3. Converging and diverging Gaussian beams and associated transverse


spatial Wigner functions. Contour plots of the Wigner functions (a),(b), and pictorial
representations of the transverse beam profiles (c),(d) for a divergent beam (a),(c)
and convergent beam (b),(d). The thick solid vertical lines represent the plane (z
coordinate) in which the Wigner functions are determined.

87
A very useful property of the transverse spatial Wigner function for optical fields is
that they have quite simple transformation properties through linear optical systems.
In particular, a paraxial beam with initial transverse spatial Wigner distribution
W0 (r, q), passing through a linear optical system characterized by an impulse response
function h (r, r0 ; ), leads to an output transverse spatial Wigner function [15]
1
Wf (r, q) =
2

2 0

dr

d2 q 0 K (r, q; r0 , q0 ) W0 (r, q) ,

(3.34)

where the kernel function K (r, q; r0 , q0 ) is given by


0

K (r, q; r , q ) =

d2 h (r + , r0 + ; ) h (r , r0 ; ) e2i(qq ) .
(3.35)

For a first-order optical system, which can be characterized by the ABCD-matrices of


the system [125, 127], the input-output relationship for the transverse spatial Wigner
function simplifies significantly to the form
Wf (r, q) = W0 (Ar + Bq, Cr + Dq) ,

(3.36)

where A, B, C, D are the components of the complete-system ABCD-matrix, which


are assumed to be the same for both the x and y directions, but can be generalized
to the non-symmetric situation [15, 134].

Parity-inverting Sagnac Interferometer

In order to characterize the transverse spatial state of light, assuming constant


polarization, (to second-order correlation in the field) in a given plane defined by

88
the longitudinal coordinate z, one must measure the transverse spatial coherence
function (SCF) (1) (r1 , r2 ; z), or equivalently the transverse spatial Wigner function
WT (r, q; z) discussed above. The first published quantitative measurement of the
transverse SCF used tomography and the inverse Radon transform to reconstruct the
transverse spatial Wigner function from intensity measurements [20]. Improvements
upon this method were presented by Iaconis and Walmsley [22] and by Cheng, Raymer
and Heier [23], in which they introduced common-path (or Sagnac) interferometers
capable of directly measuring the real and imaginary parts of the transverse SCF.
These methods relied on imaging optics and imaging array detectors, limiting not
only their spatial resolution and wave front quality, but also the overall strength of
the fields which could be characterized.

Measurements of extremely weak fields, that is, fields having only a single photon
in the measurement apparatus on average, at any given time, has proven to be quite
difficult. Losses, detector size, dark count rates, and other practical difficulties have
not allowed reconstruction or measurement of the transverse spatial state at the
single-photon level. The measurement method that we introduce below is the first
to address the characterization of the transverse spatial state at the single-photon
level. The method can also be used to characterize classical fields with much higher
intensities than the single-photon level. The measurement scheme, based upon the
fact that the Wigner distribution function is the expectation value of the displaced

89
parity operator [24], directly measures the transverse spatial Wigner function locally
in phase space.
The basic idea behind the measurement is to create two copies of the transverse
field amplitude (or transverse spatial wave function) that have been displaced in
opposite directions in phase space, and combine them. In other words, the field to be
characterized should be transformed into a superposition of the displaced fields
(r) (r + s) eik(r+s) + (r s) eik(rs) ,

(3.37)

where r = (x, y) is the two-dimensional position vector, s = (sx , sy ) is a twodimensional shear of the wave front, and k = (kx , ky ) is a two-dimensional transverse
wave vector. One then observes the average count rate R (s, k), given by the modulus
squared of this amplitude, integrated over the detector surface area (A)
Z
R (s, k) =



(r + s) eik(r+s) + (r s) eik(rs) 2 d2 r.

(3.38)

For a single-photon wave packet, R (s, k) is proportional to the probability for a


photo-count per trial. If the detector area is large compared to the fall-off of the field
(wave function), then the limits on the integral in Eq. (3.38) may be taken to infinity
(implied below), to give
Z

i2kr 2

(r s) (r + s) e

R (s, k) = 2C + 2Re

dr ,

(3.39)

where the constant C is


Z
C=

| (r)|2 d2 r.

(3.40)

90
The second term on the right hand side of Eq. (3.39) is just the Wigner distribution
function for the state (r), (up to an overall proportionality constant). Thus we see
that the average intensity, or count rate, as a function of the phase-space displacement
R (s, k), for a device that can employ this kind of measurement, gives a constant plus
the Wigner function. Note that this is for the case of a fully coherent field or pure
state. For a partially-coherent field or a mixed state, one takes the ensemble average
of the modulus squared, leading to the following expressions for the average intensity
or count rate R (s, k)
Z
R (s, k) = 2Ccl + 2

(1) (r s, r + s) ei2kr d2 r,

(3.41)

in the case of partially-coherent light and


Z
R (s, k) = 2Cqm + 2

(r s, r + s) ei2kr d2 r,

(3.42)

for a mixed single-photon quantum state. Here the constants are similar to the
constant defined in Eq. (3.40), and are given by
Z
Ccl =

(1) (r, r) d2 r,

(3.43)

and
Z
Cqm =

(r, r) d2 r.

(3.44)

Note that we have dropped the z-dependence in the correlation function and density
matrix here to simplify the notation.
The measurement method that we introduce below, is based upon the recent
characterization scheme developed by Mukamel et. al. [123], denoted MBWD, and

91
utilizes a Sagnac (common-path) interferometer to directly measure the transverse
spatial Wigner function. In the MBWD scheme, sketched in Fig. 4, a light beam is
directed into a Sagnac interferometer consisting of a beam splitter, three in-plane
mirrors, and a Dove prism rotated by 45 about the propagation axis. The Dove prism
in the interferometer performs a relative rotation of 180 and a mirror inversion on the
wave fronts of the two counter-propagating beams. In effect, the Dove prism performs
a two-dimensional parity operation on one of the beams relative to the other. The
fields are recombined at the beam splitter and detected on a large-area linear-response
photodiode. The mean photocurrent is directly proportional to the transverse spatial
Wigner function at a phase-space point that is set by the tilt and translation of a
mirror external to the interferometer, labeled M1 in the figures, which controls the
input beam transverse position and propagation direction. This setup is designed
to scan in only one dimension, labeled by x. For a one-dimensional field, that is a
field that varies in only one transverse spatial dimension, then this is sufficient to
characterize the complete transverse spatial state. However, if the field of interest
varies in both transverse directions, then the setup measures the transverse spatial
Wigner function W (sx , sy = 0, kx , ky = 0) in the (sy , ky ) = (0, 0) subspace of the full
transverse phase space.

92

FIGURE 4. Parity-inverting Sagnac interferometer with a Dove prism. Diagram of


the MBWD scheme to measure the transverse spatial Wigner function using a Dove
prism. A steering mirror, M1, directs the wave front to be characterized into the
Sagnac interferometer. The counter-propagating wave fronts pass through a Dove
prism inside the interferometer resulting in two displaced transverse copies of the
input at the output as depicted above.

This direct measurement in phase space eliminates any reconstructive errors that
may occur in the inverse Radon transform used to reconstruct the transverse spatial
Wigner function required with previous methods [20]. The method does not rely
on a CCD array, which leads to spatial-resolution limitations through pixel size,
and requires imaging techniques through the interferometer. Such imaging adds
errors due to aberrations, astigmatisms, and difficulties associated with alignment. In

93
addition, CCD arrays do not work well at the single-photon level (they have high dark
count rates, and are not a well developed technology at present). In spite of these
advantages, the MBWD method of measuring the transverse spatial Wigner function
suffers from astigmatism of the optical wave fronts, due to the Dove prism used
inside the interferometer leading to measurement errors. Furthermore, the ability
of the interferometer to collect divergent light, which is measured by the devices
numerical aperture (NA 0.01), is severely limited by the cross section of the Dove
prism. This also decreases the spatial resolution one can obtain.

Our improved scheme [135] utilizes an all-reflecting Sagnac interferometer as


sketched in Fig. 5. The Dove prism and two of the mirrors inside the interferometer
are replaced by three mirrors in what we call a top-mirror configuration. We refer to
the set of three mirrors as the top-mirror. The action of the top-mirror on the field
amplitudes propagating through it is sketched in Fig. 6 below. If the beam enters the
top-mirror from the left hand side, the wave front of the beam is rotated by +90 .
Whereas if the beam enters the top mirror from the right hand side the wave front is
rotated by 90 . This effectively produces a parity inversion between the left-hand
and right-hand propagating fields. It is interesting to note that the rotation is due to
a geometric (or Berry) phase introduced by directing the beam out of, and back into,
the plane of the table, where the beam usually travels [136]. The rotation angle of the
wave front is given by the solid angle swept out on the unit helicity sphere [136]. The
angles that the beam must make while propagating through the top mirror are given

94
by = 45 and = 60 , leading to a rotation angle of = 90 . In addition to these
changes, our scheme also replaces the photodiode detector with a large-active-area
photon-counting photomultiplier tube (PMT) module. The need for a large active
area device is due to the inherent transverse walk-off of the beam at the output of
the interferometer.

FIGURE 5. Parity-inverting Sagnac interferometer with the top mirror. Schematic


depiction of the experimental setup used to measure the transverse spatial Wigner
function using the top-mirror. A steering mirror, M1, directs the wave front to be
characterized into the all-reflecting Sagnac interferometer. The counter-propagating
wave fronts pass through the top mirror, directing the beams out of, and back into
the plane of the interferometer, resulting in two parity-inverted copies of the input
transverse field at the output.

95
With the top mirror in place the new design no long suffers from the astigmatism
introduced by the Dove prism. Additionally, by using large-diameter mirrors in
the construction of the interferometer, one can significantly increase the numerical
aperture of the device, allowing greater ability in characterizing a wider variety of
light. In particular, it is better suited for the study of highly divergent light that
arise from scattering processes.

FIGURE 6. Rotation of the transverse spatial state caused by the top-mirror. A


clockwise propagating transverse field is rotated by +90 (left) while a counterclockwise propagating transverse field is rotated by 90 (right).

Sagnac Interferometer Diffraction Theory

To see that the Sagnac interferometer depicted in Fig. 5 effectively yields the
output count rate given by Eq. (3.38), one must perform the appropriate linear
optical analysis of the clockwise and counter clockwise propagating fields through

96
the interferometer. Consider the schematic drawing in Fig. 7, which shows the
all-reflecting interferometer with propagation distances, and linear optical elements
relevant to the calculation. To describe the transverse state of the field and determine
in what plane the transverse spatial Wigner function is measured, we start with an
arbitrary transverse spatial state in (x, y), at the z = 0 plane as shown in the figure.
This field undergoes free-space propagation until it encounters the steering mirror
labeled by M1. The field just prior the steering mirror 0 (x, y) is related to the input
field in (x, y) by the Fresnel propagation integral, Eq. (3.10), which we shall denote
0 . Here the subscript refers to the distance d0 traveled from the
with the operator P
input plane to the output plane, that is,

ikd0
0 in (x, y) = ke
0 (x, y) = P
2id0

dx0

0 2
0 2
dy 0 e{ik[(xx ) +(yy ) ]2d0 } in (x0 , y 0 ) .

(3.45)
We shall continue to use operators to describe the evolution of the transverse spatial
states. We developed this operator approach to the diffraction theory of light in
order to simplify the discussion of our Sagnac interferometer. This was done prior to
learning that a similar approach had been introduced in the early 1980s [137].
Continuing to follow the input field, we find that just after the steering mirror,
which has been sheared a distance sx and tilted by the small angle x in the x-z plane,
leads to the field 1 (x, y), which is related to the field just before the steering mirror
the shear operator S
(sx ), and
0 (x, y), through the mirror reflection operator M,

97

FIGURE 7. Parity-inverting Sagnac interferometer. Diagram of parity-inverting


Sagnac interferometer with distances and linear optical elements necessary to
calculate the transverse plane in which the Wigner function is measured. M1 is
the steering mirror.
(kx ), by
the tilt operator T
S
(sx ) T
(kx ) 0 (x, y)
1 (x, y) = M
(3.46)
= 0 ( (x + sx ) , y) exp [ikx (x + sx )] ,
where kx kx is a transverse wave vector, which is proportional to the tilt angle of
the steering mirror x , and the total wave number of the field k. The effect of these
operators on a transverse field are given by
(x, y) eikx x = (x, y) eikx x ,
M

(3.47)

98
for the mirror reflection operator, which takes x x and kx kx , with the
mirrors assumed to reflect in the x-z plane. The action of the shear and tilt operators
are given by
(sx ) (x, y) = (x + sx , y) ,
S

(3.48)

(kx ) (x, y) = (x, y) exp (ikx x) .


T

(3.49)

and

After the steering mirror, the field is propagated a distance d1 to the entrance of the
interferometer, so that the field going into the interferometer is

1M
S
(sx ) T
(kx ) P
0 in (x, y) .
2 (x, y) = P

(3.50)

At the input to the interferometer the field is split by a beam splitter (BS) into
transmitted and reflected fields, which propagate in a clockwise and counter-clockwise
sense through the interferometer respectively. At the output of the interferometer the
transmitted and reflected fields are recombined to give the output field out (x, y),
which equals the sum of the two fields. This implies that we may treat the two cases
(transmitted and reflected) separately, and add the results together at the end to
arrive at the output field. Thus consider the two linear-optical systems in Fig. 8,
which depict the evolution of the transmitted and reflected fields separately.
From Fig. 8 we see that the transmitted and reflected fields at the output of the
interferometer t (x, y) and r (x, y) are given by the following expressions in terms

99

FIGURE 8. Linear-optical evolution of the transmitted (top) and reflected (bottom)


j is the
fields that pass through the parity-inverting Sagnac interferometer. Here P
is the mirror reflection operator, S
is
free-space propagator for a distance dj , M

the shear operator, T is the tilt operator, BS is the beam splitter characterized by
is the top-mirror field rotation
transmission and reflection coefficients t and r, and R
operator for fields propagating through the interferometer in the CCW (+) and CW
() sense.
of the linear-optical-element operators
2M
P
3R
P
3 tP
1M
S
T
P
0 in (x, y) ,
t (x, y) = tP
(3.51)
P
3R
+P
3M
P
2 rM
P
1M
S
T
P
0 in (x, y) ,
r (x, y) = rM
where t (r) is the amplitude transmission (reflection) coefficient of the BS. The minus
sign on the last BS reflection arises from the BS phase shift upon reflection [25],
is the top-mirror image-rotation operator for fields propagating through the
and R
interferometer in the CCW (+) and CW() sense. The top-mirror field-rotation
acting on a field (x, y) gives
operator R
(x, y) = (y, x) ,
R

(3.52)

as can be seen from Fig. 6. Next we note the following properties regarding the
commutation relations of these operators. The mirror-reflection, shear, and field-

100
rotation operators all commute with the free-space propagation operator. This can
be represented symbolically by
h

i
Q
=P
Q
Q
P
= 0,
P,

n
o
M,
S,
R
.
Q

(3.53)

This can be determined by considering the field after propagation and application of
one of these operators. For example it does not matter if one rotates the field and then
propagates a given distance, or propagates the field and then rotates. The resultant
field is the same. Similar analysis leads to the conclusion that the mirror-reflection,
shear and tilt operators all commute with one another.
Also note that the mirror reflections, shearing, and tilting do not commute with
the field-rotation operator, and the tilt operator does not commute with the free-space
propagation operator. That is,
h

i
T
=P
T
T
P
6= 0,
P,
i
h

Q
Q
R
6= 0,
R , Q = R

(3.54)
n
o

Q M, S, T .

Another useful property of these operators is that they are unitary, that is,
Q
= I,
Q

n
o

Q M, S, T, R ,

(3.55)

is the Hermitian conjugate of the operator Q.


For the free-space propagation
where Q
operator, this amounts to the fact that propagating a field in the positive direction a
distance d, followed by propagation in the negative direction the same distance should
=M
so that
give the exact same field. Also note that M
2 = I.
M

(3.56)

101
With these operator relations we can now simplify the output field of the parityinverting Sagnac interferometer. The transmitted and reflected fields in Eq. (3.51)
may be written as

M
S
T
P
0 in (x, y) ,
P
1P
2P
23 R
t (x, y) = t2 M

(3.57)

P
1P
2P
23 R
+M
S
T
P
0 in (x, y) .
r (x, y) = r M
2

The output field out (x, y), is the sum of the transmitted and reflected fields after
propagation through the interferometer

out (x, y) = t (x, y) + r (x, y)




S
T
P
0 in (x, y) .
P
1P
2P
23 t2 R
r2 R
+ M
=M

(3.58)

The detected count rate R, of a large-area detector placed at the output plane of the
interferometer measures the modulus squared of this amplitude, integrated over the
detector area (A)

Z
R

dxdy |out (x, y)|2 .

(3.59)

The detector area may be taken to infinity if the field amplitude falls off significantly
before reaching the edge of the detector (a good approximation in most cases). Using
Eq. (3.58) in Eq. (3.59) to calculate the count rate at a detector placed at the output

102
of the interferometer yields



i
2
2
2

R dxdy MP1 P2 P3 t R r R+ MSTP0 in (x, y)


h


i
r2 R
+ M
S
T
P
0 in (x, y)
P
1P
2P
23 t2 R
M


Z

  

2
2
2

1M
2P
3 P
= dxdy in (x, y) P0 T S M t R r R+ P
h

h


i
P
1P
2P
23 t2 R
r2 R
+ M
S
T
P
0 in (x, y)
M
Z



2
2
= dxdy in (x, y) P0 T S M t R r R+


r2 R
+ M
S
T
P
0 in (x, y)
t2 R
Z



r2 R
+ t2 R
r2 R
+ 1 (x, y)
= dxdy 1 (x, y) t2 R
Z

i
h


R
+ + R
+ R
1 (y, x)
= dxdy t4 + r4 |1 (x, y)|2 r2 t2 1 (x, y) R
Z



= dxdy t4 + r4 |1 (x, y)|2 2r2 t2 Re {1 (y, x) 1 (y, x)} ,
(3.60)
S
T
P
0 in (x, y), is the field just after the steering mirror M1.
where 1 (x, y) = M
This implies that the transverse spatial Wigner function measured by this setup is
defined in the plane of the steering mirror. We have made liberal use of the unitarity
of the operators, and in the last line have used the action of the field-rotation operator
acting on 1 (x, y). Writing out explicitly what 1 (x, y) is in terms of the field
R
just before the steering mirror, Eq. (3.46), the interference term (the last term on

103
the final line of Eq. (3.60)) can be written as
Z

dxdy 1 (y, x) 1 (y, x) =

dxdy 0 ((y sx ) , x) 0 ( (y + sx ) , x)

exp [ikx (y sx )] exp [ikx (y + sx )]


Z
=

dxdy 0 (y sx , x) 0 ( (y + sx ) , x) exp [2ikx y] ,


(3.61)

which is the transverse spatial Wigner function W (sx , sy = 0, kx , ky = 0), measured


in the transverse plane of the steering mirror M1.

Experimental Setup

The transverse spatial Wigner function measurement scheme detailed in Fig. 9


consists of the steering mirror, denoted by M1, the parity-inverting, all-reflecting
Sagnac interferometer, and the detection system. The steering mirror (1 inch diameter,
Thorlabs BB1 E03 mirror) was mounted on translation and rotation stages (Newport
Model 433 and Newport Model RS65) which could be positioned using computer
interfaced actuators (Newport 850G actuators, Newport ESP6000 interface board,
and Newport DCIB control box), which also serve as position indicators via internal
encoders in the actuators. The actuators have an encoder resolution of 0.051m, and
backlash less than 20 m. The steering-mirror reflection surface is centered over the
rotation stage pivot axis by precisely machined mirror mounts (in-house construction)
ensuring proper tilt of the wave fronts. The steering mirror changes the input field

104
entering the interferometer from (r) = (x, y), to (x + sx , y) exp [ikx (x + sx )], as
shown in Fig. 10.

FIGURE 9.. Experimental setup to measure the transverse spatial Wigner function.

The Sagnac interferometer was constructed from a 2-inch-diameter, anti-reflection


(AR) coated beam splitter (Thorlabs BSW14) mounted in a standard optical mount,
and 2-inch-diameter mirrors (Thorlabs BB2 E03) to ensure maximum numerical
aperture (NA) of the device. The top-mirror was constructed from aluminum blocks
machined to position three mirror mounts (Lees mount, Siskiyou IMX200.C3, and
IMX200.C3L) at the appropriate angles, and to capture the largest possible solid
angle (and thus largest NA). At the output of the interferometer an AR-coated, 2-inch-

105

FIGURE 10. Steering mirror motion control setup. Steering mirror without shear or
tilt (left) and with finite shear and tilt (sx , kx ) (right).

diameter lens with a 100 mm focal length (Thorlabs LA1050-B) collects the emerging
light, and focused it onto the active area of the photon-counting photomultiplier tube
(Hamamatsu H7421-50) located at the focus of the lens. As the steering mirror is
tilted the output of the interferometer splits into two different directions, which focus
onto two different positions in the focal plane of the collection lens. This walk-off
must be smaller than half the diameter of the detector active area, thus setting the
minimum focal length of the collection lens.
The photon-counting photomultiplier tube (PMT) module (Hamamatsu H742150) was chosen for its large active area (5 mm-diameter active area with better than
85% spatial uniformity over the entire surface) and quantum efficiency in the near
infrared (approximately 10% at a wavelength of 633 nm). Typical photon-counting

106
experiments use solid-state avalanche photodiodes (APDs) which have much higher
quantum efficiencies (around 85% in the near infrared). However, these detectors
cannot be currently built with active-area diameters much larger than a few hundred
micrometers. This is due to dark counts associated with the larger surface area. The
dark counts of the PMT detectors are typically around 50 counts per second.
The electronics used to count the output of the photon-counting module are shown
in Fig. 11 below. The photon-counting module (Hamamatsu H7421-50) outputs
a transistor-transistor-logic (TTL) electrical pulse, which is approximately 30ns in
duration. The dead time associated with the PMT modules (Hamamatsu H7421-50)
is about 70ns, with a maximum count rate of 1.5 106 counts per second. The TTL
pulses are directed into a BNC breakout box (National Instruments BNC-2121) which
is connected to a counter board (National Instruments NI-PCI-6602) placed inside a
desktop computer (Dell Dimension 2400).

FIGURE 11.. Detection electronics for single-photon Wigner measurements.

To scan a transverse spatial Wigner function in one transverse dimension (say the

107
x-direction) we assume the wave front is symmetric in the other transverse dimension,
(y). A computer program written in LabVIEW executes a raster scan through the
transverse phase space as depicted in Fig. 12. The user enters in the scanning range
for shear Sx , and tilt Tx , given in units of millimeters that the physical actuators are
to move in the positive and negative directions from their initial positions (assumed
to be the (0,0) coordinate in phase space). The number of shear Ns and tilt Nt
points desired (typically a 50-by-50 grid is sufficient), and the time to average each
data point tav are also entered into the program. The program controls the motion
of the actuators, moving them to the negative limit in phase space, defined by
(Sx , Kx = kx ), where k = 2/ is the central wave number of the incident
beam, and x is the maximum tilt angle (in radians) of the steering mirror. The
tilt angle is given by x = Tx /b, where b is the tilt-stage lever-arm distance that
is, the distance from the center of the steering-mirror pivot axis, to the point where
the tilt actuator contacts the tilt rotation stage. The program then takes data from
the counter card, and averages for the time tav , yielding the average count rate Rav
at each phase-space coordinate. The average count rate and position in phase space
is recorded in a buffer, which is to be exported to a spreadsheet file at the end of
the scan. The program then steps to the next point in phase space to take another
data point, and repeats the raster scan back and forth across the phase space until
all Ns Nt -data points have been taken. The raster scan goes back and forth between
the negative and positive limits in shear, but steps through the tilt only once. This

108
scanning scheme was chosen because the tilt actuator moves a much smaller distance
than the shear actuator (0.1 mm for the tilt actuator compared with 1.0 mm for the
shear actuator). This leads to smaller backlash errors for the shear as compared to
what would occur for the tilt. The data is output to a text file at the end of the scan,
which is then plotted in a graphics program.

FIGURE 12.. Raster scan of phase space for a one-dimensional field.

Experimental Results

To demonstrate the ability of the Sagnac interferometer to measure transverse


spatial Wigner functions at the single-photon level and with a large numerical aperture
we examined several fields derived from a highly-attenuated helium-neon (HeNe)

109
laser [135]. For simplicity only one-dimensional fields are considered, that is only
fields that vary in only one spatial dimension, which we denote x. This can be
done by considering only fields that can be factored into functions that depend on
orthogonal coordinates (x, y). If this is not done, then the transverse spatial Wigner
function measured is defined in the phase-space subspace in which (sy , ky ) = (0, 0).
In each measurement that follows the HeNe laser was greatly attenuated, such that
on average, there was at most a single photon in the interferometer at any given time.
We present four different transverse spatial Wigner function measurements below,
along with the corresponding theoretical predictions. We first consider a Gaussian
field whose transverse amplitude is given by

(x) = A exp[(

k 2
x2
1
+
i
)x
]
=
A
exp(ik
),
w2
2R
2q

(3.62)

where A is a normalization factor, w is known as the Gaussian beam waist, R is called


the radius of curvature, and q is known as the complex Gaussian beam parameter
1
2i
1
=
,
q
R kw2

(3.63)

all of which depend on the z-plane in which they are measured. The transverse spatial
Wigner function for this particular field is given by
|A|2 1
1
(kx q + isx )(kx q isx )
WG (sx , kx ) = ( + ) exp[2
]
q + q
2 q q
(ksx kx R)2 w2
w|A|2
2s2
=
exp[ 2x
].
2
w
2R2

(3.64)

The experimental and theoretical transverse spatial Wigner functions for the Gaussian

110
field are shown in Fig. 13. Note that the Wigner function is tilted to the right in the
contour plot, indicating the beam in the plane of measurement is diverging.

FIGURE 13. Experimental (left) and theoretical (right) plots of the transverse spatial
Wigner function of a diverging Gaussian beam.

The next field to be considered is a Hermite-Gaussian (HG), transverse spatial


mode, which is a solution to the paraxial Helmholtz equation (3.13) above. The
various HG-mode amplitudes are given by

k
l,m (x, y) = Nl,m HGl ( 2x/w)HGm ( 2y/w) exp[i (x2 + y 2 )],
2R

(3.65)

where Nl,m is a normalization constant, and w and R are the characteristic Gaussian
beam width and radius of curvature. The subscripts l, m label the order of the mode,

111
and HGl (x) is the l -order Hermite-Gaussian function

HGl (x) = Hl (x) exp(x2 /2).

(3.66)

Here the Hl (x) are the l-order Hermite polynomials. The particular mode that we
examined was the HG10 mode (where we use the notation HGlm for l,m (x, y)), whose
xy-intensity distribution, and field amplitude through the y = 0 slice are plotted in
Fig. 14.

We again only examine the one-dimensional transverse spatial Wigner function


in the x direction, and ignore the y dependence. This leads to a transverse spatial
Wigner function of the form

WHG10 (sx , kx ) w[4

2
2s2
(k Rksx )2 w2
s2x
[ 2x x
]
2w
w
2R2

1
+
(k
R

ks
)
]e
.
x
x
w2
R2

(3.67)

The theoretical and experimental data for this transverse spatial Wigner function are
shown in Fig. 15. Note again that the overall Wigner function shape is tilted in the
positive direction, indicating a diverging beam.

112

FIGURE 14.. Intensity (left) and field amplitude (right) of the HG10 mode.

FIGURE 15. Experimental (left) and theoretical (right) plots of the transverse spatial
Wigner function of a HG10 mode.

113
To demonstrate the interferometers ability to characterize divergent light fields,
we consider the top-hat field t.h. . This field is derived from a collimated, nearlyplane-wave field incident upon a transmitting single slit of width a, placed just before
the scanning mirror, as sketched in Fig. 16. In the experiment, the slit width was
a = 0.4 mm. The top-hat field is described by the step function (x), which is 0
when x < 0, and 1 when x 0
1
t.h. (x) = [(x + a/2) (x a/2)],
a

(3.68)

where a is the full width of the single slit. This leads a transverse spatial Wigner
function of the form

Wt.h. (sx , kx ) =

0, s < a/2

sin [2kx (sx + a/2)] / (kx a) , a/2 s 0

sin [2kx (sx a/2)] / (kx a) ,

0, s > a/2.

(3.69)

0 < s a/2

We plot the experimental and theoretical transverse spatial Wigner functions in


Fig. 17. The triangular peak along the kx = 0 section represents the autocorrelation
of the top-hat function, and the large values of transverse wave vector kx along the
edges (sx = a/2) correspond to diffraction from the edges of the slit. The angular
spread x in the top-hat field is given by the ratio of the range of transverse wave
vector kxmax to the overall wave vector magnitude k, i.e., x = kxmax /k, which is 3
milliradians for the top-hat field.

114

FIGURE 16. Top-hat field amplitude (left) and experimental arrangement for its
construction (right).

FIGURE 17. Experimental (left) and theoretical (right) plots of the transverse spatial
Wigner function for a propagated top-hat field.

115
Note that the plots are slightly skewed. This is due to the fact that the field
at the steering mirror is not precisely a top-hat field given by Eq. (3.68), but such
a field propagated a small distance along the beam axis. This is due to the fact
that we cannot experimentally place the slit exactly at the steering mirror, where
the Wigner function is determined. The propagation distance in the experiment was
40 mm. The Wigner function for the translated field is easily calculated using the
ABCD-matrix formalism given by using Eqs. (3.36) and (3.69) with the free-space
ABCD-propagation matrix for a distance d

1 d
,
Mf.s. (d) =

0 1

(3.70)

which leads to the propagated single-slit transverse spatial Wigner function

Wt.h. (sx , kx ) =

0, sx + kx d < a/2

sin [2kx (sx + kx d + a/2)] / (kx a) , a/2 sx + kx d 0

sin [2kx (sx + kx d a/2)] / (kx a) , 0 < sx + kx d a/2

0, sx + kx d > a/2.
(3.71)

The final field that we consider is a displaced, double-top-hat field d.t.h. . This
field is derived from a collimated, nearly plane wave field incident upon a transmitting
double-slit (slit width a = 0.06 mm, and slit spacing d = 0.28 mm) just before the
scanning mirror, as sketched in Fig. 18. The double top-hat field is given by the sum

116
of a pair of displaced top-hat functions, and can be written as
1
d.t.h. (x) = [t.h. (x + d/2) + t.h. (x d/2)] .
2

(3.72)

This leads to the following transverse spatial Wigner function


Wd.t.h. (sx , kx ) Wt.h. (sx + d/2, kx ) + Wt.h. (sx d/2, kx ) + f (sx , kx ) ,
where the function f (sx , kx ) is given by

0, sx < a/2

2 sin [2kx (sx + a/2)] cos (kx d) / (kx a) , a/2 sx 0


f (sx , kx ) =

2 sin [2kx (sx a/2)] cos (kx d) / (kx a) , 0 < sx a/2

0, sx > a/2.

(3.73)

(3.74)

The experimental and theoretical transverse spatial Wigner functions for the double
top-hat field are shown in Fig. 19. In the plot there are two single-slit peaks of
width a = 0.06 mm separated by the slit displacement d = 0.28 mm. The sinusoidal
interference region between the peaks is a sign of the mutual coherence between the
two displaced top-hat field sources, and would be absent if they were incoherent,
that is, if they had random relative phases. The plots are tilted here, as in the tophat distribution, due to the fact that the transverse Wigner distribution measured is
not exactly at the plane of the slits, but rather at the plane of the steering mirror,
which is a distance z = 25 mm after the slits. Again this can be accounted for
by using the ABCD-matrix for propagation. The fact that the experimental plots
are slightly smeared is due to the spacing between successive data points, and a

117
smoothing algorithm used to process the data. The smoothing algorithm averages
the nearest-neighbor count rates at a phase-space point to give the count rate for that
point.

FIGURE 18. Displaced double-top-hat field amplitude (left) and experimental


arrangement for its construction (right).

FIGURE 19. Experimental (left) and theoretical (right) transverse spatial Wigner
functions for the displaced top-hat field.

118
All of the experimental plots presented were constructed from the raw photodetector count rates, which averaged around 100,000 counts per second. The smoothing
algorithm is first applied. Then the constant background, taken to be an average of
the count rates beyond the Wigner function structure in phase space, was subtracted
from the raw data. After subtraction, the resultant data is multiplied by 1 (from
the beam-splitter phase shift) and numerically normalized to unity.

119

CHAPTER 4

TWO-PHOTON TRANSVERSE SPATIAL-STATE CHARACTERIZATION

Introduction

Single-photon states and the single-photon wave function discussed in the last
chapters are interesting from a fundamental viewpoint. However multi-photon states
hold unique properties such as the quantum coherence between different photons,
known as entanglement. Indeed, entanglement between photons has been utilized
to demonstrate several of the salient features of quantum theory, such as the nonlocality of quantum correlations through the violation of Bell inequalities [13, 32].
More recently, entangled two-photon states produced from spontaneous parametric
down conversion have been utilized in numerous experiments demonstrating the basic
elements of linear-optics quantum computing and quantum cryptography systems
[31, 138]. The theoretical description and experimental characterization of such multiphoton states is essential for proper understanding of these experiments. In many of
these experiments, the photon polarization degree of freedom of the electromagnetic
field is used. There have been other experiments that have examined entanglement in
the frequency-time [139141] and the spatial-momentum [142, 143] degrees of freedom.

120
In this chapter we introduce the two-photon transverse spatial wave function
and its associated two-photon transverse spatial Wigner function. We discuss the
characterization of the two-photon transverse spatial Wigner function by using a
pair of parity-inverting Sagnac interferometers introduced in Chapter III. Then we
examine the special case of spontaneous parametric down conversion, which has
been the workhorse of many proof-of-principle quantum mechanics and quantuminformation experiments. In particular, we discuss the photonic state emitted from
a down-conversion source, and its transverse spatial wave and Wigner functions.
We briefly mention the Hong-Ou-Mandel two-photon interference experiments as a
method for characterizing the temporal coherence of the down-conversion source.

This is followed by a discussion of ghost imaging and ghost diffraction, which


utilize the momentum-entangled photons emerging from a down-conversion source to
perform non-local, correlation-imaging experiments. The term ghost implies that
the image or diffraction pattern of an object placed in the path of one photon
from the down-conversion source is obtained by scanning a point-like detector in
the path of the other photon. The two different photon paths are spatially separated,
and the count rates of the spatially-separated, individual detectors remain relatively
constant, but the coincidence count rate maps out the ghost-image (diffraction)
pattern. We generalize this concept to the idea of remote-state preparation, and
measurement of the transverse ghost Wigner function using a parity-inverting
Sagnac interferometer.

121
This leads us to a discussion of the transmission of photon pairs, entangled in
their transverse spatial states, through a turbulent atmosphere. We examine the
case of photon pairs, entangled in orbital angular momentum, propagating through
a turbulent atmosphere. The entanglement can be described by a quantity called
the concurrence [144], which can be calculated analytically for the particular case we
discuss. Then we describe an experimental setup that enables one to measure the
concurrence as the turbulent atmosphere changes (increasing turbulence), and thus
the disentanglement of the pair of photons.
We discuss how one may use a pair of parity-inverting Sagnac interferometers to
measure the joint two-photon transverse spatial Wigner function for two photons. In
particular we explain how one may use such an experimental setup to violate a Bell
inequality.

Two-photon Transverse Wave Function and Wigner Function

In many two-photon experiments it is necessary to consider only photon-pair


sources in which the pairs are emitted into two different, spatially-separated beams
as sketched in Fig. 20. Such sources are often encountered because momentum
conservation results in pairs of photons emitted in opposite directions. The twophoton wave function introduced in Chapter II, as a tensor product of single-photon
wave functions, can be put into a paraxial form in much the same way as was done in
Chapter III for the single-photon wave function. We assume that the source emits a

122

FIGURE 20.. Paraxial two-photon source.

pure, two-photon state. This can be generalized to mixed states by constructing the
appropriate density matrix from the pure states. Focusing on the transverse spatial
state, we assume that the photons are in identical polarization states, implying that
only a scalar field is necessary to describe the two-photon state. This is equivalent
to choosing only a single, diagonal component of the two-photon wave function Eq.
(2.81), in Chapter II. Pairs with non-identical polarizations may also be treated, but
this does not add to the discussion of the transverse spatial state. We assume the two
photons propagate in a beam-like geometry along the z-axis in opposite directions
as sketched in Fig. 20. To further simplify the situation and bring out the key
features of the two-photon transverse spatial state, we take the photons to be in the
same longitudinal (i.e., temporal wave-packet) state. Under these approximations,
neglecting the polarization and longitudinal parts, the two-photon wave function can
be written as
(2) (r1 , r2 ; z1 , z2 ) =

X
j,k

cij j (r1 ; z1 ) k (r2 ; z2 ),

(4.1)

123
where {j (r; z)} is a complete set of single-photon transverse spatial wave functions
evaluated at (r; z) = (x, y; z), in which r = (x, y) is the transverse position coordinate
in the plane defined by z. One may then define a related two-photon transverse spatial
Wigner function in terms of this two-photon transverse spatial wave function as
W

(2)

1
(r1 , q1 , r2 , q2 ; z1 , z2 ) = 4

d 1

d2 2 (2) (r1 1 , r2 2 ; z1 , z2 )
(4.2)

(2) (r1 + 1 , r2 + 2 ; z1 , z2 ) exp [2i (q1 1 + q2 2 )] .


This treatment may be generalized to any number of photons in a straightforward
manner.

Transverse Spatial Disentanglement of a Photon Pair

To illustrate the utility of the two-photon wave function and its relation to classical
coherence theory discussed in Chapter II, we consider the propagation through a
turbulent atmosphere of two photons that are initially entangled in their spatial
degrees of freedom, as depicted in Fig. 21. In particular, we assume the photons
are emitted from the source in opposite directions in one of two orbital-angularmomentum (OAM) states, described by the Laguerre-Gauss transverse spatial wave
functions [125, 145, 146] ,

p,l (r, ) = Rlp (r) exp (il) / 2,

(4.3)

where r and are cylindrical coordinates. Here Rlp (r) is the radial wave function,

exp (il) / 2 is the angular wave function, l is the OAM quantum number, and

124
p is the radial quantum number. To simplify the discussion and calculation we
assume that both photons, labeled A and B, have the same polarization and quasimonochromatic frequency, which are neglected here, and radial quantum number,
p = 0. Furthermore, we consider orbital quantum numbers of equal magnitudes, |l|,
separately. We take the standard two-photon basis as
B
A
,
p,l
1AB = p,l

3AB
A(B)

where p,l

A
p,l

B
p,l
,

B
A
,
p,l
2AB = p,l

4AB

A
p,l

(4.4)

B
p,l
,

is evaluated at the position of photon A(B). For concreteness, we treat

the initial two-photon pure state




(2)
in = 2AB + 3AB / 2.

(4.5)

The photon pair travels through independent, thin, Gaussian phase-randomizing


atmospheres, modeled by a quadratic structure function [147149]. The structure
function describes the refractive index fluctuations of the atmosphere, and the quadratic
nature allows us to explicitly calculate the output. The density matrix elements at the
output of the turbulence are determined by integrating the radial power distributions

2
r Rlp=0 (r) , for each photon multiplied by the circular-harmonic transform of the
phase correlation function, C , which describes the effect of the atmosphere on the
state [150]. In our model, the phase correlation function for each atmosphere is


i
1 h A(B)
C (r, ) = exp D (r, ) ,
2

(4.6)

125
where
A(B)


2
(r, ) = 2r sin () /rA(B) ,

(4.7)

is the quadratic phase structure function [148] of the aberrations in atmosphere


A(B). Here rA(B) is the transverse length scale of the turbulence (index of refraction
fluctuations) in atmosphere A(B). We find a closed form expression for the circularharmonic transform
C (r, m) =

C (r, ) exp [im] d


0

(4.8)


 

= exp 2 (r1 /rA )2 Im 2 (r1 /rA )2
where m is 0 or 2l, and Im (x) is an mth-order modified Bessel function.

FIGURE 21. Entangled photon pair transmission through a turbulent atmosphere.


The photon pair, initially entangled in their OAM states, l = 1, 2, 3 are emitted
by the source and pass through turbulent atmospheres labeled by A and B.

Being interested in only two OAM states, p,l , for each photon at the output,
we may treat each photon as a qubit, that is a two-state system labeled by l.
Other photon states can be considered to be loss channels [151]. We examine the
loss of entanglement by calculating the concurrence C [144], from the normalized
output density matrix, as a function of the ratio of the optical beam waist to the

126
characteristic turbulence length scale, w/r0 [150]. The concurrence is defined as [144]
C () = max {0, 1 2 3 4 } ,

(4.9)

where the i s are the square roots of the eigenvalues, in decreasing order, of the
non-Hermitian matrix
. Here is the spin-flipped density matrix given by
= (y y ) (y y ) ,

(4.10)

where y is the usual Pauli spin-flip matrices, and implies complex conjugation.
Note that each i is a non-negative, real number. For a maximally entangled state,
C = 1, and for a non-entangled state, C = 0. We assume that the atmosphere
is unmonitored, so that any independent information about its fluctuations is lost,
leading to loss of entanglement. The concurrence verses the ratio of the characteristic
beam waist to turbulence coherence length, w/r0 , is plotted in Fig. 22, for various
initial OAM quantum numbers. This result shows that for a beam waist much
smaller than the turbulence length, w << r0 , the entanglement is more robust to
the turbulent atmosphere. Physically this reflects the fact that the photons will
experience few phase distortions across their wave fronts. These results also indicate
that entangled states with larger OAM values experience less disentanglement through
a turbulent atmosphere. This appears to be due to the fact that scattering from one
OAM state to another depends only on the change in OAM, ~l [20], which must be
supplied by the atmosphere. The atmosphere can, on average, change the OAM of
the light only by a particular amount set by the spatial fluctuations that characterize

127

FIGURE 22. Concurrence and transmission fidelity of an OAM-entangled photon pair


after propagation through a turbulent atmosphere. (a) Concurrence as a function
of the ratio, w/r0 , of beam waist to turbulence length scale, for three different
magnitudes of input OAM. (b) Fidelity of the output state with respect to the input
pure state.
it. We can also calculate the fidelity of the output two-photon state relative to the
input state. For a pure-state input the fidelity is
F (in , out ) = hin | out |in i ,

(4.11)

which is plotted in Fig. 22 for the input state given above in Eq. (4.5). This result
indicates that states with small OAM values, and thus small rms beam width [148],
have better overall transmission than do states with large OAM values. We should
stress that the overall transmission of the OAM states depends on the beam waist.
We conclude that entangled states with smaller waists and larger OAM quantum
numbers will be more robust to turbulence.
To measure the density matrix elements for a given OAM magnitude |l|, one
may employ a quantum state tomography method that uses amplitude holograms to
spatially separate photons with differing OAM values [152]. The basic experimental

128
setup is depicted in Fig. 23. The two OAM-entangled photons are emitted in different
directions from the source and pass through a turbulent atmosphere. Assuming most
of the photon amplitude is scattered in the forward direction (weak-scattering limit),
each beam is directed onto an amplitude, Laguerre-Gauss hologram (LGH), which
we show in Fig. 23. The LGH scatters photon amplitudes with OAM value l in
different well-defined directions, while the zero-OAM beam passes straight through.
The signals from photon-counting detectors positioned to detect the scattered lOAM states are sent to coincidence-logic circuitry. The coincidence rates for different
combinations of detector pairs, and different LGHs, give the density matrix elements
needed to calculate the concurrence.

FIGURE 23. Measuring the OAM density matrix using Laguerre-Gauss holograms.
Schematic depiction of an experiment to measure the decoherence of OAM-entangled
photon pairs traversing a turbulent atmosphere (top). Example Laguerre-Gauss
hologram (bottom).

129
Spontaneous Parametric Down Conversion

The first two-photon experiments to probe the inherently non-classical behavior


of entangled photon pairs utilized photon-pair emission from atomic cascade decay
processes of excited atoms and ions [2, 3, 153155]. These experiments are difficult to
perform for several reasons, not the least of which is the great deal of care necessary
to isolate and store single atoms or ions for the relatively long period of time required
to perform the two-photon experiments many times.
Spontaneous parametric down conversion (SPDC), is an alternative entantled
photon-pair source that first appeared in the mid-1960s in unseeded optical parametric
amplifiers (OPAs) [156159]. The process of parametric down conversion (parametric
amplification) is a nonlinear three-wave mixing process. In the standard OPA, an
intense beam of high-frequency light (frequency p , typically in the blue range of
the spectrum), known as the pump beam, is incident on a nonlinear optical material
(usually a nonlinear optical crystal). The nonlinear medium is pumped to a virtual
energy state by the pump beam as shown in Fig. 24. A second optical beam, called the
signal, with lower frequency s , can be amplified when it is incident on the nonlinear
medium in certain directions, while the pump beam is depleted. Energy conservation
requires that a third beam, known as the idler, be generated at the same time with a
frequency equal to the difference i = p s , as in Fig. 24. Parametric amplification
was known in the radio and microwave regimes prior to the advent of the laser, and

130
was first demonstrated in the optical regime in 1965 [160]. The technique is now
widely used as a tunable source of coherent optical radiation.

FIGURE 24. Parametric amplification process. Energy level diagram (left) and beam
geometry of the process (right).

The quantum-theoretical description of parametric amplification was known prior


to the advent of the laser [161], and a close inspection of this treatment leads to
the prediction of spontaneous parametric down conversion, that is, signal and idler
radiation emitted from the nonlinear medium when neither is present at the input as
sketched in Fig. 25. In the quantum treatment of the down-conversion process, the
interaction Hamiltonian describing the three-wave mixing process is
Z
1

HI (t) =
d3 x PN L (x, t) Ep (x, t)
2
Z
0
()
=
d3 x(2) Ei (x, t) Es() (x, t) Ep (x, t).
2

(4.12)

Here PN L is the nonlinear polarization, Ep is the pump field, typically taken to be a


()
classical field, Es() and Ei are the negative frequency parts of the signal and idler

electric field operators, and (2) is the second-order dielectric susceptibility of the
medium, which couples the pump to the signal and idler fields. The average number

131
of photons in the signal and idler field modes at the output of the medium are [161]

ns (t) = ns0 cosh2 (gt) + (1 + ni0 ) sinh2 (gt)


(4.13)
2

ni (t) = ni0 cosh (gt) + (1 + ns0 ) sinh (gt) ,


where ns0 (ni0 ) is the photon number of the signal (idler) beam at the input. The
parametric-amplification gain coefficient g, is proportional to the second-order nonlinear susceptibility, the medium length, and the pump field amplitude. Notice that
for zero input signal and idler photons ns0 = ni0 = 0, one still obtains, on average,
non-zero signal and idler photons at the output due to the 1 in the second terms of
Eq. (4.13). This extra photon per mode, due to vacuum fluctuations, can be viewed
as vacuum-seeded parametric down conversion, in which a single pump photon is
annihilated, and a signal and idler photon pair are created. Energy and momentum
conservation are usually satisfied in this process, and are written as

p = s + i ,
(4.14)
kp = ks + ki ,
where kj , is the wave vector associated with beam j, (j = p, s, i), which is related to
the photon momentum by p = ~k. Momentum conservation is often referred to as
phase matching. The SPDC process has been theoretically treated quite thoroughly,
and the details are beyond the scope of this discussion. A detailed treatment can be
found in Appendix B and references therein. Below we just highlight the SPDC
treatment.

132

FIGURE 25. Spontaneous parametric down-conversion process. Energy level diagram


(left) and beam geometry of the process (right).

The SPDC process is said to be either type-I or type-II, depending on whether the
down-converted photons (the signal and idler) are emitted with the same polarization,
or with polarizations orthogonal to one another. Additionally, in SPDC experiments
the down-converted photons emerge from the nonlinear medium (usually a nonlinear
optical crystal) in either the same direction as the pump (collinear), or directions
different from the pump (non-collinear). This leads to four possible SPDC schemes
as depicted in Fig. 26.

133

FIGURE 26. Four possible spontaneous parametric down-conversion configurations.


The pump beam, with frequency p and wave vector kp is incident on the nonlinear
crystal (NLC). The signal (s) and idler (i) photons have frequency j , and wave
vector kj , (j = s, i). The signal and idler photons may emerge from the crystal with
wave vectors parallel (bottom) or perpendicular (top) to the pump beam. The signal
and idler may also emerge with polarizations parallel (type-I, left), or perpendicular
(type-II, right) to each other.

A pump photon incident on the nonlinear crystal has a small probability (on the
order of 107 ) to be annihilated and converted into a pair of down-converted photons.
The energy and momentum conservation obeyed in this process Eq. (4.14), lead to
spectral and spatial entanglement of the daughter photons created. The standard
theoretical approach taken to determine the output state of light from the SPDC
process involves treating the electromagnetic field as a quantum field, which interacts
with a bulk, classical nonlinear crystal. One performs a first-order, time-dependent
perturbation expansion in the interaction picture [100, 139, 142, 162, 163] to give the
output state of the signal and idler modes of the electromagnetic field
Z
|f i |vaci +

d ks

d3 ki C (ks , ki ) |1ks i |1ki i ,

(4.15)

134
where |vaci is the vacuum state of the electromagnetic field, |1k i is a single-photon
state with wave vector k = (kx , ky , kz ) (see Appendix B). Here we are assuming type-I
down conversion in which the photons have identical polarizations, orthogonal to the
pump polarization. From this point on, the polarization of the fields is suppressed in
our calcuations. In the paraxial approximation, with the z-axis as the propagation
direction, one can write the two-photon output state as
Z
|f i |vaci +

d qs

Z
ds

d qi

Z
di A (qs , qi ; s , i ) |1qs ,s i |1qi ,i i ,

(4.16)

where |1q, i is a single-photon state with transverse wave vector q = (qx , qy ), and
frequency . The two-photon amplitude A (qs , qi ; s , i ) is determined by the pump
amplitude and spectrum, the pump transverse momentum distribution, and the
phase-matching conditions governed by the nonlinear-medium dispersion properties.
The phase-matching conditions are contained in the refractive indices of each beams
polarization. Typically one uses birefringent nonlinear optical crystals to achieve
phase matching (momentum conservation) by angle tuning [164166]. The twophoton amplitude can be expressed in the following way
A (qs , qi ; s , i ) = eikz L/2 sinc (kz L/2)
p (p (s + i )) up (qs + qi ) ,

(4.17)

where is a conversion efficiency, which is proportional to the nonlinear susceptibility,


the pump amplitude and the length of the crystal L. Here
p () is the pump spectral
amplitude, p is the pump central frequency, up (q) is the two-dimensional Fourier
transform of the pump transverse beam amplitude, and sinc (x) = sin (x) /x. The

135
longitudinal wave-vector mismatch kz is given by
kz = kpz ksz kiz ,

(4.18)

where kjz is the component of the wave vector kj along the z-axis (j = p, s, i). The
longitudinal wave vector can be expressed as a function of frequency , and transverse
wave vector q as
kjz (, q) =

r 
2
kj2 () q 2 =
nj ()2 q 2 ,
c

(4.19)

where nj () is the index of refraction for mode j, which generally depends on the
polarization of mode j, and q = |q| is the magnitude of the transverse wave vector.
It is clear that the wave vector mismatch depends on the linear dispersion properties
of the medium, as seen from the explicit appearance of the refractive index in the
longitudinal wave vector. To determine the output down-conversion lights behavior
one typically expands the wave vector mismatch about the central pump frequency
p , the degenerate down-conversion frequency p /2, and zero transverse wave vector.
This leads to the following expression for the two-photon amplitude (see Appendix
B)
ikz L/2

A (qs , qi ; s , i ) = e

p (s + i ) up (qs + qi ) sinc


Dps L
(s + i ) ,
2

(4.20)

where s (i ) is the difference frequency between the actual down-conversion signal


(idler) frequency, and the phase-matched frequency (the degenerate frequency for us)
j = j j .

(4.21)

136
Experimental Down-conversion Setup

In the previous section the theoretical treatment of SPDC was discussed. From
this treatment it appears that in order to realize SPDC all one must do is shine
blue light on a nonlinear crystal and look for the down-converted photons. This is
essentially all that is required. However, when put into practice it becomes much
more difficult. In this section we detail the SPDC setup used in the laboratory.
The most basic break down of the components required for the SPDC setup are
sketched in Fig. 27 below. The setup consists of an intense source of high-frequency
(blue) pump photons necessary to amplify the vacuum noise, a nonlinear crystal
needed to mediate the amplification process, and a photo-detection system that is
sensitive at the single-photon level.

FIGURE 27.. Basic components for SPDC setup.

When considering the blue pump source, one has two basic options, either a
quasi-monochromatic, continuous-wave (CW) pump, or a broad-band, pulsed pump,
contrasted in Fig. 28. Note that because a CW pump intensity is constant in time,
the down conversion arising from a CW pump can be created at any instant in

137
time. However, in the case of a pulsed pump, the down-converted photons will
only be created while the pump pulse is inside the nonlinear crystal (given by the
coherence time of the pulse t), leading to well-timed, photon-pair creation. This is
one motivation for the use of a pulsed pump for linear-optical quantum computing
and quantum cryptography schemes. In our experiments a pulsed source is used.

FIGURE 28. Contrast between continuous-wave (CW) and pulsed pumps. The CW
(top) pump has a constant temporal profile (left), while the pulsed pump (bottom)
is confined to a short time window t (left). The spectra (right) for the two cases
are related via inverse Fourier transforms.

The pulsed-pump source of high-frequency photons used in our experiments is


derived from a home-built, self-mode-locked, femto-second (fs), titanium-sapphire
(TS) laser. The TS laser outputs ultra-short laser pulses (approximately 60 fs in

138
duration) centered on 810 nm wavelength in the near-infrared (NIR) region of the
spectrum. These pulses are sent into a second-order nonlinear optical crystal (other
than the down conversion crystal) in which blue pulses are created through a process
known as second-harmonic generation (SHG), as described below.

The TS laser is pumped by a 6.25 Watt, CW argon-ion laser (Spectra-Physics


Beamlok 2060). The central wave length of the TS laser spectrum can be tuned from
around 650 nm to 1100 nm. The bandwidth associated with the spectrum of the
self-mode locked beam can be quite broad (between 1 nm and 140 nm full-width at
half maximum (FWHM)), giving extremely short temporal pulses (ranging from a few
pico-seconds (ps) to 30 fs FWHM in duration). The repetition rate of a self-mode
locked TS laser can range from 70 MHz to 100 MHz, with an average output power
between 200 mW and 600 mW. Design and operation of TS lasers has been treated
in great detail, and we refer the reader to the abundant literature covering this topic
[126]. The relevant information needed for our discussion is the output characteristics
of the TS laser.

Our TS laser emits horizontally- (H-) polarized, nearly-transform limited, 60 fs


NIR pulses centered on 810 nm wavelength, at a repetition rate of about 90 MHz. The
average output power is about 500 mW. The transverse beam profile is a Gaussian
with a width of approximately 1.2 mm FWHM, and a divergence angle of about 0.8
milli-radians. The actual values can vary with the weather due to fluctuations in the

139
temperature and humidity in the lab, changing the cavity length, and thus the lasing
characteristics. This also leads to instability of self mode-locking in the laser.
The output of the TS laser is used to create the blue beam needed for the
down conversion through the second harmonic generation (SHG) process.

SHG

can be thought of as the reverse process of parametric down conversion (PDC).


Recall that in PDC a single blue photon passing through nonlinear crystal can be
annihilated, resulting in the creation of two red photons. In SHG two red photons
in a beam(refered to as the fundamental beam) passing through a nonlinear crystal
can be annihilated, leading to creation of a single blue photon in a beam called the
second-harmonic (SH) beam, as sketched in Fig. 29. Just as in PDC, energy and
momentum must be conserved in this process, leading to phase-matching conditions
for the SHG process. The SHG also comes in two different configurations type-I,
in which the two fundamental (NIR) photons have the same polarization and the SH
(blue) has polarization perpendicular to the fundamental polarization, and type-II, in
which the fundamental photons have orthogonal polarizations, and the SH is parallel
to one of them. The theoretical description of SHG can be described in a purely
classical manner, and does not require the quantum theory of light [164166].
A simplified theory of the SHG process starts from the nonlinear wave equation
in a medium
h

n
c

i
t2 E = 0 t2 P N L ,

(4.22)

where n is the index of refraction in the medium, and E is the electric field amplitude

140

FIGURE 29. Second-harmonic generation. Energy-level diagram (left), and two


possible geometries of the process collinear (middle) and non-collinear (right).
(with the vector nature suppressed for simplicity). Here 0 is the permittivity of free
space, and P N L is the nonlinear polarization induced by the incident fundamental
(NIR) field, at frequency . We write the electric field amplitude as a sum of plane
waves with different amplitude coefficients
E=


1
E (z) ei(k zt) + E2 (z) ei(k2 z2t) + c.c. .
2

(4.23)

Here E is the amplitude of the incident fundamental field, E2 is the amplitude of


the second-harmonic field, and k (k2 ) is the wave vector inside the medium of the
NIR(SH) field.
We can expand P N L in a power series of the incident electric field amplitude, and
keeping only the lowest-order terms (second-order in this case) with frequencies at
twice the fundamental frequency, i.e. 2, we find
1
P N L = 0 (2) E2 (z) ei(2k z2t) + c.c..
4

(4.24)

Here 0 is the permittivity of free space, (2) is the nonlinear susceptibility of the
crystal, and c.c. stands for complex conjugate. Using Eqs. (4.23) and (4.24) in Eq.
(4.22), keeping only the positive-frequency part, we arrive at the equation of motion

141
for the SH field amplitude

2 2 (2) 2
z2 + 2ik2 z E2 (z) =
E (z) eikz ,
c2

(4.25)

where the wave vector mismatch k, is


k = 2k k2 .

(4.26)

In the slowly-varying-envelope approximation one can neglect the z2 term and arrive
at the standard result for the SHG field amplitude
z E2 (z) = i

(2) 2
E (z) eikz .
2cn2

(4.27)

This equation can be directly integrated along the length of the crystal to give the
output SH field
(2)
E2 (L) = i
2cn2

ZL

E2 (z) eikz dz.

(4.28)

This expression simplifies to the following form if the incident fundamental field is
not significantly depleted by much, allowing it to be pulled out of the integral,
(2) L 2
E (0) eikL/2 sinc
E2 (L) = i
2cn2

kL
2


,

(4.29)

where sinc (x) = sin (x) /x is the phase-matching function. This is optimized when
the wave-vector mismatch vanishes, k = 0. This condition is known as the phasematching condition.
The 60 fs NIR (fundamental) pulse centered on 810 nm wavelength from our TS
laser is focused into a thin (L = 1mm), second-order nonlinear birefringent crystal

142
(beta-barium borate, -barium borate, or BBO) aligned to optimize the SHG process,
producing 405 nm wavelength blue pulses. The crystal length was chosen to give the
maximum down-conversion efficiency, and minimize longitudinal walk-off effects for
the down conversion (not the SHG). The longitudinal walk off, as sketched in Fig.
30, occurs because the fundamental and newly-created SH pulses have different group
velocities, resulting in the fundamental pulse running away from the SH. Thus when
they no longer overlap longitudinally, the fundamental pulse can no longer contribute
to the SH generation. This results in a broadening of the SH pulse. Only for crystal
lengths
N IR

L  LSHG = 1
1 ,
vSH vN
IR

(4.30)

can the group-velocity mismatch be ignored. Here N IR is the NIR pulse duration,
and vN IR (vSH ) is the group velocity of the NIR (SH) pulse. If L  LSHG , the pulse
duration is determined by the longitudinal walk off, and approaches a value given by
[126]
1

1
SH = L vSH
vN
IR .

(4.31)

The group velocities for the pulses can be calculated from the dispersion properties of
the BBO crystal [167], often given in terms of the Sellmeier equations for ordinary and
extraordinary beam propagation. For the wavelengths we use, N IR = 810 nm and
SH = 405 nm, we find that the group velocities are vN IR = 1.780108 m/s and vSH =
1.724 108 m/s. Typically this is quoted in terms of the group-velocity mismatch
1

v v 1 , which is approximately 160 fs/mm for our crystal and wavelengths.
SH
N IR

143
Thus our SHG length is LSHG = 60 fs/(160 fs/ mm) = 0.375 mm, which is much
smaller than our crystal length, implying that the generated SH pulses have a duration
of approximately SH = 160 fs.

FIGURE 30. Walk-off effects in second-harmonic generation. Transverse walk off


(left), and longitudinal walk off (right).

To generate the second-harmonic beam, a short focal-length lens, f = 25 mm


(Thorlabs AC127-025-B) was used. One reason for choosing a short focal-length
lens is because the SHG efficiency goes like 1/w2 , where w is the beam waist of the
fundamental pulse. Thus the shorter the focal length, the tighter the focused beam
will be in the crystal (until one reaches the diffraction-limited spot size). However,
there is a trade off as to how short a focal length one may use. If focused too tightly,
the SH will transversely walk off from the fundamental as sketched in Fig. 30. This
leads to a non-Gaussian transverse beam profile that is difficult to manipulate and
use. Thus one must choose a focal length to give the smallest possible beam waist

144
without significant transverse beam walk off. The walk-off angle, given by [166]
1

= tan




1
1 2
1
n (2, OA ) 2

sin (2OA ) ,
2 e
ne (2) n20 (2)

(4.32)

is approximately = 3 for BBO, leading to a transverse beam walk off of about 0.05
mm for a 1 mm long crystal. A good rule of thumb is to choose a lens that gives a
Rayleigh range z0 = w02 / that is half the crystal length L/2, as shown in Fig. 31
[165]. Here w0 is the beam waist at the focus of the fundamental beam.

FIGURE 31. Optimal focusing for second-harmonic generation. The optimal focusing
condition into a crystal of length L occurs when the Rayliegh-range z0 L/2. This
implies that the beam waist does not change significantly through the crystal.

The phase-matching condition k = 0, is often difficult to achieve in lossless


(transparent) materials because the index of refraction typically displays normaldispersion behavior. That is, the index of refraction increases as the frequency of light
increases. This is why birefringent nonlinear crystals are often used. Birefringence is
the dependence of the refractive index on the polarization direction of the light [164,
168]. Phase matching in a birefringent nonlinear crystal is achieved by precise angular

145
orientation of the crystal with respect to the propagation direction and polarization of
the incident light, as sketched in Fig. 32. BBO is a negative uniaxial crystal. Uniaxial
crystals are characterized by a particular direction known as the optic axis (OA). Light
polarized perpendicular to the plane containing its propagation vector k and the OA,
is said to have ordinary (o) polarization, and propagates with the ordinary refractive
index no (). Light polarized in the plane containing its propagation vector and the
OA, is said to have extraordinary (e) polarization, and propagates with a refractive
index ne (, OA ), that depends on the angle OA , between the wave vector k and the
OA [164166]
sin2 (OA ) cos2 (OA )
1
=
+
.
n2e (, OA )
n2e ()
n2o ()

(4.33)

Here ne () is called the principle value of the extraordinary refractive index. Phase
matching is achieved by adjusting the angle OA so that k = 0. Note negative
uniaxial implies that ne < no . From Eq. (4.26), we can rewrite the phase-matching
condition in terms of the refractive indices as (here we assume type-I phase matching
with the fundamental having ordinary polarized)
no () = ne (2, OA ) ,

(4.34)

which can be solved to give the phase-matching angle for our wavelengths OA = 28.7 .

The nearly-transform limited output blue pulses, with a central wavelength of


405 nm, and bandwidth of approximately 3 nm, from the SHG crystal are vertically
(V) polarized. The blue pulses are collimated using a fused-silica lens (125 mm

146

FIGURE 32. Phase-matching in second-harmonic generation by angle tuning. The


fundamental beam propagates along the z axis with vertical-polarization vector eF
and wave vector kF . The second-harmonic (SH) light propagates collinear to the
fundamental in this example, which horizontal-polarization vector eSH and wave
vector kSH . The optic axis (OA) angle OA , between the OA and fundamental wave
vector is adjusted to obtain optimal phase matching, k = 0.

focal length, Thorlabs LA4236-UV) to ensure that the lens does not fluoresce. The
fundamental and SH beams are separated from one another using a series of four
dichroic mirrors (CVI Laser BSR-48-1025 and Edmund Optics NT47-264) that reflect
the SH, and pass the fundamental light, as shown in Fig. 33.
The blue pulses are then directed into the down-conversion crystal, which is an
identical BBO crystal to that used in the SHG, cut so that light incident perpendicular
to the surface makes an angle OA = 31.25 with the optic axis. This particular crystal

147

FIGURE 33. Experimental second-harmonic generation setup. Pulses from the TS


laser are focused (f1 ) into the BBO crystal to produce SHG. The SHG is collimated
(f2 ), and then filtered from the fundamental TS light using a series of four dichrioc
mirrors (DM), and beam dumps (BD) to capture the TS radiation.

cut was chosen to produce non-collinear, degenerate type-I SPDC at a wavelength


= 810 nm, with the down-converted photons emerging from the crystal at an angle
of = 7.5 , away from the pump beam. The non-collinear geometry was chosen to
ensure easy separation of the blue pump and down-converted light, as shown in Fig.
34 below. After the down-conversion crystal, the blue pump beam is absorbed by a
beam dump (BD). The two down-converted beams are passed through anti-reflection
(AR) coated, color-glass filters (CVI W2-CG-RG-715-2X2-2.5-633-1064-0), and 10 nm
bandwidth interference filters centered on 810 nm wavelength (CVI F10-810.0-4-0.50)
prior to hitting the detectors. Depending upon which experiment is being run, we
have two different pairs of photon-counting detectors that can be employed to collect
the down conversion. The first detector pair is a set of photon-counting avalanche
photodiode (APD) detector modules (EG&G SPCM-200 CD2027), powered by home

148

FIGURE 34. Experimental down-conversion setup. The blue SHG pulses are directed
into a BBO crystal to produce the down-converted light. The BBO crystal is cut for
type-I, non-collinear phase matching. The down-converted light passes through colorglass filters (CGF) and interference filters (IF) prior to absorption by the detectors.
built power supplies. The APD detectors have a quantum efficiency of 50% at 810 nm
(70% at 633 nm), and a 200 m-diameter active area, with greater than 90% spatial
uniformity over the active area. The dark count rate of the APD detectors is about
40 counts per second.
The second detector pair is a set of photon-counting photo-multiplier tube (PMT)
modules (Hamamatsu H7421-50). The PMT detectors have a quantum efficiency
of 15% at 810 nm, and a circular active area with 5 mm diameter, with spatial
uniformity greater than 85% over the entire active area. The dark count rate of the
PMT detectors is about 50 counts per second.
The electronics used to analyze the signals coming from the detectors are sketched
in Fig. 35 for the two different detector pairs. When purchased (c. 1992), the APD
detectors output transistor-transistor-logic (TTL) pulses. The internal electronics
have since failed, and now only a weak, non-standard electrical pulse is emitted (100 mV across a 50 resistor, approximately 4 ns in duration). The weak electrical

149
pulse output from the APD detectors can pass for nuclear-instrumentation-module
(NIM) pulses, which our equipment can utilize. The electrical-pulse output from
each APD module is connected to a 50 terminator prior to being sent through a
coaxial cable (40 ft. in length). The cables are connected to another 50 terminator
before connecting to a short (1 ft. long) BNC-to-Lemo cable (US Ultratek BNCLemo 00), whose opposite end is connected to a NIM, pulse discriminator (Phillips
Scientific NIM Model 708). The need for the short BNC-to-Lemo cable is due to a
double-pulsing effect that arises from the reflection of the pulse at the discriminator.
Without the short cable, a pulse reflects from the discriminator due to impedence
mismatch, which travels to the detector and back, making it appear as if there are
two pulses instead of one. The discriminator threshold is adjusted to output a NIM
pulse into two different channels for every APD pulse (labeled channels 1 and 2).
These pulses are stretched from 3 ns to about 10 ns by the discriminator unit. The
stretching is done to ensure the coincidence logic will work properly. One stretched
NIM pulse from each discriminator channel is sent into the coincidence counting
(C.C.) logic unit (Phillips Scientific NIM Model 754), which registers a coincidence
count if the two pulses overlap temporally. The coincidence-count signal from the
coincidence logic unit, along with the second set of pulses from the discriminator
(called the singles counts for channels 1 and 2) are sent to a pulse translator. The pulse
translator (Phillips Scientific NIM Model 726) converts NIM pulses to TTL pulses,
which are compatible with our computer boards. The TTL pulses are sent to 50

150
terminated BNC connectors on a breakout box (National Instruments BNC-2121),
that is connected to a counter board (National Instruments NI-PCI-6602), mounted
inside a desktop computer (Dell Dimension 2400). A computer program (written in
LabVIEW) collects the average count rates from the two individual detectors, denoted
singles 1 and singles 2, and the coincidence count rate.

FIGURE 35. Detection electronics used with the APD detectors (top) and PMT
detectors (bottom).

151
The electronic setup used with the PMT modules (Hamamatsu H7421-50) is nearly
identical to the one used with the APDs, but there are notable differences. The
electrical pulses output from the PMT modules are standard TTL. The output of
each PMT module is connected to a 50 terminator prior to being sent through a
coaxial cable (40 ft. in length). The same double pulsing occurs in this setup, and thus
the cables are connected to another 50 terminator before connecting to a short (1
ft. long) BNC-to-Lemo cable, whose opposite end is connects to the logic translator.
There are two outputs from the translator, one TTL and one NIM, for each detector
channel. The TTL outputs for each channel are sent directly to the counter board
in the computer to register as the singles counts for each channel. The NIM outputs
are sent to the discriminator to be stretched, and sent to the coincidence logic unit.
The output of the coincidence logic unit is then sent to the translator (NIM-TTL),
and on to the counter to register as the coincidence count signal.

To verify that we do indeed have entangled photon pairs from our down-conversion
source, we performed a Hong-Ou-Mandel-interference (HOMI) experiment [113]. The
experimental setup used to observe HOMI is shown in Fig. 36 below. The downconverted photons emitted from the BBO crystal at an angle of 7.5 from the pump
beam axis are directed onto a moveable 50:50 non-polarizing beam splitter (Thorlabs
BS014) by two mirrors (Thorlabs BB1 E03). The down-conversion beams are aligned
so that they spatially overlap on the beam splitter. Each beam from the output ports
of the beam splitter is directed to a photon-counting PMT module. Placed in front

152
of each PMT detector, a 10 nm bandwidth interference filter, centered at 810 nm
wavelength, and two broadband color-glass filters define the spectral properties of
the detected photons. The output count rate of each detector, and the coincidence
count rate were recorded by a computer, while the position of the beam splitter was
scanned and recorded.

FIGURE 36. Hong-Ou-Mandel-interference experiment setup. The down-converted


light is directed onto a moveable non-polarizing beam splitter (BS) using two mirrors
(M). Interference filters (IF) determine the bandwidth of the photons detected. The
detector signals are sent to a coincidence counting (C.C.) logic unit. The coincidence
count rate as a function of the BS position ( ) is recorded.

The probability for joint-detection events at the two detectors D1 and D2 , at times
t and t + respectively, for the HOMI experiment is [113]
D
E
()
()
(+)
(+)
P12 ( ) = E1 (t) E2 (t + ) E2 (t + ) E1 (t) .

(4.35)

()
Here characterizes the detector efficiencies, and Ej (t) is the positive- (negative-)

frequency parts of the electric field operators at detector Dj , (j = 1, 2). For the

153
experimental setup described above, this leads to the following expression for the
coincidence-count rate [163]

Rcc ( ) = C 1

2RT
sinc
2
R + T2

x1
L


sinc

x2
L

( )2


.

(4.36)

Here C is a constant, R and T are the intensity transmission and reflection coefficients
of the beam splitter, xj is the width of a pinhole placed in front of detector Dj ,
(j = 1, 2) (or the diameter of the detector if there is no pinhole), and is the
bandwidth of the interference filters placed in front of the detectors. The characteristic
length L is proportional to 1/, where is the angle between the two beams after
superposition on the beam splitter. Explicit form of the characteristic length L, is

L=

2c

=
,

(4.37)

where () is the central frequency (wavelength) of the interference filters. The


sinc-functions in Eq. (4.36), arise from the distinguishability of the two photons
from different propagation directions, and degrades the visibility of the HOMI dip
when not equal to unity. The Gaussian of width , comes from the interference
filters (assumed to have a Gaussian profile), which limit the spectrum of the detected
photons, and thus their temporal coherence length. The time delay between the
detection events is introduced by shifting the beam splitter a distance s = c .
Figure 37 shows the results of our HOMI experiment. The visibility of the HOMI dip,
approximately 77%, is quite good for free-space (non-fiber coupled) pulsed SPDC.

154

FIGURE 37.. Hong-Ou-Mandel-interference experiment results.

Alignment of the HOMI experiment depends critically upon the path lengths
between the crystal and the two beam-splitter ports, as well as the spatial mode
matching into the beam splitter. The alignment of the HOMI experiment was accomplished
by performing non-collinear SHG with the femto-second pulses from our TS laser.

155
When the path lengths are matched within the pulse width, on the order of a 180 m
for a 60 fs pulse, and the birefringent BBO crystal is properly oriented, one can
observe, and optimize the blue, second-harmonic radiation. This optimized SHG
may be used as a guide for the down-conversion pump beam.

Ghost Imaging and Ghost Wigner Distribution

The HOMI experiment described above verifies that the SPDC setup in our lab
is working properly, but only characterizes the temporal correlations of the downconverted photons. The main interest of the work presented in this dissertation is
focused on transverse spatial states of light. The study of the transverse spatial
state of the photon pairs created in SPDC can be traced back to the studies by
Klyshko [158], and Madge [157] in 1967. The first theoretical treatments of SPDC
[158, 161, 162, 169, 170], although adequate to describe the down-conversion state,
do not lend themselves to very insightful or useful forms when the transverse spatial
state is the central concern.
Experimental and theoretical investigations centered on the transverse spatial
state from the SPDC source, carried out in the mid-1990s [26, 27, 29, 100, 101, 142,
171], resulted in a new formalism in which to treat SPDC [100, 101, 171]. Several of
the experiments carried out [26, 27, 29, 142], known collectively as quantum-imaging
experiments, rely on the transverse-momentum entanglement between photon pairs
produced in the SPDC process. The spatial entanglement needed in these experiments

156
arises from momentum conservation in the down-conversion process, which can be
modified by the phase-matching conditions [100], and pump beam angular distribution
[101, 142]. Recall that in the process of SPDC, an intense laser beam, called the pump,
is incident on a nonlinear optical crystal. Through the nonlinear, optical interaction,
a pump photon has a small, but non-zero probability to split into a pair of daughter
photons, traditionally called the signal and idler photons. Energy and momentum are
typically conserved in the interaction, resulting in the correlation, or entanglement,
between the daughter photons.

There are two basic quantum-imaging experiments, refered to as ghost-imaging


and ghost-diffraction. These can be considered variations on the same theme, which
we depict in Fig. 38. In the ghost-imaging experiment, the the entangled photons
from the SPDC source are directed along two different, spatially-separated paths.
In one path, say the idler-photon path, a transmitting aperture (amplitude mask)
followed by a fixed large-area photon-counting detector, also called a bucket detector,
collects any photon that passes through the aperture. The bucket does not give any
distinguishing position or momentum information about the photon other than it
passed through the aperture. In the other path, the signal-photon path, a lens,
followed by a small, point-like photon-counting detector collects the down-converted
light. As the imaging, point-like detector is scanned in the plane perpendicular to
the beam axis, the coincidence count rate between the two detectors is recorded.
The coincidence rate as function of the point-like detectors position maps out the

157
aperture, which is located in the other detectors path. Each individual detectors
count rates remain essentially constant as the point-like detector is scanned, and no
image is observed in the so-called singles count rates. Because the coincidence image
is formed by scanning the point-like detector, which registers photons that do not
pass through the aperture, the term ghost-imaging was coined to emphasize the
non-local nature of the technique. In the ghost-diffraction experiments, the large-area
photon-counting detector is replaced by a lens after the aperture, which is followed
by another point-like detector. This additional lens is placed its focal length away
from the aperture, and the point-like detector is placed in the Fourier plane of the
lens. This selects out a single transverse momentum of the photon passing through
the aperture plane. The coincidence image that appears as the imaging detector is
scanned, reproduces the far-field diffraction pattern of the aperture.

FIGURE 38. Quantum-imaging experimental setups. Ghost-imaging (left) employs a


fixed bucket detector to collect photons passing through the aperture A, while ghostdiffraction relies on a Fourier transform lens f 0 , and fixed point-like detector to filter
photons passing through the aperture.

158
In the standard ghost-imaging experiment, depicted in Fig. 38, the transmitting
aperture, is placed a distance da from the nonlinear down-conversion crystal in the
idler-photon path. The aperture is characterized by the transmission function t (r),
where r = (x, y) is the transverse coordinate in the aperture plane. The largearea bucket detector D1 is assumed to be placed just after the aperture to simplify
calculations. The imaging lens, with focal length f , is placed in the signal-photon
path a distance db from the nonlinear crystal. The point-like imaging detector D2 , is
placed a distance d0b from the imaging lens. The image plane in which the imaging
detector should be located is determined by a thin-lens-like equation [26, 27, 171]
1
1
1
+ = ,
so si
f

(4.38)

where the object distance is so = da + db , and the image distance is so = d0b .


The coincidence-count rate as a function of the imaging-counter position RCC (r2 )
is proportional to the modulus squared of the transmission function
RCC (r2 ) |t (r2 /m)|2 ,

(4.39)

where m = si /so is the image magnification. This experiment and the imaging
condition can be explained by using the Klyshko, advanced-wave picture [170, 171].
In the advance wave formulation one fictitiously replaces either detector with an
incoherent source, emitting omnidirectional, incoherent light over a broad spectral
range, propagating backward towards the crystal, as depicted in Fig. 39. Due to the
transverse-momentum anti-correlations of the actual down conversion, the nonlinear

159
crystal acts like a planar mirror, reflecting the virtual source towards the ghostimaging plane. In particular, if the down-conversion crystal is pumped with a planewave pump, the crystal acts like a planar mirror. Whereas if the down-conversion
crystal is pumped with a focused beam, the crystal acts like a curved mirror, due
to the modification of the down-conversion state by the pump transverse momentum
distribution. Tracing the fictitious advanced wave through the imaging lens, we find
that the image of the incoherently illuminated aperture appears in the imaging plane
defined by the thin-lens-like equation (4.38). These results may be obtained, with the
proper approximations, from the full SPDC theory (see Appendix B).

The standard ghost-diffraction experiment utilizes the same setup as the ghostimaging, with the exception of the fixed bucket detector being replaced with a Fourier
transform lens, with focal length f 0 , and fixed point-like detector. The observed
coincidence image reveals the Fourier transform of the aperture transmission function,
T (r/f 0 ), where f 0 is the focal length of the collection lens, and is the central
wavelength of the light (see Appendix B).

160

FIGURE 39. Advance-wave description of quantum imaging. Either detector acts as


a virtual source, emitting incoherent, broadband light in all directions. These virtual
photons propagate toward the nonlinear crystal, which acts like a planar mirror (left)
if the crystal is pumped with a plane wave, or a curved mirror (right) when pumped
by a focused beam. The virtual photons then propagate toward the second detector.

We used the ghost-imaging technique in the laboratory for alignment diagnostics


of the down-conversion, and verification that the two beam axes were well correlated
spatially. To implement the ghost-imaging experiment we used an array of five parallel
wires (1 mm diameter) displaced 1.5 mm apart for our aperture. This aperture was
placed in the idler-beam, a distance da = 45cm, from the nonlinear crystal, as depicted
in Fig. 38. The light passing through the multi-wire aperture was collected by a 50
mm diameter, 100 mm focal-length lens (Thorlabs LA1050-B), and focused onto a
large-area, photon-counting photomultiplier tube module (Hamamatsu H7421-50). In
the signal beam, the imaging lens, a 25 mm diameter, 300 mm focal-length, doublet
lens (Thorlabs AC254-300-B), was placed a distance db = 45cm, from the nonlinear
crystal. The point-like, imaging detector, a photon-counting avalanche photo-diode
(EG&G), was placed a distance d0b = 45cm from the imaging lens. This gives an
image magnification of m = 1/2 (the minus sign implies that the image is inverted).
A 10nm bandwidth (FWHM) interference filter, centered on 810nm wavelength, was

161
placed in front of each detector. A computer program controled the vertical position
of the point-like, imaging detector, which also collected the average coincidence and
singles count rates of the detectors. The results of the experiment, shown in Fig. 40,
match the predicted distances quite well.

The Klyshko advanced-wave explanation of ghost imaging also lends itself to the
description of conditionally prepared single-photon states from SPDC [172, 173]. In
these experiments, the idler photon is passed through a spatial filter (single-mode
optical fiber, for example) and spectral filter (narrow-band interference filter) to
conditionally prepare the signal photon in a particular spatio-temporal state, as in
Fig. 41. The state purity approaches unity as the filters become infinitely narrow
band, at the cost of diminishing heralding efficiency [173, 174]. Following the Klyshko
advanced-wave picture, a point-like detector at the fiber tip acts as a virtual, reversepropagating, broad-band, incoherent source, which is coupled into the optical fiber.
The optical fiber acts like a spatial and spectral filter, allowing only certain modes
to propagate. After the fiber, a spectral filter further narrows the allowed spectrum
of the virtual source. Again, the nonlinear crystal acts as a planar mirror (for a
plane wave pump) due to the transverse-momentum anti-correlations of the down
conversion. The conditionally-prepared single photon then occupies a spatio-temporal
state, which is an incoherent mixture of pure, spatio-temporal states, whose purity is
defined by the filtering process [173, 174].

162

Singles counts in 10 sec.

80000
70000
60000
50000

Fixed detector
Scanning detector

40000
30000
20000
10000
0
-2

-1

-2

-1

Coincidence counts in 10 sec.

2200
2000
1800
1600
1400
1200
1000

Scanning detector position [mm]

FIGURE 40. Experimental ghost-imaging results. The singles counts for both
detectors (top) are relatively constant, while the coincidence counts (bottom) map
out the ghost-image of five wires.

163

FIGURE 41. Conditional single-photon state preparation setup. The trigger photon
passes through an interference filter (IF) prior to being focused into a single-mode
optical fiber (SMF), which acts like a spatial-mode filter. The output of the SMF
is directed to the trigger detector. A click at the trigger detector heralds a photon
in the other down-conversion path in a spatio-temporal mode defined by the IF and
SMF.

Suppose one replaces the scanning, ghost detector (D2 ) in Fig. 38 by a parityinverting Sagnac interferometer introduced in Chapter III, as depicted in Fig. 42
below. Initially, one may think that the coincidence-count rate will lead to the
transverse Wigner function for the aperture, a ghost-Wigner function. However,
the coincidence-count rate between the Sagnac and the bucket detector measures the
transverse spatial Wigner function of the incoherently-illuminated aperture. The
transverse Wigner function for an incoherently-illuminated single-slit aperture of
width a, is given by
in
(sx , kx ) (2sx ) sinc (kx a/2) ,
WSS

(4.40)

where (x) is the Dirac delta function and the sinc (x) = sin (x) /x is the Fourier

164
transform of the slit. If the transverse coherence length of the down conversion at the
single slit is lT C , then the delta function in Eq. (4.40) is widened to a Gaussian of
width lT C , which for our experiments is on the order of 10m. Experimentally, this is
difficult to detect because it requires alignment of the down-conversion beam axis into
the Sagnac to within 10m, and at the same time, angular alignment of the beam axis
within /a. This incoherently-illuminated single-slit Wigner function, differs from the
coherently-illuminated single-slit Wigner function found in Chapter III significantly
in its transverse position distribution, but not the transverse momentum.
One may bypass the problem of the incoherent ghost-Wigner distribution, by
using only photon pairs from the down conversion that have coherent, single-mode,
transverse spatial states. That is, by using conditionally prepared single photons.
This may be accomplished by replacing the bucket detector after the aperture, in
Fig. 38, by a collection lens, coupled to a single-mode fiber. The output of the
fiber is focused onto a photon-counting avalanche photodiode (APD) detector. In the
Klyshko advanced-wave picture, the APD detector acts like a point-like, broad-band,
incoherent source, which is coupled into the fiber. The emitted light is collimated and
sent back, through the aperture, which gets imaged back to the Sagnac interferometer,
as shown in Fig. 42. At the time of this writing, we are currently engaged in
performing this experiment in our laboratory.
Several years after the initial ghost-imaging investigations, the question was raised
as to whether classically-correlated light, that is, non-entangled, or separable light

165

FIGURE 42. Incoherent (left) and coherent (right) ghost-Wigner experimental


setups. Replacing the scanning detector of the ghost-imaging setup with a parityinverting Sagnac interferometer (Sagnac) gives the ghost-Wigner function of the
illuminated aperture A by spatially incoherent light. The coherent ghost-Wigner
of the aperture can be obtained by using the conditionally prepared single-photon
trigger setup Fig. 41 instead of the bucket detector.

166
sources, can also reproduce such effects [175180]. Indeed, it was found that a classical
light source could produce coincidence images [178180]. It was only recently that
the fundamental distinctions between entangled-source and classical-source ghostimaging were clarified [181] for the case of Gaussian state sources. Thus the differences
between classical- and quantum-coincidence (rather than ghost) imaging remain at the
forefront of modern research. For the Gaussian-input states, the ghost-imaging effect
was shown to be mostly classical, arising from non-zero intensity cross correlations
between the detected fields [181]. When considering only two-photon fields, entangled
sources gain a factor of

2 spatial resolution in near-field imaging, and a

2 increase

in the field of view in far-field diffraction, over their classically-correlated counterparts.


In addition, the image visibility is also increased [178, 180, 181].

Two-photon Transverse Spatial Wigner Function Measurement and Bell Inequalities

The ghost-imaging experiments of the previous section can measure probability


distributions in either transverse momentum (ghost-diffraction) or transverse position
(ghost-imaging), depending upon the experimental setup. In order to completely
characterize the transverse state of a down-conversion source from such probability
distribution measurements, one must perform a tomographic-reconstruction of the
transverse Wigner distribution [20]. A more realistic method to characterize the
two-photon state involves characterization of the complex, second-order correlation

167
function (also called the fourth-order correlation function)

(2) (r1 , r01 ; r2 , r02 ) = (r01 ) (r02 ) (r2 ) (r1 ) ,

(4.41)

where rj = (xj , yj ), (j = 1, 2), is a transverse position in the signal or idler beam


respectively, the s are scalar wave functions of the electromagnetic field (suppressing
the vector nature for simplicity). The brackets hi imply ensemble averaging over all
realizations of the fields.
The second-order correlation function is also equivalent to the transverse spatial
Wigner function in the sense that knowledge of one can be used to reconstruct the
other. We choose to focus on an interferometric characterization of the two-photon
transverse spatial Wigner function in a certain pair of transverse planes, as defined
in Eq. (4.2),
W

(2)

1
(r1 , q1 , r2 , q2 ) = 4

(2)

d 1

d2 2 (2) (r1 1 , r2 2 )
(4.42)


(r1 + 1 , r2 + 2 ) exp [2i (q1 1 + q2 2 )] ,

where we have suppressed the longitudinal coordinates that define the transverse
planes. Measurement of the two-photon transverse spatial Wigner function may be
accomplished by using two parity-inverting Sagnac interferometers, as shown in Fig.
43. The interferometers differ slightly from those introduced in Chapter III, in that
each has an additional beam splitter used to collect light exiting the interferometer
from the second port of the interferometer. Spatially separated photons emitted
from the source are directed into the two interferometers, placed equidistant from the
source. In order to measure the full, two-photon Wigner function, light from both

168
exit ports of the interferometer must be collected. This is accomplished by placement
of an additional beam splitter just after the steering mirror, prior to the entrance into
each interferometer, that directs the light exiting the interferometer at the entrance
port to a photon counting PMT module.

FIGURE 43. Experimental setup to measure the two-photon transverse spatial


Wigner function. Each photon from the two-photon source (1 or 2) is directed into
a separate parity-inverting Sagnac interferometer with an additional beam splitter
(BS). The ouputs of each interferometer (primed or unprimed) are directed to photon
counting detectors, which register the intensity Ij and Ij0 . The photo-currents from
the two detectors of each interferometer are subtracted (-), and the differnce signal,
Ij Ij0 , is sent to a coincidence logic unit. The average coincidence rate is proportional
to the two-photon transverse spatial Wigner function evaluated at the phase-space
point determined by the steering mirror positions (s1 , k1 ) and (s2 , k2 ).

169
The detector signals from the two output ports of each individual interferometer
are subtracted, indicated by the circle with a minus symbol in Fig. 43, and the
coincidence rate between the two difference signals is collected. The coincidence rate
between the difference signals is proportional to the two-photon, transverse spatial
Wigner function, as can be seen from the expectation value for coincidences

RCC

D 

 E
0
0

: I1 I1 I2 I2 : ,

(4.43)

()
(+)
where the colons :: imply normal ordering [87], and Ij = Ej (rj ) Ej (rj ) , j =

1, 2 are the electric field operators at the outputs of the beam splitters of each
interferometer. In terms of the input field operators of each interferometer, the two
output fields are given by
h
i
(+)
(+)
(+)
Ej (xj ) = j0 j2 Ejin (xj + sj ) exp (ikj xj ) rj2 Ejin (xj sj ) exp (ikj xj ) ,
(4.44)
for the unprimed field output and
h
i
(+)
(+)
(+)
Ej (xj ) = j0 rj0 j rj Ejin (xj + sj ) exp (ikj xj ) + Ejin (xj sj ) exp (ikj xj ) ,
(4.45)

at the primed output. Here rj , j rj0 , j0 are the amplitude reflection and transmission
coefficients of the Sagnac (extra) beam splitter. The difference signals are given by






0

Ij aIj = j02 1 14 + r14 a10 r12 12 rj02 Ij




(sj , kj ) .
2 j j2 rj2 + aj j0 rj2 j2 rj02 W

(4.46)

170

Here aj is a real-valued proportionality constant to be determined, j j0 is the

quantum efficiency of detector Dj Dj0 , j = 1, 2, Ij is a constant given by
Z

Ij =

()
(+)
dxEjin (x) Ejin (x),

(4.47)

(sj , kj ) is the single-field, Wigner-function operator


and W
(sj , kj ) =
W

()
(+)
dxEjin (x sj ) Ejin (x + sj ) exp (i2kj x).

(4.48)

Choosing the proportionality constant aj , to cancel the first term on the right-hand
side of Eq. (4.46), gives
aj =

j j4 + rj4
j0 rj2 j2 rj0 2


.

(4.49)

This leads to the following value for the difference signal







2
0

(sj , kj )
Ij aj Ij = 2j0 j j2 rj2 + j4 + rj4 W
=

2j0 j W

(4.50)

(sj , kj ) ,

where we have used the identity j2 + rj2 = 1. Thus we see that the expectation value
of the product of the difference signals is proportional to the two-photon, transverse
spatial Wigner function
D

I1 a1 I10



I2 a2 I20

E

= (2)2 10 1 20 2 W (s1 , k1 ; s2 , k2 ) ,

(4.51)

where the two-photon transverse spatial Wigner function is given by


Z
Z
1
W (s1 , k1 ; s2 , k2 ) = 2 dx1 dx2 exp [i2 (k1 x1 + k2 x2 )]

D
E
()
()
(+)
(+)
E1in (x s1 ) E2in (x s2 ) E2in (x + s2 ) E1in (x + s1 ) .

(4.52)

171
The ability to experimentally determine the two-photon transverse spatial Wigner
function, as propose above, would be a useful tool in the study of continuous-variable
entanglement and disentanglement. In addition, the technique allows for one to test
a Bell inequality for contious variables, as prescibed by Banaszek and Wodkiewicz
[33]. This Bell-inequality test is based upon the fact that the Wigner function is
the expectation value of the displaced parity operator [24]. Choosing four different
appropriately-placed points in phase space, one can violate a Bell inequality.

172

CHAPTER 5

CONCLUSIONS
Light continues to be of great importance in the study of nature, both as a probe
of matter and a fundamental system of interest itself. From implementations of
new quantum-information protocols to greater-precision metrology, the better light
is understood, the more efficient use can be made of it. In the case of single-photon
and two-photon states, such as those relied upon in various quantum technologies, it
is necessary to have complete knowledge of these states. Quantum state tomography
(QST), also know as quantum state determination, provides an experimental means
to characterize the state of such quantum systems.
In this dissertation, we have presented experimental techniques to characterize
the spatial state of light in a plane transverse to its propagation direction (assuming
a paraxial beam). The first method that is presented measures the transverse spatial
Wigner function
1
W (r, q) = 2

hd2 x (r x) (r + x)ie2iqx ,

(5.1)

of a beam with well-defined polarization, in the transverse plane, that is, the plane
perpendicular to the propagation (z) axis, at a particular longitudinal coordinate
z. Here r = (x, y) and q = (qx , qy ) are the two-component transverse-position and
transverse-wave vectors in the transverse plane, and (r) is the transverse wave

173
function of the single-photon field. The angle brackets h i imply the ensemble average
over all realizations of the fields. In the case of classical light, the method can be used
to characterize the first-order correlation function (1) (r, r0 ). The new technique has
several advantages over previous methods to characterize the transverse spatial state
of light, not the least of which is its ability to characterize single-photon states.
Another unique feature that was tested is the ability of our experimental setup
to measure highly divergent sources of light, which is useful to study decoherence
of scattered light. We demonstrated the ability of our method by measuring the
transverse spatial Wigner function of four different light sources derived from an
attenuated helium-neon laser.
The second experimental technique proposed to measure the two-photon transverse
spatial Wigner function in two transverse planes defined by longitudinal coordinates
z1 and z2
W

(2)

1
(r1 , q1 , r2 , q2 ; z1 , z2 ) = 4

d 1

d2 2 h (2) (r1 1 , r2 2 ; z1 , z2 )
(5.2)

(2) (r1 + 1 , r2 + 2 ; z1 , z2 )i exp [2i (q1 1 + q2 2 )] .


In the classical-domain, the method determines the second-order correlation function
(2) (r1 , r01 ; r2 , r02 ). We have shown in theory how this technique may be used to violate
a Bell inequality, by using the fact that the Wigner function is proportional to the
expectation value of the displaced parity operator [24, 33]. The source of entangled
photon pairs needed to carry out such experiments is provided by the spontaneous
parametric down-conversion (SPDC) source.

We have presented theoretical and

174
experimental results characterizing our SPDC source, including Hong-Ou-Mandel
interference with a visibility of 77%, and a demonstration of ghost-imaging. In the
discussion of ghost-imaging, which is described in terms of the Klyshko, advancedwave picture [89], we proposed a ghost-Wigner experiment, in which the photons
passing through the aperture are collected by a single-mode optical fiber to produce
a spatially-coherent advance-wave source. We also point out that this is closely
related to remote state preparation. The area of quantum-imaging is still an active
area of research [181]. These are good steps towards experimental realizations of the
two-photon Wigner function measurements and the Bell-inequality violation, which
are being performed in our lab as this is being written.

In discussing these experiments, we have found it useful to introduce the concept


of photon wave mechanics (PWM), in which single-photon and two-photon states
are thought of as the states of quantum particles. These states are represented by
wave functions, which evolve in time according to a prescribed wave equation. In
order to talk about multi-photon states, we had to extend the single-photon theory
developed by Bialynicki-Birula and Sipe [59, 60, 68, 73] to a multi-particle theory,
and in doing so, discovered several useful relationships between PWM and other wellknown theories of electromagnetism, such as classical coherence theory. The wave
functions in this PWM theory and the generalized photo-detection amplitudes are
found to be equivalent, and to obey the same equations of motion. We showed how
to translate between the standard quantum-field-theory (QFT) approach to quantum

175
optics and our theory of PWM, noting that there is a direct relationship between the
modes of QFT and the states of PWM.
These developments naturally led us to discuss the application of this new theory
to a set of example calculations. One calculation, in which we analyzed the so-called
measurement-induced interaction of photons incident on a beam splitter, showed the
explicit dependence of the mode-matching, or wave-packet overlap in the language
of PWM, necessary for maximal interaction strength. This example highlights the
correspondence between modes and states of the QFT and PWM. Another calculation
examined entangled photon pairs traveling through a realistic turbulent atmosphere.
Here we considered photon pairs, entangled in their orbital-angular-momentum (OAM)
degrees of freedom, and characterized the loss of entanglement as the photons traversed
the turbulent atmosphere. We found that as the characteristic turbulence size, which
is associated with the refractive-index fluctuation length scale, decreases with respect
to the beam width of the light, the photon pair disentangles more rapidly. We also
found that for large OAM values, the disentanglement is less severe than for small
OAM values. This can be explained by the fact that scattering from one OAM state
to another depends only on the change in OAM, ~l, which must be supplied by
the atmosphere. This calculation demonstrates the utility of the connection between
PWM and CCT.

176
Future Directions

In this dissertation we have focused on the description of light at the single-photon


and two-photon levels, from both a theoretical and experimental vantage. The photon
wave mechanics developed in the dissertation shows that the scalar product between
photon wave functions is not the standard, local-integral form, but rather takes on
a non-local character. This non-local scalar product differs significantly from the
local scalar product only when one considers non-quasi-monochromatic light, such as
that encountered in ultra-short pulse phenomena. This is one area in which further
theoretical and experimental research should be focused, and could provide fruitful
discovery. For example, the form of the state overlap characterizing the measurementinduced interaction strength in the example calculation of Chapter II, suggests that
there are observable differences between quasi-monochromatic and broad-band light.
Relativistic quantum information and quantum theory from the PWM viewpoint
could also provide an interesting avenue of study in itself. The inherent relativistic
nature of the photon is captured in the transformation property of the wave function,
the relativistic wave equations, and non-local scalar product of PWM. These may
become useful tools in the pursuit to understanding the connections between relativity
and quantum information. Thus making the theoretical description and experimental
characterization method of photon wave functions worthwhile areas of investigation.

177
This could also branch out to the interface between general-relativity and quantum
information [182].
The measurement of the transverse spatial Wigner function puts the photon wave
function, and the PWM theory, on a solid experimental foundation. Pushing these
experiments to measure the ghost-Wigner function and two-photon Wigner function
will further solidify the PWM theory. This also provides a means to characterize the
transverse spatial state of entangled photon pairs, which presents an arena in which
the fundamental nature of entanglement and decoherence may be studied.
A theme in this dissertation has been that light is a driving force in our comprehension
of the world in which we live. There is no doubt that quantum state determination
techniques, such as those presented here, will play an ever increasing role not only in
our understanding of light, but also in the applications and technologies of the future.
From the fundamentals of quantum theory, to the colors of a rainbow, light will keep
us wondering.

178

APPENDIX A

SINGLE-PHOTON WAVE FUNCTION LORENTZ TRANSFORMATION


PROPERTIES

The single-photon wave function has been thoroughly treated by Bialynicki-Birula


[59, 68, 73] and by Sipe [60]. There was, and still remains, some controversy regarding
the photon wave function in coordinate space. This controversy stems from the
confusion regarding the position operator and localizability of the photon.

The

localizability of elementary quantum particles was first posed (and solved) by


Newton and Wigner in 1949 [41]. It was later refined by Wightman [42], and revisited
by Amrein in 1969 [44]. The outcome of these investigations are essentially equivalent
zero-mass particles with spin (intrinsic angular momentum) s 1/2 are strictly
non-localizable. It should be noted that even for massive particles, in the relativistic
limit their energy, mass, and charge densities are not equal (i.e. there is not a good
definition of a point particle in this limit) [68]. The lack of a position operator
for such elementary particles follows from the result that they are not localizable.
In an attempt to rectify the lack of position operator for the photon, Hawton and
Baylis [56] put forth a non-standard position operator, and an associated photon wave
function that differs from that presented by Bialynicki-Birula [59, 68, 73] and by Sipe
[60].

179
The single-photon wave function has been discussed quite thoroughly [59, 60, 68,
73], but some of the details of the Lorentz invariance were not clearly elucidated in
those treatments. Here we briefly review the photon wave function, and show the
Lorentz invariance of the inner product.
The photon wave function in coordinate space may be defined in terms of the
positve frequency Riemann-Silberstein (RS) vector, which in free space is

(+)

r
(x, t) =


0
E(+) (x, t) + icB(+) (x, t) .
2

(A.1)

This describes a positive-helicity state only. However, the negative helicity may be
treated by simply changing the +i to a i in the definition Eqs. (A.1). In the case
of plane waves, positive and negative helicity coincide with right-circular and leftcircular polarization, respectively. For concreteness, suppose a plane, electromagnetic
wave is propagating in the z direction, with the right-circular (R) or left-circular (L)
polarization (RCP or LCP). Then the corresponding positive frequency parts of the
electric and magnetic-induction fields can be written in the plane wave expansion
x
i
y
(+)
ER(L) (x, t) = E (x, t) ei(kzt)
2
(
z (
x i
y)) i(kzt)
(+)

BR(L) (x, t) = E (x, t)


e
c 2
i (
x i
y) i(kzt)

= E (x, t)
e
.
c 2

(A.2)

Here E(x, t) is the scalar electric field amplitude. The positive-frequency Riemann(+)

Silberstein vectors for the positive and negative helicities, denoted F (x, t), are then

180
given by the plane wave expansion
r 

0
(+)
(+)
(+)
F (x, t) =
ER(L) (x, t) icBR(L) (x, t)
2
r
0
(
x i
y) i(kzt)
i (
x i
y) i(kzt)

E (x, t)
e
icE (x, t)
e
=
2
2
c
2
r
0
(
x i
y) i(kzt)
2E (x, t)
=
e
.
2
2

(A.3)

This shows the explicit connection between LCP (RCP) with positive and negative
helicity.
The definition of the photon wave function in coordinate space that accounts for
both helicities can be written as a six-component, complex-valued vector related to
the RS vector by [68, 73]

(+)
F (x, t)
,

(x, t) =

F() (x, t)

(A.4)

where the F() (x, t) are defined as the positive- and negative-frequency components
of the positive-helicity RS vector, Eq. (A.1), i.e.,
F() (x, t) =

() (x, ) eit d.
F

(A.5)

() (x, ) is the Fourier coefficient of the original (positive-helicity) RS vector


Here F
r
0
F (x, t) =
(E (x, t) + icB (x, t))
2
Z
(A.6)
it

it

=
(F (x, ) e + F (x, ) e
).

() (x, ) are defined only over the interval (0, ). This is because
Note that the F
the negative part of the frequency spectrum is redundant, and contains no additional
information (see Mandel and Wolf [87] for a discussion of analytic signals).

181
The wave function defined in equation (A.4) obeys the complex form of the
Maxwell equations
i~t (x, t) = 3 (i~s ) (x, t) ,
and

(A.7)

(+)

F (x, t) 0
= .
(x, t) =


F() (x, t)
0

(A.8)

Here 3 is the 6 6 block diagonal Pauli-z type matrix [68, 73], and s, is the threevector composed of the 3 3 spin-1 rotation matrices

0 0 0

,
sx =
0
0
i

0 i 0

0 0 i

,
sy =
0
0
0

i 0 0

0 i 0

.
sz =
i
0
0

0 0 0

(A.9)

The reason that sz is not diagonal here is due to the fact that the spin-1 matrices act
on the Cartesian vector components of the wave function, and not the components
labeled by the spin eigenvalues. The spin matrices si , operate separately on upper
and lower components of (x, t)

(+)

si F (x, t)
.
si (x, t) =

()
si F
(x, t)

(A.10)

The three Pauli-type matrices i , act on the wave function (x, t), as follows

()
(r, t)
F

,
1 (r, t) =

(+)
F (r, t)

()
(r, t)
iF

,
2 (r, t) =

(+)
iF (r, t)

(+)
F (r, t)

.
3 (r, t) =

F() (r, t)

(A.11)

182
A component the photon wave function has dimensions of [energy/volume]1/2 , when
defined in this way. This has implications when examining the Lorentz invariance of
inner products, and transformation properties of the wave function.
Under a Lorentz transformation characterized by the boost parameter = v/c,
where v is the relative velocity between two reference frames (unprimed and primed
coordinate systems), the positive- and negative-helicity RS vectors in the two different
frames, F and F0 , are related by
F0 (x0 , t0 ) = [F (x, t) i F (x, t)]

2
[ F (x, t)] ,
+1

(A.12)

where the fields on the right-hand side are evaluated at (x, t), which are functions of
the primed coordinate system through the standard Lorentz transformation
t = t0 + cx0k
x=

x0

x0k


(A.13)
0


+ t /c /.

p
Here is the standard relativistic factor = 1/ 1 2 , and the the position vector
x0 has been broken up into components parallel x0k , and perpendicular x0 to the boost
parameter . The transformations for the RS vector can be easily derived from the
electric and magnetic-induction field transformations [183]
2
[ E (x, t)]
+1
2
B0 (x0 , t0 ) = [B (x, t) E (x, t) /c]
[ B (x, t)] .
+1
E0 (x0 , t0 ) = [E (x, t) + c B (x, t)]

(A.14)

Note that this transformation preserves the square of the RS vector


2

(F0 ) = (F)2 ,

(A.15)

183
but not the modulus squared
2

|F0 | 6= |F|2 .

(A.16)

The Lorentz transformation of the RS vector (and equivalently, the wave function)
can be given as an orthogonal transformation

B 0
(x, t) ,
0 (x0 , t0 ) =

0 B

(A.17)

where B is a three-dimensional, complex orthogonal matrix, and B is its complex


conjugate. The wave function on the right-hand side is again take to be evaluated
at the unprimed coordinates, which are functions of the primed coordinates. The B
matrices characterize the Lorentz transformation, including spatial rotations. Note
that orthogonal implies BT B = 1, where BT is the transpose of the matrix B. The
form of B can be infered from the transformations in Eq. (A.12) by writing them out
in tensor notation
 0 0 0 
2
F (x , t ) j = [F (x, t) i F (x, t)]j
[ [ F (x, t)]]j
+1
2
= [F (x, t)]j i [ F (x, t)]j
j [ F (x, t)]
+1
2
= jk [F (x, t)]k ijkl l [F (x, t)]k
j k [F (x, t)]k
+1


2
= jk ijkl l
j k [F (x, t)]k .
+1

(A.18)

Here jkl is the totally anti-symmetric symbol of three indices, and we have made use
of the vector identities
A B = Aj Bj ,

(A.19)

184
and
A B = jkl Aj Bk = jkl Ak Bj .

(A.20)

We have implicitly used the Einstein summation convention, where repeated indices
are summed over. The orthogonal transformation matrix is given in tensor notation
by the term in square brackets on the last line of Eq. (A.18)
Bjk = jk + ijkl l
This can be written out in matrix form as

2
j k .
+1

(A.21)

z y
0
1 0 0
x x x y x z

+ i
.
B=
0

0
1
0

y x y y y z
z
x

+1

y x
0
0 0 1
z x z y z z
(A.22)

185
We now use these transformation properties to explicitly show the invariance of
the square of the RS vector under Lorentz transformations
2

2
( F)
(F ) = (F i F)
+1
 2 2

3
= 2 (F i F)2 +
2 ( F)2 2
(F i F) ( F)
+1
+1
4
2
2
2
2
2
2
= (F) ( F) 2i +
2 ( F)
( + 1)
3


( + 1)
2
2
(

F)

i
(

F)
( + 1)2
4 2 2 3 ( + 1)
( F)2
= 2 (F)2 2 ( F)2 +
( + 1)2


2 ( 2 1) 2 2
2
2
2
2
= (F) ( F) +
( F)
( + 1)2


2 + 2 + 1
2
2
2
2
= (F) ( F)
( F)
( + 1)2


= 2 (F)2 ( F)2 ( F)2
0 2



= 2 (F)2 2 (F)2


= 2 1 2 (F)2
= (F)2 .
(A.23)
We have made liberal use of the fact that the cross product of two vectors is orthogonal
to the two vectors, the dot product of two orthogonal vectors is zero, and the definition
of , which leads to the identity 1 2 = 2 . Another way to see that (F )2 is
invariant under Lorentz transformations is to note that it can be written in terms

of the scalar invariant, S = 0 E2 c2 B2 /2, and pseudo-scalar invariant, P =
p
0 /0 E B, of the electromagnetic field (F )2 = S iP .

186
Next we perform a Fourier expansion of the positive- and negative-helicity RS
vectors

d3 k
(2)3
Z

d3 k
F (x, t) = ~c
(2)3
F+ (x, t) =

~c



f+ (k) ei(kxt) + f (k) ei(kxt) ,
(A.24)


f (k) ei(kxt) + f+ (k) ei(kxt) .

Here = c|k| is the frequency associated with the wave vector k. The fact that
B+ and F are complex conjugates of each other is reflected in the relationship
between the expansion coeficients (note that they are not independent). The vector
dependence is explicitly accounted for with unit helicity vectors, which can be extracted
from the vector expansion coefficients

f+ (k) = e (k) f (k, 1) ,


(A.25)
f (k) = e (k) f (k, 1) .

The two complex functions f (k, ), where = 1, describe the independent degrees
of freedom of the free electromagnetic field. The RS vectors for the two different
helicities can then be expressed as



d3 k
i(kxt)

i(kxt)
F+ (x, t) = ~c
e
(k)
f
(k,
1)
e
+
f
(k,
1)
e
,
(2)3
Z d3 k


i(kxt)

i(kxt)
F (x, t) = ~c
e
(k)
f
(k,
1)
e
+
f
(k,
1)
e
.
(2)3
Z

(A.26)

The free space energy, momentum, angular momentum, and moment of energy of

187
the electromagnetic field can be expressed inthe momentum-space representation [68]
E=

X Z
{1,1}

d3 k
~f (k, ) f (k, ),
(2)3 |k|

d3 k
~kf (k, ) f (k, ),
3
(2) |k|
{1,1}


Z
X
1
d3 k
k
~k Dk + ~
M=
f (k, ) f (k, ),
3
i
|k|
(2)
|k|
{1,1}
X Z
d3 k
N=
i~ Dk f (k, ) f (k, ),
3
(2) |k|
{1,1}
P=

X Z

(A.27)

where
Dk = k + i (k) ,

(A.28)

(k) = l1 (k) k l2 (k) l2 (k) k l1 (k) .

(A.29)

and

Here the li (k) are real unit vectors that make up the polarization unit vector e (k)
e (k) =

(l1 (k) + il2 (k))

.
2

(A.30)

Together with the unit vector in the k direction given by n (k) = k/ |k|, these form
an orthonormal triad
li (k) lj (k) = ij ,
n (k) li (k) = 0,

(A.31)

l1 (k) l2 (k) = n (k) .


We may write the coordinate-space photon wave function Eq. (A.4), in terms
of the Fourier-space amplitudes functions f (k, ). This can be done by first noting
that f (k, ) describe field amplitudes with definite energy, momentum and helicity,

188
given by ~, ~k, and = 1, respectively. The functions, f (k, ), have a dual
interpretation. In the classical theory they yield full information about the electromagnetic
field. In the wave mechanics of a single photon, the functions are the components of
the photon wave function in the momentum representation with polarization vector
e(k) or e (k). Following Bialynicki-Birula [68], we distinguish the two cases by
changing f (k, ) to (k, ) to distinguish the photon wave function in momentum
space.

Note that this differs slightly from the approach taken by Sipe [60], in

which the momentum-space wave function is a true probability amplitude, whereas


in the Bialynicki-Birula treatment, the momentum-space wave function is an energy
amplitude. Sipes momentum-space wave function S (k, ), is directly proportional
to the Bialynicki-Birula momentum-space wave function (k, )
1
S (k, ) = p (k, ) .
|k|

(A.32)

Thus we are led to the following expansion of the photon wave function in the planewave basis

(x, t) =

Z
~c

e (k, 1) (k, 1)

d3 k

exp (ik x it).


3

(2)
e (k, 1) (k, 1)

(A.33)

Here e (k, 1) = e (k), and e (k, 1) = e (k) are the unit polarization vectors corresponding
to positive and negative helicity, respectively. The unit polarization vectors are
orthogonal to the wave vector k

k e (k, ) = 0.

(A.34)

189
Note that the wave function in Eq. (A.33) can be written in terms of the positivefrequency components of the real and imaginary parts (for each helicity), which
resemble the electric and magnetic-induction fields

r
(x, t) =

E(+) (x, t) + icB(+) (x, t)


0

2 (+)
E (x, t) icB(+) (x, t)

(A.35)

This analogy between the photon wave function and classical electromagnetism is
only formalistic. One should not interpret the real and imaginary parts for the photon
wave function as the electric and magnetic-induction field of the photon. Rather, the
electric and magnetic fields may be thought of as emergent quantities when many
photons are present.
We now examine the Lorentz transform properties of the photon wave function
utilizing its Fourier expansion in momentum-space wave functions (k, ), Eq. (A.33).
In particular, we are interested in the transformation properties of the expansion
coefficients (k, ).

We can easily determine that the only affect of a Lorentz

transform on the expansion coefficients is a phase shift. The form of that phase shift
is not as easily determined, but is irrelevant in determining the Lorentz invariance of
scalar products. To see that under Lorentz transformations the (k, )s undergo a
only phase shift, first recall that an energy-momentum four vector (E, p), transforms
as

0
E
E E cpk

=
,


p
p0
pk E/c / + p

(A.36)

190
Here we have broken up the momentum into components parallel pk , and perpendicular,
p , to the boost parameter ,
pk = p /,

(A.37)

p = p p /.
This can also be done for the frequency-wave-vector four vector, (, k). We examine
the free-space expressions for the energy and momentum of the photon, Eq. (A.27),
in terms of the momentum-space wave functions, (k, )
X Z

E=

{1,1}

X Z

P=

{1,1}

d3 k
~ (k, ) (k, ),
3
(2) |k|
d3 k
~k (k, ) (k, ).
3
(2) |k|

(A.38)

Performing a Lorentz transformation to the primed basis gives


X Z

E =

{1,1}

X Z

P =

{1,1}

d3 k 0
~ 0 0 (k0 , ) 0 (k0 , ),
(2)3 |k0 |
d3 k 0
~k0 0 (k0 , ) 0 (k0 , ).
3
0
(2) |k |

(A.39)

The integrand volume element d3 k 0 / |k0 |, is Lorentz invariant, and using the transforms
of 0 and k0 explicitly, we get
X Z

E =
0

P =

{1,1}
3 0

X Z
{1,1}


d3 k
~ ckk 0 (k, ) 0 (k, ),
3
(2) |k|




dk
~ kk /c / + k 0 (k, ) 0 (k, ).
3
(2) |k0 |

(A.40)

Using the decomposition of the wave vector in terms of parallel and perpendicular

191
components along the boost axis, p = pk / + p , we see
X Z
d3 k
0
~ 0 (k, ) 0 (k, )
E =
3
(2) |k|
{1,1}
Z
X
d3 k
c
~kk 0 (k, ) 0 (k, ),
3
(2)
|k|
{1,1}
X Z
d3 k 0
P0 = /
~kk [(/c)] 0 (k, ) 0 (k, )
3
0
(2) |k |
{1,1}
X Z
d3 k 0
/c
~0 (k, ) 0 (k, )
3
0|
(2)
|k
{1,1}
X Z
d3 k 0
~k 0 (k, ) 0 (k, ).
+
3
0
(2) |k |
{1,1}

(A.41)

These expressions for energy and momentum must transform like four-vector components,
as in Eq. (A.36), which implies that
X Z
0
E =
{1,1}

d3 k
~kk 0 (k, ) 0 (k, )
3
(2) |k|

X Z

d3 k
~ 0 (k, ) 0 (k, )
3
(2) |k|

{1,1}

(A.42)

= E cpk ,
and
0

P = /

X Z
{1,1}

/c

X Z
{1,1}

X Z
{1,1}

d3 k 0
~kk 0 (k, ) 0 (k, )
3
0
(2) |k |
d3 k 0
~0 (k, ) 0 (k, )
(2)3 |k0 |

(A.43)

d3 k 0
~k 0 (k, ) 0 (k, )
3
0
(2) |k |

= (/) pk (/c) E + p .
This can only occur if the modulus squared of the photon wave function in momentum
space is invariant, that is
0 (k, ) 0 (k, ) = (k, ) (k, ) .

(A.44)

192
If this were not the case, then the expressions for the electromagnetic energy and
momentum would not transform like components of a four-vector. This means that at
most, a Lorentz transformation will cause the scalar momentum-space wave function
(k, ) to gain a phase factor
0 (k0 , ) = (k, ) ei(k,) ,

(A.45)

where the phase function, (k, ), depends on the Lorentz transformation, . The
specific form of the phase function can be determined from a group-theoretic treatment
of the Poincare group [44], but is rather obscure and does not add to the discussion
of Lorentz invariance of the scalar product.
In wave mechanics one must have a properly defined scalar product, h1 | 2 i,
that is to be used in calculation of transition probabilities. The modulus squared
of the scalar product of two normalized wave functions, | h1 | 2 i |2 , determines the
probability of finding the system in state 1 , when the system is known to be in
state 2 . This probability is a number, and thus must be invariant under all Poincare
transformations. The standard definition of the scalar product
Z
h1 | 2 i =

d3 x1 (x, t) 2 (x, t),

(A.46)

cannot be used because it is not Poincare invariant, nor is it dimensionally correct.


The true scalar product for the photon wave function is found by examining the
wave functions in momentum space. This is motivated by noting that in quantum
mechanics, each photon carries energy ~, and momentum ~k. From this statement

193
one can determine that the total number of photons, N , in the electromagnetic field
(quantum or classical) is obtained by dividing the integrand of the total-energy
expectation value, Eq. (A.38), by ~
N=

X Z
{1,1}

d3 k
(k, ) (k, ).
3
(2) |k|

(A.47)

In the case of a single-photon field, we must have N = 1. Thus, normalized photon


wave functions in momentum space, (k, ), must satisfy the normalization condition
d3 k
(k, ) (k, ) = 1.
3
(2) |k|

X Z
{1,1}

(A.48)

From this expression for the photon wave function normalization, one can deduce the
proper form of the scalar product for single-photon states formulated in momentum
space
(1 ||2 ) =

X Z
{1,1}

d3 k
1 (k, ) 2 (k, ).
3
(2) |k|

(A.49)

We have introduced the non-standard notation (1 ||2 ) for the scalar product to
emphasize that this is not the usual expression. To express this scalar product in
coordinate space, we use the inverse Fourier transform of the plane wave expansion,
Eq. (A.33), which gives the momentum-space wave functions (k, ), in terms of the
coordinate space wave functions
1
(k, ) = e (k, )
~c

d3 x exp (ik x) (x, t).

(A.50)

Here the scalar (dot) product with e (k, ) is evaluated separately for upper and lower
components of (x, t), and for each , only one of the dot products is non-vanishing.

194
We will also make use of the following properties of the unit vectors e (k, ),
e (k, ) e (k, 0 ) = ,0 ,
X
ki k j
ei (k, ) ej (k, ) = ij 2 .
k

(A.51)

Here the second equation shows the transverse nature of the photon wave function,
and equals the transverse delta function in momentum-space [90]. Combining Eqs.
(A.49), (A.50), (A.50) we find the coordinate space scalar product
d3 k
1 (k, ) 2 (k, )
3
(2)
|k|
{1,1}
Z
Z
Z
X
1
d3 k
0
3 0
=
dx
d3 x e[ik(xx )]
3
~c
(2) |k|
{1,1}
X Z

(1 ||2 ) =

e (k, ) 1 (x, t) e (k, ) 2 (x0 , t)


Z
Z
Z
d3 k
1
0
3 0
dx
d3 x e[ik(xx )]
=
3
~c
(2) |k|
X



[ei (k, ) 1i (x, t)] ej (k, ) 2j (x0 , t)


1
=
~c

{1,1}
3

dk
(2)3 |k|

3 0

d3 x e[ik(xx )] 1i (x, t) 2j (x0 , t)

dx

(A.52)
X 


ei (k, ) ej (k, )
{1,1}

1
=
~c

d3 k
(2)3 |k|

3 0

dx
Z

1
=
" ~c

[ik(xx0 )]

d xe

d3 k
(2)3 |k|

3 0

dx

1i



ki k j
(x, t) 2j (x , t) ij 2
k
0

d3 x e[ik(xx )]

0
k

(x,
t)
k

(x
,
t)
2
1
1 (x, t) 2 (x0 , t)
k2
Z
Z
Z
1
d3 k
0

0
3 0
3
=
d x1 (x, t) 2 (x , t)
dx
e[ik(xx )] I
3
~c
(2) |k|
Z
Z

1
(x, t) 2 (x0 , t)
= 2
d3 x0 d3 x 1
I.
2 ~c
|x x0 |2

195
Here we have used the following Fourier transform pair
1
1
,

2
r
4k

(A.53)

where the Fourier transform is defined by


Z

d3 r eikx g (x),
Z
d3 k ikx
1
g (x) = F {G (k)} =
e
G (k).
(2)3

G (k) = F {g (x)} =

(A.54)

The final term I, on the right-hand side of the bottom line of Eq. (A.52), is
1
I=
~c

d3 k
(2)3 |k|

3 0

dx

[ik(xx0 )] k

d xe

1 (x, t) k 2 (x0 , t)
.
k2

(A.55)

This term actually vanishes, as can be seen in the following calculation


Z
Z
Z

0
1
d3 k
3 0
3
[ik(xx0 )] k 1 (x, t) k 2 (x , t)
I=
d
x
d
x
e
~c
k2
(2)3 |k|
Z
Z
Z
Z
Z
d3 k 0
d3 k
d3 k 00
0
0
00 0
=
d3 x0 d3 x e[ik(xx )] e[i(k xk x )]
3 3
3
3
(2) k
(2)
(2)
X

k e (k0 , 0 ) 1 (k0 , 0 ) k e (k00 , 00 ) 2 (k00 , 00 )


0 , 00
3 0

d3 k 00 X
k e (k0 , 0 ) 1 (k0 , 0 ) k e (k00 , 00 ) 2 (k00 , 00 )
(2)3 0 ,00
Z
Z
0
3 0 [i(k00 k)x0 ]
d xe
d3 xe[i(k k)x]
Z 3 0 Z 3 00 X
Z
dk
dk
d3 k
k e (k0 , 0 ) 1 (k0 , 0 ) k e (k00 , 00 ) 2 (k00 , 00 )
=
3 3
3
3
(2) k
(2)
(2) 0 ,00
Z

d3 k
(2)3 k 3

dk
(2)3

(2)3 (3) (k0 k) (2)3 (3) (k00 k)


Z
d3 k X
=
k e (k, 0 ) 1 (k, 0 ) k e (k, 00 ) 2 (k, 00 )
3 3
(2) k 0 ,00
= 0.
(A.56)

196
In the second to last line of Eq. (A.56) we have used the orthogonality relation Eq.
(A.34), between the unit polarization vectors e (k, ) and the wave vector, k.
Note that the scalar product is non-local in coordinate space, and leads to the
following non-local, coordinate-space normalization
2

d3 k
(k, ) (k, )
3
(2) |k|
{1,1}
Z
Z
1
(x, t) (x0 , t)
3 0
= 2
dx
d3 x
.
2 ~c
|x x0 |2

N = kk = h | i =

X Z

(A.57)

The non-locality in coordinate space has a simple interpretation the photon number
density is a non-local object in coordinate space, i.e., there is no local expression for
the photon probability density in coordinate space [51].
In conlusion, the Lorentz invariance of the photon-wave-function scalar product
Eq. (A.48), can be explicity shown by examining the Lorentz-transformation properties
of the energy-momentum four vector. This gives the result that the norm of the wave
function is invariant under Lorentz transformations, which we expect on the physical
grounds that numbers are Lorentz invariant. The Lorentz-invariant scalar product
is a local integral in momentum space, which upon transformation to coordinate
space gives a non-local expression. This non-local expression for the photon number
is consistent with the well-known absence of a localized coordinate-space photonnumber operator [51].

197

APPENDIX B

SPONTANEOUS PARAMETRIC DOWN CONVERSION

A key resource in many quantum optics experiments, spontaneous parametric


down conversion (SPDC) has been utilized in experiments that probe the fundamental
behavior of quantum-mechanical systems. These include experiments violating the
Bell inequality, quantum-interference experiments such as the Hong-Ou-Mandel effect,
and more recently quantum-cryptography proof-of-principle experiments. Spontaneous
parametric down conversion, also called spontaneous parametric scattering was first
predicted in 1961 [161], and first experimentally observed in 1967 [156, 157, 159, 184].
The outcome of standard SPDC is the production of an entangled photon pair.
In the typical SPDC experiment, a strong (assumed classical) beam of light known as
the pump, is incident on a second-order nonlinear optical crystal (usually birefringent
for reasons detailed below). Through the second-order nonlinear process of differencefrequency generation, seeded by the vacuum, a single pump photon is annihilated,
and two daughter photons, known historically as the signal and idler, are created.
The entanglement of the signal and idler photons arises from energy and momentum
conservation in the SPDC process. Note that energy is conserved in this process
because the medium is non absorbing in the spectral region considered.

There

are two possible geometries for momentum conservation to be satisfied. Collinear

198
SPDC involves the signal and idler photons emerging from the nonlinear crystal
in the same direction as the pump. In non-collinear SPDC, the signal and idler
photons emerge from the crystal with wave-vector directions different from the pump.
The process comes in two types type-I, in which the signal and idler photons
have the same polarization, and type-II, in which the signal and idler photons are
orthogonally polarized to each other. This leads to four different possible experimental
configurations shown in Fig. 26 of Chapter IV.
The standard theoretical approach to determine the output state of light from
the SPDC process treats the electromagnetic field as a quantum field, which interacts
with a bulk, classical nonlinear crystal. A first-order, time-dependent perturbation
expansion in the interaction picture [100, 139, 142, 162] gives the output state of
the signal and idler modes of the quantized electromagnetic field. The nonlinear
interaction gives rise to the interaction Hamiltonian.
To obtain the general form of the interaction Hamiltonian, we consider the classical
problem of nonlinear optical interaction, which may then be quantized. The interaction
Hamiltonian is found by considering the electromagnetic field Hamiltonian [183]

1
H=
2

(D E + B H) d3 x,

(B.1)

where the integral is take over the volume of interest (the crystal volume for our
problem). The constitutive relations between the displacement field D, and the

199
electric field E, and the magnetic field H, and the magnetic induction B are
D = 0 E + P,
(B.2)
1
H = B M.
0
Here P and M are the polarization and magnetization vectors of the medium, and 0
and 0 are the vacuum permittivity and permeability. Most common optical materials
are non magnetic, and thus we set M = 0. The medium polarization P, is expanded
in powers of the incident electric field [164166], which for a single vector component
(0)

Pi

(x, t), (i = x, y, z), gives


(0)

Pi (x, t) = Pi
(0)

where Pi

(1)

(x, t) + Pi

(2)

(x, t) + Pi

(x, t) + ,

(B.3)

(x, t) is a constant polarization of the medium, independent of the incident


(1)

field (indicative of ferroelectric materials). Pi

(x, t) is the medium polarization

that is linearly dependent on the applied field, known as the first-order polarization.
(2)

Pi

(x, t) is the medium polarization that depends quadratically on the incident field,

known as the second-order polarization, and so on. The material response of a medium
is characterized by its susceptibility tensors [164166]. If we assume the medium
polarization induced by an external field at some point x inside the medium is only
affected by the value of the applied field at that point, which is usually a good
approximation, we say the response of the material is local. When the material
response is local, the induced polarization of order n is
(n)
Pi

Z
(x, t) = 0

Z
dt1

(n)

dtn ij1 jn (t t1 , , t tn ) Ej1 (x, t1 ) Ejn (x, tn ),

(B.4)

200
where we employ the Einstein summation convention on repeated indices. Here
(n)

ij1 jn (t t1 , , t tn ) is known as the nth-order time-domain polarization response


tensor, or the time-domain susceptibility tensor. The need for time integrals stems
from the fact that the material response is not instantaneous, and the medium has
a memory of the applied field. In most cases, the higher-order tensors (larger
n) are increasingly small. The SPDC process arises in materials with a secondorder nonlinear susceptibility (2) , which occur only in materials without inversion
(0)

symmetry [166]. For non-ferroelectric materials Pi

(x, t) = 0, without inversion

symmetry, the induced polarization up to second order is [164166]


Z
(1)
Pi (x, t) = 0
dt1 ij (t t1 ) Ej (x, t1 )
Z
+0

Z
dt1

(2)
dt2 ijk

(B.5)

(t t1 , t t2 ) Ej (x, t1 ) Ej (x, t2 ).

The first term on the right-hand side of Eq. (B.5) is the linear material response,
which characterizes the medium refractive indices and absorption properties. The
second term is the quadratic nonlinear medium response that gives rise to three-wave
mixing processes. Decomposing the electric field into positive- and negative-frequency
(+)

parts, (Ei

()

(x, t) and Ei

(x, t)), a component of the electric field is written (i =

x, y, z)
(+)

Ei (x, t) = Ei

()

(x, t) + Ei

(x, t) ,

(B.6)

where
()
Ei

Z
(x, t) =
0

Ei (x, ) exp (it).

(B.7)

201
Inserting Eq. (B.7) into Eq. (B.5) we get the induced polarization

Z
Pi (x, t) = 0
Z
+0

Z
d

d
0

0
0

(2)

d 0 ijk (, 0 ) Ej (x, ) Ej (x, 0 ) ei(+ )t + c.c.

(B.8)

Z
+0

(1)

d ij () Ej (x, ) eit + c.c.

(2)

d 0 ijk (, 0 ) Ej (x, ) Ej (x, 0 ) ei( )t + c.c.,

where c.c. means complex conjugate. Here we have defined the susceptibility tensors
as the Fourier transforms of the time-domain polarization response tensors

(1)
ij

Z
() =

(2)

ijk (, 0 ) =

(1)

dt1 ij (t t1 ) ei(tt1 ) ,
Z

Z
dt1

(B.9)
i 0 (tt2 )

(2)

dt2 ijk (t t1 , t t2 ) ei(tt1 ) e

Combining the polarization in Eq. (B.8) with the constitutive relations Eq. (B.2),
assuming a non-magnetic material, and inserting the result in the electromagnetic
Hamiltonian Eq. (B.1), we find the electromagnetic Hamiltonian expression


Z 

1
1
(1)
(2)
H=
0 E + P + P
E + B B d3 x
2
0
V

Z 
Z


1
0
1  (2)
(1)
3
=
1+
EE+
BB d x+
P E d3 x
2
0  0
2
V

= H0 + HI ,

(B.10)

202
where the interaction Hamiltonian, H I , is

Z
Z
Z
0
0
(2)
d3 x Ei (x, t) d d 0 ijk (, 0 ) Ej (x, ) Ej (x, 0 ) ei(+ )t .
HI (t) =
2
0
0

Z
Z
0
(2)
+ d
d 0 ijk (, 0 ) Ej (x, ) Ej (x, 0 ) ei( )t
0

+c.c..
(B.11)
Upon quantization in the paraxial approximation (the z-axis being the direction
of propagation), the electric field is split into positive and negative frequency parts,
and quantized in terms of creation and annihilation operators
Z
XZ
2
(+)
(x, t) =
d q d E eq,, ei[kxt] a
(q, ) ,
E

() (x, t) =
E

XZ

dq

(B.12)

Z
d

E eq,, ei[kxt] a

(q, ) .

Here

(q, ) and a
(q, ) are the creation and annihilation operators for electro-

magnetic field excitations with transverse wave vector q, frequency , and polarization
denoted by . The eq,, , ( = 1, 2) are orthonormal polarizations unit vectors,
associated with wave vector k = (q, kz (q, )). The longitudinal component of the
wave vector kz (q, ), is
q
kz (q, ) = k ()2 |q|2 .

(B.13)

Here the wave vector magnitude is


k () =

n () ,
c

(B.14)

where n () is the refractive index inside the crystal for light with frequency , which
generally depends on the polarization and propagation direction. The field amplitudes

203
E , are
r
E = i

~
.
20

(B.15)

The annihilation and creation operators obey the canonical bosonic commutation
relations
h

a
(q, ) , a
0

i
(q , ) = ,0 (2) (q q0 ) ( 0 ) ,
0

(B.16)

with all other commuatators zero. With this paraxial, continuous-mode quantization
scheme, the interaction Hamiltonian operator is
(2)

I (t) =
H

0 ef f
2

dx

d qp

i[(kp ks ki )x(p s i )t]

Z
dp

d qs

a
p (qp , p ) a
s

Z
ds

(qs , s ) a
i

d qi

di Ep Es Ei

(B.17)

(qi , i ) + H.c.,

where H.c. stands for Hermitian conjugate. We have assumed that the nonlinear
susceptibility is non dispersive in the frequency range of interest. This enables us to
pull it out of the frequency integrals, and define the effective nonlinearity
(2)

ef f =


(2)
ijk ep i (es )j (ei )k .

(B.18)

p ,s ,i

This expression accounts for the polarization vectors of the three interacting fields.
The spatial integration is taken over the quantization (crystal) volume.
In the interaction picture, the state evolution of a system is governed by the
unitary operator

1
U (t, t0 ) = exp
i~

Zt

I (t0 ) dt0 .
H

(B.19)

t0

The state of the system at time t, is related to the state of the system at an earlier

204
time t0 , through this unitary operator by
| (t)i = U (t, t0 ) | (t0 )i .

(B.20)

If the interaction is weak, that is, the interaction Hamiltonian is small, we can expand
the unitary operator in Eq. (B.19) to lowest order

|(t)i 1 +

1
i~

Zt

I (t0 ) dt0 |(t0 )i .


H

(B.21)

t0

In the treatment that follows, we assume a classical pulsed pump with central
frequency p , traveling in the z direction. The pump has a temporal pulse envelope
p (t), transverse spatial mode up (x, y), and is assumed to be linearly polarized in the
x direction. We can make this approximation since the incident pump is an intense
laser pulse, and the interaction is quite small, effectively leading to no loss of the
pump field. In the paraxial approximation, the electric field of the pump is
Z
Ep (x, y, z, t) = Ap ex

Z
dp

d2 qp
p (p + p ) up (qp )
(B.22)

exp {i [k (p + p ) (p + p ) t]} ,
where we have Fourier expanded the temporal pump envelope about its central
frequency p , and the pump transverse profile in transverse wave vector qp . Here

p (p ), is the pump spectral amplitude given by the Fourier transform of the pump
temporal envelope p (t). The pump transverse wave vector spectral amplitude up (qp )
is given by the two-dimensional Fourier transform of the pump transverse envelope
up (x, y).

205
We assume only the pump field to have non-zero amplitude at the crystal input.
The pump can be described by the coherent state |p i, characterized by the Fourier
transforms of the pulse-envelope, and transverse mode functions,
p (p ), and up (qp ),
respectively. The signal and idler fields are assumed to have no field excitations at
the input, that is, they are in the vacuum state |vaci. The initial state of the system
is then
|in i = |p i |vacis |vacii .

(B.23)

We are interested in the signal and idler fields after the pump pulse has passed
through the nonlinear medium, and thus may take the time limits in Eq. (B.21) to
without trouble. Under the above assumptions, the output state of the system
is found by inserting Eqs. (B.17), (B.22), and (B.23) into Eq. (B.21)
Z
|f i |vaci +

d qs

Z
ds

d qi

Z
di A (qs , qi ; s , i ) |1qs ,s i |1qi ,i i .

(B.24)

The two-photon amplitude A (qs , qi ; s , i ), is determined by the pump spectral


amplitude, the pump transverse-momentum distribution, and the phase-matching
conditions, (which are governed by the mediums linear dispersion properties)
Z
Z
(2)
i0 ef f Ep Es Ei
2
A (qs , qi ; s , i ) =
d qp dp
p (p p ) up (qp )
2~

Z
Z

dt d3 x ei[(kp ks ki )x(p s i )t] .

(B.25)

Here the spatial integration is taken over the volume of the nonlinear crystal, V . The
time integral yields a delta function in frequencies (p s i ), ensuring energy
conservation. The spatial integral can be broken into transverse and longitudinal

206
parts
Z

d3 x ei(kp ks ki )x =

ZL

dz eikz z

d2 eiq

(B.26)

= Leikz L/2 sinc (kz L/2) (2) (q) .


The wave vector mismatches are
kz = kpz ksz kiz ,

(B.27)

for the longitudinal wave-vector mismatch and


q = qp qs qi ,

(B.28)

for the transverse wave-vector mismatch. Here kjz and qj , (j = p, s, i), are the
longitudinal and transverse components of the wave vector. For the longitudinal part
we have taken the crystal length to be L, giving the sinc-function sinc (x) = sin (x) /x,
which is known as the phase-matching function. If the pump transverse beam waist
is much smaller than the transverse dimensions of the crystal, (a good approximation
in most experimental setups), then the transverse integral limits are taken to , as
done in Eq. (B.28). This gives the delta function of transverse wave-vector mismatch
q. (Note that the sinc-function or phase-matching function, is maximized when the
difference in longitudinal wave vectors is zero. This condition is known as optimal
phase matching or the phase-matching condition.) The expression for the two-photon
amplitude in Eq. (B.25) is further simplified by using Eq. (B.26), giving
(2)

i0 ef f Ep Es Ei ikz L/2


Le
sinc (kz L/2)
A (qs , qi ; s , i ) =
2~

p (p s + i ) up (qs + qi ) .

(B.29)

207
It is clear that if the longitudinal wave-vector mismatch is zero, that is, if kz =
0, then the two-photon amplitude depends only on the pump beam properties. In
general, kz is non zero. The next step is to calculate the wave vector mismatch
kz , for a specific experimental geometry.
Recall that experimentally there are two phase-matching types in birefringent
nonlinear optical crystals, known as type-I and type-II phase matching. In typeI phase matching, the pump beam propagates with extraordinary (e) polarization,
and the down-converted beams both have ordinary (o) polarization orthogonal to
the pump polarization. In type-II phase matching the pump beam again typically
has extraordinary polarization, but the two down-converted beams have opposite
polarizations. One down-converted photon has ordinary polarization and the other
extraordinary polarization. The down-converted fields produced with these two types
of phase matching have similar spatial and temporal behaviors in the nearly-collinear
phase-matching regime (that is when the pump, signal and idler beams propagate in
nearly the same direction). We shall only treat type-I phase matching here, because
this is what is employed in our lab. The generalization to type-II is straightforward
[100].
Consider the non-collinear, type-I down-conversion experiment depicted in Fig. 26
of Chapter IV. Efficient down conversion occurs when the phase-matching condition
is met, kz = 0. It is often difficult to achieve optimal phase matching in lossless
(transparent) materials because the refractive index typically displays normal-dispersion

208
behavior. That is, the index of refraction increases as the frequency of light increases.
This is why birefringent nonlinear crystals are often used. Birefringence is the dependence
of the refractive index on the polarization direction of the light [164, 168]. Phasematching in a birefringent nonlinear crystal is achieved by precise angular orientation
of the crystal with respect to the propagation direction and polarization of the
incident light, as sketched in Fig. 31. We use a beta-barium borate (BBO) nonlinear
crystal, which is a negative-uniaxial crystal. Uniaxial crystals are characterized by a
particular direction known as the optic axis (OA). Light polarized perpendicular to
the plane containing its propagation vector k and the OA, is said to have ordinary
(o) polarization, and propagates with the ordinary refractive index no (). Light
polarized in the plane containing its propagation vector and the OA, is said to have
extraordinary (e) polarization, and propagates with a refractive index ne (, OA ) that
depends on the angle OA , between the wave vector k, and the OA [164166]
sin2 (OA ) cos2 (OA )
1
=
+
.
n2e (, OA )
n2e ()
n2o ()

(B.30)

Here ne () is called the principle value of the extraordinary refractive index. Phase
matching is achieved by adjusting the angle OA , so that k = 0. Note negative
uniaxial implies that ne < no .
The type-I down-conversion setup depicted in Fig. 44 has the optic axis (OA) of
the nonlinear, birefringent crystal in the x-z plane, making an angle OA with the
z axis. The pump beam propagates with a wave vector kp , having magnitude Kp ,
which is assumed to be collinear with the z axis. Assuming the crystal has been

209
cut so that degenerate (s = i = p /2), collinear phase matching occurs when the
pump beam propagates along the z direction, normal to the surface of the crystal
face, the signal and idler will have wave vectors along the z axis with magnitude Ko .
To determine the spatial and spectral behavior of the non-collinear down conversion
when the phase matching is not quite met (kz 0), we perform a perturbation
expansion in the transverse wave vectors, and frequencies of all three fields (pump,
signal and idler), about this ideal phase-matching case [100].

FIGURE 44. Diagram of degenerate, type-I, non-collinear SPDC. The pump beam
propagates with wave vector kp , which is nearly collinear with the z axis. The
degenerate signal and idler photons are emitted into a cone with opening half-angle
0 . The optic axis (OA) of the crystal makes an angle OA with the z axis, which is
adjusted for optimal phase matching.

We begin by expanding the longitudinal wave-vector mismatch, kz , about the


ideal phase-matching condition. We assume that the phase-matching condition is met

210
when the pump has frequency p and wave vector Kp in the z direction. The signal
and idler in this ideal phase-matching case are emitted at the degenerate frequency
p /2, with wave vectors collinear to the pump, with magnitude Ko . The pump beam,
which is e-polarized, and the signal and idler beams, which are o-polarized, have the
following z-component wave vectors
kpz =

r

2
ne (p , p ) |qp |2 ,

(B.31)

for the pump and


kjz =

r

2
no (j ) |qj |2 ,

j = s, i,

(B.32)

for the signal and idler. Here j , (j = p, s, i), is the perturbed frequency of the
pump, signal or idler, and qj , (j = p, s, i), is the transverse wave-vector perturbation
about the z axis. The z-component of the wave vectors are expanded about the ideal
phase-matching conditions p = p , s(i)p = p /2 and qj = 0
kpz = Kp +

p 1 0 2
|qp |2
+ Dp p Np qpx
,
up 2
2Kp

(B.33)

where
Kp =

p
ne (p , OA ) ,
c

(B.34)

for the pump. The unperturbed pump wave-vector magnitude Kp , is evaluated at the
pump central frequency (p ), when propagating along the z axis, and p = p p
is the pump expansion frequency. Here up is the group velocity of the pump
up =

!1

kp
,
p p =p ,qp =0

(B.35)

211
and

ne (p , p )
1
Np =
,

ne (p , OA )
p
p =OA

(B.36)

is an expansion parameter that comes from the fact that the pump propagates as an
extraordinary beam. We expanded to second order in frequency because the firstorder terms cancel when considering expansion about the degenerate, collinear phase
matching case, as we are doing. The group-velocity dispersion parameter Dp 0 , is given
by
Dp0

=
p

1
up (p )


.

(B.37)

The z-components of the signal and idler wave vectors are expanded about the
degenerate down-conversion frequency o = p /2, and the zero transverse wave vector
qs = qi = 0. The longitudinal wave-vector components for the signal and idler are
then
ks(i)z

s(i) 1 0
|qs(i) |2
2
= Ko +
+ Do s(i)
,
uo
2
2Ko

(B.38)

p
no (p /2) ,
2c

(B.39)

where
Ko =

is the magnitude of the unperturbed wave vector. The frequency perturbation is

s(i) = s(i) p /2.

(B.40)

Here uo and Do0 are the signal and idler group velocity and group-velocity dispersion
at the degenerate frequency. Combining Eqs. (B.27), (B.33), and (B.38) gives the

212
longitudinal wave-vector mismatch
1 0 2 1 0 2
|qs |2 + |qi |2 |qp |2
2
kz = Dpo (s + i ) + Dp p Do (s + i ) Np qpx +
+
, (B.41)
2
2
2Ko
2Kp
where
Dpo =

1
1
,
up u o

(B.42)

is a group-velocity mismatch parameter between the pump and down conversion. We


also used the frequency delta function to eliminate the pump frequency perturbation
p . Recall that the down-conversion amplitude is optimized when the longitudinal
wave-vector mismatch is zero. With this in mind, note that for a monochromatic
plane-wave pump propagating along the z axis (qs = qi , qp = 0, p = 0), the
degenerate down conversion (s = i = ) is emitted in a cone with transverse wave
vectors determined by the optimal phase-matching condition

kz = Do0 2 +

|qs |2
= 0.
Ko

(B.43)

In the general case, when the pump is not a mono-chromatic plane wave, the phasematching condition above is modified

kz = Dpo + Np q+x +

|q+ |2 + |q |2 |q+ |2
+
.
4Ko
2Kp

(B.44)

Here we have defined the frequency and transverse wave-vector sum and difference
values
= s i ,

q = qs qi ,

(B.45)

213
and made use of the delta function in Eq. (B.26) to eliminate the pump variables.
Using Eqs. (B.41) and (B.45) in the two-photon amplitude equation (B.29), we obtain
(2)

i0 ef f Ep Es Ei ikz L/2


Le

p (+ ) up (q+ )
A (q+ , q ; + , ) =
2~


Np L
(|q+ |2 + |q |2 )L |q+ |2 L
Dpo L
+
q+x +
+
.
sinc
2
2
8Ko
4Kp

(B.46)

This is the amplitude that we shall use to discuss the ghost imaging experiments.

214

APPENDIX C

GHOST IMAGING

The concepts of ghost imaging and diffraction, known collectively as quantum


imaging, were first proposed by Klyshko [170], and experimentally realized by Pittman
et. al. [26], and Ribeiro et. al. [29], respectively. Both ghost imaging and ghost
diffraction rely on the spatially-entangled photon pairs produced in spontaneous
parametric down conversion. In these experiments, the image of an object is nonlocally retrieved by correlating the intensities at two distant detectors. The general
quantum-imaging experimental setup is sketched in Fig. 45. Each photon, historically
called the signal (s) and idler (i) photons, emitted from the nonlinear crystal propagates
through separate linear optical systems characterized by an impulse response function
hj (rj , r, ), or transfer function Hj (qj , q, ), j = 1, 2 (see Appendix D). In one
photon path, the linear optical system contains an object (transmitting aperture) and
a fixed detector, while in the other photon path a detector is scanned to reproduce
a sharp image in the coincidences between the fixed and scanning detectors, of the
object or its Fourier transform (i.e., far-field diffraction pattern).

215

FIGURE 45. Schematic diagram of a standard ghost-imaging experiment. The


photons emerge from the nonlinear crystal source plane at z = 0. Propagation of the
electromagnetic field operators from the crystal through two different linear optical
systems is characterized by the impulse response functions h1 (r1 , r, ) and h2 (r2 , r, )
of the two linear optical systems.

In one linear optical system, an aperture that is to be imaged is placed. The


ghost image or diffraction pattern, depending on the experimental configuration,
appears in the coincidence detection rate between the two detectors, labeled D1 and
D2 . The detection rate of each individual detector remains approximately constant,
as shown in Fig. 40 of Chapter IV. There are two basic imaging systems that may be
employed to perform ghost imaging. One utilizes a plane wave pump beam [26], while
the other uses a focused pump beam [142]. The simplest configuration to implement
experimentally, and treat theoretically, is the situation involving a plane-wave pump
incident on the nonlinear crystal. We shall treat this example here, and refer the
reader to the literature for treatment of the alternative approach [142].

216
Consider the linear-optical system sketched in Fig. 46 below. Here a planewave pump beam propagating along the z-axis is incident on a thin, nonlinear optical
crystal. The signal beam propagates a distance da prior to encountering a transmitting
aperture to be imaged. The aperture is characterized by its transmission function
t (r). All of the light that passes through the aperture in the signal beam path is
collected by a large-area bucket detector D1 . The bucket detector does not distinguish
photon momentum or position, it only indicates if the signal photon passed through
the aperture. The idler beam propagates freely a distance db from the crystal before
passing through the imaging lens, (focal length f ). After the imaging lens, the
idler beam undergoes free-space propagation a distance d0b to the point-like, scanning
detector, D2 . Narrow-band interference filters are placed in front of each detector
(not shown) to select out a particular frequency from the broad-band down-conversion
light.
The coincidence counting rate between detectors D1 and D2 is proportional to the
squared modulus of two-photon detection amplitude
(+)
(+)
(2) (r1 , t1 , r2 , t2 ) = hvac| E1 (r1 , t1 ) E2 (r2 , t2 ) |SP DC i .

(C.1)

Here rj is a transverse position (relative to the beam (z) axis) in the plane of detector
(+)
Dj , and Ej (rj , tj ), (j = 1, 2) is the positive-frequency electric-field operator at

detector j.

The electric field operators at the detectors are determined by the

impulse response functions associated with the signal and idler paths. The downconversion two-photon state |SP DC i, is treated in Appendix B. Note that the two-

217
photon amplitude is equivalent to the two-photon wave function defined in Chapter
II when the detectors are sensitive to only the electric field, and not the magnetic
field.

FIGURE 46. Ghost imaging experimental setup. The signal beam propagates a
distance da freely from the crystal to aperture plane. All light that passes through the
aperture, which is represented mathematically by the aperture transmission function
t (r), is detected by a fixed bucket detector, D1 . The idler beam propagates freely
through a distance db prior to passing through a lens of focal length f and then
undergoes free space propagation a distance d0b , prior to being detected by a pointlike detector D2 , located at the transverse position r2 , which is scanned in the imaging
plane. An interference filter is placed in front of each detector (not shown).

The two-photon state emerging from the down-conversion crystal is (see Appendix
B and references therein)
Z
|SP DC i |vaci +

d qs

Z
ds

d qi

di A (qs , qi ; s , i )
(C.2)

as

(qs , s ) a
i

(qi , i ) |vaci ,

where |vaci is the vacuum state of the electromagnetic field, a


j (qj , j ) is the creation
operator for an excitation in the signal (j = s), or idler (j = i) field with transverse

218
wave vector qj , and frequency j , (j = s, i). The constant characterizes the probability
that a down-conversion event occurs. The two-photon amplitude A (qs , qi ; s , i ) is
related to the phase-matching conditions, the pump spectrum, and transverse pump
profile as discussed in Appendix B.
The positive-frequency electric-field operators at detectors D1 and D2 can be
expressed in terms of the field operators at the output plane of the nonlinear crystal
with the use of Fourier optics (see Appendix D). Thus we adopt a Heisenberg Picture
approach to the problem, in which the operators evolve while the states remain
constant. The generic form of the field operators at the detection planes are
Z
Z
(+)
2

E1 (r1 , t1 ) = d qs ds eis t1 Hs (r1 , qs ; s ) a


s (qs , s ) ,
Z
Z
(+)
E2 (r2 , t2 ) = d2 qi di eii t2 Hi (r2 , qi ; i ) a
i (qi , i ) ,

(C.3)

where the transfer function Hl (rm , q; ) (l = s, i and m = 1, 2) describes the propagation


of a photon (l) in the (q, ) mode, from the nonlinear crystal output plane to the
detection plane (m).

The form of the transfer function depends on the optical

elements placed in the beam path. With this formulation of field operator propagation,
we see that the general two-photon detection amplitude in Eq. (C.1) may be written
in terms of the two-photon down-conversion amplitude A (qs , qi ; s , i ) combining by
Eqs. (C.1), (C.2), and (C.3) to give
Z
Z
Z
Z
(2)
2
2
(r1 , t1 , r2 , t2 ) = d qs ds d qi di A (qs , qi ; s , i )

Hs (r1 , qs ; s ) Hi (r2 , qi ; i ) eis t1 eii t2

+ Hs (r1 , qi ; i ) Hi (r2 , qs ; s ) eii t1 eis t2 .

(C.4)

219
Note that for a symmetric two-photon amplitude, that is, for A (qs , qi ; s , i ) =
A (qi , qs ; i , s ), the two-photon detection amplitude in Eq. (C.4) simplifies to
Z
Z
Z
Z
(2)
2
2
(r1 , t1 , r2 , t2 ) = 2 d qs ds d qi di A (qs , qi ; s , i )
(C.5)
is t1 ii t2

Hs (r1 , qs ; s ) Hi (r2 , qi ; i ) e

The specific form of the transfer functions depends on the optical elements employed
in the imaging setup. The transfer functions for various ghost-imaging setups are
easily determined by using an unfolded schematic drawing, as shown in Fig. 47.
The unfolded diagram of the experimental setup depicted in Fig. 47 is given in Fig.
48

FIGURE 47. Unfolded diagram of ghost-imaging experiment. Unfolded schematic


drawing of the experimental ghost-imaging setup sketched in Fig. 47 above. The
source emits momentum-anti-correlated photon pairs in all possible directions, which
pass through an aperture with transmission function t(r1 or a lens before being
detected.

We see that the signal transfer function Hs (r1 , qs ; s ), is composed of free-space


propagation a distance da , followed by the aperture transmission function, and the

220
interference filter. The frequency filter, which is characterized by the frequency filter
function F1 (s ), is implied but not shown. We assume that the fixed bucket detector
D1 , is placed immediately after the aperture. This simplifies the calculation in that
further propagation after the aperture does not need to be considered. Thus the
signal impulse response function is
hs (r1 , r; s ) = F1 (s ) t (r1 )

s
2
is i s da i 2cd
e c e a |rr1 | ,
2cda

(C.6)

When Fourier transformed with respect to r, Eq. (C.6) gives the signal transfer
function

Hs (r1 , qs ; s ) = F1 (s ) ei
= F1 (s ) ei

s
d
c a

s
d
c a

s
2

ei 2cda r1 exp i

eicda |q|



qs

/(2s ) iqs r1


s
r
cda 1

2 (s /c)

da
t (r1 )

(C.7)

t (r1 ) .

From Fig. 48, the idler transfer function (impulse response function) is comprised of
free-space propagation a distance db , followed by a thin lens of focal length f . After
the lens, the idler freely propagates a distance d0b . The light then passes through
an interference filter characterized by the filter function F2 (i ), and is subsequently
registered by the point-like detector D2 . Just prior to the lens, the impulse response
function is
hF S (rl , r; i ) =

i
ii i i db i 2cd
|rrl |2
e c e b
.
2cdb

(C.8)

The lens adds a quadratic phase factor to the impulse response function in Eq. (C.8),
so that just after the lens the impulse response function is
h (rl , r; i ) =

i
ii i i db i 2cd
|rrl |2 i i |rl |2
e c e b
.
e 2cf
2cdb

(C.9)

221
After undergoing free-space propagation a distance d0b , and passing through the
interference filter, the impulse response function for the idler becomes
Z

d2 rl

hi (r2 , r; i ) = F2 (i )

(ii /2c)2 i ci (db +d0b ) i 2cd


|rrl |2 i i |rl |2 i 2cdi |rl r2 |2
2cf
b
b
e
e
.
e
e
db d0b

(C.10)
Upon Fourier transforming with respect to r, we arrive at the transfer function for
the idler path
(ii /2c)2 i ci (db +d0b )
e
F2 (i )
db d0b
Z
Z
i
2
2

i(q i r )r i i |r|2 +|rl |2 ) i i |rl |2 i 2cd0 (|rl | +|r2 | 2rl r2 )


2
e 2cf
e b
.
d rl d2 r e i cdb l e 2cdb (
Hi (r2 , qi ; i ) =

(C.11)

Carrying out the r and rl integrals in Eq. (C.11) gives




1
Hi (r2 , q; i ) = 0
db

1
1

0
db f

1

i 2cdi0 |r2 |2 [1 d10 ( d10 f1 )1 ]

F2 (i ) ei c (db +db ) e

(C.12)

c
i 2
|qi |2 [db +( d10 f1 )1 ] iqi r2 d10 ( d10 f1 )1

The two-photon detection amplitude, which is proportional to the joint-probability


amplitude for both detectors D1 and D2 to click, is obtained by combining Eqs. (C.5),
(C.7), and (C.12)
(2)

Z
(r1 , t1 , r2 , t2 ) = 2

d qs

ds

eis t1 eii t2 F1 (s ) F2 (i )
icda |qs |2 /(2s )

t (r1 ) e

1
d0b

d qi

di A (qs , qi ; s , i )
1
i
s
1
1
0

ei c da ei c (db +db )
0
db f

eiqs r1 e

c
i 2
|qi |2 [db +( d10 f1 )1 ] iqi r2 d10 ( d10 f1 )1

(C.13)

i 2cdi0 |r2 |2 [1 d10 ( d10 f1 )1 ]

We next assume that the interference filters are very narrow-band This allows them to
be approximated by -functions centered on the degenerate down conversion frequency,

222
which we denote as . This leads to the following two-photon detection amplitude

(2)

i(T1 +T2 )

(r1 , t1 , r2 , t2 ) = 2e

t (r1 ) eicda |qs |

1
d0b
2

1
1

0
db f

1 Z

d qs

d2 qi A (qs , qi ; , )

1
1
1 1
2
/(2) iqs r1 i 2cd0b |r2 | [1 d0b ( d0b f ) ]

c
i 2
|qi |2 [db +( d10 f1 )1 ] iqi r2 d10 ( d10 f1 )1

.
(C.14)

Here Tj = tj zj /c is the retarded time at detector Dj , (j = 1, 2), with zs = da and


zi = db + d0b . The two-photon amplitude for a paraxial, non-monochromatic pump
can be written as (see Appendix B)

A (q+ , q ; + , ) =
p (+ ) up (q+ )


Np L
Dps L
+
q+x .
sinc
2
2

(C.15)

Here
p (+ ) is the pulsed-pump spectral amplitude, and up (q+ ) is the pump transverse
wave-vector amplitude. The sinc-function, sinc (x) = sin (x) /x, arises from the phasematching conditions, and L is the length of the nonlinear crystal. The parameters
Dps and Np are related to the crystal dispersion properties (see Appendix B). We
have made a change of variables

= s i ,

q+ = qi + qs ,

q = (qi qs ) /2,

(C.16)

where j = j j , is the frequency perturbation away from the optimal down


conversion frequency j , (j = s, i). By assuming a plane wave pump, that is, up (q) = (2) (q) ,

223
we get the following two-photon detection amplitude

1 Z
Z
1
1
(2)
i(T1 +T2 ) 1
2
(r1 , t1 , r2 , t2 ) = 2
p (0) e

d qs d2 qi (2) (qs + qi )
0
0
db db f
t (r1 ) eicda |qs |

1
1
1 1
2
/(2) iqs r1 i 2cd0b |r2 | [1 d0b ( d0b f ) ]

c
i 2
|qi |2 [db +( d10 f1 )1 ] iqi r2 d10 ( d10 f1 )1

.
(C.17)

Changing to the q variables gives


(2)

i(T1 +T2 )

(r1 , t1 , r2 , t2 ) = 2
p (0) e

1 1
1
( 0 )1
0
db db f

d q+

d2 q (2) (q+ )

2
i 2cd
0 |r2 | (1n)

t(r1 )eiq+ (r1 +nr2 )/2 eiq (r1 nr2 ) e


c
i 2
(|q+ |2 /2+|q |2 )[da +db +( d10 f1 )1 ] (i

cq+ q
)[da db ( d10 f1 )1 ]
2
b

,
(C.18)

where n = (1/d0b 1/f )

d0b . The integral over q+ collapses the -function, and

gives the two-photon detection amplitude


i

|r2 |2 (1n)

t (r1 )
(2) (r1 , t1 , r2 , t2 ) = 2
p (0) ei(T1 +T2 ) ne 2cdb
Z
i c |q |2 [da +db +( d10 f1 )1 ]
b
d2 q eiq (r1 nr2 ) e 2
.

(C.19)

Note that if
1

da + db + (1/d0b 1/f )

= 0,

(C.20)

then the q -integral leads to a -function (2) (r1 nr2 ). Equation (C.20) is the
ghost-imaging condition, which can be recast into a Gaussian thin-lens-like equation
1
1
1
+ = ,
so si
f

(C.21)

where the object and image distances are so = da + db and si = d0b respectively. When
the ghost-imaging condition, Eq. (C.20), is met, the two-photon detection amplitude

224
becomes
(2) (r1 , t1 , r2 , t2 ) = 2
p (0) ei(T1 +T2 ) ne

i 2cd
0 |r2 | (1n)
b

t (r1 ) (2) (r1 nr2 ) .

(C.22)

The expected coincidence rate for the experiment sketched in Fig. 47 can now
be calculated. Detector D1 is assumed to collect all the light passing through the
aperture, and thus we must integrated over its active area. The active area of the
bucket detector D1 is assumed much larger than the aperture so that the integration
limits may be taken to infinity. Detector D2 is assumed to be a point-like detector
located at position r2 in the imaging plane. We do not integrate over r2 because
of the point-like nature of detector D2 . The coincidence count rate between the two
detectors RCC , is proportional to the integrated square modulus of the two-photon
detection amplitude (over detector D1 s active area and the detector response times)
Z
RCC (r2 )

Z
dt1

Z
dt2


2
d2 r1 (2) (r1 , t1 , r2 , t2 ) .

(C.23)

When Eqs. (C.22) and (C.23) are combined, they lead to the coincidence rate
2

Z
Z
Z

i(T +T ) i 2cd 0 |r2 |2 (1n)
(2)
2
1
2
t (r1 ) (r1 nr2 )
RCC (r2 ) dt1 dt2 d r1 2e
ne b
4 2 |n
p (0)|2 t (nr2 ) .
(C.24)

2
Here we have used the fact that t (r1 ) (2) (r1 nr2 ) = t (r1 ) (2) (r1 nr2 ) for a
transmission aperture. Thus, we see that coincidence count rate as a function of the
point-like detector position r2 , is proportional to the aperture transmission function
t (nr2 ). Here the constant n plays the role of a magnification factor. This is the ghostimaging effect first suggested by Klyshko [170, 171], and experimentally demonstrated

225
by Pittman et. al. [26]. Note that the resolution of the image depends on the overall
active area of the scanning detector. This is given by the point-like detectors active
area. Diffraction effects such as the finite size of the lenses, and the fact that the
pump beam is not a perfect plane wave, also contribute to the image resolution.

226

APPENDIX D

FOURIER OPTICS
The principle idea behind Fourier optics is to utilize the response of linear optical
systems to plane wave inputs for the calculation of generalized outputs. Most often
one is interested in the paraxial approximation in which plane waves traveling along
chosen axis are considered. Choosing the z axis as the propagation direction, with
wave vector amplitude k, we consider small transverse deviations kx , ky  k. In
this approximation, one can characterize the response of a linear optical system in
momentum space by its transfer function H (kx , ky ), which is the system response to
input plane waves exp [i (kx x + ky y)]. A linear optical system can also be characterized
in paraxial approximation by its impulse-response function h (x, y), the spatial domain.
The impulse response function is the system response to an impulse, or point source
(2) (r), at the input plane. Here r = (x, y) is a transverse position coordinate. Given
the input field f (x, y), at the plane z = 0, one may determine the output field g (x, y),
at the plane z = d, of a linear optical system if either the transfer, or impulse response
function is known. This is shown schematically in Fig. 49 below. In particular, the
output field is given by the convolution of the input field with the impulse response
function
Z
g (x, y) =

dx0 dy 0 f (x0 , y 0 ) h (x x0 , y y 0 ).

(D.1)

227
From the convolution theorem, the Fourier transform of the output field G (kx , ky ), is
G (kx , ky ) = H (kx , ky ) F (kx , ky ) ,

(D.2)

where F (kx , ky ) is the Fourier transform of the input field at the plane z = 0, and
H (kx , ky ) is the transfer function of the linear system. The relationships in Eqs.
(D.1) and (D.2) imply that the impulse response function and transfer function are
Fourier transform pairs, which we show explicitly below. Note that we define the
Fourier transforms as
Z
G (kx , ky ) =

1
g (x, y) =
(2)2

ei(kx x+ky y) g (x, y) dx dy,


Z

(D.3)
ei(kx x+ky y) G (kx , ky ) dkx dky .

FIGURE 48. Linear optical system in the paraxial approximation. Diagram of a


linear optical system in which a quasi-monochromatic, paraxial scalar field f (r0 ) in
the z = 0 transverse plane is propagated through a linear optical system characterized
by its impulse response function h (r, r0 , ), to give the output g (r) at the z = d
transverse plane.

228
Thus the problem of calculating the output field for given the input field is reduced
to determination of the transfer function, or impulse response function for the linear
optical system, and performing one of the above integrals.
We shall explicitly examine two important cases of linear optical systems, freespace propagation and a thin lens. In the case of free-space propagation, the transfer
function is most easily found by considering a plane wave input with transverse wave
numbers k0x , k0y  k, given by
f (x, y) = ei(k0x x+k0y y) .

(D.4)

The corresponding transverse wave number spectrum F (kx , ky ) is given by the Fourier
transform of Eq. (D.4)
F (kx , ky ) = (2)2 (kx k0x ) (ky k0y ) .

(D.5)

In the paraxial approximation, where k0x , k0y  k, one may expand the plane wave
amplitude traveling along the z-direction
h

exp i k

kx2

1/2
ky2

 
 
kx2 + ky2
z .
z exp ik 1
2k 2
i

(D.6)

The transfer function of free space is then approximated by


ikz

H (kx , ky ; z) e



kx2 + ky2
exp i
z .
2k

(D.7)

Thus, the output transverse wave number spectrum is


ikz

G (kx , ky ) = e



kx2 + ky2
exp i
z (2)2 (kx k0x ) (ky k0y ) .
2k

(D.8)

229
The output field is given by the Fourier transform of Eq. (D.8),
Z
1
g (x, y) =
ei(kx x+ky y) G (kx , ky ) dkx dky
2
(2)



Z
kx2 + ky2
ikz
i(kx x+ky y)
=e
e
exp i
z (kx k0x ) (ky k0y ) dkx dky
2k


2 
2
+ k0y
k0x
ikz
= e exp i
z ei(k0x x+k0y y) .
2k

(D.9)

So we see that the output field is equal to the input field multiplied by the transfer
function evaluated at the appropriate wave numbers. This is a special case, and is
not a general result. As we shall see, when considering inputs other than plane waves,
Eq. (D.9) leads to different results.
The impulse response function of free space may be determined by the inverse
Fourier transform of the free-space transfer function. Combining Eqs. (D.1), (D.2),
and (D.3) gives
Z Z
g (x, y) =

dx0 dy 0 f (x0 , y 0 ) h (x x0 , y y 0 )

1
(2)2

Z Z

dkx dky ei(kx x+ky y) G (kx , ky )

(D.10)

Z
1
=
dkx dky ei(kx x+ky y) H (kx , ky ) F (kx , ky )
(2)2

Z Z
Z Z
dkx dky i(kx x+ky y)
0
0
0
0
0 0
=
dx dy f (x , y )
H (kx , ky ) ei(kx x +ky y ) .
2 e
(2)

We can see that by equating the final terms on the left-hand-side of the top line
and the bottom line of Eq. (D.10), the impulse response function of a general linear
system h(x, y) is related to the transfer function of that system by the following

230
inverse Fourier transform relationship
Z

h (x x0 , y y 0 ) =

dkx dky i[kx (xx0 )+ky (yy0 )]


e
H (kx , ky )
(2)2

(D.11)

The free-space transfer function Eq. (D.7), leads to the free-space impulse response
function
h (x x0 , y y 0 ) =

ikz

"

0 2

0 2

e
(x x ) + (y y )
exp ik
iz
2z

#
,

(D.12)

where = 2/k is the wavelength of the light.


The case of a lens is quite simple, and follows from the excess phase accumulated
at each point of the wave front due to the lens thickness variation. This leads to a
quadratic-phase-factor impulse response function

ik (x02 + y 02 )
h (x x , y y ) = exp
(x x0 ) (y y 0 ) .
2f
0

(D.13)

Thus the transfer function for a lens is



ik (x2 + y 2 )
H (kx , ky ; f ) = exp
,
2f


(D.14)

which is independent of the transverse wave number. Here we have ignored the actual
thickness of the lens (the thin-lens approximation).

231

BIBLIOGRAPHY
[1] J. S. Bell, Phys. 1, 195 (1964), reproduced in: Speakable and Unspeakable
in Quantum Mechanics, 2nd edition, J. S. Bell, (Cambridge University Press,
Cambridge, 2004) pp. 14-21.
[2] A. Aspect, P. Grangier, and G. Roger, Phys. Rev. Lett. 47, 460 (1981).
[3] A. Aspect, J. Dalibard, and G. Roger, Phys. Rev. Lett. 49, 1804 (1982).
[4] M. Beck, D. T. Smithey, and M. G. Raymer, Phys. Rev. A 48, R890 (1993).
[5] O. Alter and Y. Yamamoto, Phys. Rev. Lett. 74, 4106 (1995).
[6] G. M. DAriano and H. P. Yuen, Phys. Rev. Lett. 76, 2832 (1996).
[7] W. K. Wootters and W. H. Zurek, Nature 299, 802 (1982).
[8] W. Heisenberg, Zeit. fur Phys. 43, 172 (1927), English translation in: Quantum
Theory and Measurement, edited by J. A. Wheeler and W. H. Zurek, (Princeton
University Press, Princeton, 1983), pp.62-84.
[9] W. Pauli, in Handbuch der Physik (Springer, Berlin, 1933), vol. 24, English
translation: General Principles of Quantum Mechanics (Springer, Berlin, 1980).
[10] U. Fano, Rev. Mod. Phys. 29, 74 (1957).
[11] K. Vogel and H. Risken, Phys. Rev. A 40, 2847 (1989).

232
[12] E. Wigner, Phys. Rev. 40, 749 (1932).
[13] U. Leonhardt, Measuring the Quantum State of Light (Cambridge University
Press, Cambridge, 1997).
[14] M. G. A. Paris and J. Rehacek, Quantum State Estimation (Springer-Verlag,
New York, 2004).
[15] M. J. Bastiaans, J. Opt. Soc. Am. A 3, 1227 (1986).
[16] D. Huang, E. A. Swanson, C. P. Lin, J. S. Schuman, W. Stinson, W. Chang,
M. R. Hee, T. Flotte, K. Gregory, C. A. Puliafito, et al., Science 254, 1178
(1991).
[17] M. Francon and S. Mallick, in Progress in Optics VI, edited by E. Wolf (NorthHolland, Amsterdam, 1967), pp. 73104.
[18] D. F. McAlister, Ph.D. thesis, University of Oregon, Eugene, Oregon (1999).
[19] C.-C. Cheng, Ph.D. thesis, University of Oregon, Eugene, Oregon (1999).
[20] D. F. McAlister, M. Beck, L. Clarke, A. Mayer, and M. G. Raymer, Opt. Lett.
20, 1181 (1995).
[21] A. Wax and J. E. Thomas, Opt. Lett. 21, 1427 (1996).
[22] C. Iaconis and I. A. Walmsley, Opt. Lett. 21, 1783 (1996).
[23] C.-C. Cheng, M. G. Raymer, and H. Heier, J. Mod. Opt. 47, 1237 (2000).

233
[24] A. Royer, Phys. Rev. A 15, 449 (1977).
[25] R. Loudon, The Quantum Theory of Light, 3rd edition (Oxford University Press,
New York, 2000).
[26] T. B. Pittman, Y. H. Shih, D. V. Strekalov, and A. V. Sergienko, Phys. Rev. A
52, R3429 (1995).
[27] D. V. Strekalov, A. V. Sergienko, D. N. Klyshko, and Y. H. Shih, Phys. Rev.
Lett. 74, 3600 (1995).
[28] M. DAngelo, A. Valencia, M. H. Rubin, and Y. Shih, Phys. Rev. A 72, 013810
(2005).
[29] P. H. S. Ribeiro, S. Padua, J. C. Machado da Silva, and G. A. Barbosa, Phys.
Rev. A 49, 4176 (1994).
[30] G. Jaeger and A. V. Sergienko, in Progress in Optics 42, edited by E. Wolf
(Elsevier, Amsterdam, 2001), pp. 277324.
[31] I. A. Walmsley and M. G. Raymer, Science 307, 1733 (2005).
[32] H. Weinfurter, J. Phys. B 38, 579 (2005).
[33] K. Banaszek and K. Wodkiewicz, Phys. Rev. A 58, 4345 (1998).
[34] A. Einstein, Ann. Phys. 17, 132 (1905), reproduced in : The Old Quantum
Theory, D. ter Haar, (Pergamon Press, New York, 1967) pp. 91-107.

234
[35] P. Lenard, Ann. der Phys. 313, 149 (1902).
[36] A. H. Compton, Phys. Rev. 21, 483 (1923).
[37] P. A. M. Dirac, Proc. R. Soc. London, Series A 114, 243 (1927).
[38] E. Fermi, Rev. Mod. Phys. 4, 87 (1932).
[39] G. N. Lewis, Nature 118, 784 (1926).
[40] W. E. Lamb, App. Phys. B 60, 77 (1995).
[41] T. D. Newton and E. P. Wigner, Rev. Mod. Phys. 21, 400 (1949).
[42] A. S. Wightman, Rev. Mod. Phys. 34, 845 (1962).
[43] J. M. Jauch and C. Piron, Helv. Phys. Acta 40, 559 (1967).
[44] W. O. Amrein, Helv. Phys. Acta 42, 149 (1969).
[45] I. H. Deutsch and J. C. Garrison, Phys. Rev. A 43, 2498 (1991).
[46] J. Mourad, Phys. Lett. A 182, 319 (1993).
[47] C. Adlard, E. R. Pike, and S. Sarkar, Phys. Rev. Lett. 79, 1585 (1997).
[48] I. Bialynicki-Birula, Phys. Rev. Lett. 80, 5247 (1998).
[49] R. J. Glauber, Phys. Rev. Lett. 10, 84 (1963).
[50] R. J. Glauber, Phys. Rev. 130, 2529 (1963).

235
[51] L. Mandel, Phys. Rev. 144, 1071 (1966).
[52] M. O. Scully and M. S. Zubairy, Quantum Optics (Cambridge University Press,
Cambridge, 1997).
[53] A. Muthukrishnan, M. O. Scully, and M. S. Zubairy, Proc. SPIE 5866, 287
(2005).
[54] M. H. L. Pryce, Proc. Roy. Soc. London A195, 62 (1948).
[55] M. Hawton, Phys. Rev. A 59, 954 (1999).
[56] M. Hawton and W. E. Baylis, Phys. Rev. A 64, 012101 (2001).
[57] M. Hawton (2006), quant-ph/0609086.
[58] L. D. Landau and R. J. Peierls, Z. Phys. 62, 188 (1930).
[59] I. Bialynicki-Birula, Acta Phys. Pol. A 86, 97 (1994).
[60] J. E. Sipe, Phys. Rev. A 52, 1875 (1995).
[61] M. Hawton, Phys. Rev. A 59, 3223 (1999).
[62] K. W. Chan, C. K. Law, and J. H. Eberly, Phys. Rev. Lett. 88, 100402 (2002).
[63] B. Ritchie, Opt. Comm. 262, 229 (2006).
[64] R. J. Cook, Phys. Rev. A 25, 2164 (1982).
[65] R. J. Cook, Phys. Rev. A 26, 2754 (1982).

236
[66] T. Inagaki, Phys. Rev. A 49, 2839 (1994).
[67] T. Inagaki, Phys. Rev. A 57, 2204 (1998).
[68] I. Bialynicki-Birula, in Progress in Optics XXXVI, edited by E. Wolf (Elsevier,
Amsterdam, 1996), pp. 245294.
[69] G. Moliere, Ann. der Phys. 441, 146 (1949).
[70] R. H. Good, Phys. Rev. 105, 1914 (1957).
[71] C. L. Hammer and R. H. Good, Phys. Rev. 108, 882 (1957).
[72] H. E. Moses, Phys. Rev. 113, 1670 (1959).
[73] I. Bialynicki-Birula, in Coherence and Quantum Optics VII, edited by J. H.
Eberly, L. Mandel, and E. Wolf (Plenum, New York, 1996), p. 313.
[74] O. Keller, Phys. Rev. A 60, 1652 (1999).
[75] P. D. Drummond, Phys. Rev. A 60, R3331 (1999).
[76] D. H. Kobe, Found. Phys. 29, 1203 (1999).
[77] S. Esposito, Found. Phys. Lett. 12, 533 (1999).
[78] O. Keller, Phys. Rev. A 62, 022111 (2000).
[79] A. Gersten, Found. Phys. 31, 1211 (2001).
[80] J. Zalesny, Int. J. Theor. Phys. 43, 2093 (2004).

237
[81] O. Keller, Phys. Rep. 411, 1 (2005).

[82] P. D. Drummond, J. Phys. B 39, S573 (2006).

[83] B. J. Smith and M. G. Raymer, Phys. Rev. A 74, 062104 (2006).

[84] G. G. Lapaire and J. E. Sipe (2006), quant-ph/0607008.

[85] J. R. Oppenheimer, Phys. Rev. 38, 725 (1931).

[86] R. P. Feynman (1965), The Development of the Space-Time View of Quantum


Electrodynamics, http://nobelprize.org/nobel prizes/physics/
laureates/1965/feynman-lecture.html.

[87] L. Mandel and E. Wolf, Optical Coherence and Quantum Optics (Cambridge
University Press, Cambridge, 1995).

[88] U. M. Titulaer and R. J. Glauber, Phys. Rev. 145, 1041 (1966).

[89] D. N. Klyshko, Photons and Nonlinear Optics (Gordon and Breach, New York,
1988).

[90] C. Cohen-Tannoudji, J. Dupont-Roc, and G. Grynberg, Photons and Atoms Introduction to Quantum Electrodynamics (Wiley, New York, 1997).

[91] W. Greiner, Relativistic Quantum Mechanics: Wave Equations, 3rd edition


(Springer-Verlag, New York, 2000).

238
[92] M. E. Peskin and D. V. Schroeder, An Introduction to Quantum Field Theory
(Westview Press, Boulder, 1995).
[93] L. H. Ryder, Quantum Field Theory, 2nd edition (Cambridge University Press,
Cambridge, 1996).
[94] R. Shankar, Principles of Quantum Mechanics, 2nd edition (Cambridge
University Press, Cambridge, 1994).
[95] S. Weinberg and E. Witten, Phys. Lett. B 96, 59 (1980).
[96] E. Merzbacher, Quantum Mechanics, 3rd edition (Wiley, New York, 1997).
[97] G. A. Smith, S. Chaudhury, A. Silberfarb, I. H. Deutsch, and P. S. Jessen, Phys.
Rev. Lett. 93, 163602 (2004).
[98] A. Silberfarb, P. S. Jessen, and I. H. Deutsch, Phys. Rev. Lett. 95, 030402
(2005).
[99] G. A. Smith, A. Silberfarb, I. H. Deutsch, and P. S. Jessen, Phys. Rev. Lett.
97, 180403 (2006).
[100] M. H. Rubin, Phys. Rev. A 54, 5349 (1996).
[101] C. H. Monken, P. H. S. Ribeiro, and S. Padua, Phys. Rev. A 57, 3123 (1998).
[102] S. P. Walborn, A. N. de Oliveira, S. Padua, and C. H. Monken, Phys. Rev.
Lett. 90, 143601 (2003).

239
[103] J. J. Sakurai, Modern Quantum Mechanics, Revised edition (Addison Wesley,
Reading, MA, edited by S. F. Tuan, 1994).
[104] B. E. A. Saleh, M. C. Teich, and A. V. Sergienko, Phys. Rev. Lett. 94, 223601
(2005).
[105] B. M. Escher, A. T. Avelar, T. M. da Rocha Filho, and B. Baseia, Phys. Rev.
A 70, 025801 (2004).
[106] K. J. Resch, Phys. Rev. A 70, 051803 (2004).
[107] K. Sanaka, Phys. Rev. A 71, 021801 (2005).
[108] K. Sanaka, K. J. Resch, and A. Zeilinger, Phys. Rev. Lett. 96, 083601 (2006).
[109] E. Knill, R. Laflamme, and G. J. Milburn, Nature 409, 46 (2001).
[110] H. Fearn and R. Loudon, Opt. Comm. 64, 485 (1987).
[111] R. Loudon, Phys. Rev. A 58, 4904 (1998).
[112] S. M. Barnett, J. Jeffers, A. Gatti, and R. Loudon, Phys. Rev. A 57, 2134
(1998).
[113] C. K. Hong, Z. Y. Ou, and L. Mandel, Phys. Rev. Lett. 59, 2044 (1987).
[114] R. J. Glauber, Phys. Rev. 131, 2766 (1963).
[115] E. C. G. Sudarshan, Phys. Rev. Lett. 10, 277 (1963).

240
[116] L. Mandel, E. C. G. Sudarshan, and E. Wolf, Proc. Phys. Soc. London 84, 435
(1964).

[117] E. Wolf and C. L. Mehta, Phys. Rev. Lett. 13, 705 (1964).

[118] L. Mandel and E. Wolf, Rev. Mod. Phys. 37, 231 (1965).

[119] A. G. White, D. F. V. James, P. H. Eberhard, and P. G. Kwiat, Phys. Rev.


Lett. 83, 3103 (1999).

[120] C. Iaconis and I. A. Walmsley, Opt. Lett. 23, 792 (1998).

[121] D. Kane and R. Trebino, IEEE J. Quantum Electron. 29, 571 (1993).

[122] A. I. Lvovsky, H. Hansen, T. Aichele, O. Benson, J. Mlynek, and S. Schiller,


Phys. Rev. Lett. 87, 050402 (2001).

[123] E. Mukamel, K. Banaszek, I. A. Walmsley, and C. Dorrer, Opt. Lett. 28, 1317
(2003).

[124] I. Kimel and L. R. Elias, IEEE J. Quantum Electron. 29, 2562 (1993).

[125] A. E. Siegman, Lasers (University Science Books, Sausalito, CA, 1986).

[126] J.-C. Diels and W. Rudolph, Ultrashort Laser Pulse Phenomena: Fundamentals,
Techniques, and Applications on a Femtosecond Time Scale (Academic Press,
San Diego, CA, 1995).

241
[127] B. E. A. Saleh and M. C. Teich, Fundamentals of Photonics (Wiley-Interscience,
New York, 1991).
[128] M. A. Bandres and J. Gutirrez-Vega, Opt. Lett. 29, 144 (2004).
[129] M. A. Bandres and J. Gutirrez-Vega, J. Opt. Soc. Am. A 21, 873 (2004).
[130] H. Kogelnik and T. Li, App. Opt. 5, 1550 (1966).
[131] M. Hillery, R. F. OConnell, M. O. Scully, and E. P. Wigner, Phys. Rep. 106,
121 (1984).
[132] E. Joos, H. Zeh, C. Kiefer, D. Giulini, J. Kupsch, and I.-O. Stamatescu,
Decoherence and the Appearance of a Classical World in Quantum Theory
(Springer-Verlag, New York, 1997).
[133] L. Ryzhik, G. Papanicolaou, and J. B. Keller, Wave Motion 24, 327 (1996).
[134] D. Dragoman, in Progress in Optics 43, edited by E. Wolf (Elsevier, Amsterdam,
2002), pp. 433496.
[135] B. J. Smith, B. Killett, M. G. Raymer, I. A. Walmsley, and K. Banaszek, Opt.
Lett. 30, 3365 (2005).
[136] E. J. Galvez and C. D. Holmes, J. Opt. Soc. Am. A 16, 1981 (1999).
[137] M. Nazarathy and J. Shamir, J. Opt. Soc. Am. 72, 356 (1981).
[138] P. P. Rohde, T. C. Ralph, and M. A. Nielsen, Phys. Rev. A 72, 052332 (2005).

242
[139] W. P. Grice and I. A. Walmsley, Phys. Rev. A 56, 1627 (1997).
[140] W. P. Grice, A. B. URen, and I. A. Walmsley, Phys. Rev. A 64, 063815 (2001).
[141] I. A. Khan and J. C. Howell, Phys. Rev. A 73, 031801 (2006).
[142] T. B. Pittman, D. V. Strekalov, D. N. Klyshko, M. H. Rubin, A. V. Sergienko,
and Y. H. Shih, Phys. Rev. A 53, 2804 (1996).
[143] R. S. Bennink, Y. Liu, D. D. Earl, and W. P. Grice, Phys. Rev. A 74, 023802
(2006).
[144] W. K. Wootters, Phys. Rev. Lett. 80, 2245 (1998).
[145] L. Allen, M. J. Padgett, and M. Babiker, in Progress in Optics 39, edited by
E. Wolf (Elsevier, Amsterdam, 1999), pp. 291372.
[146] G. F. Calvo, A. Picon, and E. Bagan, Phys. Rev. A 73, 013805 (2006).
[147] J. C. Leader, J. Opt. Soc. Am. 68, 175 (1978).
[148] T. Shirai, A. Dogariu, and E. Wolf, J. Opt. Soc. Am. A 20, 1094 (2003).
[149] S. C. H. Wang and M. A. Plonus, J. Opt. Soc. Am. 69, 1297 (1979).
[150] C. Paterson, Phys. Rev. Lett. 94, 153901 (2005).
[151] M. W. Mitchell, C. W. Ellenor, S. Schneider, and A. M. Steinberg, Phys. Rev.
Lett. 91, 120402 (2003).

243
[152] N. K. Langford, R. B. Dalton, M. D. Harvey, J. L. OBrien, G. J. Pryde,
A. Gilchrist, S. D. Bartlett, and A. G. White, Phys. Rev. Lett. 93, 053601
(2004).
[153] C. A. Kocher and E. D. Commins, Phys. Rev. Lett. 18, 575 (1967).
[154] S. J. Freedman and J. F. Clauser, Phys. Rev. Lett. 28, 938 (1972).
[155] J. F. Clauser and A. Shimony, Rep. Prog. Phys. 41, 1881 (1978).
[156] S. E. Harris, M. K. Oshman, and R. L. Byer, Phys. Rev. Lett. 18, 732 (1967).
[157] D. Magde, R. Scarlet, and H. Mahr, Appl. Phys. Lett. 11, 381 (1967).
[158] D. N. Klyshko, Pisma Zh. Eksp. Teor. Fiz. 6, 490 (1967), [JETP Lett. 6, 23
(1967)].
[159] R. L. Byer and S. E. Harris, Phys. Rev. 168, 1064 (1968).
[160] J. A. Giordmaine and R. C. Miller, Phys. Rev. Lett. 14, 973 (1965).
[161] W. H. Louisell, A. Yariv, and A. E. Siegman, Phys. Rev. 124, 1646 (1961).
[162] C. K. Hong and L. Mandel, Phys. Rev. A 31, 2409 (1985).
[163] Z. Y. Ou, Ph.D. thesis, University of Rochester, Rochester, New York (1990).
[164] R. W. Boyd, Nonlinear Optics, 2nd edition (Academic Press, New York, 2003).
[165] A. Yariv, Quantum Electronics, 2nd edition (Wiley, New York, 1975).

244
[166] P. N. Butcher and D. Cotter, The Elements of Nonlinear Optics (Cambridge
University Press, Cambridge, 1990).
[167] D. N. Nikogosyan, App. Phys. A 52, 359 (1991).
[168] D. A. Roberts, IEEE J. Quantum Electron. 28, 2057 (1992).
[169] D. N. Klyshko, Zh. Eksp. Teor. Fiz. 83, 1313 (1982), [JETP 56, 753 (1982)].
[170] D. N. Klyshko, Zh. Eksp. Teor. Fiz. 94, 82 (1988), [JETP 67, 1131 (1988)].
[171] A. V. Belinkskii and D. N. Klyshko, Zh. Eksp. Teor. Fiz. 105, 487 (1994),
[JETP 78, 259 (1994)].
[172] C. K. Hong and L. Mandel, Phys. Rev. Lett. 56, 58 (1986).
[173] T. Aichele, A. Lvovsky, and S. Schiller, Eur. Phys. J. D 18, 237 (2002).
[174] Z. Y. Ou, Quant. Semiclass. Opt. 9, 599 (1994).
[175] A. Gatti, E. Brambilla, and L. A. Lugiato, Phys. Rev. Lett. 90, 133603 (2003).
[176] G. Bjork, J. Soderholm, and L. L. Sanchez-Soto, J. Opt. B 6, S478 (2004).
[177] A. Gatti, E. Brambilla, M. Bache, and L. A. Lugiato, Phys. Rev. A 70, 013802
(2004).
[178] M. DAngelo, Y.-H. Kim, S. P. Kulik, and Y. Shih, Phys. Rev. Lett. 92, 233601
(2004).

245
[179] R. S. Bennink, S. J. Bentley, and R. W. Boyd, Phys. Rev. Lett. 89, 113601
(2002).
[180] R. S. Bennink, S. J. Bentley, R. W. Boyd, and J. C. Howell, Phys. Rev. Lett.
92, 033601 (2004).
[181] B. I. Erkmen and J. H. Shapiro (2006), quant-ph/0612070.
[182] T. D. T. C. Ralph, G. J. Milburn (2006), quant-ph/0609139.
[183] J. D. Jackson, Classical Electrodynamics, 3rd edition (Wiley, New York, 1999).
[184] S. A. Akhmanov, V. V. Fadeev, R. V. Khokhlov, and O. N. Chunaev, Pisma
Zh. Eksp. Teor. Fiz. 6, 575 (1967), [JETP Lett. 6, 85 (1967)].

Вам также может понравиться