06.04.2014 Views

Dynamics of Electromagnetically Induced Transparency Optical Kerr ...

Dynamics of Electromagnetically Induced Transparency Optical Kerr ...

Dynamics of Electromagnetically Induced Transparency Optical Kerr ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Dynamics</strong> <strong>of</strong> <strong>Electromagnetically</strong><br />

<strong>Induced</strong> <strong>Transparency</strong> <strong>Optical</strong><br />

<strong>Kerr</strong> nonlinearities<br />

by<br />

Michael Vernon Pack<br />

Submitted in Partial Fulfillment<br />

<strong>of</strong> the<br />

Requirements for the Degree<br />

Doctor <strong>of</strong> Philosophy<br />

Supervised by<br />

Pr<strong>of</strong>essor John C. Howell<br />

Department <strong>of</strong> Physics and Astronomy<br />

The College<br />

Arts and Sciences<br />

University <strong>of</strong> Rochester<br />

Rochester, New York<br />

2007


To my wife,<br />

Janene Kaufman Pack,<br />

for all <strong>of</strong> her love and support.<br />

ii


iii<br />

Curriculum Vitae<br />

The author was born in Logan, Utah on the 23rd December, 1974. He graduated<br />

from Holbrook High School in 1993. He attended Brigham Young University<br />

from 1996 untill 2001 and graduated with a BS degree in physics and electrical<br />

engineering and an MS degree in electrical engineering.<br />

He worked at Sandia<br />

National Laboratory in Albuquerque, NM from 2001 to 2003. He came to the<br />

University <strong>of</strong> Rochester in the fall <strong>of</strong> 2003, and received a MA degree in physics in<br />

2004. Under the supervision <strong>of</strong> Dr. John C. Howell, he has performed his doctoral<br />

research in the fields <strong>of</strong> quantum optics and atomic, and optical physics.


CURRICULUM VITAE<br />

iv<br />

Publications<br />

M.V. Pack, P. K. Setu, and J.C. Howell, “Using slow light to improve quantum<br />

non-demolition measurements using electromagnetically-induced-transparencyenhanced<br />

<strong>Kerr</strong> nonlinearities” (submitted to Phys. Rev. A)<br />

M.V. Pack, R.M. Camacho, and J.C. Howell, “Transients <strong>of</strong> the<br />

electromagnetically-induced-transparency-enhanced refractive <strong>Kerr</strong> nonlinearity:<br />

Experiment” Phys. Rev. A (in press)<br />

M.V. Pack, R.M. Camacho, and J.C. Howell, “<strong>Electromagnetically</strong> induced<br />

transparency lineshapes for large probe fields and optically thick media” Phys.<br />

Rev. A (in press)<br />

R. M. Camacho, M. V. Pack, J. C. Howell, A. Schweinsberg, R. W. Boyd,<br />

“Wide-bandwidth, tunable, multiple-pulse-width optical delays using slow light<br />

in cesium vapor” Phys. Rev. Lett. 98, 153601 (2007).<br />

R. M. Camacho, M. V. Pack, and J. C. Howell, “Slow light with large fractional<br />

delays by spectral hole-burning in rubidium vapor” Phys. Rev. A 74, 033801<br />

(2006).<br />

M. V. Pack, R. M. Camacho, and J. C. Howell, “Transients <strong>of</strong> the<br />

electromagnetically-induced-transparency-enhanced refractive <strong>Kerr</strong> nonlinearity:<br />

Theory” Phys. Rev. A 74, 013812 (2006).<br />

R. M. Camacho, M. V. Pack, and J. C. Howell, “Low-distortion slow light<br />

using two absorption resonances” Phys. Rev. A 73, 063812 (2006).<br />

M. V. Pack, D. J. Armstrong, and A. V. Smith, “Measurement <strong>of</strong> the χ (2)<br />

tensor <strong>of</strong> GdCa 4 O(BO 3 ) 3 and YCa 4 O(BO 3 ) 3 crystals” J. Opt. Soc. Am. B 22,<br />

417 (2005).<br />

C. D. Nordquist, A. Muyshondt, M. V. Pack, P. S. Finnegan, C. W. Dyck,<br />

I. C. Reines, G. M. Kraus, T. A. Plut, G. R. Sloan, C. L. Goldsmith, and C.<br />

T. Sullivan, “An X-band and Ku-band RF MEMS switched coplanar strip filter”<br />

IEEE Microw. Wireless Compon. Lett, 14, 425 (2004).<br />

M. V. Pack, D. J. Armstrong, A. V. Smith, “Second harmonic generation with<br />

focused beams in a pair <strong>of</strong> walk<strong>of</strong>f-compensating crystals” Opt. Comm. 221, 211<br />

(2003).<br />

M. V. Pack, D. J. Armstrong, and A. V. Smith, “Measurement <strong>of</strong> the χ (2)


CURRICULUM VITAE<br />

v<br />

tensor <strong>of</strong> KTiOPO 4 , KTiOAsO 4 , KTiOPO 4 and KTiOAsO 4 crystals” Appl. Opt.<br />

43, 3319 (2004).<br />

M. V. Pack, D. J. Armstrong, and A. V. Smith, “Measurement <strong>of</strong> the χ (2)<br />

tensor <strong>of</strong> the potassium niobate crystal” J. Opt. Soc. Am. B 20, 2109 (2003).<br />

D. J. Armstrong, M. V. Pack, and A. V. Smith, “Instrument and method<br />

for measuring second-order nonlinear optical tensors” Rev. Sci. Inst. 74, 3250<br />

(2003).<br />

Michael V. Pack, Darrell J. Armstrong, Arlee V. Smith, Michael E.<br />

Amiet, “Second harmonic generation with focused beams in a pair <strong>of</strong> walk<strong>of</strong>fcompensating<br />

crystals” Opt. Comm. 221, 211 (2003).<br />

J. Peatross and M. V. Pack , “Visual introduction to Gaussian beams using a<br />

single lens as an interferometer” Am. J. Phys. 69, 1169 (2001).


CURRICULUM VITAE<br />

vi<br />

Presentations<br />

M. V. Pack, R. M Camacho, and J.C. Howell, “Large fractional delays in hot<br />

atomic vapors” SPIE Photonics West, San Jose, CA **Invited** (January 2007)<br />

M. V. Pack, R. M Camacho, and J.C. Howell, “Low-light-level all optical<br />

switching” Frontier in Optics, Rochester, NY (Oct 2006)<br />

M. V. Pack, R. M Camacho, and J.C. Howell, “<strong>Dynamics</strong> <strong>of</strong> the EIT <strong>Kerr</strong><br />

nonlinearity” Slow light OSA, Wasington, DC (June 2006)<br />

M. V. Pack, R. M Camacho, and J.C. Howell, “Low-light-level all optical<br />

switching” CLEO/QELS, Baltimore, MD (May 2005)<br />

M. V. Pack, R. M Camacho, and J.C. Howell, “Low-light-level all optical<br />

switching” SPIE Photonics West, San Jose, CA **Invited** (January 2005)<br />

M. V. Pack and D. Armstrong and A. V. Smith, “Measurements <strong>of</strong> the second<br />

order nonlinear tensors for KTiOPO 4 , KTiOAsO 4 , KTiOPO 4 , KTiOAsO 4 , KDP,<br />

and KNBO 4 ” OSA Annual Meeting, Tucsun, AZ (Oct. 2003)<br />

M. V. Pack and D. Armstrong and A. V. Smith, “The D-factory: Measuring<br />

the second order nonlinear tensor <strong>of</strong> crystals” CLEO/QELS, Baltimore, MD (May<br />

2003)


vii<br />

Acknowledgements<br />

There are many people who have made my time at the University <strong>of</strong> Rochester<br />

both enjoyable and enlightening, and I would like to thank them.<br />

First, I am thankful to my advisor John Howell for the things I have learned<br />

from him and the pleasant working environment he creates.<br />

I lucked into an<br />

advisor who’s philosophy is “happy cows give good milk”, and who’s genuine<br />

enthusiasm for physics and life is not exceeded by anyone I have met. This enthusiasm<br />

is witnessed by John’s inability to stay out <strong>of</strong> the lab in Rochester while he<br />

was on sabbatical in Italy. While the technical skills I have acquired at Rochester<br />

are important, the other aspects <strong>of</strong> research such as teamwork, public relations,<br />

and how to search out and formulate new research problems are the real gems I<br />

am taking with me. In my experience, John is singularly adept at teaching these<br />

skills, and I am thankful to have learned them from him.<br />

I have worked with a number <strong>of</strong> graduate students while at Rochester, and<br />

have benefited by our exchange <strong>of</strong> ideas and tricks <strong>of</strong> the trade. I have worked<br />

predominantly with Ryan Camacho with Ryan emphasizing the slow-light work,<br />

and me emphasizing the EIT <strong>Kerr</strong> nonlinearity work. In our frequent discussions


ACKNOWLEDGEMENTS<br />

viii<br />

<strong>of</strong> physics, Ryan has helped me sharpen my understanding and arguments by not<br />

ceding any argument until my ideas were fully developed. Ryan’s insightful questions<br />

have <strong>of</strong>ten launched our research in new directions. I am also thankful for<br />

similar interactions with other members <strong>of</strong> our research group including Praveen<br />

Setu, Irfan Ali-Khan, Curtis Broadbent, and Ben Dixon.<br />

I have also enjoyed collaborations with Kevin Wright and Nick Bigelow, as well<br />

as Aarron Schweinsberg and Bob Boyd. Each research group has had something<br />

unique to <strong>of</strong>fer, and I am glad to have been able to work with three <strong>of</strong> the best.<br />

Finally, I would never have been in a position to start much less finish this<br />

thesis except for the support <strong>of</strong> my parents and my wife. My parents provided<br />

me with the opportunities which made this possible, and Janene, my wife, has<br />

helped me realized my goals with her encouragement and support. Janene has<br />

done the little things that have made many late nights at the lab and writing<br />

papers possible, and she has also done the big things such as pro<strong>of</strong> reading this<br />

thesis.<br />

Thank you all.


ix<br />

Abstract<br />

This thesis presents theoretical derivations and experimental demonstrations<br />

<strong>of</strong> <strong>Kerr</strong> optical nonlinearities and their dynamics. The <strong>Kerr</strong> nonlinearity studied is<br />

based on electromagnetically induced transparency (EIT), and is known for having<br />

an unusually large cross-phase modulation (XPM) [1]. Also, quantum nondemolition<br />

(QND) measurements <strong>of</strong> photon number using the EIT <strong>Kerr</strong> nonlinearity are<br />

discussed.<br />

We first discuss the EIT line shapes in hot atomic vapors both with and without<br />

buffer gasses, and for the specific case <strong>of</strong> EIT line shapes on the D 1 line <strong>of</strong><br />

rubidium. Using a Bloch-vector formalism to describe EIT, we derive new expressions<br />

describing the dynamical evolution <strong>of</strong> the EIT <strong>Kerr</strong> nonlinearity and the<br />

rise-time <strong>of</strong> the EIT <strong>Kerr</strong> effect.<br />

The theory <strong>of</strong> the EIT <strong>Kerr</strong> dynamics is confirmed by experiments using hot<br />

rubidium vapor in a buffer gas. These experimental observations confirm that the<br />

rise-time <strong>of</strong> the EIT <strong>Kerr</strong> effect is linearly proportional to the optical thickness <strong>of</strong><br />

the EIT medium and inversely proportional to the width <strong>of</strong> the EIT transparency<br />

resonance.


ABSTRACT<br />

x<br />

We find that the EIT <strong>Kerr</strong> effect is relatively slow, and that its rise-time is<br />

inversely proportional to the EIT <strong>Kerr</strong> susceptibility. The implications <strong>of</strong> the slow<br />

rise-time to QND measurements is investigated, and it is found that by matching<br />

the group velocities <strong>of</strong> the probe and signal optical fields, the signal to noise ratio<br />

for QND measurements is improved. Predictions for the EIT <strong>Kerr</strong> dynamics with<br />

slow group velocity for the signal field were confirmed experimentally in rubidium<br />

vapor with group velocities <strong>of</strong> about 10 4 m/s for both probe and signal optical<br />

fields.


xi<br />

Table <strong>of</strong> Contents<br />

1 Introduction 1<br />

1.1 Introduction to <strong>Electromagnetically</strong> <strong>Induced</strong> <strong>Transparency</strong> . . . . 3<br />

1.2 EIT enhanced nonlinear optics . . . . . . . . . . . . . . . . . . . . 6<br />

1.3 EIT <strong>Kerr</strong> nonlinearity . . . . . . . . . . . . . . . . . . . . . . . . 7<br />

1.4 Other research related to EIT <strong>Kerr</strong> nonlinearities . . . . . . . . . 10<br />

1.5 Applications <strong>of</strong> the EIT <strong>Kerr</strong> nonlinearity . . . . . . . . . . . . . 16<br />

1.5.1 Number resolving photon counters . . . . . . . . . . . . . 17<br />

1.5.2 Quantum state preparation . . . . . . . . . . . . . . . . . 17<br />

1.5.3 Quantum logic gates . . . . . . . . . . . . . . . . . . . . . 18<br />

1.6 Outline <strong>of</strong> dissertation . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

2 Theory <strong>of</strong> <strong>Electromagnetically</strong> induced transparency 23<br />

2.1 Back <strong>of</strong> the envelope theory <strong>of</strong> EIT . . . . . . . . . . . . . . . . . 24<br />

2.2 Perturbative solution for EIT . . . . . . . . . . . . . . . . . . . . 28<br />

2.3 Bloch-vector solution for EIT . . . . . . . . . . . . . . . . . . . . 39


CONTENTS<br />

xii<br />

2.4 Transient Solution for EIT . . . . . . . . . . . . . . . . . . . . . . 43<br />

2.5 Doppler broadening . . . . . . . . . . . . . . . . . . . . . . . . . . 46<br />

2.6 Coherence preserving collisions in<br />

buffer gasses and wall coatings . . . . . . . . . . . . . . . . . . . . 52<br />

2.7 Magnetic sublevels and hyper-fine levels . . . . . . . . . . . . . . 56<br />

2.8 Miscellaneous experimental considerations . . . . . . . . . . . . . 62<br />

3 Theory <strong>of</strong> the dynamics <strong>of</strong> the EIT <strong>Kerr</strong> nonlinearity 67<br />

3.1 The EIT <strong>Kerr</strong> nonlinearity . . . . . . . . . . . . . . . . . . . . . . 69<br />

3.1.1 EIT <strong>Kerr</strong> dynamics . . . . . . . . . . . . . . . . . . . . . . 72<br />

3.1.2 Rise-times in optically thin media . . . . . . . . . . . . . . 74<br />

3.1.3 Rise times in optically thick media . . . . . . . . . . . . . 76<br />

3.2 QND measurements & <strong>Kerr</strong> nonlinearities . . . . . . . . . . . . . 80<br />

3.2.1 QND measurement and SNR . . . . . . . . . . . . . . . . . 81<br />

3.2.2 SNR for EIT <strong>Kerr</strong> Measurements . . . . . . . . . . . . . . 88<br />

4 EIT <strong>Kerr</strong> dynamics Experiment 93<br />

4.1 Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93<br />

4.2 CW measurements . . . . . . . . . . . . . . . . . . . . . . . . . . 99<br />

4.3 Transients Measurements . . . . . . . . . . . . . . . . . . . . . . . 105<br />

4.4 Discussion <strong>of</strong> Measurements . . . . . . . . . . . . . . . . . . . . . 111<br />

5 EIT <strong>Kerr</strong> nonlinearity with slow signal group velocity 116


CONTENTS<br />

xiii<br />

5.1 Slow Light EIT <strong>Kerr</strong> nonlinearity . . . . . . . . . . . . . . . . . . 117<br />

5.2 Experimental apparatus . . . . . . . . . . . . . . . . . . . . . . . 126<br />

5.3 Slow signal EIT <strong>Kerr</strong> measurements . . . . . . . . . . . . . . . . . 131<br />

5.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138<br />

6 Conclusions and perspective 140<br />

Bibliography 147<br />

A Appendix 1: EIT on the rubidium D1 line 158<br />

A.1 Choice <strong>of</strong> Polarizations . . . . . . . . . . . . . . . . . . . . . . . . 158<br />

A.2 Transition matrix elements . . . . . . . . . . . . . . . . . . . . . . 160<br />

A.3 EIT in rubidium 85 . . . . . . . . . . . . . . . . . . . . . . . . . . 161<br />

A.4 important parameters for rubidium . . . . . . . . . . . . . . . . . 167<br />

A.5 EIT in rubidium 87 . . . . . . . . . . . . . . . . . . . . . . . . . . 168


xiv<br />

List <strong>of</strong> Tables<br />

Table Title Page<br />

4.1 Experimental parameters for EIT <strong>Kerr</strong> measurements . . . . . . . 98<br />

A.1 Comparison <strong>of</strong> EIT for several choices <strong>of</strong> polarization . . . . . . . 171<br />

A.2 Dipole matrix elements for Rb 85 D 1 line with F = 2 . . . . . . . . 171<br />

A.3 Dipole matrix elements for Rb 85 D 1 line with F = 3 . . . . . . . . 171<br />

A.4 Dipole matrix elements for Rb 87 D 1 line with F = 1 . . . . . . . . 172<br />

A.5 Dipole matrix elements for Rb 87 D 1 line with F = 2 . . . . . . . . 172


xv<br />

List <strong>of</strong> Figures<br />

Figure Title Page<br />

1.1 Level diagram–EIT Λ system . . . . . . . . . . . . . . . . . . . . 3<br />

1.2 Three-level EIT systems–Λ, V , and ladder . . . . . . . . . . . . . 5<br />

1.3 Various EIT <strong>Kerr</strong> systems . . . . . . . . . . . . . . . . . . . . . . 9<br />

1.4 Autler-Townes splitting–dressed state picture . . . . . . . . . . . . 11<br />

1.5 Nonlinear magneto-optical rotation . . . . . . . . . . . . . . . . . 14<br />

2.1 Transformation to bright-dark basis . . . . . . . . . . . . . . . . . 25<br />

2.2 Level diagram–three level atom . . . . . . . . . . . . . . . . . . . 29<br />

2.3 Transformation to dressed state picture . . . . . . . . . . . . . . . 33<br />

2.4 Probe absorption–large single-photon detuning . . . . . . . . . . . 35<br />

2.5 Doppler integration . . . . . . . . . . . . . . . . . . . . . . . . . . 36<br />

2.6 Error bounds–Perturbative calculation . . . . . . . . . . . . . . . 39<br />

2.7 Transformation from GSC to susceptibility . . . . . . . . . . . . . 46<br />

2.8 EIT line shapes for various circumstances . . . . . . . . . . . . . . 51


LIST OF FIGURES<br />

xvi<br />

2.9 EIT linewidth VS rate <strong>of</strong> VCCPC . . . . . . . . . . . . . . . . . . 55<br />

2.10 Clebsch-Gordon coefficients for Rb 8 5 D 1 line . . . . . . . . . . . . 58<br />

2.11 Bright Rabi frequencies Rb 8 5 . . . . . . . . . . . . . . . . . . . . 60<br />

2.12 Average optical pumping rate R ave . . . . . . . . . . . . . . . . . 61<br />

2.13 Four wave mixing . . . . . . . . . . . . . . . . . . . . . . . . . . . 63<br />

2.14 Visual definition <strong>of</strong> EIT terms . . . . . . . . . . . . . . . . . . . . 66<br />

3.1 Level diagram four-level atom . . . . . . . . . . . . . . . . . . . . 68<br />

3.2 Susceptibility dynamics . . . . . . . . . . . . . . . . . . . . . . . . 73<br />

3.3 Rise time VS Raman detuning theory . . . . . . . . . . . . . . . . 74<br />

3.4 EIT <strong>Kerr</strong> rise time (thick media . . . . . . . . . . . . . . . . . . . 79<br />

3.5 Simple QND diagram . . . . . . . . . . . . . . . . . . . . . . . . . 82<br />

4.1 Schematic <strong>of</strong> EIT <strong>Kerr</strong> dynamics <strong>of</strong> experiments . . . . . . . . . . 94<br />

4.2 Level diagram EIT <strong>Kerr</strong> dynamics experiment . . . . . . . . . . . 97<br />

4.3 Probe absorption measured (T=58 ◦ C) . . . . . . . . . . . . . . . 100<br />

4.4 Measured Stark shift due to signal . . . . . . . . . . . . . . . . . . 101<br />

4.5 Measured Stark shift due to coupling/probe . . . . . . . . . . . . 102<br />

4.6 Raw data homodyne signal–CW . . . . . . . . . . . . . . . . . . . 103<br />

4.7 Processed data homodyne signal–CW . . . . . . . . . . . . . . . . 104<br />

4.8 Raw data homodyne signal–transient . . . . . . . . . . . . . . . . 105<br />

4.9 Processed data homodyne signal–transient . . . . . . . . . . . . . 106<br />

4.10 GSC dynamics–measured . . . . . . . . . . . . . . . . . . . . . . . 108


LIST OF FIGURES<br />

xvii<br />

4.11 Rise and fall times (T=58 ◦ C) . . . . . . . . . . . . . . . . . . . . 109<br />

4.12 Rise and fall times (different Temperatures) . . . . . . . . . . . . 110<br />

4.13 Rise and fall times VS optical depth . . . . . . . . . . . . . . . . 111<br />

4.14 Comparison <strong>of</strong> theory and measured–GSC dynamics . . . . . . . . 112<br />

5.1 Theory for XPM with slow signal group velocity . . . . . . . . . . 120<br />

5.2 Diagram <strong>of</strong> slow signal experiment . . . . . . . . . . . . . . . . . 127<br />

5.3 Doppler valleys <strong>of</strong> Rb 85 and Rb 87 D 1 line . . . . . . . . . . . . . . 129<br />

5.4 Beam areas for slow signal experiment . . . . . . . . . . . . . . . 130<br />

5.5 Probe absorption for EIT in slow signal . . . . . . . . . . . . . . . 131<br />

5.6 Signal absorption . . . . . . . . . . . . . . . . . . . . . . . . . . . 132<br />

5.7 Probe phase for signal pulses <strong>of</strong> various duration . . . . . . . . . . 134<br />

5.8 Probe phase with slow and fast signal pulses . . . . . . . . . . . . 135<br />

5.9 Probe phase for various signal group velocities–theory . . . . . . . 136<br />

5.10 Width <strong>of</strong> probe phase pulse VS signal group velocity . . . . . . . 138<br />

6.1 EIT <strong>Kerr</strong> experiment using an optical cavity . . . . . . . . . . . . 143<br />

A.1 Transition matrix elements . . . . . . . . . . . . . . . . . . . . . . 162<br />

A.2 Dark states in rubidium 85 . . . . . . . . . . . . . . . . . . . . . . 164<br />

A.3 Dark states in rubidium 85 . . . . . . . . . . . . . . . . . . . . . . 165<br />

A.4 Rubidium 85 D lines hyperfine splittings . . . . . . . . . . . . . . 167<br />

A.5 Rubidium 87 D lines hyperfine splittings . . . . . . . . . . . . . . 168


xviii<br />

A.6 Transition matrix elements . . . . . . . . . . . . . . . . . . . . . . 169<br />

A.7 Dark states in rubidium 87 . . . . . . . . . . . . . . . . . . . . . . 170


1<br />

Chapter 1<br />

Introduction<br />

The field <strong>of</strong> nonlinear optics explores how the interaction between light and<br />

and a medium modifies the medium’s optical characteristics. For example, the<br />

index <strong>of</strong> refraction can change as a function <strong>of</strong> the intensity <strong>of</strong> the light, which<br />

is known as the <strong>Kerr</strong> nonlinearity. Other nonlinear effects include generation <strong>of</strong><br />

harmonic frequencies, frequency mixing between multiple fields (e.g.<br />

sum and<br />

difference frequency generation), and phase conjugate mirror effects. Generally,<br />

these nonlinear effects are only significant for optical fields with large intensities.<br />

Since the invention <strong>of</strong> the laser, nonlinear optics has become an integral part <strong>of</strong><br />

optical communications, remote sensing, spectroscopy, and quantum information<br />

science.<br />

Over the past decade, there has been increasing interest in realizing low-lightlevel<br />

nonlinear optics using coherently prepared media (CPM). By decreasing<br />

the intensity required to achieve nonlinear optical effects, many new applications<br />

become possible. One intriguing application is cross-phase modulation (XPM),


2<br />

where the <strong>Kerr</strong> nonlinearity is used to perform a quantum nondemolition (QND)<br />

measurement <strong>of</strong> photon number. For a long time the largest <strong>Kerr</strong> nonlinearities<br />

were still quite small (χ (3) ∼ 10 −22 m 2 /V 2 ) limiting QND measurements to intense<br />

fields with large numbers <strong>of</strong> photons. These QND measurements are <strong>of</strong> greatest<br />

interest for small numbers <strong>of</strong> photons where individual number states can be<br />

resolved.<br />

Using electromagnetically induced transparency (EIT), Schmidt and Imamoğlu<br />

[1] showed that it is theoretically possible to achieve <strong>Kerr</strong> nonlinearities that are<br />

ten orders <strong>of</strong> magnitude larger than conventional <strong>Kerr</strong> nonlinearities. Their analysis<br />

concluded, “This shows that our scheme makes possible conditional phase shifts<br />

<strong>of</strong> the order <strong>of</strong> π with single photons, which should be beneficial for quantum nondemolition<br />

measurements <strong>of</strong> weak signals and quantum logic gate operation....The<br />

principal result here is that one can obtain arbitrarily large XPM phase shifts <strong>of</strong><br />

the probe field by arbitrarily weak signal fields.”<br />

[1] This conclusion is overly<br />

optimistic because it ignores the rise-time <strong>of</strong> the EIT <strong>Kerr</strong> effect. However, it provides<br />

a reasonable glimpse into the possibilities <strong>of</strong>fered by low-light-level nonlinear<br />

optics.<br />

The focus <strong>of</strong> this work is to better understand the limitations and potential <strong>of</strong><br />

the EIT enhanced <strong>Kerr</strong> nonlinearity by exploring the dynamics and rise-time <strong>of</strong><br />

the EIT <strong>Kerr</strong> effect.


1.1. INTRODUCTION TO ELECTROMAGNETICALLY<br />

INDUCED TRANSPARENCY 3<br />

γ<br />

2<br />

3<br />

γ<br />

2<br />

2<br />

ΩP<br />

ΩC<br />

1<br />

Figure 1.1: Simple level diagram <strong>of</strong> an EIT Λ system with probe Ω p and coupling Ω c fields in<br />

single-photon resonance and Raman resonance.<br />

1.1 Introduction to <strong>Electromagnetically</strong> <strong>Induced</strong><br />

<strong>Transparency</strong><br />

Consider the three-level system shown in Fig. 1.1 with three electronic states |1〉,<br />

|2〉, and |3〉, and two dipole allowed transitions |2〉 ↔ |3〉 and |1〉 ↔ |3〉, that<br />

are resonant with a probe optical field with Rabi frequency Ω p and a coupling<br />

optical field with Rabi frequency Ω c . If the system is in Raman resonance (i.e.<br />

the frequency difference between the optical fields ω p −ω c is equal to the frequency<br />

difference between ground state energies ω 12 ≡ (E 1 − E 2 )/), there is coherent<br />

superposition <strong>of</strong> ground states called the dark state that is defined as<br />

|−〉 ≡ Ω p|1〉 − Ω c |2〉<br />

, (1.1)<br />

Ω


1.1. INTRODUCTION TO ELECTROMAGNETICALLY<br />

INDUCED TRANSPARENCY 4<br />

where Ω ≡ √ Ω 2 c + Ω 2 p is the bright Rabi frequency. This dark state is decoupled<br />

from the excited state because probability amplitude for the electronic transition<br />

|2〉 → |3〉 destructively interferes with probability amplitude for the transition<br />

|1〉 → |3〉. Once in the dark state an electron is “trapped” and cannot be excited<br />

to state |3〉 making the medium transparent to the optical fields.<br />

There is also a superposition <strong>of</strong> ground states orthogonal to the dark state<br />

called the bright state |+〉 which is coupled to the excited state with Rabi frequency<br />

Ω. Once an atom is excited from the bright state to level |3〉 it spontaneously<br />

decays into either the bright or dark states with equal probability. This<br />

results in coherent preparation and eventual trapping <strong>of</strong> atoms in this dark state.<br />

This phenomenon was first observed in the late 1970’s by Alzetta et al. [2, 3],<br />

and was given the name <strong>of</strong> coherent population trapping (CPT) by Gray et al.<br />

[4]. This phenomenon is also commonly referred to as electromagnetically induced<br />

transparency (EIT), which is a term first coined by Harris in 1990 [5].<br />

EIT can be found in many different types <strong>of</strong> electronic level structures with<br />

different number <strong>of</strong> levels [6]. Figure 1.2 shows examples <strong>of</strong> each <strong>of</strong> these systems<br />

with with level |2〉 being the “excited” state, and the dark state given by Eq. (1.1).<br />

EIT typically implies the assumption that the probe field is much smaller than<br />

the coupling field (Ω c ≫ Ω p ). Although this assumption is <strong>of</strong>ten convenient for<br />

simplifying the mathematics, it is not necessary. In this work we discuss Λ-type<br />

EIT systems exclusively.


1.1. INTRODUCTION TO ELECTROMAGNETICALLY<br />

INDUCED TRANSPARENCY 5<br />

3<br />

2<br />

1<br />

Ω C<br />

1<br />

Ω<br />

2<br />

P<br />

Ω<br />

1<br />

C<br />

Ω P<br />

3<br />

Ω C<br />

Ω P<br />

2<br />

3<br />

(a) (b) (c)<br />

Figure 1.2: Level diagrams for three different typed os EIT systems; (a) Λ, (b) V , and (c) ladder.<br />

Both the Λ and V systems will be Doppler free when the coupling and probe fields copropagate.<br />

The ladder system is Doppler free when the coupling and probe fields counter propagate.<br />

In the past 15 years EIT have developed into a very active research area.<br />

Much <strong>of</strong> this work in the early to mid 1990’s centered on using EIT for lasing<br />

without inversion (LWI). In LWI, EIT suppresses stimulated absorption but not<br />

stimulated emission such that when population is injected into the excited state<br />

gain is achieved without inversion [7–11]. EIT has also been applied to various<br />

precession measurement tasks such as magnetometery [12–15], spectroscopy [16,<br />

17], and improving frequency standards [18]. The value <strong>of</strong> EIT for precession<br />

measurement is due to the fact that the transparency resonance can be very<br />

narrow {e.g. an EIT linewidth with a full width at half maximum (FWHM) <strong>of</strong><br />

20-100 <strong>of</strong> Hz is not uncommon [19, 20]}, making EIT a very sensitive meter for<br />

the Raman detuning and any environmental changes which affect it.<br />

The narrow transparency resonance <strong>of</strong> EIT has a correspondingly steep dis-


1.2. EIT ENHANCED NONLINEAR OPTICS 6<br />

persion curves as dictated by the Kramer-Krönig relations. This steep dispersion<br />

results in slow group velocities for pulses with frequencies near the EIT transparency<br />

resonance. In 1999 several groups used the steep dispersion <strong>of</strong> EIT to<br />

observe the slowing <strong>of</strong> light by seven orders <strong>of</strong> magnitude, reducing the group velocity<br />

to 10’s <strong>of</strong> meters per second [21–23]. Using similar EIT slow-light techniques<br />

several groups “stopped” a light pulse ∗ , and holographically stored the light for a<br />

short time before coherently retrieving the pulse [24–26]. In recent work, Longdell<br />

et al. have achieved storage times <strong>of</strong> up to ten second for light pulses in EIT media<br />

[27]. The incorporation <strong>of</strong> EIT and a CPM have been used to create single<br />

photon states, store these photons, and then retrieve them while preserving their<br />

quantum properties [28].<br />

1.2 EIT enhanced nonlinear optics<br />

Apart from the EIT <strong>Kerr</strong> nonlinearity, which we will discuss separately, there<br />

have been many advances in using EIT to provide more efficient low-light-level<br />

nonlinear optical media. The field <strong>of</strong> EIT enhanced nonlinear optics began in the<br />

mid 1990’s only a few years after the first experiments in EIT. In 1995 Hemmer et<br />

∗ In stopped light there are no photons stored in the medium, only a holographic imprint <strong>of</strong><br />

the pulse stored in the coherence <strong>of</strong> the medium. The term stopped light really refers to stopping<br />

a dark state polarition (a quasi-particle consisting <strong>of</strong> a superposition <strong>of</strong> photon and exciton and<br />

enabled by the dark state <strong>of</strong> EIT). The dark state polariton is stopped by adiabatically turning<br />

<strong>of</strong>f the cw coupling field, which continuously slows down the probe field until it “stops”. When<br />

the coupling field goes completely to zero such that the dark state polariton is stationary, there<br />

are no photons in dark state polariton, which consists entirely <strong>of</strong> the coherence between ground<br />

states. Thus, “stopped” light is essentially the same as a hologram.


1.3. EIT KERR NONLINEARITY 7<br />

al. demonstrated that EIT could be used to realize efficient phase conjugate effects<br />

[29], and in 1996 several groups demonstrated efficient frequency conversion<br />

via EIT [30, 31]. Other subjects that have been studied in connection with EIT<br />

include coherent Raman scattering [32], four-wave mixing [33–36], generating single<br />

photons [28, 37], generating correlated photons [38–40], and number squeezing<br />

<strong>of</strong> optical fields [41].<br />

1.3 EIT <strong>Kerr</strong> nonlinearity<br />

The EIT <strong>Kerr</strong> nonlinearity was originally proposed by Schmidt and Imamoğlu<br />

in 1996 [1]. The Schmidt-Imamoğlu proposal adds a fourth energy level and an<br />

additional optical field, labeled the signal field, to the Λ EIT system in Fig. 1.1,<br />

transforming the Λ into the shape <strong>of</strong> an N [see Fig 1.2(a)].<br />

The signal field<br />

induces an AC Stark shift † that perturbs the ground states perturbing <strong>of</strong> the EIT<br />

system away from Raman resonance. The combination <strong>of</strong> Stark shift and steep<br />

dispersion due to EIT result in a significant change in the index <strong>of</strong> refraction,<br />

and the resulting XPM <strong>of</strong> the optical fields is the EIT <strong>Kerr</strong> nonlinearity. The<br />

advantage <strong>of</strong> the EIT <strong>Kerr</strong> effect is that for continuous wave (cw) fields, it can<br />

be arbitrarily large by either increasing the optical thickness <strong>of</strong> the EIT medium<br />

or by decreasing the EIT linewidth making the dispersion steeper. For pulsed<br />

† The AC Stark shift is also known as the light shift and the dynamic Stark shift.


1.3. EIT KERR NONLINEARITY 8<br />

applications, which are the focus <strong>of</strong> this work, the dynamics <strong>of</strong> the EIT <strong>Kerr</strong><br />

nonlinearity limit the XPM [42].<br />

Similar to the refractive <strong>Kerr</strong> nonlinearity, Harris and Yamamoto proposed a<br />

scheme for an absorptive EIT <strong>Kerr</strong> nonlinearity in which the EIT medium “absorbs<br />

two photons, but does not absorb one photon.” [43]. Although there are<br />

technically two EIT <strong>Kerr</strong> nonlinearities (i.e. absorptive and refractive), in this<br />

work we typically refer to the refractive EIT <strong>Kerr</strong> nonlinearity as the EIT <strong>Kerr</strong><br />

nonlinearity. While we always explicitly refer to the absorptive EIT <strong>Kerr</strong> nonlinearity<br />

as being absorptive.<br />

While both EIT <strong>Kerr</strong> nonlinearities have been observed in cw conditions [44–<br />

48], only the dynamics <strong>of</strong> the absorptive <strong>Kerr</strong> nonlinearity have been observed<br />

and characterized [49–51]. The experiments described in this work are the first<br />

observations <strong>of</strong> the dynamics <strong>of</strong> the refractive EIT <strong>Kerr</strong> nonlinearity.<br />

Several other schemes for EIT <strong>Kerr</strong> nonlinearities followed the original<br />

Schmidt-Imamoğlu proposal.<br />

A few <strong>of</strong> these EIT <strong>Kerr</strong> schemes are shown in<br />

Fig. 1.3. The original four-level N systems proposed by Schmidt and Imamoğlu is<br />

shown in Fig. 1.3(a) [1]. Matsko et al. suggested that M-type systems like the one<br />

shown in Fig. 1.3(b) also exhibit large EIT <strong>Kerr</strong> nonlinearities [52, 53]. The tripod<br />

scheme shown in Fig. 1.3(c) was proposed by Petrosyan and Malakyan [54, 55].<br />

Harris and Hau showed that the nonlinear interaction between probe and signal<br />

fields in an N system is limited by temporal walk-<strong>of</strong>f. The temporal walk-<strong>of</strong>f


1.3. EIT KERR NONLINEARITY 9<br />

(a)<br />

(b)<br />

Ω S<br />

ΩS<br />

Ω C 2<br />

Ω<br />

P<br />

Ω<br />

C<br />

Ω Ω<br />

P C 1<br />

(c)<br />

(d)<br />

Ω P<br />

Ω C<br />

Ω Ω C<br />

1<br />

P<br />

Ω S<br />

Ω S<br />

Ω C2<br />

Ω S<br />

Atomic<br />

Species #1<br />

Atomic<br />

Species #2<br />

Figure 1.3: Level diagrams for various systems exhibiting EIT <strong>Kerr</strong> nonlinearities; (a) N system,<br />

(b) M system, (c) a tripod system, and (d) the matched group velocity system.


1.4. OTHER RESEARCH RELATED TO EIT KERR<br />

NONLINEARITIES 10<br />

is the result <strong>of</strong> the different group velocities <strong>of</strong> the probe and signal pulses. The<br />

probe pulse propagates with a slow group velocity due to EIT ‡ . While, the signal’s<br />

group velocity is close to c the speed <strong>of</strong> light in vacuum [56]. To overcome<br />

this temporal walk-<strong>of</strong>f and increase the interaction length several research groups<br />

proposed double Λ schemes, where the signal field would also be part <strong>of</strong> an EIT<br />

system and propagate with a slow group velocity [57, 58]. Figure 1.3(d) shows a<br />

diagram for the double Λ proposal by Lukin and Imamoğlu, in which the signal<br />

EIT is obtained using a second atomic species [57]. Petrosyan and Kurizki have<br />

also proposed a slightly more elegant double Λ scheme that requires only a single<br />

atomic system [58].<br />

1.4 Other research related to EIT <strong>Kerr</strong> nonlinearities<br />

In addition to the research discussed above there has been a tremendous amount <strong>of</strong><br />

related work that provides valuable context for our research. Due to the sheer volume<br />

<strong>of</strong> work related to quantum coherence and coherent transients in multi-level<br />

systems, our review cannot be comprehensive. We focus on the most important<br />

historical developments and those experiments that are closely related to the EIT<br />

<strong>Kerr</strong> effect.<br />

‡ The coupling field is assumed to be cw.


1.4. OTHER RESEARCH RELATED TO EIT KERR<br />

NONLINEARITIES 11<br />

2<br />

ΩP<br />

3<br />

Undressed<br />

Picture<br />

ΩC<br />

1<br />

hΩ C<br />

2<br />

Ω P<br />

Dressed State<br />

Picture<br />

3'<br />

1'<br />

Figure 1.4: Transformation to the dressed state picture in which the atom is dressed by the<br />

coupling field. In the dressed state picture the Autler-Townes splitting <strong>of</strong> the excited state<br />

becomes evident.<br />

First, there are primarily three historical ideas that laid the framework for<br />

EIT; the Hanle effect [59], Fano interference [60], and Autler-Townes splitting<br />

[61]. The field <strong>of</strong> quantum coherence and interference dates back to 1921 with the<br />

Hanle effect and quantum beats. The Hanle effect occurs when a medium with a V<br />

shaped electronic structure [e.g. see Fig. 1.2(b)] absorbs linearly polarized light.<br />

When a magnetic field Zeeman shifts the excited states, the light emitted from the<br />

medium will have a precessing linear polarization [59]). Fano interference occurs<br />

when two ionization pathways (autoionization and photoionization) interfere to<br />

inhibit absorption and ionization. [60]. The difference between EIT and Fano<br />

interference is that, in Fano interference the medium starts in a single electronic<br />

level not a coherent superposition <strong>of</strong> atomic levels.<br />

Finally, Autler-Townes splitting is very closely related to EIT [61]. Autler-<br />

Townes splitting can be understood by considering any <strong>of</strong> the three level systems


1.4. OTHER RESEARCH RELATED TO EIT KERR<br />

NONLINEARITIES 12<br />

shown in Fig. 1.2. When the atom is dressed by the coupling field, there are two<br />

dressed states |2 ′ 〉 and |3 ′ 〉 as shown in Fig. 1.4. These two dressed states are an<br />

Autler-Townes doublet with an energy difference <strong>of</strong> ∆U = |Ω c | = µ|E c |, where<br />

µ is the dipole moment and E c is the amplitude <strong>of</strong> the coupling electric field. In<br />

one paradigm for EIT, EIT is interpreted as the destructive interference between<br />

the Autler-Townes doublet. In three-level systems, like the one shown in Fig. 1.4,<br />

the distinction between EIT and Autler-Townes splitting is somewhat artificial<br />

because there is no way <strong>of</strong> isolating the two effects.<br />

In optical dense media where radiation trapping is significant it is impossible<br />

to use CPT to obtain EIT because CPT requires spontaneous emission [62–65].<br />

Radiation trapping causes the spontaneously emitted photon to be recaptured and<br />

reradiated several times before leaving the region <strong>of</strong> EIT resulting in prohibitive<br />

decoherence and destroying EIT. In these cases stimulated rapid adiabatic passage<br />

(STIRAP) can be used to obtain EIT [62, 63]. In STIRAP the medium starts in<br />

a non coherent dark state (i.e. it starts in a single electronic state) and the fields<br />

are evolved adiabatically such that the system remains trapped in the dark state<br />

as it evolves to a coherent dark state [66, 67].<br />

There is also a great deal <strong>of</strong> theoretical research studying EIT in cavities.<br />

There are two reasons why EIT in cavities is interesting. First, like EIT enhanced<br />

nonlinearities, cavities can help increase the strength <strong>of</strong> nonlinear optical inter-


1.4. OTHER RESEARCH RELATED TO EIT KERR<br />

NONLINEARITIES 13<br />

actions [68–70].<br />

Some researchers, mostly at the University <strong>of</strong> Arkansas, have<br />

performed experiments with EIT in cavities to study optical bistability [46].<br />

The propagation <strong>of</strong> optical pulses in EIT media has been studied extensively,<br />

but it seems this field is still far from maturity. We have already mentioned that<br />

EIT can be used to obtain very slow group velocities. When the coupling field is<br />

cw and the probe field is pulsed, the probe pulse will propagate at a group velocity<br />

v g,p ≈<br />

c<br />

1 + χ p /2 + ω p<br />

2<br />

∂χ p<br />

,<br />

∂ω p<br />

where ω p is the probe frequency, and χ p ≪ 1 is the probe susceptibility. However,<br />

for matched pulses § in an EIT media, all fields propagate with group velocities near<br />

c the speed <strong>of</strong> light in vacuum [71]. Also, when two pulses with sufficiently energy<br />

enter a coherently prepared EIT media, they will naturally evolve towards matched<br />

pulses[72]. There is a dark area theorem for EIT similar to the area theorem <strong>of</strong><br />

self induced transparency that describes certian aspects <strong>of</strong> this evolution towards<br />

matched pulses [73]. Various solutions for temporal soliton have also been found<br />

for EIT media; simultons [74, 75], adiabatons [76–79], etc. [80–82]. Some <strong>of</strong> these<br />

solitons propagate at a slow group velocity, while others propagate with group<br />

velocity near c [81, 82]. For the EIT <strong>Kerr</strong> nonlinearity spatial Thirring solitons,<br />

have been predicted [83].<br />

§ In matched pulses both the probe and coupling fields are pulsed with the same field amplitudes<br />

and phase modulation– i.e. Ω p (t) = κΩ c (t), where κ is a constant <strong>of</strong> proportionality.


1.4. OTHER RESEARCH RELATED TO EIT KERR<br />

NONLINEARITIES 14<br />

F e = 1<br />

Ω (σ )<br />

c<br />

+<br />

Ω p<br />

(σ −)<br />

Applying<br />

B-field<br />

F g = 1<br />

m = -1 m =0 m = +1<br />

(a)<br />

(b)<br />

µ B B z<br />

Figure 1.5: Level diagram exhibiting the relation between nonlinear magneto-optic rotation and<br />

electromagnetically induced transparency. Right- and left-hand circular polarized light interact<br />

with an atom with a ground hyper-fine level F g = 1 and excited hyperfine level F e = 1. In the<br />

absence <strong>of</strong> a magnetic field (a) Raman resonance is established leading to coherent population<br />

trapping. When a magnetic field is applied (b) Zeeman splitting shifts the ground (excited)<br />

magnetic sublevels by mµ B,g B z (mµ B,e B z ), where mu B,i is the magnetic moment for level i<br />

and B z is the magnitude <strong>of</strong> the magnetic field which is directed along the direction <strong>of</strong> optical<br />

propagation.<br />

EIT-<strong>Kerr</strong> dynamics are also closely related to a considerable amount <strong>of</strong> work<br />

that has been performed under the broad umbrella <strong>of</strong> coherent transients in multilevel<br />

systems. Park et al. and Chen et al. have observed absorption transients in<br />

Λ-EIT media induced by rapid changes to the Raman (two-photon) detuning [51,<br />

84]. Similarly, Godone et al. have observed transients due to phase modulation<br />

<strong>of</strong> Raman beams in a Λ system [85]. Many others have also studied transients in<br />

three level EIT systems under a number <strong>of</strong> different contexts such as: coherent<br />

Raman beats [86], Zeeman splitting [87–89], EIT in semiconductors [16, 90] and<br />

coherent population trapping in closed loop systems [91, 92].<br />

Nonlinear magneto-optical rotation (NMOR) shares many similarities with<br />

EIT. In fact there are some experiments that can either be explianed either as EIT


1.4. OTHER RESEARCH RELATED TO EIT KERR<br />

NONLINEARITIES 15<br />

or as NMOR with different insights coming from the different perspectives. For<br />

example, consider the atomic level structure shown in Fig. 1.5(a). The probe and<br />

coupling fields are equal strength with the same frequency and orthogonal circular<br />

polarizations.<br />

Taken together, the probe and coupling field can be considered<br />

as one field with linear polarization, and the angle <strong>of</strong> the linear polarization is<br />

determined by the relative phase between the probe and coupling field.<br />

CPT<br />

causes the atom to become trapped in the dark state |−〉 = 1/ √ 2(|g, −1〉 +<br />

|g, +1〉) . Non-zero Raman detunings can be achieved using a magnetic field in<br />

the direction <strong>of</strong> optical propagation to Zeeman shift the magnetic sublevels [this<br />

is shown in Fig. 1.5(b)]. Away from Raman resonance the coupling and probe<br />

fields experience phase shifts with opposite signs rotating the linear polarization<br />

<strong>of</strong> the combined probe/coupling field. This rotation <strong>of</strong> the optical polarization is<br />

what is meant by NMOR.<br />

One <strong>of</strong> the slowest group velocity measured using EIT was in a NMOR experiment<br />

[21]. In this experiment, which used a room temperature rubidium vapor<br />

with paraffin coated walls, the atoms were allowed to establish equilibrium with<br />

a linearly polarized field. The polarization angle was then modulated; in the language<br />

<strong>of</strong> EIT one would say the probe field was phase modulated. The phase<br />

For this system the Clebsch-Gordon coefficients for the σ ± polarized light is ∓1/ √ 2. This<br />

difference in sign for the two circular polarizations accounts <strong>of</strong> the pulse sign in the dark state<br />

superposition.


1.5. APPLICATIONS OF THE EIT KERR NONLINEARITY 16<br />

modulation <strong>of</strong> the probe and coupling fields propagated through the medium at<br />

8 m/s.For a review <strong>of</strong> NMOR see Budker [93].<br />

There are also several reviews <strong>of</strong> CPT and EIT [6, 94–96] and its subfields<br />

[66, 93, 97–100].<br />

1.5 Applications <strong>of</strong> the EIT <strong>Kerr</strong> nonlinearity<br />

As we have already mentioned there are several applications that would benefit<br />

from larger <strong>Kerr</strong> nonlinearities. For example, there are classical applications<br />

such as all optical switching for fiber communication networks.<br />

In this work<br />

we have focused more on the quantum and few photon applications <strong>of</strong> the EIT<br />

<strong>Kerr</strong> effect. In particular we focus on the possibility <strong>of</strong> QND measurements <strong>of</strong><br />

photon number with single photon resolution. The ability to detect and resolve<br />

individual optical number states would have several uses in quantum information<br />

processing (QIP) including; long-distance quantum teleportation [101], Bell-state<br />

measurements [102], quantum bit regeneration [103], optical Fock state synthesis<br />

[104], multiphoton correlation measurements [105], simultaneous closure <strong>of</strong> all <strong>of</strong><br />

Bell’s inequalities loopholes [106, 107], and better-than-diffraction-limited photolithography<br />

[108]. We elaborate on a couple <strong>of</strong> these potential uses <strong>of</strong> few-photon<br />

<strong>Kerr</strong> nonlinearities.


1.5. APPLICATIONS OF THE EIT KERR NONLINEARITY 17<br />

1.5.1 Number resolving photon counters<br />

QND measurement <strong>of</strong> photon number may be the best way to improve upon the<br />

current state <strong>of</strong> photon counters/detectos and satisfy the demanding requirements<br />

<strong>of</strong> quantum computing and QIP. The current state <strong>of</strong> the art photon counters are<br />

Gieger-mode avalanche photodiodes with quantum efficiencies in the range <strong>of</strong> 0.2-<br />

0.7, and these detectors cannot distinguish between one photon and many photons.<br />

Also, most optical QIP experiments use weak coherent states and post-selection.<br />

Because weak coherent states have finite probabilities <strong>of</strong> having multiple photons,<br />

these experiments are operated at low photon flux (i.e. coherent states with much<br />

less than one photon on average) to minimize the number <strong>of</strong> multi-photon states<br />

post-selected as single photon states. By being able to distinguish between 1, 2<br />

and more photons, a QND measurement would enable QIP experiments to operate<br />

with higher photon flux.<br />

1.5.2 Quantum state preparation<br />

<strong>Kerr</strong> nonlinearities are helpful for quantum state preparation in two ways: as<br />

projective measurements and for entangling two previously uncorrelated wavepackets.<br />

QND measurements project the signal field into an eigen state <strong>of</strong> the<br />

measurement basis. For <strong>Kerr</strong> QND measurements the measurement basis is the<br />

number states. Thus, <strong>Kerr</strong> QND measurements provide a method for obtaining<br />

optical number states form any optical source (thermal or laser). Deterministic


1.5. APPLICATIONS OF THE EIT KERR NONLINEARITY 18<br />

optical quantum computing requires single photons on demand. Being able to<br />

preselect single photon states using a QND measurement would go a long was<br />

towards satisfying this demand. EIT <strong>Kerr</strong> nonlinearities can also entangle two<br />

previously uncorrelated wavepackets [57, 58]. This is accomplished using a QND<br />

measurement without a classical measurement <strong>of</strong> the probe/meter field. Finally,<br />

<strong>Kerr</strong> QND measurements could be used to create macroscopic quantum states<br />

(i.e. Schrödinger cat states) [96].<br />

1.5.3 Quantum logic gates<br />

In their original paper Schmidt and Imamoğlu speculated the the EIT <strong>Kerr</strong> effect<br />

“should be beneficial for...quantum logic gate operation” [1]. This is an important<br />

application from the perspective <strong>of</strong> optical quantum computing because in<br />

many ways photons make ideal qubits. They experience little decoherence. It is<br />

straightforward to move photons between nodes <strong>of</strong> a quantum network or computer.<br />

Entangled photons are straightforward to obtain (i.e. via parametric down<br />

conversion). Finally, single qubit quantum logic operations are trivial (e.g. a<br />

Hadamard gate is simply a 50/50 beam splitter, and a phase shift is a wave<br />

plate). However, deterministic two-qubit quantum logic operations using photons<br />

have been elusive because photon-photon interactions are so weak (i.e. traditionally<br />

nonlinear optics has required intense optical fields). Linear optics quantum<br />

computing (LOQC) somewhat circumvents this issue using probabilistic two-qubit


1.6. OUTLINE OF DISSERTATION 19<br />

quantum gates, but LOQC has its own set <strong>of</strong> issues which are beyond the scope<br />

<strong>of</strong> this work [109, 110].<br />

Several architectures based on optical <strong>Kerr</strong> nonlinearities have been proposed<br />

for quantum computing [111–115]. For example, a <strong>Kerr</strong> nonlinearity that is strong<br />

enough to induce a π phase shift with a single photon could be used as a controlled<br />

phase gate. Even a weak <strong>Kerr</strong> nonlinearity (i.e. much less π phase shift per<br />

photon) can be used to implement a CNOT gate [116].<br />

Several groups have<br />

predicted high fidelity quantum logic gate operation using EIT <strong>Kerr</strong> nonlinearities<br />

[117–119]. Although, Shapiro found that a fully quantum theory <strong>of</strong> <strong>Kerr</strong> effect<br />

predicts poor fidelity for <strong>Kerr</strong> based quantum logic gates [120, 121], the fidelity<br />

<strong>of</strong> <strong>Kerr</strong> based quantum logic gates is still being debated.<br />

Even if high fidelity <strong>Kerr</strong> quantum logic gates are not possible, the EIT <strong>Kerr</strong><br />

effect may still benefit quantum computing by enabling improved single photon<br />

sources and detectors.<br />

1.6 Outline <strong>of</strong> dissertation<br />

The most intriguing applications <strong>of</strong> the EIT <strong>Kerr</strong> nonlinearity are pulsed applications.<br />

Thus, a good understanding <strong>of</strong> the dynamics <strong>of</strong> the EIT <strong>Kerr</strong> nonlinearity<br />

is essential for successfully realizing these applications, and understanding what<br />

are the limitations <strong>of</strong> the EIT <strong>Kerr</strong> nonlinearity.<br />

Most papers discussing the EIT <strong>Kerr</strong> nonlinearity and its applications ignore


1.6. OUTLINE OF DISSERTATION 20<br />

the rise-time <strong>of</strong> the EIT <strong>Kerr</strong> effect (i.e. treat it as though it is instantaneous). For<br />

example, Schmidt and Imamoğlu overlooked the rise-time <strong>of</strong> the <strong>Kerr</strong> nonlinearity<br />

when they predicted that the EIT <strong>Kerr</strong> effect “makes possible conditional phase<br />

shifts <strong>of</strong> the order <strong>of</strong> π with single photons” [1]. When the dynamics are considered,<br />

the largest phase shift per photon that can be achieved is about 1/100th <strong>of</strong> a<br />

radian even under the most ideal conditions. Overlooking the EIT <strong>Kerr</strong> rise-time<br />

is common even in papers explicitly discussing dynamical properties <strong>of</strong> pulsed<br />

EIT <strong>Kerr</strong> systems (see for example [56, 57, 122]). These “transient” analysies<br />

limit their discussion to group velocities and transmission bandwidths. Recently,<br />

there has been greater attention given to dynamics <strong>of</strong> the EIT <strong>Kerr</strong> nonlinearity<br />

[42, 120, 123].<br />

One <strong>of</strong> the advantages <strong>of</strong> the EIT <strong>Kerr</strong> nonlinearity is that for cw applications,<br />

it can be arbitrarily large.<br />

The EIT <strong>Kerr</strong> effect can be made larger by either<br />

decreasing the linewidth <strong>of</strong> the EIT resonance or increasing the medium’s optical<br />

depth αL, both <strong>of</strong> which are theoretically boundless (α is the absorption coefficient<br />

and L is the thickness <strong>of</strong> the medium). However, for pulsed applications the correct<br />

figure <strong>of</strong> merit is the ratio between the size <strong>of</strong> the EIT <strong>Kerr</strong> effect and its rise-time.<br />

As we will show in this work, the EIT <strong>Kerr</strong> rise-time is directly proportional to<br />

the optical depth and inversely proportional to the EIT linewidth. Thus, the ratio<br />

between the size <strong>of</strong> the EIT <strong>Kerr</strong> effect and its rise-time is a constant with respect<br />

to both the EIT linewidth and optical depth.


1.6. OUTLINE OF DISSERTATION 21<br />

The objective <strong>of</strong> this dissertation is to first explore the dynamics <strong>of</strong> the EIT<br />

<strong>Kerr</strong> effect, and consider how these dynamics relate to QND measurements <strong>of</strong><br />

photon number based on the EIT <strong>Kerr</strong> effect.<br />

Chapter 2 is a theoretical discussion <strong>of</strong> EIT. We begin by considering the<br />

experimentally naive scenario <strong>of</strong> a three-level Λ EIT system with a single velocity<br />

class. We review two methods for calculating the steady state susceptibilities in<br />

EIT; the perturbative and Bloch-vector methods. The Bloch-vector method is<br />

then applied to the problem <strong>of</strong> transient solutions <strong>of</strong> EIT. Next, we generalize the<br />

three-level Bloch-vector solution to include Doppler broadening, buffer gas, and<br />

multiple magnetic sublevels in anticipation <strong>of</strong> our experiment.<br />

Chapter 3 is a theoretical discussion <strong>of</strong> the EIT <strong>Kerr</strong> nonlinearity and QND<br />

measurements using <strong>Kerr</strong> nonlinearities. We first discuss the cw aspects <strong>of</strong> the<br />

EIT <strong>Kerr</strong> nonlinearity before deriving expressions for its dynamical evolution.<br />

Next, we consider QND measurements <strong>of</strong> photon number using <strong>Kerr</strong> nonlinearities,<br />

and derive expressions for the signal to noise ratio for these QND measurements.<br />

This analysis <strong>of</strong> the EIT <strong>Kerr</strong> dynamics and QND measurements illustrates the<br />

limitations <strong>of</strong> the EIT <strong>Kerr</strong> effect due to its relatively slow rise-time.<br />

Chapter 4 describes our experimental observations <strong>of</strong> the EIT <strong>Kerr</strong> dynamics.<br />

The experiment is performed using rubidium 85 atoms with buffer gas in a room<br />

temperature vapor cell. The measurements <strong>of</strong> EIT <strong>Kerr</strong> dynamics and rise-time


1.6. OUTLINE OF DISSERTATION 22<br />

show agreement between theory and experiment for variations in optical thickness,<br />

EIT linewidth, and size <strong>of</strong> the Stark shift.<br />

In Chapter 5 we consider a matched group velocity experiment designed to<br />

overcome some <strong>of</strong> the limitations imposed by the slow EIT <strong>Kerr</strong> rise-time. This is<br />

similar to the Lukin-Imamoğlu double Λ proposal shown in Fig. 1.3(d) [57]. We<br />

first derive a simple theory which accounts for signal and probe pulse propagation,<br />

EIT <strong>Kerr</strong> interaction, and EIT <strong>Kerr</strong> dynamics. Then, we present experimental<br />

observations <strong>of</strong> an EIT <strong>Kerr</strong> system with a slow group velocity for the signal field.<br />

Finally, we conclude by summarizing our results and their implications for<br />

potential EIT <strong>Kerr</strong> applications. We also discuss open questions related to this<br />

work and future directions.


23<br />

Chapter 2<br />

Theory <strong>of</strong> <strong>Electromagnetically</strong><br />

induced transparency<br />

<strong>Electromagnetically</strong> induced transparency (EIT) is the result <strong>of</strong> quantum interference<br />

between multiple transitions to a single final electronic state. In order<br />

to obtain this interference the initial electronic states must be mutually coherent<br />

(i.e. the medium is coherently prepared).<br />

In this chapter we provide a detailed mathematical description <strong>of</strong> EIT in various<br />

systems and circumstances. The first three sections (i.e. Secs. 2.1-2.3)<br />

present steady state solutions for EIT assuming a single velocity class ∗ and the<br />

three-level system shown in Fig. 2.2. Section 2.1 provides a physical intuitive<br />

“back-<strong>of</strong>-the-envelope” discussion <strong>of</strong> EIT, but is <strong>of</strong> little utility for comparison<br />

with experiments. The perturbative solution in Sec. 2.2 also provides valuable<br />

insight, but is limited by the assumption <strong>of</strong> small probe fields (i.e. Ω p ≪ Ω c ). Fi-<br />

∗ Room temperature atomic vapors have multiple velocity classes due to Doppler broadening.<br />

A velocity class is defined as the collection <strong>of</strong> atoms whose velocity along the direction <strong>of</strong> laser<br />

propagation lie in the range v 1 < v ≤ v 2 , such that the doppler shift associated with frequency<br />

difference ∆v = v 2 − v 1 is equal to the atoms homogeneous linewidth γ (i.e. γ = ω 0 ∆v/c).


2.1. BACK OF THE ENVELOPE THEORY OF EIT 24<br />

nally, the Bloch-vector solution in Sec. 2.3 is too complicated to be very insightful,<br />

but it is accurate.<br />

Section 2.4 derives transient solutions <strong>of</strong> EIT using the Bloch-vector formalism.<br />

Finally, Secs. 2.5, 2.6, and 2.7 generalize the Bloch-vector solution to account<br />

for Doppler broadening, buffer gas, and the magnetic sublevel and hyperfine levels<br />

associated with real atomic systems.<br />

2.1 Back <strong>of</strong> the envelope theory <strong>of</strong> EIT<br />

A simple back-<strong>of</strong>-the-envelope calculation illustrates the dominant mechanisms <strong>of</strong><br />

EIT and provides a valuable reference point before beginning a detailed theoretical<br />

treatment <strong>of</strong> EIT. Although there are other mechanisms, such as stimulated rapid<br />

adiabatic passage (STIRAP) [63], for obtaining the atomic coherence required for<br />

EIT, CPT is the most common mechanism for achieving EIT. To derive an approximate<br />

solution for CPT, it is sufficient to perform a detailed balancing analysis<br />

between pumping from the bright state to dark states (e.g. optical pumping) and<br />

pumping from dark to bright (e.g. decoherence). Figure 2.1 shows a three-level Λ<br />

EIT system in both the physical basis and the bright/dark basis. The bright and<br />

dark states are respectively defined as<br />

|+〉 ≡ Ω p|1〉 − Ω c |2〉<br />

Ω<br />

, and|−〉 ≡ Ω∗ c|1〉 + Ω ∗ p|2〉<br />

,<br />

Ω


2.1. BACK OF THE ENVELOPE THEORY OF EIT 25<br />

γ/2<br />

3<br />

γ/2<br />

3<br />

γ<br />

γ /2<br />

/2<br />

Ω P<br />

Ω C<br />

Ω<br />

2<br />

Γ<br />

(a)<br />

1<br />

+<br />

Γ<br />

(b)<br />

-<br />

Figure 2.1:<br />

To understand EIT in a three-level Λ system sometimes it is helpful to transform<br />

from the physical basis (a) to the bright/dark basis (b). Raman resonance is<br />

assumed, and γ (Γ) is the decay rate from the excited (ground) state.<br />

where Ω = √ |Ω p | 2 + |Ω c | 2<br />

is the “bright” Rabi frequency, and we recall that<br />

Ω p ≡ µ p E p / (Ω c ≡ µ c E c /) are the probe and coupling Rabi frequencies. The<br />

probe and coupling fields are cw monochromatic fields.<br />

There are two processes moving atoms from the bright to the dark state;<br />

optical pumping R and population exchange between ground states Γ. The optical<br />

pumping rate from the bright state to the dark state is<br />

R =<br />

γΩ 2 /4<br />

Ω 2 + ∆ 2 γ/γ ⊥ + γγ ⊥<br />

, (2.1)<br />

where γ is the spontaneous decay rate out <strong>of</strong> state |3〉, γ ⊥<br />

≡ γ/2 + γ ′ is the<br />

transverse decay rate (i.e. the decay rate for the coherence terms ρ 31 and ρ 32 ),<br />

and γ ′ is the decoherence due to dephasing (i.e. no change in population). Thus,


2.1. BACK OF THE ENVELOPE THEORY OF EIT 26<br />

the total rate <strong>of</strong> transfer from bright to dark state is<br />

R |+〉→|−〉 = R + Γ/2, (2.2)<br />

There are also two processes transferring atoms from the dark to bright state;<br />

non-zero Raman detuning resulting in phase change for the coherence between<br />

ground states, and population exchange between ground states Γ. Thus, the rate<br />

<strong>of</strong> population transfer from the dark state to the bright state is<br />

R |−〉→|+〉 = Γ/2 + δ 2 R/R, (2.3)<br />

where δ R ≡ ω p − ω 2 − ω c + ω 1 is the Raman detuning.<br />

Using detailed balancing between the dark and bright states we see that the<br />

number density <strong>of</strong> atoms in the bright state is<br />

N + =<br />

NR |−〉→|+〉<br />

R |−〉→|+〉 + R |+〉→|−〉<br />

, (2.4)<br />

where the total number density <strong>of</strong> atoms is N = N + + N − , and the absorption <strong>of</strong><br />

the optical fields is<br />

αL ∝ N + ∝ ΓR+δ2 R<br />

, (2.5)<br />

R 2 +ΓR+δR<br />

2<br />

where α is an absorption coefficient and L is the length <strong>of</strong> the medium.


2.1. BACK OF THE ENVELOPE THEORY OF EIT 27<br />

Two important details are illustrated by Eq. 2.5. First, the medium only<br />

becomes “transparent” if the EIT condition R ≫ Γ is satisfied. Second, when<br />

the EIT condition satisfied and the probe and coupling fields are near singlephoton<br />

resonance, the line shape <strong>of</strong> the transparency resonance is approximately<br />

Lorentzian with a linewidth <strong>of</strong> 2R for full width at half maximum (FWHM).<br />

Finally, we note that the essential resource which enables EIT is ground states<br />

coherence ρ 21 being equal to the dark ground states coherence ρ (−)<br />

21 = −Ω c Ω ∗ p/Ω 2 .<br />

In practice, it is relatively trivial to achieve the desired ratio between ground state<br />

populations (i.e. ρ 22 /ρ 11 ≈ ρ (−)<br />

22 /ρ (−)<br />

11 = Ω 2 c/Ω 2 p, where ρ (−)<br />

22 = Ω 2 c/Ω 2 and ρ (−)<br />

11 =<br />

Ω 2 p/Ω 2 are the dark state populations for levels |2〉 and |1〉). For example, mutually<br />

incoherent † probe and coupling fields will achieve the desired ratio between ground<br />

state populations as long as the EIT condition is met and δ R ≪ γ, but there will<br />

not be EIT because ρ 21 = 0. Both coherent and incoherent optical pumping<br />

processes reach equilibrium with ρ 22 /ρ 11 = Ω 2 c/Ω 2 p, but only the coherent process<br />

(i.e. CPT) has an equilibrium <strong>of</strong> ρ 21 = ρ (−)<br />

21 .<br />

† <strong>Optical</strong> fields are incoherent when they have large uncorrelated phase fluctuations.


2.2. PERTURBATIVE SOLUTION FOR EIT 28<br />

2.2 Perturbative solution for EIT<br />

To make our discussion <strong>of</strong> EIT more rigorous we consider the master equation for<br />

the three-level system shown in Fig. 2.2, which is<br />

˙ρ = 1 [H, ρ] − D, (2.6)<br />

i<br />

where the coherent evolution is determined by the Hamiltonian<br />

⎛<br />

H =<br />

⎜<br />

⎝<br />

δ R /2 0 − Ω∗ C<br />

2<br />

0 −δ R /2 − Ω∗ P<br />

2<br />

− Ω C<br />

2<br />

− Ω P<br />

2<br />

∆<br />

⎞<br />

. (2.7)<br />

⎟<br />

⎠<br />

The decay/repumping matrix is<br />

⎛<br />

D =<br />

⎜<br />

⎝<br />

⎞<br />

Γ(ρ 11 − 1/2) − γ ρ 2 33 Γ ⊥ ρ 12 γ ⊥ ρ 13<br />

Γ ⊥ ρ 21 Γ(ρ 22 − 1/2) − γ ρ 2 33 γ ⊥ ρ 23<br />

, (2.8)<br />

⎟<br />

⎠<br />

γ ⊥ ρ 31 γ ⊥ ρ 32 γρ 33<br />

where γ ⊥<br />

= (γ + γ ′ + Γ)/2 is the transverse decay rate for the excited state<br />

coherences ρ 23 and ρ 13 . The rate <strong>of</strong> population exchange between ground states<br />

Γ is due to atoms diffusing between the optical region and a reservoir with ρ 11 =<br />

ρ 22 = 1/2 and ρ 21 = 0. The decay rate for the ground state coherence (GSC) ρ 21<br />

is given by Γ ⊥ = (Γ ′ + Γ), where Γ ′ is the dephasing rate for the GSC.


2.2. PERTURBATIVE SOLUTION FOR EIT 29<br />

γ/2<br />

2<br />

−δ R /2<br />

Ω P<br />

∆<br />

Γ<br />

3<br />

γ/2<br />

Ω C<br />

1<br />

δ R<br />

/2<br />

Figure 2.2:<br />

Diagram <strong>of</strong> a three-level Λ-type atom with a strong coupling field and a weak<br />

probe field. The detunings are defined as ∆ = (∆ P + ∆ C )/2, and δ R = ∆ P − ∆ C ,<br />

where ∆ P = (ω 1 − ω 2 ) − ω P and ∆ C = (ω 1 − ω 3 ) − ω C . Also, the Rabi frequencies<br />

are defined as Ω i = µ i E i / where i = {P, C} (signifying probe and coupling fields<br />

respectively), the dipole moments µ i are assumed to be real, and E i is the the<br />

electric field. The spontaneous emission rate out <strong>of</strong> the excited state is given by<br />

γ and the ground state decoherence rate Γ is the same for both ground states.<br />

Written out explicitly the master equations are<br />

˙ρ 33 = −γρ 33 + Im {Ω ∗ 2ρ 32 + Ω ∗ 1ρ 31 } (2.9)<br />

˙ρ 22 = γ (<br />

2 ρ 33 − Γ ρ 22 − 1 )<br />

− Im {Ω ∗<br />

2<br />

2ρ 32 } (2.10)<br />

˙ρ 11 = γ (<br />

2 ρ 33 − Γ ρ 11 − 1 )<br />

− Im {Ω ∗<br />

2<br />

1ρ 31 } (2.11)<br />

˙ρ 32 =<br />

(<br />

i∆(Ω S ) − i δ )<br />

R(Ω S )<br />

− γ ⊥ ρ 32 (2.12)<br />

2<br />

+i Ω C<br />

2 (ρ 22 − ρ 33 ) + i Ω P<br />

2 ρ∗ 21


2.2. PERTURBATIVE SOLUTION FOR EIT 30<br />

˙ρ 31 =<br />

(<br />

i∆(Ω S ) + i δ )<br />

R(Ω S )<br />

− γ ⊥ ρ 31 (2.13)<br />

2<br />

+i Ω P<br />

2 (ρ 11 − ρ 33 ) + i Ω C<br />

2 ρ 21<br />

and<br />

˙ρ 21 = (iδ R (Ω S ) − Γ ⊥ )ρ 21 − i Ω P<br />

2 ρ∗ 32 + i Ω∗ C<br />

2 ρ 31. (2.14)<br />

In its most general form, the steady state analytic solution <strong>of</strong> Eq.<br />

(2.6) is<br />

extremely complicated and provides little insight, which is why approximate solutions<br />

are typically used [6, 95, 96, 124–126]. One <strong>of</strong> the more common techniques<br />

uses perturbation theory and assumes a small probe field (i.e. Ω p ≪ Ω p ) to obtain<br />

a solution that is first order accurate in Ω P and accurate to all orders for all other<br />

parameters [124–126]). This perturbative solution is simple and insightful, but it<br />

breaks down for large single-photon detuning ∆ ≫ γ ⊥ and moderately large probe<br />

fields (i.e. Ω p ∼ Ω c /10). As we will show, these limitation make the perturbative<br />

solution poorly suited for calculations <strong>of</strong> EIT in Doppler broadened medium.<br />

Thus, for Doppler broadened media we use the Bloch-vector solution from Sec.<br />

2.3 instead <strong>of</strong> the perturbative solution.<br />

The steady state solution <strong>of</strong> the master equation is solved by setting all time<br />

derivatives in Eqs. (2.9)-(2.9) equal to zero and simultaneously solving these six<br />

equations constrained by the normalization equation ρ 11 + ρ 22 + ρ 33 = 1. We do


2.2. PERTURBATIVE SOLUTION FOR EIT 31<br />

this by first rearranging the density matrix into a real vector<br />

⃗p = [ρ 11 , ρ 22 , ρ 33 , Re(ρ 21 ), Im(ρ 21 ), . . .<br />

Re(ρ 32 ), Im(ρ 32 ), Re(ρ 31 ), Im(ρ 31 )] T .<br />

The master equation can now be expressed as a simple matrix equation<br />

˙⃗p = (M + Ω P P )⃗p, (2.15)<br />

where M is a matrix containing terms independent <strong>of</strong> Ω p , and the matrix P has<br />

all the terms that are multiplied by Ω p . The the top three rows corresponding to<br />

the ˙ρ 11 , ˙ρ 22 or ˙ρ 33 equations are linearly dependent, so one <strong>of</strong> these rows must be<br />

replaced with the normalization condition ρ 11 + ρ 22 + ρ 33 = 1. The steady state<br />

solution is then<br />

⃗p (ss) = ( ˜M + Ω P ˜P ) −1 ⃗v, (2.16)<br />

where the tildes denote that one row matrix has been replaced by the normalization<br />

condition and ⃗v = [0, 0, 1, 0, 0, 0, 0, 0, 0] T arises from the 1 on the right hand<br />

side <strong>of</strong> the normalization condition. This exact steady state solution ⃗p (ss) is used<br />

for all numerical calculations discussed in this work, but it does not produce an<br />

insightful analytic result.<br />

The perturbative solution is found by defining the steady-state solution as a


2.2. PERTURBATIVE SOLUTION FOR EIT 32<br />

power series in ɛ<br />

∞∑<br />

⃗p (ss) = ɛ n ⃗p n , (2.17)<br />

n=0<br />

where ɛ is a dummy variable to indicate the order <strong>of</strong> the perturbation. Applying<br />

Eq. (2.17), the master equation becomes<br />

( ˜M<br />

∑ ∞<br />

+ ɛΩ P ˜P ) ɛ n ⃗p n = ⃗v. (2.18)<br />

n=0<br />

Multiplying out, collecting terms <strong>of</strong> the same order in ɛ, and setting ɛ = 1, we<br />

find that the perturbation solutions are<br />

⃗p 0 = ˜M −1 ⃗v,<br />

(2.19a)<br />

and<br />

⃗p n = (−Ω P ) n (<br />

˜M<br />

−1 ˜P<br />

) n<br />

⃗p0 . (2.19b)<br />

In most discussions <strong>of</strong> EIT the first order approximation ⃗p (a) = ⃗p 0 + ⃗p 1 is used.<br />

For the case <strong>of</strong> zero decoherence (Γ ⊥ = 0) the perturbative solution gives a<br />

probe susceptibility <strong>of</strong><br />

χ (a)<br />

P (δ R) = Nµ2 p<br />

2ɛ 0<br />

ρ 12<br />

≈ α 0<br />

2δ R<br />

Ω 2 C /γ + 4∆ P δ R /γ − 2iδ R<br />

, (2.20)


2.2. PERTURBATIVE SOLUTION FOR EIT 33<br />

Virtual level<br />

Primary level<br />

Ω C<br />

Ω P<br />

Ω P<br />

Undressed<br />

Picture<br />

Dressed State<br />

Picture<br />

Figure 2.3: Schematic <strong>of</strong> transformation to the dressed state picture in which the coupling field<br />

dresses the atom splitting in the excited energy level into to new energy levels; the primary<br />

energy level and a virtual virtual energy level.<br />

where α 0 = N|µ p | 2 /2ɛ 0 γ is the incoherent absorption coefficient, µ p is the dipole<br />

moment for the probe transition, N is the atomic number density, and we<br />

have assumed γ ′ = 0. Like Sec. 2.1, Eq. (2.20) also predicts an approximately<br />

Lorentzian transparency resonance when the EIT condition is satisfied and near<br />

single-photon resonance ∆ ≪ γ ⊥ .<br />

Near single-photon resonance the dominant feature <strong>of</strong> EIT is a transparency<br />

resonance in an absorbing background. However, far from single-photon resonance<br />

(i.e. ∆ ≫ γ ⊥ ) the dominant feature <strong>of</strong> EIT is an absorbing resonance in a mostly<br />

transparent background. The physics <strong>of</strong> this process is best understood using the<br />

dressed state picture shown in Fig. 2.3. When the atom is dressed by a large<br />

coupling field the excited energy level splits into two energy levels that we call the<br />

primary and virtual energy levels. Figure 2.4 shows the probe absorption due to<br />

transition to the primary and virtual energy levels. The primary energy level can<br />

be thought <strong>of</strong> as the original excited energy level ac-Stark shifted by Ω 2 C /4∆ C,


2.2. PERTURBATIVE SOLUTION FOR EIT 34<br />

and the absorption due to the primary transition can be approximated by<br />

L 1 (δ R ) =<br />

α 0 Lγ 2 ˜ρ 2 gg<br />

(δ R + ∆ C + Ω 2 C /4∆ . (2.21)<br />

C) 2 + γ 2 ˜ρ 2 gg<br />

The virtual energy level is near Raman resonance δ R = 0, and the absorption due<br />

to the virtual transition is approximately given by<br />

L 2 (δ R ) =<br />

α 0 Lγ 2 ˜ρ 2 ee<br />

(δ R − Ω 2 C /4∆ , (2.22)<br />

C) 2 + γ 2 ˜ρ 2 ee<br />

where ˜ρ ee = Ω 2 C /4∆2 C and ˜ρ gg = 1− ˜ρ ee . EIT results from the interference between<br />

these two transitions. The total absorption is approximately<br />

( [ ( )] ) √L1<br />

L T otal = (δ R ) + exp i cot −1 δR − δ √L2 2<br />

vc<br />

(δ R ) , (2.23)<br />

δ vc /4<br />

where the phase term mimics the interference responsible for EIT, and δ vc =<br />

Ω 2 C /4∆ C is the Raman detuning at the center <strong>of</strong> the virtual transition. Figure 2.4<br />

(a) shows the agreement between Eq. (2.22) and the exact steady state solution<br />

<strong>of</strong> Eq. (2.6).<br />

Several <strong>of</strong> the features exhibited in Fig. 2.4 are important for systems with<br />

multiple velocity classes such as those discussed in Secs. 2.5 and 2.6. First,<br />

the Raman detuning corresponding to the center <strong>of</strong> the virtual transition (i.e.<br />

δ vc ≈ Ω 2 /4∆) moves closer to Raman resonance for larger single-photon detunings.


2.2. PERTURBATIVE SOLUTION FOR EIT 35<br />

1<br />

0.8<br />

(a)<br />

L 2 Eq. (8)<br />

EIT exact<br />

Ω C<br />

=0<br />

Virtual<br />

Transition<br />

x10<br />

(b)<br />

2<br />

c<br />

Ω /4∆<br />

c<br />

Probe Absorption<br />

(α L=1) 0<br />

0.6<br />

0.4<br />

Primary<br />

Transition<br />

Virtual<br />

Transition<br />

Absorption<br />

FWHM<br />

2 2<br />

γ Ω<br />

c<br />

/2∆ c<br />

0.2<br />

0<br />

-50 -40 -30 -20 -10 0 10<br />

Raman Detuning (GHz)<br />

-2 -1 0 1 2<br />

Raman Detuning (GHz)<br />

Figure 2.4: EIT in the far detuned case. Grey solid line shows the probe absorption for EIT<br />

when γ = 2π × 6 MHz and ∆ C = 5γ. The dashed grey curve shows the probe absorption in<br />

the absence <strong>of</strong> a coupling field, and the black solid line shows the probe absorption for a virtual<br />

transition as calculated in Eq. (2.22).<br />

For EIT in Doppler broadened systems there will typically be a few velocity classes<br />

near single-photon resonance, but most velocity classes will have large singlephoton<br />

detunings. The absorption due to the far detuned velocity classes occur at<br />

Raman detunings in the transparency window <strong>of</strong> the near resonant velocity classes<br />

(i.e. resonant velocity classes have a transparency window for Raman detunings<br />

|δ R | ≤ Ω 2 /γ while the far detuned velocity classes have peak absorption at Raman<br />

detunings δ R = δ vc ≈ Ω 2 /4∆ ≪ Ω 2 /γ.) Thus, the EIT linewidths in Doppler<br />

broadened media are much narrower than single velocity class EIT systems.<br />

Figure 2.5 shows how the superposition <strong>of</strong> absorption peaks from different<br />

velocity classes combine to create a narrower EIT resonance. When ∆ ≫ γ ⊥ , the<br />

absorption resonances become sufficiently narrow that they can be approximated


2.2. PERTURBATIVE SOLUTION FOR EIT 36<br />

Probe Absorption (arb. units)<br />

5<br />

4<br />

3<br />

2<br />

1<br />

v=±v p<br />

v= - v p/6<br />

v= ± vp/3<br />

v= 0<br />

Doppler<br />

averaged<br />

EIT line<br />

Single<br />

velocity<br />

EIT line<br />

0<br />

-15 -10 -5 0 5<br />

Raman detuning (kHz)<br />

Figure 2.5: Probe absorption versus Raman detuning for a Doppler averaged EIT line shape<br />

and the probe absorption for several different velocity classes which contribute to the Doppler<br />

averaging. The parameters used to calculate these absorption curves are Γ = 0, γ =≈ 2π ×<br />

6 MHz, v p = 50cγ/ω 0 = 222 m/sec.<br />

by Dirac delta functions–i.e.<br />

L 2 (δ R ) ≈ πα 0LΩ 2 γ ⊥<br />

δ(δ<br />

4∆ 2 R − Ω 2 /4∆)<br />

≈ πα 0 Lγ ⊥ δ(∆ − Ω 2 /4δ R ), for ∆ ≫ γ ⊥ . (2.24)<br />

We will revisit this expression in Sec. 2.5 when we discuss Doppler broadened<br />

media.<br />

As mentioned previously the perturbative solution has limited utility (i.e. it<br />

is not valid when the probe field approaches the same order <strong>of</strong> magnitude as the<br />

coupling field and the single-photon detuning is large ∆ ≫ γ ⊥ ). For large single-


2.2. PERTURBATIVE SOLUTION FOR EIT 37<br />

photon detunings and “large” probe fields the virtual transition is still valid if Eq.<br />

(2.22) is modified to obtain<br />

L 2 (δ R ) =<br />

≈<br />

α 0 Lγ 2 ˜ρ 2 ee<br />

(δ R − FΩ 2 C /4∆ C) 2 + γ 2 ˜ρ 2 eeF , 2<br />

πα 0Lγ ⊥<br />

δ(∆ − Ω 2 /4δ R ), for ∆ ≫ γ ⊥ , (2.25)<br />

F<br />

where F = √ 1 + |Ω P | 2 |Ω C | 2 ∆ 2 /γ 2 ⊥ Ω4 . Finally, in both Eqs. (2.22) and(2.25)<br />

we have assumed no decoherence.<br />

The effects <strong>of</strong> nonzero decoherence can be<br />

approximated by a second ad-hoc change such that<br />

L 2 (Γ ≠ 0) ≈ 2R2 L 2 (Γ = 0)<br />

( √ 2R + Γ) 2 , (2.26)<br />

where we recall that R = γΩ 2 /4(∆ 2 γ/γ ⊥ + γγ ⊥ ).<br />

An error bound for the perturbative solution ⃗p (a) can be calculated by considering<br />

the convergence rate <strong>of</strong> the perturbation power series from Eq. (2.17). From<br />

Eq. (2.19b) we have the inequality<br />

||⃗p n || ≤ B n ||⃗p 0 ||, for n > 0, (2.27)<br />

where B ≡ |Ω P | max |λ j | and max |λ j | is the maximum magnitude for the eigen-


2.2. PERTURBATIVE SOLUTION FOR EIT 38<br />

values <strong>of</strong> matrix product ˜M −1 ˜P . Thus, the upper bound for the error is<br />

‖⃗p SS − ⃗p (a) ‖<br />

‖⃗p SS ‖<br />

≤ B 2 , if B < 1. (2.28)<br />

Through numerical experimentation we found that for a given probe detuning<br />

∆ p the maximum value for B is given by<br />

⎧<br />

∣<br />

⎪⎨<br />

∣ Ω P ∣∣<br />

∆P<br />

Ω C<br />

≪ γ ⊥<br />

B ≈<br />

∣ , (2.29)<br />

⎪⎩<br />

4Ω<br />

∣ P Ω C ∆ P ∣∣<br />

∆P ≫ γ<br />

Ω 2 C γ+8√ 2∆ 2 P Γ ⊥<br />

where ∆ C = ∆ P − Ω 2 C /4∆ P for ∆ p > γ. There are a couple features <strong>of</strong> Eq. (2.28)<br />

worthy <strong>of</strong> mention. First, Ω p ≪ Ω c is typically given as the validity condition<br />

for the perturbative solution, which is correct near resonance. However, far from<br />

resonance B is large when Ω p ≪ Ω c and Γ ≈ 0. This fact makes the perturbative<br />

solution poorly suited to Doppler broadened media where the far detuned velocity<br />

classes play the dominant role in determining the EIT line shape and linewidth.<br />

In Doppler broadened media, the error bound is determined by the velocity<br />

class with the largest value <strong>of</strong> B (i.e. B max ≡ max(B)). Figure 2.6 shows how the<br />

ratio Ω p /B max varies as a function <strong>of</strong> the coupling amplitude for a medium with<br />

80 velocity classes. Two different decoherence rates are considered.


2.3. BLOCH-VECTOR SOLUTION FOR EIT 39<br />

0.03<br />

Large Probe<br />

0.02<br />

______<br />

Ω P Ω C<br />

FWHM≈<br />

3.5 γ<br />

=250 Hz Γ ⊥<br />

ΩP<br />

max<br />

0.01<br />

FWHM<br />

≈Ω(Γ /γ) 1/2<br />

Γ ⊥<br />

=0<br />

Weak Probe<br />

0<br />

0 0.5 1 1.5 2<br />

Ω C<br />

γ<br />

/<br />

FWHM ≈<br />

Ω 2<br />

____<br />

N v<br />

γ<br />

Figure 2.6: Limiting value <strong>of</strong> the Ω P /B max versus coupling Rabi frequency for calculating<br />

accuracy limits using Eq. 2.29. The limits <strong>of</strong> the Doppler integration are ±40γc/ω ◦ , i.e.<br />

∆ = {−40γ, 40γ}.<br />

2.3 Bloch-vector solution for EIT<br />

The Bloch-vector method for calculating EIT susceptibilities has several advantages<br />

over the perturbative calculation.<br />

First, Bloch-vector solution makes no<br />

assumptions about the relative strengths <strong>of</strong> the probe and coupling field. Second,<br />

it becomes more accurate for large single-photon detunings.<br />

Third, it is<br />

straightforward to use the Bloch-vector formalism to derive transient solutions for<br />

EIT.<br />

The Bloch-vector method uses the fact that in most experimentally relevant<br />

EIT scenarios, the excited state coherences ρ 23 and ρ 13 can be adiabatically eliminated,<br />

and the master equation can be reduced to a Bloch-vector equation. This


2.3. BLOCH-VECTOR SOLUTION FOR EIT 40<br />

approach to solving EIT problems was pioneered by Bennink [127]. The justification<br />

for adiabatically eliminating the excited state coherences is strengthened<br />

significantly by large single-photon detunings making it ideal for calculations in<br />

Doppler broadened media.<br />

The excited state coherences are adiabatically eliminated by setting the time<br />

derivatives in Eqs. (2.12) and (2.13) equal to zero and solving to obtain<br />

Ω P2<br />

(ρ 22 − ρ 33 ) + Ω C<br />

ρ 32 ≈ −<br />

2<br />

ρ ∗ 21<br />

(2.30)<br />

∆ + iΓ ⊥<br />

and<br />

Ω C2<br />

(ρ 11 − ρ 33 ) + Ω P<br />

ρ 31 ≈ −<br />

2<br />

ρ 21<br />

. (2.31)<br />

∆ + iΓ ⊥<br />

The density matrix can be rewritten as a Bloch-vector equation for the two ground<br />

states<br />

d<br />

{<br />

dt ⃗ρ = (R + Γ) −⃗ρ + T ⃗ × ⃗ρ + (1 − 3ρ 33 ) F ⃗ }<br />

, (2.32)<br />

and a first-order differential equation for the excited state population<br />

d<br />

dt ρ 33 = −γρ 11 − (R + Γ)⃗ρ · ⃗F + (1 − 3ρ 33 )R. (2.33)


2.3. BLOCH-VECTOR SOLUTION FOR EIT 41<br />

In the Eqs. (2.32) and (2.33) have used the following definitions<br />

⎛<br />

⎞<br />

⎛<br />

⎞<br />

⃗ρ =<br />

⎜<br />

⎝<br />

u<br />

v<br />

w<br />

≡<br />

⎟ ⎜<br />

⎠ ⎝<br />

2 Re{e −iθ ρ 21 }<br />

2 Im{e −iθ ρ 21 }<br />

ρ 22 − ρ 11<br />

, (2.34)<br />

⎟<br />

⎠<br />

⃗F ≡ R (1 − 3ρ 33)<br />

R + Γ<br />

⎛<br />

⎜<br />

⎝<br />

2e −iθ ρ (−)<br />

21<br />

0<br />

ρ (−)<br />

22 − ρ (−)<br />

11<br />

⎞<br />

, (2.35)<br />

⎟<br />

⎠<br />

⃗T ≡<br />

δ R<br />

R + Γŵ + ∆<br />

F<br />

Γ (1 − 3ρ 33 ) ⃗ , (2.36)<br />

R ≡ Ω2 Γ ⊥ /4<br />

, (2.37)<br />

∆ 2 + Γ 2 ⊥<br />

where ŵ = (0, 0, 1) T is the unit vector in the direction <strong>of</strong> w. The phase θ is defined<br />

as θ ≡ arg(ρ (−)<br />

21 ).<br />

It has been shown in Ref. [127] that the steady state solutions to Eqs. (2.32)<br />

and (2.33) are<br />

and<br />

⃗ρ ss = (1 − 3ρ 33 ) ⃗ F + ⃗ T × ⃗ F + ⃗ T ( ⃗ T · ⃗F )<br />

1 + T 2 (2.38)<br />

ρ ss<br />

33 =<br />

R − (R + Γ) F ⃗ ∣<br />

· ⃗ρ<br />

(<br />

γ + 3 R − (R + Γ) F ⃗ ∣<br />

· ⃗ρ<br />

∣<br />

ρ33 =0<br />

∣<br />

ρ33 =0<br />

) (2.39)<br />

where T = | ⃗ T |.


2.3. BLOCH-VECTOR SOLUTION FOR EIT 42<br />

In anticipation <strong>of</strong> the fact that ρ 33<br />

≈ 0 for the experimental conditions in<br />

chapters 4 and 5, we set ρ 33 = 0 throughout the remainder <strong>of</strong> this work.<br />

Finally, it is straightforward to relate the atomic parameters to the susceptibilities<br />

for the fields. We recall that the probe and coupling susceptibilities are<br />

χ P = N|µ|2<br />

ɛ 0 Ω p<br />

ρ 32 , (2.40)<br />

and<br />

χ C = N|µ|2<br />

ɛ 0 Ω c<br />

ρ 31 . (2.41)<br />

By assuming δ R ≪ γ ⊥ and ρ 22 /ρ 11 = |Ω c /Ω p | 2‡ , we can rewrite the expressions<br />

for the excited state coherences (i.e. Eqs. 2.30 and 2.31 in a more intuitive form;<br />

ρ 32 ≈<br />

(<br />

ρ 22 Ω P 1 − ρ 21 /ρ (−)<br />

21<br />

2(∆ − δ R /2 + iΓ ⊥ )<br />

) ∗<br />

(2.42)<br />

and<br />

ρ 31 ≈<br />

(<br />

)<br />

ρ 11 Ω C 1 − ρ 21 /ρ (−)<br />

21<br />

2(∆ − δ R /2 + iΓ ⊥ ) . (2.43)<br />

The crucial term in Eqs.<br />

2.42 and 2.42 is the ground state coherence (GSC)<br />

ρ 21 . In future sections it becomes evident that EIT, the EIT <strong>Kerr</strong> effect and its<br />

dynamics are all tied to ρ 21 and its evolution. The central role <strong>of</strong> the GSC is<br />

‡ The first assumption coupled with the EIT condition implies the second assumption. Recall<br />

that both coherent and incoherent optical pumping achieve the same ratio between ground state<br />

populations. Also, these assumptions are well justified for the experimental parameters used in<br />

this work.


2.4. TRANSIENT SOLUTION FOR EIT 43<br />

partially revealed by plugging Eq. (2.42) into Eq. (2.40) to obtain<br />

χ p (ρ 21 ) ≈ Nρ 22|µ| 2<br />

2ɛ ◦ <br />

(<br />

1 −<br />

) ∗<br />

ρ 21 /ρ (−)<br />

21<br />

. (2.44)<br />

∆ − δ R /2 + iγ ⊥<br />

2.4 Transient Solution for EIT<br />

The time dependent Bloch-vector equation in Eq. (2.32) can be decoupled into<br />

three independent ordinary differential equations:<br />

⎧<br />

( d<br />

2<br />

dτ + 2 d ) ⎪⎨<br />

2 dτ + 1 + T 2<br />

⎪⎩<br />

u ′ − u ′ f<br />

v ′ − v ′ f<br />

⎫<br />

⎪⎬<br />

= 0 (2.45)<br />

⎪⎭<br />

and<br />

( ) d (w<br />

dτ + 1 ) ′ − w f<br />

′ = 0, (2.46)<br />

where τ = (R + Γ)t and we have used the transformation <strong>of</strong> variables<br />

⎛<br />

⎜<br />

⎝<br />

⎞<br />

u ′<br />

v ′<br />

= 1<br />

⎟ T<br />

⎠<br />

w ′<br />

⎛<br />

⎜<br />

⎝<br />

⎞ ⎛<br />

T w 0 −T u<br />

0 T 0<br />

⎟ ⎜<br />

⎠ ⎝<br />

T u 0 T w<br />

u<br />

v<br />

w<br />

⎞<br />

. (2.47)<br />

⎟<br />

⎠<br />

The initial conditions are given by ⃗ρ(t = 0) = ⃗ρ i ,<br />

˙u ′ (0) = −T v ′ i − u ′ i + (1 + T 2 )u ′ f, (2.48)


2.4. TRANSIENT SOLUTION FOR EIT 44<br />

and<br />

˙v ′ (0) = T u ′ i − v ′ i. (2.49)<br />

The subscript f denotes the final steady state values which are derived from the<br />

steady state solution Eq. (2.38) with ρ 33 = 0 (we are assume that for t ≥ 0 all<br />

parameters are constant.).<br />

The solutions to Eqs. (2.45) and (2.46) are<br />

u ′ (τ) − u ′ f = Ae −τ cos(T τ + φ), (2.50)<br />

v ′ (τ) − v ′ f = Ae −τ sin(T τ + φ), (2.51)<br />

and<br />

w ′ (τ) − w ′ f = ( w ′ i − w ′ f)<br />

e −τ , (2.52)<br />

where<br />

tan φ = v′ f − v′ i<br />

u ′ f − , and A = u′ i − u ′ f<br />

u′ i<br />

cos φ<br />

= v′ i − v ′ f<br />

sin φ . (2.53)<br />

When u, v, and w are plotted parametrically as a function <strong>of</strong> time, these equations<br />

describe an exponentially decaying spiral.<br />

For future discussions <strong>of</strong> the field susceptibilities it is convenient to transform<br />

from a Bloch-vector back to a density matrix. This is simplest if ∆ = 0, in which


2.4. TRANSIENT SOLUTION FOR EIT 45<br />

case<br />

ρ 21 (t) = u+iv = ρ 21,f +(ρ 21 (0)−ρ 21,f ) exp [(iT − 1)(R + Γ)t] , for t ≥ 0; (2.54)<br />

where ρ 21,f = lim t→∞ ρ 21 (t) is the final steady state GSC. Also, for future reference<br />

we note that for ∆ = 0 the steady state GSC is given by<br />

ρ (ss)<br />

21 (δ R ) =<br />

ρ (−)<br />

21 R<br />

(R + Γ) − iδ R<br />

, (2.55)<br />

which is the equation for a circle when the real and imaginary parts <strong>of</strong> the GSC<br />

are plotted parametrically as a function <strong>of</strong> the Raman detuning. Finally, we note<br />

that thinking in terms <strong>of</strong> the GSC is an equivalent alternative to discussing EIT in<br />

terms <strong>of</strong> Lorentzian absorption lines with corresponding dispersion lines. Figure<br />

2.7 is a visual representation <strong>of</strong> the mapping between GSC and susceptibilities. We<br />

will change between these two representations depending on which representation<br />

yields greater insight. The dashed line in Fig. 2.7(a) is an example <strong>of</strong> a transient<br />

spiral calculated from Eq. (2.54) for the case when ρ 21 (t = 0) = ρ (ss)<br />

21 (δ R = 0)<br />

and ρ 21,f = ρ (ss)<br />

21 (δ R = 1.2R). The choice <strong>of</strong> other parameters can be found in the<br />

figure caption.


2.5. DOPPLER BROADENING 46<br />

0.5<br />

(a)<br />

Ground State<br />

Coherence<br />

21<br />

/ ρ<br />

(-)) 21<br />

(b)<br />

1<br />

Probe<br />

Susceptibility<br />

0.5 Re(χ ) ∝ -Im (<br />

Im<br />

ρ<br />

( 21 (-)<br />

/ρ 21<br />

)<br />

0<br />

-0.5<br />

ρ (ss) (1.2R)<br />

21<br />

ρ (ss) (0 )<br />

ρ<br />

(ss) (δ<br />

21 R<br />

)<br />

ρ 21<br />

(t)<br />

0 0.5 1<br />

Re<br />

ρ<br />

( 21 (-)<br />

/ρ 21<br />

)<br />

21<br />

ρ<br />

Im(χ ) ∝ Re( 1-<br />

P<br />

0.5<br />

0<br />

-4 -2 0 2 4 -0.5<br />

δ R /2R<br />

0<br />

P )<br />

ρ21<br />

/ρ /<br />

21<br />

(-))<br />

Figure 2.7: (a) Steady state (solid line) and transient (dashed line) ground state coherence are<br />

plotted in the complex plane, while (b) the probe susceptibility is plotted as a function <strong>of</strong> the<br />

Raman detuning δ R . All plots are theory with R ≈ 5Γ, and all variables except the Raman<br />

detuning are held constant. The transient ground state coherence curve (dashed) started with<br />

steady state for δ R = 0, then suddenly the Raman detuning changed to 2R. To show the<br />

mapping between ground state coherence and probe susceptibility four shapes corresponding to<br />

four Raman detunings have been plotted on both the steady-state ground state coherence and<br />

susceptibility curves.<br />

2.5 Doppler broadening<br />

Up to this point our model for EIT has only three energy levels, two dipole allowed<br />

transition, and one velocity class. The experimental systems discussed in Chapters<br />

4 and 5 are significantly more complicated than this, but the above theory is still<br />

applicable if we understand how it applies it to the real physical system. § In this<br />

section and the next two sections we generalize the theory to include Doppler<br />

broadening, velocity changing coherence preserving collisions (VCCPC), and the<br />

§ Many researchers work hard to make their physical system similar to the simplified theory.<br />

We have adopted the approach <strong>of</strong> embracing the complexity <strong>of</strong> our experiment and modifying<br />

our theory to accommodate our experiment.


2.5. DOPPLER BROADENING 47<br />

complex structure <strong>of</strong> magnetic sublevels and hyperfine levels encountered in real<br />

atoms.<br />

In hot atomic vapors the velocity <strong>of</strong> the atoms parallel to the propagation <strong>of</strong><br />

the optical fields shifts the atomic frequencies due to the Doppler effect. For a<br />

given velocity class <strong>of</strong> atoms (i.e. the set <strong>of</strong> atoms with velocity v in the range<br />

v 1 < v ≤ v 2 such that ∆vω 0 /c = 2γ ⊥ , where ∆v = v 2 − v 1 .) the detunings become<br />

δ R (v) = δ R (0) + vω 21 /c (2.56)<br />

∆(v) = ∆(0) + vω 0 /c, (2.57)<br />

where ω 0 = ω 3 − (ω 2 + ω 2 )/2, ω 21 = ω 2 − ω 1 , and ω n = E n / with E n being the<br />

energy <strong>of</strong> level n. We have assumed that the EIT fields are perfectly copropagating.<br />

For Doppler broadened media we must average quantities such as the probe<br />

susceptibility over all velocity classes. For example,<br />

χ p (δ R ) =<br />

=<br />

∫ ∞<br />

−∞<br />

∫ ∞<br />

−∞<br />

dvχ P [δ R (v), ∆(v)] exp(−v2 /v 2 p)<br />

v p<br />

√ π<br />

d∆χ P [δ R (∆), ∆] 2 ln 2 exp [−(∆ − ∆ 0) 2 4 ln 2/∆ 2 D ]<br />

∆ D<br />

√ π<br />

, (2.58)<br />

where δ R is shorthand for δ R (v = 0), ∆ 0 = ∆(v = 0), v p = √ 2k B T/m is the most<br />

probable velocity in the Maxwellian velocity distribution, m is the atomic mass,<br />

and ∆ D = 2 √ ln 2v p ω 0 /c is the Doppler broadened FWHM.


2.5. DOPPLER BROADENING 48<br />

For simplicity, we will restrict our discussion to the case when ∆ 0 = 0. As<br />

we began discussing in Sec. 2.2 Doppler broadening results in a narrowing <strong>of</strong> the<br />

EIT resonance because absorption from the far detuned velocity classes fills in the<br />

transparency wings <strong>of</strong> the near resonant velocity classes making the transparency<br />

resonance narrow.<br />

In Doppler broadened media with greater than 20 velocity<br />

classes, the EIT line shape is almost entirely determined by the far detuned velocity<br />

classes. These far detuned velocity classes have narrow absorption spikes<br />

that can be approximated by Dirac delta functions [see Eqs. (2.24) and (2.25)].<br />

In order to obtain the most accurate for EIT line shapes in Doppler broadened<br />

media we use the Bloch-vector solution. The steady state GSC is given by<br />

ρ 21 = F u<br />

2<br />

{<br />

1 +<br />

[ (<br />

¯δ<br />

i −<br />

1 + T 2<br />

¯δ + F )]}<br />

w∆<br />

. (2.59)<br />

γ<br />

Applying this result to Eqs. (2.40) and (2.42) yields<br />

χ p ≈ − Nµ2 pρ 22<br />

2ɛ 0 <br />

[<br />

Γ⊥ + ¯δR (<br />

1+T −i + ¯δ + 2 Fw ∆/γ ) ] ∗<br />

, (2.60)<br />

(R + Γ ⊥ )(∆ − δ R /2 − iγ ⊥ )<br />

where ¯δ R = δ R /(R + Γ ⊥ ), and we have used the identity F u = 2Rρ (−)<br />

23 /(R + Γ ⊥ )<br />

in the last line. Under the assumptions <strong>of</strong> ω 21 = 0, Γ ⊥ = 0 and F w ≫ F u , we find


2.5. DOPPLER BROADENING 49<br />

that the imaginary part <strong>of</strong> the susceptibility is<br />

Im(χ p ) ≈<br />

32α 0 δ 2 R γ2 ⊥<br />

( )<br />

∆ 2<br />

+ 1<br />

γ⊥<br />

2<br />

[ ( )<br />

Ω 4 4δR (∆ 2 +γ⊥ 2 2 (<br />

)<br />

Ω 2 γ ⊥<br />

− ∆ γ ⊥<br />

+ Fu 2 ∆2 + 1<br />

γ⊥<br />

2<br />

)], (2.61)<br />

where we have defined α 0 = Nµ 2 /4ɛ 0 γ ⊥ and we have used the definition <strong>of</strong> the<br />

optical pumping rate R = Ω 2 γ ⊥ /4(∆ 2 + γ⊥ 2 ). In the limit <strong>of</strong> large single-photon<br />

detunings Eq. (2.61) describes a “virtual” absorption peak centered at Ω 2 /2∆<br />

with a peak height <strong>of</strong><br />

h = 2Nσ p /(1 + F 2 u∆ 2 /γ 2 ⊥)<br />

and the FWHM is<br />

w = Ω 2 γ ⊥<br />

√1 + F 2 u∆ 2 /γ 2 ⊥ /4∆2 .<br />

This absorption peak has an approximately Lorentzian line shape, and its area is<br />

given by<br />

A(∆) =<br />

≈<br />

∫ 2Ω 2 /∆<br />

−2Ω 2 /∆<br />

dδ R Im[χ p (δ R )] ≈ π 2 h w (2.62)<br />

πα 0 Ω 2 γ ⊥<br />

4∆ 2√ 1 + F 2 u∆ 2 /γ 2 ⊥<br />

.


2.5. DOPPLER BROADENING 50<br />

Thus, the Doppler averaged absorption line is approximately<br />

Im(χ p )<br />

≈<br />

≈<br />

∫ ∞<br />

(<br />

A(∆) exp<br />

− ∆2 4 ln 2<br />

∆ 2 D<br />

) (<br />

δ<br />

δ R − Ω2<br />

4∆<br />

d∆<br />

√<br />

−∞<br />

∆ D π/ ln 2/2<br />

√ ( )<br />

2Nσ p γ ⊥ π ln 2 exp<br />

ln 2Ω4<br />

−<br />

4δR 2 ∆2 D<br />

√ . (2.63)<br />

∆ D 1 + F<br />

2<br />

u Ω 4 /16δR 2 γ2 ⊥<br />

)<br />

This expression for the Doppler averaged line shape is valid for both the largeand<br />

weak-probe limits. In the weak-probe limit the absorption area in Eq. (2.62)<br />

is A(∆) ≈ πα 0 Ω 2 γ ⊥ /4∆ 2 and the line shape simplifies to<br />

Im(χ p ) ≈ 2Nσ pγ ⊥<br />

√<br />

π ln 2<br />

∆ D<br />

)<br />

ln 2Ω4<br />

exp<br />

(− , (2.64)<br />

4δR 2 ∆2 D<br />

and the EIT linewidth (FWHM) is Ω 2 /∆ D . In the weak-probe limit the effects<br />

<strong>of</strong> non-zero decoherence (i.e. Γ ⊥ ≠ 0) can be accounted with the substitution<br />

∆ D → ˜∆, where<br />

√<br />

1<br />

˜∆ =<br />

1<br />

∆ 2 D<br />

+ Γ√ 2<br />

Ω 2 γ . (2.65)<br />

This substitution becomes necessary when not all velocity classes satisfy the EIT<br />

condition creating a coherence cut<strong>of</strong>f such that far detuned velocity classes with<br />

∆ > ˜∆ do not contribute to the line shape <strong>of</strong> EIT.<br />

In the large-probe limit the absorption area in Eq.<br />

(2.62) is A(∆) ≈


2.5. DOPPLER BROADENING 51<br />

L / )<br />

Absorption ( α max(α L)<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

U-shape<br />

0.2<br />

weak probe<br />

V-shape<br />

large probe<br />

0<br />

1.5 1 0.5 0 0.5 1 1.5<br />

FWHM )<br />

Raman Detuning ( δ R/<br />

Lorentzian<br />

single velo.<br />

Figure 2.8: Absorption line shapes for the probe field in different regimes <strong>of</strong> EIT. The solidblack<br />

curve is a perfect Lorentzian corresponding to a single velocity class. The solid-grey and<br />

dashed-black curves show respectively the large-probe and weak-probe limits for the case <strong>of</strong><br />

degenerate ground states (i.e. ω 21 = 0) and Doppler broadening with N v = 40. All curves<br />

assume single-photon resonance (i.e. ∆ 0 = 0) and Γ ⊥ = 0.<br />

πα 0 Ω 2 γ 2 ⊥ /4F u∆ 3 , and the line shape becomes<br />

Im(χ p ) ≈<br />

2Nσ p γ ⊥<br />

√<br />

π ln 2<br />

∆ D<br />

√<br />

1 + F<br />

2<br />

u Ω 4 /16δ 2 R γ2 ⊥<br />

. (2.66)<br />

√<br />

with a linewidth (FWHM) <strong>of</strong> Ω P Ω C /γ ⊥ 3.<br />

Figure 2.8 shows the probe absorption for EIT under three different sets <strong>of</strong><br />

conditions: a single velocity class and multiple velocity classes in the weak- and<br />

large-probe limits. All three transparency lines are normalized to have the same<br />

FWHM in order to compare the line shapes. The Lorentzian shape for a single<br />

velocity class provides a reference for U and V shaped lines corresponding to the<br />

weak- and large-probe limits with multiple velocity classes.


2.6. COHERENCE PRESERVING COLLISIONS IN<br />

BUFFER GASSES AND WALL COATINGS 52<br />

2.6 Coherence preserving collisions in<br />

buffer gasses and wall coatings<br />

One <strong>of</strong> the primary obstacles in EIT related experiments is reducing the decoherence<br />

to tolerable levels. In hot atomic vapors this means reducing the transit-time<br />

decoherence. For example, room temperature rubidium atoms have a velocity <strong>of</strong><br />

about v p − 250 m/s, resulting in a transit-time decoherence rate <strong>of</strong> 100 kHz for<br />

a beam diameter <strong>of</strong> 2.5 mm. Typically, either a buffer gas or a wall coating is<br />

used to keep atoms in the optical region without losing coherence. EIT atoms like<br />

rubidium can undergo thousands <strong>of</strong> velocity changing collisions with other buffer<br />

gas atoms or coated walls before they lose their GSC that is crucial for EIT.<br />

These velocity changing coherence preserving collisions (VCCPC) have the<br />

secondary effect <strong>of</strong> causing the GSC created in the near resonant velocity classes<br />

(i.e. the optical pumping rate R is largest near resonance) to diffuse to the far<br />

detuned velocity classes. If the collision rate R c is fast enough that on average<br />

each atom experiences all <strong>of</strong> the velocity classes in the coherence time τ c = 1/2Γ ⊥<br />

(i.e. R c ≥ N v /τ c , where N v = ∆ D /2γ ⊥ is the number <strong>of</strong> velocity classes), then<br />

each atom will see the same average optical pumping rate<br />

R vccpc = 2√ ln 2<br />

∆ D<br />

√ π<br />

∫ ∞<br />

−∞<br />

d∆ R(∆) exp(−∆ 2 4 ln 2/∆ 2 D)<br />

≈ R(0)/N v (2.67)


2.6. COHERENCE PRESERVING COLLISIONS IN<br />

BUFFER GASSES AND WALL COATINGS 53<br />

regardless <strong>of</strong> its current velocity class. This leads to a uniform distribution <strong>of</strong><br />

GSC, and several interesting effects for the EIT line shape. When the condition<br />

R c ≥ N v /τ c is satisfied we call the system a VCCPC system.<br />

To understand the implications <strong>of</strong> a uniform ground state coherence on the EIT<br />

line shape we use the Bloch-vector formalism <strong>of</strong> Sec. 2.3 with the assumptions<br />

Γ = 0 and Ω C ≥ 5Ω P and ∆(v = 0) = 0. Using Eq. (2.30) the GSC for all<br />

velocity classes can be expressed as<br />

ρ 21 =<br />

ρ (−)<br />

21<br />

1 − iδ R /R vccpc<br />

. (2.68)<br />

The probe susceptibility is found by plugging Eq. (2.68) into Eq. (2.44) to obtain<br />

(<br />

) ∗<br />

χ p (∆) ≈ − Nµ2 p<br />

−iδ R<br />

, (2.69)<br />

2ɛ 0 (R vccpc − iδ R )(∆ − iγ ⊥ )<br />

where we have assumed δ R ≪ γ ⊥ . Finally, integrating over all velocity classes the<br />

Doppler averaged probe susceptibility is<br />

δ R<br />

χ p ≈ −A<br />

, (2.70)<br />

R vccpc + iδ R


2.6. COHERENCE PRESERVING COLLISIONS IN<br />

BUFFER GASSES AND WALL COATINGS 54<br />

where A is a constant<br />

A = Nγ √<br />

⊥µ 2 p ln 2/π<br />

2ɛ 0 ∆ D<br />

≈<br />

∫ ∞<br />

0<br />

d∆<br />

(<br />

exp<br />

− ∆2 4 ln 2<br />

∆ 2 D<br />

∆ 2 + γ 2 ⊥<br />

Nµ2 p 1<br />

. (2.71)<br />

4ɛ 0 γ ⊥ N v<br />

)<br />

In contrast to EIT in non-VCCPC Doppler broadened media, the farthest detuned<br />

velocity classes contribute least to the EIT line shape because for far detuned<br />

velocity classes (i.e. for |v| ≫ cγ ⊥ ω 0 ) the probe susceptibility has the symmetry<br />

χ p (v) ≈ −χ p (−v) such that χ p (v) ≈ +χ p (−v) ≈ 0.<br />

Equation (2.70) is simply a Lorentzian shaped transparency resonance with<br />

FWHM <strong>of</strong> 2R ave . It is as though the blurring <strong>of</strong> velocity classes has made Doppler<br />

broadened line into a single homogeneous line with linewidth 2N v γ ⊥<br />

= ∆ D<br />

(FWHM). Empirical studies show that VCCPC systems are effectively approximated<br />

as a single homogeneous line for conditions deviating from the case we<br />

have considered, such as when the decoherence is non-zero (Γ ≠ 0), the system<br />

is away from resonance (∆(v = 0) ≠ 0), and the ground energy levels are not<br />

degenerate (ω 21 ≠ 0).<br />

Figure 2.9 shows how the linewidth and line shape change as a function <strong>of</strong> the<br />

collision rate. The black curve shows the FWHM <strong>of</strong> the transparency resonance<br />

and the grey curve shows the root mean square (rms) difference between the<br />

Approximating a Doppler broadened VCCPC system as a single homogeneous absorption<br />

line over looks the fact that Doppler-broadened lines is gaussian and not Lorentzian, but discrepancies<br />

resulting from this difference are typically small.


2.6. COHERENCE PRESERVING COLLISIONS IN<br />

BUFFER GASSES AND WALL COATINGS 55<br />

20<br />

0.1<br />

FWHM (kHz)<br />

16<br />

12<br />

Non-VCCPC<br />

8<br />

VCCPC<br />

4<br />

0<br />

0.1 1 10 10 10 10<br />

Velocity Changing Collision Rate (kHz)<br />

0.08<br />

0.06<br />

0.04<br />

2 3 4 0 0.02<br />

RMS difference from Lorentzian<br />

Figure 2.9: EIT linewidth (black) versus the rate <strong>of</strong> velocity-changing, coherence-preserving<br />

collisions (VCCPC). The grey curve shows the RMS difference between the EIT line shape and<br />

a Lorentzian line shape with the same amplitude and FWHM.<br />

numerical simulation <strong>of</strong> a system with velocity changing collisions and a Lorentzian<br />

transparency window <strong>of</strong> the same width and depth. The parameters chosen for<br />

this simulation are in the weak-probe limit. In the non-VCCPC region (i.e. the<br />

VCCPC condition R c ≥ N v /τ c is far from being satisfied), the system is a typical<br />

Doppler broadened system with linewidth Ω 2 / ˜∆. Once the VCCPC condition is<br />

satisfied there is no longer a coherence cut<strong>of</strong>f and the EIT linewidth decreases to<br />

Ω 2 /∆ D . Also, in the VCCPC region the line shape is essentially Lorentzian for<br />

collision rates satisfying the VCCPC condition.


2.7. MAGNETIC SUBLEVELS AND HYPER-FINE LEVELS 56<br />

2.7 Magnetic sublevels and hyper-fine levels<br />

In Chapters 4 and 5, the experiments use the D 1 line <strong>of</strong> rubidium which has many<br />

more electronic states than the three-level EIT systems considered up to this<br />

point. Both Rb 85 and Rb 87 isotopes are used in the experiments but only Rb 85 is<br />

considered here (Rb 87 is discussed briefly in the Appendix). The D 1 line <strong>of</strong> Rb 85<br />

has four hyperfine, 24 magnetic sublevels and 66 dipole allowed transitions making<br />

it substantially more complicated than the three-level EIT system we previoiusly<br />

considered. However, the essence <strong>of</strong> EIT in this system is the same as the threelevel<br />

Λ system (i.e. optical pumping and decoherence between bright and dark<br />

states). The theory developed for three-level systems is applicable to the D 1 line<br />

<strong>of</strong> rubidium if averaged quantities are used (e.g. the average optical pumping rate<br />

R ave and an effective homogeneous linewidth γ eff ; both are derived shortly).<br />

In anticipation <strong>of</strong> the experiments <strong>of</strong> Chapters 4 and 5, we consider orthogonal<br />

linear polarization for the probe and coupling fields ‖ . For the choice <strong>of</strong> quantization<br />

axis parallel to the coupling beam polarization, Fig. 2.10 shows the level<br />

diagram for Rb 85 with the coupling coefficients for the probe and coupling transitions.<br />

There are other dipole allowed transitions not shown in Fig. 2.10. The<br />

‖ There are several reason why the probe and coupling fields have orthogonal linear polarization,<br />

and all <strong>of</strong> them are important for achieving EIT with good contrast. First, choosing<br />

the the fields to have circular polarizations results in there being non-coherent dark state (i.e.<br />

dark states which are individual magnetic sublevels not coherent superpositions <strong>of</strong> sublevels for<br />

an appropriate choice <strong>of</strong> quantization axis). These non-coherent dark-states make the medium<br />

transparent regardless <strong>of</strong> the Raman detuning, and destroying the EIT contrast. Second, for<br />

parallel linear polarizations there are no dark states that are simultaneously dark for transitions<br />

to both excited hyper-fine levels. Finally, experimentally it is convenient to have different<br />

polarization for the two fields so that they can be separated using polarization.


2.7. MAGNETIC SUBLEVELS AND HYPER-FINE LEVELS 57<br />

probe and coupling fields’ interaction with these transitions creates an ac-Stark<br />

shifts the ground hyperfine levels and changes the hyperfine splitting. The energy<br />

shift to the hyperfine splitting is<br />

∆U =<br />

Ω2 p<br />

4∆ hf<br />

+<br />

Ω2 c<br />

4∆ hf<br />

=<br />

Ω2<br />

4∆ hf<br />

, (2.72)<br />

such that the hyperfine splitting becomes ∆ hf (Ω) ≈ ∆ hf (0) + ∆U and for Rb 85<br />

we have ∆ hf (0) = 3.035 GHz. For now, we ignore these transitions, but we will<br />

consider them again in chapters 4 and 5.<br />

Although it is far from obvious by looking at the level diagram in Fig. 2.10,<br />

there is exactly one coherent superposition <strong>of</strong> ground states that is completely<br />

decoupled from both F ′ = 2 and F ′ = 3 excited hyperfine levels. This is the<br />

only truly dark state.<br />

There are also four other partially dark states that are<br />

only weakly coupled to the excited states, and seven other bright states that are<br />

strongly coupled to the excited states. Details about the completely dark state<br />

can be found in the Appendix.<br />

The coupling between the ground and excited states can be calculated by<br />

solving for the eigenvalues <strong>of</strong> the time independent Hamiltonian. The time independent<br />

Hamiltonian is obtained using a rotating reference frame with the rotating<br />

wave approximation, and by only considering those dipole allowed transitions<br />

explicitly shown in Fig. 2.10. The eigenvalues and eigenvectors are evaluated


2.7. MAGNETIC SUBLEVELS AND HYPER-FINE LEVELS 58<br />

hω P<br />

2<br />

27<br />

1<br />

3<br />

1<br />

3<br />

2<br />

27<br />

5<br />

8<br />

1<br />

8<br />

5<br />

27<br />

27<br />

3<br />

27<br />

27<br />

2<br />

−1<br />

−1<br />

2<br />

−<br />

−<br />

27<br />

3<br />

3<br />

27<br />

hω C<br />

m F =-3 m F =-2 m F =-1 m F =0 m F =1 m F =2<br />

m F =3<br />

F -<br />

(a)<br />

F’=3<br />

F’=2<br />

F=3<br />

F=2<br />

hω P<br />

(b)<br />

15<br />

10<br />

6<br />

1<br />

1<br />

27<br />

27<br />

27<br />

3<br />

27<br />

−1<br />

4<br />

−1<br />

1<br />

4<br />

1<br />

−<br />

0<br />

3<br />

27<br />

27<br />

27<br />

27<br />

3<br />

1<br />

27<br />

1<br />

3<br />

6<br />

27<br />

10<br />

27<br />

15<br />

27<br />

m F =-3 m F =-2 m F =-1 m F<br />

=0<br />

m F =1 m F =2<br />

m F =3<br />

F’=3<br />

F’=2<br />

hω C<br />

F=3<br />

F=2<br />

Figure 2.10: Transitions excited by the vertically polarized probe laser (grey lines) and<br />

horizontally polarized coupling laser (black lines). The quantization axis is chosen along<br />

the horizontal polarization. Also shown are the hyperfine matrix elements for each<br />

excited transition, which are related to the Clebsch-Gordon coefficients by the factor<br />

(−1) √ { }<br />

F ′ +J+1+I J J<br />

(2F ′ ′<br />

1<br />

+ 1)(2J + 1)<br />

F ′ , where I = 3/2 is the nuclear spin, J = J<br />

F I<br />

′ =<br />

1/2 are the orbital plus spin angular momenta for the ground and excited state respectively,<br />

and we have use the Wigner 6-j symbol (see Ref. cite for details).


2.7. MAGNETIC SUBLEVELS AND HYPER-FINE LEVELS 59<br />

numerically for the case when δ R = 0 and ∆ ≫ Ω + , where<br />

Ω + = 2µ D1<br />

√E 2 P + E2 C /√ 3, (2.73)<br />

is the bright Rabi frequency for the D 1 line and for rubidium µ D1 = 2.54 ×<br />

10 −29 Cm. The smallest eigenvalues are the Stark shifted energies for the ground<br />

states, and these energy shifts (i.e. ∆U i with i = 1, 2, . . . 12) can be calculated as<br />

a Rabi frequency for each ground state<br />

Ω i ≡ √ −4∆U i ∆, i = {1, 2, . . . , 12}. (2.74)<br />

The degree <strong>of</strong> coupling between ground and excited states is given by the ratio<br />

Ω i /Ω + , such that Ω i /Ω + = 1 is a bright state and Ω i /Ω + = 0 is a completely dark<br />

state.<br />

Figure 2.11 shows the degree <strong>of</strong> coupling for each <strong>of</strong> ground state energy eigen<br />

states plotted as a function <strong>of</strong> the ratio between probe and coupling amplitudes<br />

E P /E C . The experiments discussed in Chapters 4 and 5 lie in the region 0.2 ≥<br />

Ω P /Ω C ≤ 0.6, and the decoherence rate Γ in Chapters 4 and 5 is large enough<br />

that the completely dark state is not appreciably darker than the partially dark<br />

states.<br />

The effective optical pumping rate can be approximated as the average <strong>of</strong> the<br />

optical pumping rates out <strong>of</strong> each bright state and into one <strong>of</strong> the five “dark”


2.7. MAGNETIC SUBLEVELS AND HYPER-FINE LEVELS 60<br />

Rabi Frequency (Ω /Ω )<br />

i +<br />

Bright States<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

Quasi-dark States<br />

0.2<br />

0<br />

Completely Dark State<br />

0 0.2 0.4 0.6 0.8 1<br />

E P<br />

/ E C<br />

Figure 2.11: Normalized Rabi frequencies Ω i /Ω + plotted versus the ratio E P /E C . There are 12<br />

Rabi frequencies; 7 bright states, 4 quasi dark states, and 1 completely dark state.<br />

states. The optical pumping rate out <strong>of</strong> bright state m is given by<br />

R m =<br />

5∑ ∑12<br />

P<br />

n=1<br />

l=1<br />

(|e〉 l<br />

∣ ∣∣|+〉m<br />

)<br />

R |e〉l →|−〉 n<br />

, (2.75)<br />

∣ ) ∣∣|+〉m<br />

where P<br />

(|e〉 l<br />

is the conditional probability that in steady state the atom<br />

will be in the excited state |e〉 l , given that it is in either an excited state or bright<br />

state |+〉 m . R |e〉l →|−〉 n<br />

is the spontaneous decay rate from excited state |e〉 l to dark<br />

state |−〉 n . The effective optical pumping rate is then<br />

R ave = 1 7∑<br />

R m , (2.76)<br />

7<br />

m=1<br />

where we have assumed that all bright states are equally populated (in reality


2.7. MAGNETIC SUBLEVELS AND HYPER-FINE LEVELS 61<br />

0.42<br />

2<br />

B=R ave γ eff /Ω +<br />

0.40<br />

0.38<br />

5/12<br />

0.36<br />

0 0.2 0.4 0.6 0.8 1<br />

E P /EC<br />

Figure 2.12: Plot <strong>of</strong> B = R ave γ eff /Ω 2 + versus the ratio E P /E C . Also plot is the value 5/12=0.417,<br />

which is the ratio <strong>of</strong> the number <strong>of</strong> dark states to ground states. Over the range 0.2 ≥ E P /E C ≤<br />

0.6 B is approximately 3/8.<br />

there are small variations in the relative populations <strong>of</strong> bright states but this is a<br />

small correction).<br />

Figure 2.12 shows the numerical calculation <strong>of</strong> R ave γ eff /Ω 2 + for the D 1 line<br />

<strong>of</strong> Rb 85 as a function <strong>of</strong> E P /E C . Naively one might expect R ave = 5γ eff /12Ω 2 +<br />

because there are five dark states out <strong>of</strong> a total <strong>of</strong> twelve ground states. However,<br />

calculations reveals that the optical pumping rate is slightly smaller than this<br />

because the branching ratios from those excited states coupled to the bright states<br />

are weighted towards spontaneous decay into the bright state. Evaluating Eq.<br />

(2.76) we find that over the range 0.2Ω C ≤ Ω p ≥ 0.6Ω C<br />

R ave ≈ 3Ω2 +<br />

8γ eff<br />

, (2.77)


2.8. MISCELLANEOUS EXPERIMENTAL CONSIDERATIONS 62<br />

where γ eff is the effective homogeneous linewidth that arises due to a combination<br />

<strong>of</strong> Doppler broadening and buffer gas effects.<br />

Finally, we need to calculate the effective homogeneous linewidth γ eff in order<br />

to apply the simple three-level EIT theory to rubidium. It can be calculated using<br />

the definition<br />

1<br />

γ eff<br />

≡<br />

∫<br />

γ ∞<br />

√ ⊥<br />

π/ ln 2∆D<br />

−∞<br />

(<br />

exp<br />

− ∆2 4 ln 2<br />

∆ 2 D<br />

∆ 2 + γ 2 ⊥<br />

)<br />

d∆, (2.78)<br />

where 2γ ⊥ is the homogeneous FWHM, and ∆ D is the inhomogeneous FWHM.<br />

For Rb 85 with a N 2 buffer gas the measured increase in homogeneous linewidth is<br />

16.3 MHz per Torr N 2 [128–130]. In the experiments, the rubidium vapor cells have<br />

20 Torr <strong>of</strong> N 2 gas resulting in a homogeneous linewidth <strong>of</strong> 2γ ⊥ ≈ 330 MHz. The<br />

inhomogeneous broadening is due to Doppler broadening and the Doppler broadened<br />

FWHM is ∆ D = 530 MHz for rubidium vapor at temperature <strong>of</strong> 350 K ∗∗ .<br />

Thus, the effective linewidth is γ eff ≈ 650 MHz.<br />

2.8 Miscellaneous experimental considerations<br />

There are various other considerations which affect real experimental systems.<br />

Some <strong>of</strong> these are four-wave mixing, various forms <strong>of</strong> decoherence, and spatial<br />

varying field amplitudes. We mention these briefly to make the reader aware <strong>of</strong><br />

them, but will not go into detail.<br />

∗∗ For high enough buffer gas densities Doppler broadening can actually be reduced due to<br />

Dicke narrowing [131], but our experiments do not approach these densities.


2.8. MISCELLANEOUS EXPERIMENTAL CONSIDERATIONS 63<br />

5P 3/2<br />

F=3<br />

F=2<br />

Ω C<br />

↔<br />

b<br />

b<br />

Ω<br />

P<br />

(a)<br />

∆ 5P 1/2<br />

Ω 4<br />

hf<br />

W<br />

↔ Ω C<br />

5S 1/2<br />

↔<br />

Ω<br />

Ω C<br />

P<br />

b<br />

F=3<br />

F=2<br />

5P 1/2<br />

(b)<br />

∆ S<br />

↔<br />

b<br />

Ω S<br />

Ω 4W<br />

5S 1/2<br />

Figure 2.13: Level diagrams for four-wave mixing created in our experimental EIT system.<br />

First, in the experiments there are two systems that exhibit four wave mixing,<br />

and these are diagramed in Fig. 2.13. In Fig. 2.13(a) only the probe and coupling<br />

beams are necessary to excite the four-wave mixing process and create a third field<br />

with Rabi frequency Ω 4W and frequency ω 4W = 2ω c − ω p . The four wave mixing<br />

process in Fig. 2.13(b) requires the probe and coupling fields with the signal field<br />

discussed in the next chapter, and it results in a fourth field with Rabi frequency<br />

Ω 4W and frequency ω 4W = ω p + ω s − ω c . In both <strong>of</strong> these cases the Maxwell wave<br />

equation can be written using the slowly varying envelope approximation (SVEA)<br />

as<br />

where in Fig.<br />

( ∂<br />

c∂t + ∂ )<br />

Ω 4W ∝ ρ 21γ ⊥<br />

, (2.79)<br />

∂z ∆ i<br />

2.13(a), ∆ i is the hyperfine splitting <strong>of</strong> the ground states, and<br />

in Fig. 2.13(b), ∆ i is the signal detuning. In both cases the intensity <strong>of</strong> the


2.8. MISCELLANEOUS EXPERIMENTAL CONSIDERATIONS 64<br />

four-wave mixing product Ω 4W<br />

grows quadratically with optical depth. In Fig.<br />

2.13(a), the probe field also experiences some gain due to four-wave mixing, and<br />

in Fig. 2.13(b), the probe experiences absorption. The four-wave mixing gain is<br />

proportional to the magnitude <strong>of</strong> the GSC in contract to EIT which is proportional<br />

to Re(1−ρ 21 /ρ (−)<br />

21 ). Thus, four-wave mixing gain has a slightly wider FWHM than<br />

EIT and a non-Lorentzian line shape. Finally, as a function <strong>of</strong> Raman detuning<br />

There are several forms <strong>of</strong> decoherence that must be overcome if one is to obtain<br />

good EIT. Some <strong>of</strong> the sources <strong>of</strong> decoherence are limited dwell time in the beam †† ,<br />

magnetic fields, spin exchange relaxation (i.e. Rb-Rb collision), buffer gas collisions<br />

(i.e. Rb-N 2 collisions), and imperfectly aligned probe and coupling beams ‡‡ .<br />

For the experiments discussed in this work, the dominant form <strong>of</strong> decoherence<br />

is the limited dwell time in the beams. The diffusion coefficient for rubidium in<br />

a nitrogen buffer gas at standard temperature and pressure is D 0 = 0.15 cm 2 /s<br />

[128, 130, 132]. It is possible to calculate that the Rb-N 2 collision cross-section is<br />

σ c ≈ 3 × 10 −15 cm 2 by using the approximate equation for the diffusion coefficient<br />

D ≈<br />

π¯v<br />

8N N σ c<br />

,<br />

where ¯v = √ 8k B T/πm is the mean velocity <strong>of</strong> the atoms, N N<br />

is the number<br />

density <strong>of</strong> nitrogen.<br />

For 20 Torr <strong>of</strong> nitrogen, we see that the mean free path<br />

†† This is the same thing as transit-time decoherence.<br />

‡‡ If the probe and coupling beams are not perfectly copropagating parallel to one another,<br />

then they will see the same atom as having different velocities and different Doppler shifts.


2.8. MISCELLANEOUS EXPERIMENTAL CONSIDERATIONS 65<br />

length is l = 1/N N σ c ≈ 5 µm, and the collision rate is R c = ¯vl ≈ 10γ, where γ =<br />

5.75 MHz is the natural linewidth for the D 1 line <strong>of</strong> rubidium. The diffusion time<br />

for an atom to move from the center to out <strong>of</strong> the optical beam is approximately<br />

t d ≈ r 2 /l 2 R c , where r is the beam radius. For the above parameters and a 2 mm<br />

beam radius the average dwell time in the beam is about 0.5 ms corresponding to<br />

a decoherence rate <strong>of</strong> about 2 kHz. In Chapters 4 and 5, we see this decoherence<br />

rate corresponds well with the measured decoherence rates. The Rb-Rb collision<br />

cross-section is about 2×10 −14 cm 2 .which is larger than the N 2 -Rb collision crosssection.<br />

However, for the rubidium density discussed in this work is sufficiently<br />

small that the decoherence rate due to spin exchange collisions never exceeds the<br />

10’s <strong>of</strong> Hz.<br />

As a consequence <strong>of</strong> the decoherence encountered in experimental systems<br />

we are typically more concerned with the depth <strong>of</strong> the EIT ˜αL = αR/(R + Γ).<br />

Figure 2.14 shows experimental probe absorption data from Chapter 4 with our<br />

definitions for the linewidth FWHM, probe absorption αL = ρ 22 α 0 L, and EIT<br />

depth ˜αL, where α 0 = N|µ p | 2 /2ɛ 0 γ is the probe absorption coefficient if all the<br />

population were in the probe ground state. These definitions are used throughout<br />

the remainder <strong>of</strong> this work.<br />

The optical pumping rate has a spatial dependence due to spatial variations in<br />

the field intensities (i.e. the beams are approximately gaussian). Thus, quantities<br />

that depend on the optical pumping rate such as the rise-times and transparency


2.8. MISCELLANEOUS EXPERIMENTAL CONSIDERATIONS 66<br />

6<br />

Probe absorption (αL)<br />

4<br />

2<br />

0<br />

αL<br />

αL<br />

∼<br />

FWHM<br />

Raman<br />

resonance<br />

-20 -10 0 10 20 30<br />

Raman Detuning (kHz)<br />

Figure 2.14: Visual representation <strong>of</strong> various terms used to describe EIT.<br />

FWHM must be averaged spatially. This is discussed in detail towards the end <strong>of</strong><br />

chapter 4.


67<br />

Chapter 3<br />

Theory <strong>of</strong> the dynamics <strong>of</strong> the<br />

EIT <strong>Kerr</strong> nonlinearity<br />

The EIT <strong>Kerr</strong> nonlinearity arises when a signal field Ω s (see Fig. 3.1) ac-Stark<br />

shifts one or both <strong>of</strong> the EIT ground states perturbing the EIT Λ system away<br />

from Raman resonance. Perturbations <strong>of</strong> the Raman resonance change the index<br />

<strong>of</strong> refraction for both probe field Ω p<br />

and coupling field Ω c , resulting in phase<br />

modulation <strong>of</strong> the fields proportional to the intensity <strong>of</strong> the signal field.<br />

This type <strong>of</strong> <strong>Kerr</strong> nonlinearity has several advantages over more conventional<br />

<strong>Kerr</strong> systems. First, the cross-phase modulation can be large with minimal selfphase<br />

modulation, in contrast most other <strong>Kerr</strong> media which necessarily have large<br />

self-phase modulation.<br />

Second, the cross-phase modulation is one-directional.<br />

The signal intensity creates a phase modulation on the probe, but the probe does<br />

not phase modulate the signal. Third, for cw fields the size <strong>of</strong> the cross-phase<br />

modulation is tunable by changing the slope <strong>of</strong> the EIT dispersion. This slope<br />

can be increased by either making the EIT linewidth narrower or increasing the


68<br />

γ<br />

2<br />

3<br />

∆<br />

γ<br />

2<br />

4<br />

∆ S<br />

γ<br />

Ω C<br />

Ω S<br />

2<br />

Ω P<br />

δ R<br />

2<br />

Γ<br />

δ R<br />

2<br />

1<br />

Figure 3.1: Simple EIT <strong>Kerr</strong> system in a four-level N-type system. The detunings are defined<br />

as ∆ P = (ω 3 − ω 2 ) − ω P , ∆ C = (ω 3 − ω 1 ) − ω C , ∆ S = (ω 4 − ω 2 ) − ω S , ∆ = (∆ P + ∆ C )/2, and<br />

δ R = ∆ C − ∆ P . Also, the Rabi frequencies are defined as Ω i = µ i E i / where i = {P, C, S}, µ i<br />

is the dipole moment, and E i is the the electric field. The spontaneous emission rate out <strong>of</strong> the<br />

excited state is given by γ and Γ is the decoherence rate for the ground states.<br />

optical depth <strong>of</strong> the EIT medium (e.g. by making the medium longer or more<br />

dense). Theoretically the EIT <strong>Kerr</strong> effect can be arbitrarily large for cw fields.<br />

Large optical nonlinearities are interesting because they open the door to quantum<br />

nonlinear optics (i.e. nonlinear optics with wavepackets at the few-photon<br />

level) and other low-light-level applications. Many <strong>of</strong> the most intriguing applications<br />

require a strong nonlinear response to pulses with small energies. This<br />

implies that the nonlinearities must be both fast and large. In these applications<br />

the correct figure <strong>of</strong> merit is not the size <strong>of</strong> the nonlinear susceptibility, but the<br />

ratio between the size and rise-time <strong>of</strong> the nonlinear susceptibility.<br />

The EIT <strong>Kerr</strong> effect is slow compared to parametric nonlinear processes. Also,<br />

the rise-time and size <strong>of</strong> the EIT <strong>Kerr</strong> susceptibility are directly proportional


3.1. THE EIT KERR NONLINEARITY 69<br />

making their ratio a constant. Thus, increasing the size <strong>of</strong> the cw EIT <strong>Kerr</strong> effect<br />

does not increase the figure <strong>of</strong> merit for pulsed EIT <strong>Kerr</strong> effects.<br />

This chapter is divided into two sections. First, in Sec. 3.1 we present the<br />

steady state and transient theory for the EIT <strong>Kerr</strong> nonlinearity. In Sec. 3.2, we<br />

discuss the application <strong>of</strong> <strong>Kerr</strong> nonlinearities to QND measurements and derive<br />

the signal to noise ratio for these measurements. In this chapter we only consider<br />

EIT <strong>Kerr</strong> QND measurements when the signal group velocity is approximately<br />

c the speed <strong>of</strong> light in vacuum. However, it is possible to generalize the theory<br />

to include arbitrary group velocities for the signal, which is done in Sec. 5.1 <strong>of</strong><br />

Chapter 5.<br />

3.1 The EIT <strong>Kerr</strong> nonlinearity<br />

Figure 3.1 shows a typical system for refractive EIT <strong>Kerr</strong> nonlinearities. Various<br />

parameters such as the Raman detuning δ R , single-photon detuning ∆ and Rabi<br />

frequencies Ω i where i = {P, C, S} are defined in the figure caption. The EIT <strong>Kerr</strong><br />

system consists <strong>of</strong> a Λ EIT system ∗ discussed in chapter 2 and a signal field Ω s<br />

which is detuned by ∆ s from the |1〉 → |4〉 transition. The signal field Stark shifts<br />

ground state |1〉 by Ω 2 S /4∆ S, which changes the Raman detuning and ground state<br />

to<br />

δ R (Ω S ) = δ R (0) + |Ω S | 2 /4∆ S . (3.1)<br />

∗ Figure 3.1 uses the same definitions for fields and detunings as Fig. 2.2 from chapter 2.


3.1. THE EIT KERR NONLINEARITY 70<br />

The decoherence rate <strong>of</strong> ground state is modified,<br />

Γ(Ω S ) = Γ(0) + |Ω S | 2 γ/4∆ 2 S, (3.2)<br />

due to weak optical pumping to state |4〉 followed by spontaneous decay back<br />

to the ground states. Typically, the signal detuning is chosen to be large (i.e.<br />

∆ s ≫ γ) such that the change in Raman detuning is significant while the change<br />

in decoherence is negligible and can be ignored. Throughout the remainder <strong>of</strong> this<br />

chapter we ignore the signal field contribution to the decoherence.<br />

The nonlinear susceptibility for the EIT <strong>Kerr</strong> effect is defined<br />

χ (3)<br />

p,ss ≡ ∂χ ∣<br />

p ∣∣∣Es<br />

. (3.3)<br />

∂|E s | 2 =0<br />

Returning to the theory from chapter 2, we recall that on single-photon resonance<br />

∆ = 0 the probe susceptibility is<br />

χ p ≈ αc<br />

ω p<br />

δ R + iΓ<br />

, (3.4)<br />

R + Γ − iδ R<br />

where α = ρ 22 α 0 is the probe absorption coefficient and ρ 22 ≈ |Ω c | 2 /Ω 2 c is population<br />

in state |2〉. Combining Eqs. (3.1), (3.5), and (3.4) we see that the third-order


3.1. THE EIT KERR NONLINEARITY 71<br />

susceptibility “<strong>Kerr</strong>” susceptibility is<br />

χ (3)<br />

p,ss<br />

≈<br />

|µ| 2<br />

2∆ s 2 ω p<br />

c<br />

v g,p<br />

, (3.5)<br />

where v g,p<br />

= 2R/˜α is the group velocity <strong>of</strong> the probe, and we have assumed<br />

δ R (0) = 0 † .<br />

As long as the Raman detuning stays in the linear regime (i.e. δ R (Ω s ) ≪ R)<br />

the cross phase modulation <strong>of</strong> the probe field will be<br />

φ = Lω p χ (3)<br />

p,ss|E s | 2 /2c, (3.6)<br />

where L is the length <strong>of</strong> the EIT medium, and we assume χ p ≪ 1 (this assumption<br />

is consistent with atomic vapors). However, this cross-phase modulation (XPM)<br />

can be written more generally as<br />

φ ≈ ω pL<br />

2c Re [χ p(Ω S ) − χ p (0)] . (3.7)<br />

Our primary focus is the linear regime for the EIT <strong>Kerr</strong> effect because <strong>of</strong> its<br />

applications to QND measurements, but there are also some interesting effects in<br />

the nonlinear regime.<br />

† The assumption <strong>of</strong> Raman resonance in the absence <strong>of</strong> a signal field is continued throughout<br />

this chapter.


3.1. THE EIT KERR NONLINEARITY 72<br />

3.1.1 EIT <strong>Kerr</strong> dynamics<br />

The EIT <strong>Kerr</strong> transients and rise times result from the slow evolution <strong>of</strong> the GSC<br />

ρ 21 when the signal field perturbs the Raman detuning. The dynamics <strong>of</strong> an EIT<br />

system were derived in Sec. 2.4. For the purpose <strong>of</strong> discussing the EIT <strong>Kerr</strong><br />

dynamics we restrict our discussion to the case <strong>of</strong> zero single-photon detuning<br />

∆ = 0 and a single homogeneous line. We also assume γ ≫ 5Ω ≥ δ R .<br />

For optically thin media (i.e. ˜αL ≪ 1), the EIT <strong>Kerr</strong> dynamics can be determined<br />

by a straightforward application <strong>of</strong> the EIT transient theory <strong>of</strong> Sec. 2.4.<br />

To determine the rise and fall dynamics <strong>of</strong> the EIT <strong>Kerr</strong> effect we consider two<br />

cases. First, at time t = 0 the atoms are in steady state for Raman resonance.<br />

Then a signal field is suddenly turned on, Stark shifting the system away from<br />

Raman resonance. We call this the rising dynamics. The second case is simply<br />

the reverse. At time t = 0 the system is in steady state with the signal field on,<br />

and then the signal is suddenly turned <strong>of</strong>f. This is the falling dynamics. In both<br />

cases the transients <strong>of</strong> the GSC are given by<br />

ρ 21 (t) = ρ 21,f + (ρ 21 (0) − ρ 21,f )e t(iδ R−R−Γ) , for t ≥ 0, (3.8)<br />

where the initial GSC is ρ 21 (0), the final steady state GSC is ρ 21,f , and δ R is the


3.1. THE EIT KERR NONLINEARITY 73<br />

0.5<br />

Im<br />

ρ<br />

(-)<br />

( 21<br />

/ρ 21<br />

)<br />

0<br />

(a)<br />

ρ (ss) (1.2R)<br />

21<br />

ρ<br />

(ss) (δ<br />

21 R<br />

)<br />

ρ 21<br />

(t)<br />

ρ (ss) (0 )<br />

21<br />

0 0. 5 1<br />

Re<br />

ρ<br />

( 21 (-)<br />

/ρ 21<br />

)<br />

Porbe Susceptibility<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

(b)<br />

Im(χ p<br />

)<br />

Re(χ p<br />

)<br />

0<br />

0 2 4 6 8 10<br />

Time (t /R)<br />

Figure 3.2: The transients <strong>of</strong> the EIT <strong>Kerr</strong> nonlinearity plotted looking at (a) the ground state<br />

coherence (GSD) and (b) the probe susceptibility. In (a) the real part <strong>of</strong> GSC is plotted versus<br />

the imaginary part <strong>of</strong> the GSC parametrically as functions <strong>of</strong> time. The dashed red (green) line<br />

shows the rise (fall) dynamics, and the black solid line shows all possible steady state values.<br />

In (b) the rise dynamics for the real and imaginary parts <strong>of</strong> the probe susceptibility are plotted<br />

as a functions <strong>of</strong> time. The parameters used for these calculations are R ≈ 5Γ and for t ≤ 0<br />

δ R = 0, while for t > 0 δ R = 1.2R).<br />

Raman detuning for t > 0. We also recall that the steady state GSC is given by<br />

ρ (ss)<br />

21 (δ R ) =<br />

ρ (−)<br />

21 R<br />

(R + Γ) − iδ R<br />

. (3.9)<br />

Finally, we recall from chapter 2 that χ p ∝ i(1 − ρ 21 /ρ (−)<br />

21 ).<br />

Figure 3.2 shows two different ways <strong>of</strong> understanding the EIT <strong>Kerr</strong> dynamics.<br />

In Fig. 3.2(a) the rise <strong>of</strong> the EIT <strong>Kerr</strong> effect (red dashed line) is seen as the GSC<br />

spiraling from δ R = 0 to its new steady state δ R = 1.2R. The fall <strong>of</strong> the EIT<br />

<strong>Kerr</strong> effect (green dashed line) is a straight line back towards the dark state GSC.<br />

In Fig. 3.2(b) the rising transients for the probe susceptibility are plotted as a<br />

function <strong>of</strong> time. Both the real and imaginary parts <strong>of</strong> the susceptibility oscillate


3.1. THE EIT KERR NONLINEARITY 74<br />

Rise tiem (τR )<br />

10 -1<br />

10 0 δ<br />

/ R<br />

τ Ref,rise<br />

τ Abs,rise<br />

10 -2<br />

10 -1 10 0 10 1<br />

Raman detuning ( ) R<br />

Figure 3.3: Solid lines show the 1/e rise time for absorption and refraction resulting from the<br />

EIT <strong>Kerr</strong> nonlinearity. Rise times are plotted logarithmically as a function <strong>of</strong> the magnitude<br />

<strong>of</strong> the Raman detuning δ R . All quantities are dimensionless and normalized by the optical<br />

pumping rate R, and Γ/R = .2. For small detunings the rise times asymptotically approach<br />

τ Abs = 2.15/(R + Γ) and τ Ref = 1/(R + Γ). For large two-photon detunings the rise times<br />

asymptotically approach τ Abs = 1.2/δ R and τ Ref = 0.63(R + Γ)/δR 2 . These asymptotes are<br />

shown as the dashed lines.<br />

as damped simple harmonic oscillators. For the probe susceptibility the falling<br />

transients, which are not shown, would be a decaying exponential.<br />

3.1.2 Rise-times in optically thin media<br />

From transient curves, like those in Fig. 3.2, it is possible to find the 1/e rise<br />

times for the refractive and absorptive parts <strong>of</strong> the EIT <strong>Kerr</strong> nonlinearity. Figure<br />

3.3 shows these rise times as a function <strong>of</strong> Raman detuning. Also shown as dashed<br />

lines are the asymptotic values for the rise times times. The fall times correspond<br />

to rise times with δ R ≪ R because regardless <strong>of</strong> the initial Raman detuning for


3.1. THE EIT KERR NONLINEARITY 75<br />

the fall dynamics, the final steady state is always in Raman resonance. As Fig.<br />

3.3 shows there are two regimes for which the asymptotic rise times are accurate:<br />

a nonlinear regime with δ R ≫ R and a linear regime with δ R ≪ R.<br />

For most applications <strong>of</strong> the refractive EIT <strong>Kerr</strong> nonlinearity, it is desirable to<br />

work in the linear regime near Raman resonance. This is the case we primarily<br />

focus on. Besides the obvious advantages <strong>of</strong> being linear, the linear regime also<br />

provides the maximum nonlinear susceptibility with the minimum absorption. In<br />

the linear regime, the asymptotic expressions for the 1/e rise and fall times are<br />

given by<br />

lim τ Ref, rise = lim τ Ref, fall = 1<br />

δ R →0 δR →0 R + Γ<br />

(3.10)<br />

and the absorptive rise and fall times are<br />

lim τ Abs, rise = 1 + c 1<br />

δ R →0 R + Γ<br />

(3.11)<br />

and<br />

lim τ Abs, fall = 1<br />

δ R →0 R + Γ , (3.12)<br />

where c 1 = ln{2 + ln[2 + ln(2 + . . .)]}, and the subscripts Ref and Abs refer rise<br />

times for the refractive and absorptive changes.<br />

These rise and fall times can be understood intuitively as the ratio ∆θ/ ˙θ, where<br />

∆θ is the total total phase change <strong>of</strong> the GSC and ˙θ is the rate at which the GSC


3.1. THE EIT KERR NONLINEARITY 76<br />

phase changes ‡ . The rate <strong>of</strong> GSC phase change is always ˙θ = δ R [this can be<br />

deduced from Eq. (3.8)]. The total phase change can be deduced by considering<br />

that in the linear regime Eq. (3.9) simplifies to<br />

( )<br />

21 (z) ≈ ρ(−) 21 R<br />

R + Γ exp δ R<br />

i , (3.13)<br />

R + Γ<br />

ρ (ss)<br />

such that ∆θ = δ R /(R + Γ). Combining these two results we expect the rise time<br />

to be approximately<br />

τ ∼ ∆θ/ ˙θ = 1/(R + Γ).<br />

In the nonlinear regime the above analysis is more challenging an less insightful.<br />

In this regime the asymptotic rise times are<br />

lim τ Abs, rise = arccos (e−1 + (R + Γ)/δ R )<br />

(3.14)<br />

δ R →∞ δ R<br />

and<br />

lim τ Ref, rise = (1 − e−1 )(R + Γ)<br />

. (3.15)<br />

δ R →∞<br />

δ 2 R<br />

3.1.3 Rise times in optically thick media<br />

It is more challenging to calculate the EIT <strong>Kerr</strong> rise times in optically thick<br />

media because both the fields and atoms are simultaneously evolving. Adding to<br />

‡ For consistency throughout this work, we use θ for the phase <strong>of</strong> the GSC, and φ for the<br />

phase <strong>of</strong> the probe field


3.1. THE EIT KERR NONLINEARITY 77<br />

the complication is the fact that the evolution <strong>of</strong> the fields (atoms) is conditional<br />

on the state <strong>of</strong> the atoms (fields). This interplay between atoms and fields results<br />

in a longer rise time. We will show that the rise time is proportional to the optical<br />

thickness <strong>of</strong> the medium.<br />

In the linear regime, the rise time in optically thick media can be estimated<br />

as τ ≈ ∆θ/ ˙θ, similar to the optically thin case. To do this we must first calculate<br />

the steady state change GSC phase.<br />

We recall that the susceptibilities are<br />

χ p ≈ −i α (<br />

)<br />

0c<br />

∗<br />

ρ 22 1 − ρ 21 /ρ (−)<br />

21<br />

(3.16)<br />

ω p<br />

and<br />

χ c ≈ −i α 0c<br />

ω c<br />

ρ 11<br />

(<br />

1 − ρ 21 /ρ (−)<br />

21<br />

)<br />

. (3.17)<br />

In the SVEA, the Maxwell equations for the fields are<br />

( ∂<br />

∂z + 1 )<br />

( )<br />

∂<br />

Ω P = − α 0ρ 22<br />

1 − ρ∗ 21<br />

Ω<br />

c ∂t<br />

2 ρ (−)∗ P (3.18)<br />

21<br />

and<br />

( ∂<br />

∂z + 1 )<br />

( )<br />

∂<br />

Ω C = − α 0ρ 11<br />

1 − ρ 21<br />

Ω<br />

c ∂t<br />

2 ρ (−) C . (3.19)<br />

21<br />

Assuming steady state and negligible probe absorption, we can make the approx-


3.1. THE EIT KERR NONLINEARITY 78<br />

imation |Ω P (z)| ≈ |Ω P (0)|, and Eq. (3.18) can be integrated in space to obtain<br />

Ω P (z) ≈ Ω P (0)e − α 0 ρ(−) 22<br />

hz−Rz<br />

2 0 ρ∗ 21 (z′ )/ρ (−)∗<br />

(<br />

≈<br />

Ω P (0) exp<br />

− zα 0ρ (−)<br />

22<br />

2<br />

iδ R<br />

R + Γ<br />

21 (z ′ ) dz ′i<br />

)<br />

, (3.20)<br />

where we have used the fact that in steady state ρ 21 (z)/ρ (−)<br />

21 (z) ≈ exp(δ R /(R+Γ)).<br />

Similarly,<br />

(<br />

)<br />

zα 0 ρ (−)<br />

11 iδ R<br />

Ω C (z) ≈ Ω C (0) exp<br />

. (3.21)<br />

2 R + Γ<br />

Finally, the ground state coherence as a function <strong>of</strong> position is<br />

(<br />

ρ 21 (z) ≈ ρ (−)<br />

21 (0) exp i (1 + αz/2)δ )<br />

R<br />

. (3.22)<br />

R + Γ<br />

From Eq. (3.22) we see that ∆θ(z) = (1 + α 0 z/2)δ R /(R + Γ) making the rise time<br />

approximately<br />

τ ≈ ∆θ/ ˙θ = 1 + α 0z/2<br />

R + Γ .<br />

In order to obtain a more accurate expression for the EIT <strong>Kerr</strong> rise time we<br />

numerically simulated the EIT <strong>Kerr</strong> effect in an optically thick system and calculated<br />

the rise time as a function <strong>of</strong> optical depth. The results <strong>of</strong> this simulation<br />

are shown in Fig. 3.4. By curve fitting the numerical derived rise times we found


3.1. THE EIT KERR NONLINEARITY 79<br />

8<br />

7<br />

6<br />

Absorp.<br />

rise-time<br />

Rise Time τ(R+Γ)<br />

5<br />

4<br />

3<br />

2<br />

1<br />

Absorp.<br />

fall-time<br />

Refractive<br />

rise & fall times<br />

0<br />

0 2 4 6 8 10<br />

<strong>Optical</strong> Depth (αL)<br />

Figure 3.4: Numerical simulations <strong>of</strong> the EIT <strong>Kerr</strong> 1/e rise time as a function <strong>of</strong> optical depth.<br />

Rise times were calculated using the fields.<br />

that the optically thick equivalents <strong>of</strong> Eqs. (3.10)-(3.12) are<br />

lim τ Ref, rise = lim τ Ref, fall = 1 + ˜α pz/4<br />

δ R →0 δR →0 R + Γ , (3.23)<br />

and the absorption rise and fall times are<br />

lim τ Abs, rise = 1 + (1 + ˜α pz/2)c 1<br />

δ R →0 R + Γ<br />

(3.24)<br />

and<br />

lim τ Abs, fall = 1 + 3c 1 ˜α p z/8<br />

. (3.25)<br />

δ R →0 R + Γ<br />

The EIT <strong>Kerr</strong> dynamics for optically thick media in the nonlinear regime


3.2. QND MEASUREMENTS & KERR NONLINEARITIES 80<br />

δ R ∼ R are not considered here. We also do not consider the effects <strong>of</strong> the signal<br />

group velocity, but they are discussed in Chapter 5.<br />

3.2 QND measurements & <strong>Kerr</strong> nonlinearities<br />

Quantum non-demolition (QND) measurement <strong>of</strong> photon number is only one <strong>of</strong><br />

several application <strong>of</strong> the EIT <strong>Kerr</strong> nonlinearity. However in evaluating the effects<br />

<strong>of</strong> the EIT <strong>Kerr</strong> rise time on potential applications, it is insightful to focus on<br />

QND measurements because all potential EIT <strong>Kerr</strong> applications are either based<br />

on QND measurements or are based on similar operating principles.<br />

Some <strong>of</strong> the EIT <strong>Kerr</strong> applications based on QND measurements are creating<br />

optical number states, high efficiency photon counting detection, preselection for<br />

single photons for single-photon-on-demand sources (a prerequisite for deterministic<br />

QIP), and as a trigger for quantum bit regeneration [103, 104].<br />

There are also several related applications <strong>of</strong> the EIT <strong>Kerr</strong> effect that do not<br />

directly involve QND measurements but operate on nearly identical principles.<br />

These applications include: quantum computing with optical nonlinearities, twoqubit<br />

quantum gates such as the controlled phase gate, quantum entanglement <strong>of</strong><br />

two optical wavepackets, and quantum state preparation for macroscopic quantum<br />

states (i.e. schrodinger cat states).<br />

Each <strong>of</strong> these applications makes similar demands <strong>of</strong> the EIT <strong>Kerr</strong> nonlinearity<br />

to what would be required for a QND measurement capable <strong>of</strong> resolving a single


3.2. QND MEASUREMENTS & KERR NONLINEARITIES 81<br />

photon. QND measurements for large photon numbers have been observed previously<br />

[133, 134], but a QND measurement capable <strong>of</strong> resolving a small photon<br />

number has not been realized. Although we have not actually performed a QND<br />

measurement in this work, the measurements discussed here provide valuable insight<br />

into the limitations and potential <strong>of</strong> the EIT <strong>Kerr</strong> nonlinearity for QND<br />

measurements.<br />

3.2.1 QND measurement and SNR<br />

Before discussing EIT <strong>Kerr</strong> nonlinearities and QND measurements, we consider<br />

the simpler case <strong>of</strong> an idealized <strong>Kerr</strong> nonlinearity with instantaneous rise time.<br />

The primary figure <strong>of</strong> merit for the QND measurements is the signal to noise ratio<br />

(SNR) to be defined shortly. There are also several other criteria that must be<br />

satisfied in order for the measurement to be a QND measurement [135, 136]. We<br />

will show that these criteria can be expressed in terms <strong>of</strong> the SNR.<br />

Figure 3.5 shows an experimental arrangement for a photon number QND<br />

measurement using a <strong>Kerr</strong> medium. The signal and probe fields come together<br />

and interact in a <strong>Kerr</strong> medium, resulting in the phase modulation <strong>of</strong> the probe<br />

[i.e.<br />

E p (L, t) = E p (0, t) exp(iφ(t)),<br />

where φ(t) = ω p χ (3)<br />

p,ss|E s (L, t)| 2 /2c is the phase modulation]. The field amplitudes


3.2. QND MEASUREMENTS & KERR NONLINEARITIES 82<br />

Probe<br />

Beam<br />

Signal<br />

Pulse<br />

PBS<br />

<strong>Kerr</strong><br />

Medium<br />

Local<br />

Oscillator<br />

PD<br />

PBS<br />

i sig(t)<br />

50/50<br />

BS<br />

PD<br />

PD<br />

i (t)<br />

diff<br />

Balanced<br />

Homodyne<br />

Detection<br />

Figure 3.5: Simplified diagram <strong>of</strong> a QND measurement <strong>of</strong> photon number using an optical <strong>Kerr</strong><br />

nonlinearity and balanced homodyne detection.<br />

for the signal and probe fields are<br />

E l (z, t) = E l (z, t)e iω l(z/c−t)<br />

l = {s, p}, (3.26)<br />

where E l (z, t) are the slowly varying amplitudes <strong>of</strong> the fields. The probe field is<br />

a monochromatic cw field and the signal is a transform limited pulse that is zero<br />

for times t < 0 and z > 0. Also, we assume that the signal pulse has a temporal<br />

width τ sig that is long compared the length <strong>of</strong> the <strong>Kerr</strong> medium (i.e. τ sig ≫ L/c).<br />

The probe phase is measured via balanced homodyne detection, resulting in a<br />

difference current <strong>of</strong><br />

i diff (t) = 2ηq eleA cɛ 0 |E p E LO |<br />

sin[φ(t)], (3.27)<br />

ω p 2<br />

where A is the area <strong>of</strong> the beams, E LO is the amplitude <strong>of</strong> the local oscillator,


3.2. QND MEASUREMENTS & KERR NONLINEARITIES 83<br />

η is the quantum efficiency <strong>of</strong> the photo diodes (PD), and q ele is the charge per<br />

electron. The signal <strong>of</strong> interest is the time integral <strong>of</strong> the difference current, and<br />

this charge can be expressed in terms <strong>of</strong> quantum mechanical operators as<br />

∫ t<br />

〈 ˆQ〉(t) = dt ′ i diff (t ′ ) (3.28)<br />

t−T int<br />

= ηq [(<br />

ele<br />

¯n LO + ¯n p + 2 √¯n LO¯n p 〈sin<br />

2<br />

ˆφ〉<br />

) (<br />

− ¯n LO + ¯n p − 2 √¯n LO¯n p 〈sin ˆφ〉<br />

)]<br />

,<br />

≈ 2ηq ele<br />

√¯nLO¯n p C 1¯n s ,<br />

where we have assumed the small angle approximation (sin φ ≈ φ), and the integration<br />

time is T int § . The other parameters in Eq. (3.28) are defined as<br />

¯n p ≡ 〈ˆn p 〉 = T int I p /Aω p and ¯n LO ≡ 〈ˆn LO 〉 = ¯n P /κ 2 (3.29)<br />

are the expectation value for the number <strong>of</strong> photon in the cw probe and local<br />

oscillator fields over a time interval T int . The local oscillator is much larger than<br />

the probe (i.e. κ 2 ≪ 1). C 1 ≡ φ/¯n s is the proportionality constant between the<br />

phase modulation φ averaged over the time interval [t − T int , t] and ¯n s ≡ 〈ˆn s 〉<br />

is the expectation value for the number <strong>of</strong> signal photons. C 1 ∝ χ (3)<br />

pss should be<br />

thought <strong>of</strong> as the “size” <strong>of</strong> the <strong>Kerr</strong> effect (i.e. in cw operation or if the <strong>Kerr</strong><br />

effect is instantaneous then C 1<br />

is proportional to the nonlinear susceptibility).<br />

§ In order to achieve the maximum SNR the integration time which should be chosen such<br />

that T int ∼ (τ sig + τ <strong>Kerr</strong> ), where τ <strong>Kerr</strong> being the rise time <strong>of</strong> the <strong>Kerr</strong> nonlinearity.


3.2. QND MEASUREMENTS & KERR NONLINEARITIES 84<br />

The variance <strong>of</strong> the charge is given by<br />

√〈(∆ ˆQ) 2 〉 = q ele<br />

√ η<br />

[¯n p<br />

(1 + 1 κ 2<br />

)<br />

] 1/2<br />

+ 2¯n2 p<br />

C 2 1 〈(∆ˆn s ) 2 〉 , (3.30)<br />

κ 2<br />

such that the SNR is<br />

SNR =<br />

〈<br />

√<br />

ˆQ〉<br />

√<br />

≈ 2C 1¯n s η¯np , (3.31)<br />

〈(∆ ˆQ) 2 〉<br />

where it is assumed that the probe and local oscillator fields are both coherent<br />

states (i.e. they obey the statistics 〈(∆n i ) 2 〉 = ¯n i for i = p, LO, the signal field is<br />

a number state (i.e. 〈(∆n s ) 2 〉 = 0), and ∆X ≡ X − 〈X〉 .<br />

There are several quantifiable criteria that must be simultaneously satisfied<br />

in order for a measurement to be a QND measurement. Also, many <strong>of</strong> the EIT<br />

<strong>Kerr</strong> applications require a certain number resolution. Fortunately, it is relatively<br />

straightforward to express the resolution and QND criteria in terms <strong>of</strong> the SNR.<br />

The problem <strong>of</strong> resolving individual number states using a QND <strong>Kerr</strong> measurement<br />

was considered by Haus in Ref. [136]. It was found that given the signal<br />

field is in a particular number state n s , the probability for correctly identifying<br />

the photon number is given by<br />

P (n s |n s ) ≈ erf(SNR/2 √ 2n s ), (3.32)<br />

This SNR does not include technical noise.


3.2. QND MEASUREMENTS & KERR NONLINEARITIES 85<br />

where erf(x) = ∫ x<br />

0 dy2 exp(−y2 )/ √ π is the error function [136]. A SNR <strong>of</strong> 4 will<br />

provide 95% probability <strong>of</strong> correct detection, and a SNR <strong>of</strong> 10 gives a probability<br />

<strong>of</strong> a detection error is below 10 −6 . Because the resolution <strong>of</strong> a QND measurement<br />

is exponential in the SNR, QND measurements may be the ideal method to satisfy<br />

the demanding requirements for detection efficiency in linear optics quantum<br />

computing [109].<br />

There are three criteria that must be satisfied for a measurement to be a QND<br />

measurement.<br />

1. A QND measurement should be non-demolition. This means that the signal<br />

should not be destroyed by its interaction with the probe. Tracing over the<br />

probe, the signal observable at the output should be highly correlated with<br />

the signal observable at the input. In our case, this signal observable is the<br />

number <strong>of</strong> signal photons. Expressed mathematically, the “nondemolition”<br />

condition is C s<br />

defined as<br />

= C(∆ˆn in<br />

s , ∆ˆn out ) ≈ 1, where the correlation function is<br />

s<br />

C(∆ ˆX, ∆Ŷ ) ≡ 〈∆<br />

√<br />

ˆX∆Ŷ 〉 . (3.33)<br />

〈∆ ˆX 2 〉〉〈∆Ŷ 2 〉〉<br />

2. The QND measurement should be a measurement (i.e. when we measure the<br />

probe observable, we should obtain information about signal observable). In<br />

our case, the probe observable is the charge ˆQ which is proportional to the<br />

probe phase. This “measurement” condition can be expressed mathemati-


3.2. QND MEASUREMENTS & KERR NONLINEARITIES 86<br />

cally using another correlation function (i.e. C m = C(∆ˆn in<br />

s , ∆ ˆQ) > 0). For<br />

a perfect QND measurement C m = 1.<br />

3. Finally, a QND measurement should be quantum (i.e. the probe and signal<br />

observables should be quantum correlated). In other words, the conditional<br />

variance between signal and probe observables should be below the standard<br />

quantum limit. This characteristic <strong>of</strong> QND measurements is <strong>of</strong>ten referred<br />

to as the quantum state preparation property, and it implies that if neither<br />

signal or probe observable is squeezed, then the probe and signal will be<br />

non-separable. This “quantum” condition can be expressed mathematically<br />

by the inequality<br />

V s|p = V (∆ˆn out<br />

s |∆ ˆQ) < 1 (3.34)<br />

(<br />

= 〈∆ˆn out2<br />

s 〉 1 − C 2 (∆ˆn out<br />

s , ∆ ˆQ)<br />

)<br />

/〈ˆn out<br />

s < 1〉,<br />

where V s|p is a conditional variance, and the last line is only true for observables<br />

with Gaussian statistics. Also, we follow the convention <strong>of</strong> Refs.<br />

[135, 137] by normalizing the conditional variance to the standard quantum<br />

limit, such that quantum correlations are implied by V s|p < 1.<br />

A measurement is considered to be QND if both C s + C m > 1 and V s|p < 1.<br />

With the exception <strong>of</strong> the first criterion, these QND criteria can be expressed<br />

in terms <strong>of</strong> the SNR. Satisfying the first criterion, non-demolition <strong>of</strong> the signal,


3.2. QND MEASUREMENTS & KERR NONLINEARITIES 87<br />

requires knowledge <strong>of</strong> the specifics <strong>of</strong> the <strong>Kerr</strong> measurement. For the EIT <strong>Kerr</strong><br />

effect, the signal intensity after the medium is<br />

(<br />

I s (L) ≈ I s (0) exp −<br />

α 0Lγ<br />

∆ s (1 + κ 1 )<br />

)<br />

,<br />

where κ 1 ≡ |Ω c /Ω p | 2 and ∆ s ≫ γ. Thus, the correlation function is C s ≈<br />

exp[−α 0 Lγ/∆ s (1 + κ 1 )]. As long as α 0 Lγ ≪ ∆ s (1 + κ 1 ) the first condition C s ≈ 1<br />

is satisfied. This also enables the simplification<br />

C m = 1/ √ 1 + ¯n 2 s/SNR 2 〈(∆n s ) 2 〉, (3.35)<br />

and the conditional variance is<br />

V s|p = (1 − C m 2 ). (3.36)<br />

This implies that both the measurement condition C m ≈ 1 and the quantum state<br />

preparation condition are satisfied when SNR/¯n s ≫ 1/ √ 〈(∆n s ) 2 〉 ‖ . Assuming<br />

the signal, probe, and local oscillator fields are all coherent states, the conditions<br />

‖ All three <strong>of</strong> the QND measurement criteria seem to have been developed under the assumption<br />

that the input noise for both signal and probe observables is at the standard quantum<br />

limit-i.e. 〈(∆n s ) 2 〉 = 〈n s 〉 and 〈(Q) 2 〉 = 〈Q〉. All <strong>of</strong> the conditions lose their utility when one <strong>of</strong><br />

the variances goes to zero. This is somewhat unfortunate because one <strong>of</strong> the primary conditions<br />

<strong>of</strong> a QND measurement is that repeated QND measurements obtain the same results. However,<br />

according to the three criteria given for QND measurements repeated non-ideal QND measurements<br />

become progressively more poor because the variance <strong>of</strong> the input signal is smaller with<br />

each repeated measurement.


3.2. QND MEASUREMENTS & KERR NONLINEARITIES 88<br />

for a good EIT <strong>Kerr</strong> QND measurement simplify to α 0 Lγ ≪ ∆ s (1 + κ 1 ) and<br />

SNR ≫ √¯n s .<br />

3.2.2 SNR for EIT <strong>Kerr</strong> Measurements<br />

In order for the results <strong>of</strong> Sec. 3.2.1 to be useful, we must apply these results to the<br />

EIT <strong>Kerr</strong> nonlinearity, which does not have an instantaneous rise time. In fact,<br />

the EIT <strong>Kerr</strong> rise time is typically rather slow (in the msec. to µsec. range for the<br />

experiments we discuss). Understanding the EIT <strong>Kerr</strong> rise time and dynamics are<br />

essential in optimizing the SNR and determining the limitations <strong>of</strong> the EIT <strong>Kerr</strong><br />

effect. We consider only the linear regime δ R ≪ R, and in this section we focus<br />

primarily on optically-thin media ˜αL ∼ 1 although some <strong>of</strong> the results also apply<br />

to optically thick media. Most <strong>of</strong> the discussion <strong>of</strong> EIT <strong>Kerr</strong> QND measurements<br />

in optically thick media is reserved for chapter 5. Finally, in anticipation <strong>of</strong> the<br />

experimental systems in Chapters 4 and 5, we consider the specific case that the<br />

probe, signal, and coupling transitions are all on the D 1 line <strong>of</strong> an alkali metal<br />

vapor (i.e. |µ D1 | 2 = πγɛ 0 c 3 /ω0, 3 and ω 0 ≈ ω i with i = {c, p, s}) ∗∗ .<br />

For both optically thin and thick media the phase modulation <strong>of</strong> the probe<br />

∗∗ In the chapter 4 experiment the signal is on the D 2 line not the D 1 . The only significant<br />

change would be that the D 2 and D 1 dipole moments are related by |µ D2 | 2 = 2|µ D1 | 2 such that<br />

for chapter 4 C 1 in the last line <strong>of</strong> Eq. (3.41) should be a factor <strong>of</strong> two larger.


3.2. QND MEASUREMENTS & KERR NONLINEARITIES 89<br />

field can be written as<br />

φ(L, t) = ω P χ (3) ∫ t<br />

p,ssL<br />

c 2 ɛ 0 0<br />

dt ′ h(t ′ )I s (0, t − t ′ ), (3.37)<br />

where h(t) is the normalized impulse response function for the <strong>Kerr</strong> nonlinearity †† .<br />

For optically thin media the EIT <strong>Kerr</strong> impulse response function is<br />

⎧<br />

⎪⎨<br />

exp[−t/τ eit ]<br />

τ <strong>Kerr</strong><br />

t ≥ 0<br />

h thin (t) ≈<br />

⎪⎩ 0 otherwise<br />

, (3.38)<br />

where it is assumed L/v g,s ≪ τ eit and the signal’s propagation delay through the<br />

cell can be ignored [42]. The difference current then becomes<br />

i diff (t) ≈ 2ηq elen p ω P χ (3)<br />

p,ssL<br />

√ × . . . (3.39)<br />

T int κ1 c 2 ɛ 0<br />

⎧<br />

( ) ∫ ⎪⎨ exp −<br />

t t<br />

τ <strong>Kerr</strong> 0 dt′ I s(t ′ )<br />

τ <strong>Kerr</strong><br />

τ <strong>Kerr</strong> ≫ τ sig<br />

.<br />

⎪⎩ I s (t) τ <strong>Kerr</strong> ≪ τ sig<br />

The maximal SNR is achieved when τ eit ≫ τ sig and T int = 5τ eit /4.<br />

Even though the time dependent current i diff is highly dependent on the EIT<br />

<strong>Kerr</strong> dynamics, the integrated charge 〈 ˆQ〉 is independent <strong>of</strong> the dynamics if the<br />

integration time is sufficiently long.<br />

In the limit <strong>of</strong> extremely long integration<br />

†† The assumption <strong>of</strong> an impulse response function implies that the <strong>Kerr</strong> response is a linear<br />

in the signal intensity i.e. the linear regime with δ R ≪ R.


3.2. QND MEASUREMENTS & KERR NONLINEARITIES 90<br />

times the charge becomes<br />

p,ssL<br />

lim 〈Q〉 = 2ηq elen p ω s χ (3)<br />

√<br />

T int →∞ T int κ1 c 2 ɛ 0<br />

n s ω s<br />

A . (3.40)<br />

As long as the integration time is long enough to satisfy the inequality T int ≥<br />

(τ sig + τ eit ), we can reasonably approximate 〈Q〉 ≈ lim Tint →∞〈Q〉. Throughout the<br />

remainder <strong>of</strong> this chapter, we assume T int = τ eit and τ sig ≪ τ eit . The fact that the<br />

signal 〈 ˆQ〉 is independent <strong>of</strong> the dynamics is important for ensuring the accuracy<br />

<strong>of</strong> the QND measurement. However, the SNR is not independent <strong>of</strong> the dynamics,<br />

which raises the question <strong>of</strong> how it can be maximized and what is its maximum<br />

value.<br />

Using Eq. (3.40) we find that for an EIT <strong>Kerr</strong> nonlinearity<br />

C 1 ≈ ω2 sLχ (3)<br />

p,ss<br />

2Ac 2 ɛ 0 T int<br />

,<br />

≈<br />

πγ<br />

2∆ s A k 2 0<br />

˜αL<br />

4RT int<br />

, (3.41)<br />

where in the last line we have used the D 1 dipole moment µ D1 to simplify the<br />

expression. From the definition for ¯n p in Eq. (3.29) we have<br />

n p = (1 + ˜αL/4)Ak2 0<br />

π(1 + κ 1 )<br />

T int<br />

τ eit<br />

, (3.42)<br />

where we have used the fact that τ eit = (1 + ˜αL/4)/(R + Γ). The SNR is largest


3.2. QND MEASUREMENTS & KERR NONLINEARITIES 91<br />

for optically thick media, in which case the SNR simplifies to<br />

SNR<br />

n s<br />

≈<br />

√ √ √ πη 1 γ ˜αL τeit<br />

√ . (3.43)<br />

2 k 0 A ∆ s 1 + κ 1 T int<br />

There are three parameters that can be optimized to increase the SNR; optical<br />

depth ˜αL, beam area A, and signal detuning ∆ s . The beam area has a lower limit<br />

<strong>of</strong> A ≥ 2π/k0 2 due to diffraction. Also, the optical depth and signal detuning are<br />

not truly independent. We recall that the nondemolition criterion required that<br />

the inequality ∆ S /γ ≫ √ α 0 L/(1 + κ 1 ) be satisfied.<br />

Thus, the nondemolition<br />

criterion and measurement criterion have competing objectives. It is not possible<br />

to adjust the optical depth and signal detuning to increase the SNR without also<br />

increasing the absorption <strong>of</strong> the signal.<br />

Let us suppose that in a given experiment it is possible to focus down to the<br />

diffraction limit A = 2π/k 2 0<br />

and that 5% absorption <strong>of</strong> the signal field can be<br />

tolerated (i.e. αL/(1 + κ 1 ) ≤ 0.05∆ 2 s/γ 2 ). Also, assume that τ eit ≫ τ sig such<br />

that T int ≈ τ eit . Then, the SNR can be no larger than 0.16 √ η. The maximum<br />

phase shift per signal photon achievable under these circumstances is φ/¯n s =<br />

C 1 = γ/4∆ s ≪ √ 1 + κ 1 /4 √ α 0 L. Although the phase shift could theoretically<br />

be made large by decreasing the signal detuning, for practical reasons the signal<br />

detuning is typically ∆ s<br />

≥ 10γ, making the maximum phase shift per signal<br />

photon about two hundredth <strong>of</strong> a radian ‡‡ . From this discussion, it may seem<br />

‡‡ This result is similar to the result found by Harris and Hau [56] for a pulsed signal and


3.2. QND MEASUREMENTS & KERR NONLINEARITIES 92<br />

that the prospects are not good for single-photon resolution using EIT <strong>Kerr</strong> QND<br />

measurements. However, there are some work-arounds to this problem that are<br />

discussed in Chapters 5 and 6.<br />

For example, it is possible to decrease the integration time T int by slowing<br />

down the group velocity <strong>of</strong> the signal field.<br />

When the group velocities <strong>of</strong> the<br />

probe and signal are matched, the phase modulation on the probe will be much<br />

shorter temporally and a shorter integration time is required. The details <strong>of</strong> this<br />

process are discussed in Chapter 5.<br />

pulsed probe. In the case considered by Harris and Hau the signal and probe pulses had a<br />

limited interaction length due to temporal walk-<strong>of</strong>f arising from the different group velocities for<br />

probe and signal fields. The probe has a very slow group velocity due to EIT, and the signal<br />

group velocity is approximately c. In our case the signal is pulsed but the probe is cw, and<br />

the limitation on the maximum phase shift is due to the dependence <strong>of</strong> the EIT <strong>Kerr</strong> rise time<br />

on optical thickness. There are greater similarities between the two cases than are immediately<br />

apparent, and these are discussed in chapter 5


93<br />

Chapter 4<br />

EIT <strong>Kerr</strong> dynamics Experiment<br />

Over the past ten years there has been considerable interest in the EIT <strong>Kerr</strong><br />

nonlinearity and its potential applications. We have shown that the dynamics <strong>of</strong><br />

the EIT <strong>Kerr</strong> effect play an important role in determining the limitation <strong>of</strong> the<br />

EIT <strong>Kerr</strong> effect for pulsed applications. This chapter discusses the first reported<br />

experiments measuring the dynamics <strong>of</strong> the EIT <strong>Kerr</strong> nonlinearity.<br />

The first section details the experimental parameters and apparatus. We then<br />

present steady state measurements <strong>of</strong> the EIT <strong>Kerr</strong> effect in Sec.<br />

4.2, before<br />

presenting the transient EIT <strong>Kerr</strong> measurements in Sec. 4.3. Finally, in Sec.<br />

4.4 we discuss the agreement between theory and experiment, and explain why<br />

certain small discrepancies exist.<br />

4.1 Experiment<br />

Using the Mach-Zehnder interferometer as shown in Fig. 4.1, we observed the<br />

transients <strong>of</strong> EIT on the D 1 line <strong>of</strong> rubidium 85. The transients are created


4.1. EXPERIMENT 94<br />

ECDL2<br />

AOM<br />

80 MHz<br />

Signal<br />

50/50<br />

BS<br />

ECDL1<br />

Homodyne<br />

Detection<br />

λ<br />

2<br />

PBS<br />

Pump<br />

Probe<br />

AOM<br />

1.5 GHz<br />

Computer<br />

DAQ<br />

50/50 BS 50/50 BS<br />

Spectral<br />

filter<br />

PBS<br />

Mach-Zehnder<br />

Interferometer<br />

Rb Cell<br />

PZT<br />

Magnetic Shielding<br />

PBS<br />

Figure 4.1: Schematic <strong>of</strong> the experimental apparatus. The probe beam is sent through a Mach-<br />

Zehnder interferometer to measure changes in the relative phase between the two arms <strong>of</strong> the<br />

interferometer. The coupling and signal beams interact with the probe in the rubidium cell before<br />

being separated from the probe via polarization and spectral filtering. A computer controlled<br />

data acquisition system records and averages the difference signal from balanced homodyne<br />

detection.<br />

by pulsing the signal beam that was about 2 GHz red detuned from the Rb 85<br />

D 2 line. Previous experiments have used frequency modulation spectroscopy to<br />

observe similar Stark shifts and refractive EIT <strong>Kerr</strong> nonlinearities [47], but such<br />

experiments are unable to observe the transient behavior because they require<br />

lock-in amplification. Thus, this is the first direct observation <strong>of</strong> the dynamics<br />

and rise times <strong>of</strong> the refractive EIT <strong>Kerr</strong> nonlinearity.<br />

Mutually coherent coupling and probe fields were obtained from the same external<br />

cavity diode laser (ECDL1) with wavelength λ = 795 nm. ECDL1 was


4.1. EXPERIMENT 95<br />

tuned to the coupling frequency. The probe frequency was obtained by double<br />

passing through a 1.5 GHz acousto-optic modulator (AOM) such that the frequency<br />

difference between probe and coupling was 3 GHz, which is the ground<br />

hyperfine splitting for Rb 85 . Deriving both EIT fields from a single laser ensured<br />

the excellent phase coherence necessary for good EIT. Additionally, we are able<br />

to control the frequency difference between probe and coupling precisely via the<br />

microwave frequency synthesizer driving the AOM.<br />

The signal was obtained from a second laser (ECDL2) with wavelength λ =<br />

780 nm, and pulses were created using a 80 MHz AOM with a rise time <strong>of</strong> 50 ns.<br />

Both ECDL lasers are New Focus Vortex lasers. The signal and coupling beams<br />

are combined on a 50/50 beam splitter (BS) and are both horizontally polarized in<br />

order to be transmitted through a polarizing beam splitter (PBS) into the probe<br />

arm <strong>of</strong> the Mach-Zehnder interferometer (MZI).<br />

The MZI provides measurements <strong>of</strong> relative phase changes between the two<br />

arms <strong>of</strong> the interferometer. The MZI has a strong reference/local-oscillator arm<br />

and a weaker probe arm ∗ . After the probe has propagated through the rubidium<br />

cell, both polarization and spectral filtering are used to isolate the probe field<br />

from the coupling, signal, and any four-wave mixing products. The mirror after<br />

∗ Both arms <strong>of</strong> the interferometer have the same power after the first 50/50 beam splitter<br />

(BS), but the probe arm becomes significantly weaker due to attenuation from the rubidium<br />

cell, spectral filter, and various other optics.


4.1. EXPERIMENT 96<br />

the rubidium cell is mounted on a piezo-electric actuator to control the relative<br />

phase between the MZI arms.<br />

At the output <strong>of</strong> the Mach-Zehnder interferometer the homodyne signal is<br />

measured using two reverse biased Hamamtsu S2830 photodiodes soldered for<br />

differential detection. After the photodiodes a 1 MHz low-noise transimpedance<br />

amplifier prepares the signal for computer controlled data acquisition (DAQ). CW<br />

measurements are recorded using the computer, and the fast transient data are<br />

recorded and averaged using a 1.5 GHz bandwidth oscilloscope. The homodyne<br />

signal is given by<br />

Q ∝ |E LO | |E p | sin [φ pzt + Re(χ p )k 0 L/2] , (4.1)<br />

where E LO is the amplitude <strong>of</strong> the field in the local-oscillator arm, E p is the<br />

probe field amplitude immediately before the last 50/50 BS, φ pzt<br />

is the piezo<br />

controlled phase difference between the MZI arms. The other parameters have<br />

been defined in early chapters. If amplitude measurements are desired instead <strong>of</strong><br />

phase measurements, the local-oscillator beam and one <strong>of</strong> the photo diodes can<br />

be blocked.<br />

The apparatus for the rubidium cell consists <strong>of</strong> three layers <strong>of</strong> magnetic shielding<br />

with a long solenoid inside the innermost layer <strong>of</strong> magnetic shielding.<br />

In<br />

addition to containing 20 torr <strong>of</strong> nitrogen buffer gas, the rubidium cell contains


4.1. EXPERIMENT 97<br />

5P 3/2<br />

5P 1/2 λ = 780 nm<br />

ΩC↔ΩS<br />

ΩP<br />

λ = 795 nm<br />

F=3<br />

F=2<br />

↔↔<br />

5S 1/2<br />

y-position (mm)<br />

2<br />

1<br />

0<br />

-1<br />

-2<br />

-2 -1 0 1 2<br />

x-position (mm)<br />

Figure 4.2: Level diagram <strong>of</strong> rubidium 85 D 1 and D 2 lines with the primary transitions excited<br />

by the probe, coupling, and signal fields. The inset shows the approximate beam sizes and<br />

shapes for the probe (w p,x = 1 mm and w p,y = 1.8 mm), coupling (w c = 2.1 mm), and signal<br />

(w s = 2.2 mm) fields.<br />

both rubidium isotopes in their natural abundance † .<br />

The cell is cylindrical in<br />

shape with a length <strong>of</strong> 30 mm and a diameter <strong>of</strong> 25.4 mm. The number density<br />

<strong>of</strong> rubidium atoms is controlled by heating the cell with electrical strip heaters<br />

positioned between the outer and middle layers <strong>of</strong> magnetic shielding.<br />

Figure 4.2 shows the hyperfine levels for the Rb 85 D 1 and D 2 lines with laser<br />

transitions that are driven in the experiment. The hyperfine splittings ‡ for the<br />

5P 1/2 (361 MHz) and 5P 3/2 (220 MHz) hyperfine levels are much smaller than<br />

the Doppler width (600 MHz), such that the individual hyperfine levels cannot<br />

be resolved. This is convenient because we can treat the excited states as a single<br />

† The natural abundance <strong>of</strong> rubidium isotopes are 27.8% Rb 87 and 72.2% Rb 85<br />

‡ The hyperfine splitting for Rb 85 and Rb 87 D 1 and D 2 lines can be found in Appendix 1.


4.1. EXPERIMENT 98<br />

hyper-fine level. Finally, the signal field Stark shifts both ground hyperfine levels<br />

resulting in a total change in Raman detuning <strong>of</strong><br />

δ R (Ω S ) = δ R (0) +<br />

Ω 2 ∆ hf<br />

4∆ S (∆ S + ∆ hf ) , (4.2)<br />

where ∆ hf = 2π × 3.0357 GHz is the hyperfine splitting <strong>of</strong> the ground states.<br />

The inset in Fig. 4.2 shows the approximate 1/e 2 intensity beam shapes for the<br />

coupling, probe, and signal fields (in the experiment the beams were not perfectly<br />

centered). The coupling and signal beams were essentially identical with beam<br />

radius <strong>of</strong> w c ≈ w s ≈ 2.2 mm, while the probe beam was smaller and slightly<br />

elliptical w p,x ≈ 1.0 mm and w p,y ≈ 1.8 mm.<br />

Table 4.1: Several measured parameters for the EIT line shapes with no signal<br />

present. The beam powers were measured immediately before the PBS in front<br />

<strong>of</strong> the vapor cell. αL is the maximum absorption experienced far from Raman<br />

resonance. ˜αL is the difference between maximum absorption and the absorption<br />

minimum near Raman resonance, and EIT FWHM is self-explanatory.<br />

Temper- Power αL ˜αL EIT<br />

ature Coupl. Total EIT Depth FWHM)<br />

58 ◦ C 340(20) µW 5.2(3) 3.7(1) 11.5(7) kHz<br />

58 ◦ C 1.0(1) µW 5.3(2) 4.3(1) 31.9(1.2) kHz<br />

58 ◦ C 1.9(1) µW 5.5(2) 4.6(1) 58.1(1.8) kHz<br />

48 ◦ C 1.0(1) µW 2.0(1) 1.5(1) 31.9(7) kHz<br />

52 ◦ C 1.0(1) µW 3.3(1) 2.6(1) 33.2(7) kHz<br />

72 ◦ C 1.0(1) µW 20.4(9) 16.4(7) 41(3) kHz


4.2. CW MEASUREMENTS 99<br />

4.2 CW measurements<br />

To verify the dependence <strong>of</strong> the EIT <strong>Kerr</strong> dynamics on optical depth, EIT<br />

linewidth, and Raman detuning, we measured the transient and CW response<br />

<strong>of</strong> EIT to the signal field while varying the cell temperature (optical thickness<br />

αL), coupling intensity (EIT linewidth (R + Γ)), and signal intensity (Stark shift<br />

δ R ). The different parameters at which data was taken is summarized in Table<br />

4.1. Each row in Table 4.1 corresponds to a different set <strong>of</strong> data in which both<br />

cw and transient measurements were taken for several signal powers ranging from<br />

0.2 mW to 4.9 mW. In all measurements the probe power was 40 µW, and the<br />

signal detuning was ∆ S = 2π × 1.9 GHz.<br />

Figure 4.3 shows the cw probe absorption as a function <strong>of</strong> Raman detuning for<br />

several different signal powers, with a cell temperature <strong>of</strong> 58 ◦ C and a coupling<br />

power <strong>of</strong> 0.34 mW (first row in Table 4.1). Figures 4.6-4.11 also use this same<br />

data set. In addition to the shift in absorption minimum resulting from the Stark<br />

shift, the transparency also becomes broader and less deep with increasing signal<br />

power. These effects are mostly due to the additional decoherence created by the<br />

signal field (i.e. Γ(Ω S ) = Γ(0) + Ω 2 S γ/4∆2 S ). This additional decoherence is most<br />

noticeable when the optical pumping is <strong>of</strong> the same order <strong>of</strong> magnitude as the<br />

decoherence rate R ∼ Γ, which is the case for the data corresponding to row 1<br />

<strong>of</strong> Table 4.1. For all other sets <strong>of</strong> parameters in Table 4.1 the optical pumping R<br />

is a factor <strong>of</strong> 3-6 times larger than for row 1 and the broadening is significantly


4.2. CW MEASUREMENTS 100<br />

5<br />

Probe absorption (αL)<br />

4<br />

3<br />

2<br />

P s = 4.5 mW<br />

P s = 2.5 mW<br />

P s = 1.5 mW<br />

P s = 0.4 mW<br />

P = 0<br />

s<br />

-20 -10 0 10 20 30<br />

Raman Detuning (kHz)<br />

Figure 4.3: Probe absorption versus Raman detuning for different cw signal powers. The signal<br />

detuning is ∆ S = 2π×1.9 GHz and the coupling (probe) power is 340 µW (40 µW). The fact<br />

that with no signal field Raman resonance is already shifted away from the expected zero is due<br />

mostly to the 20 Torr N 2 buffer gas shifting the hyperfine splitting (∆ω hf = 2π × 240 Hz per<br />

Torr N 2 ). The probe and coupling also cause a small (about 1 kHz) Stark shift at these powers.


4.2. CW MEASUREMENTS 101<br />

Stark Shift (kHz)<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

Strark Shift due to<br />

Signal Field Only<br />

theory<br />

αL=1.7, P C<br />

= 1 mW<br />

αL=2.6, P = 1 mW<br />

C<br />

αL=3.8, P C<br />

=340 µ W<br />

αL=4.3, P = 1 mW<br />

C<br />

αL=4.6, P =1.9 mW<br />

C<br />

αL=16.4, P C<br />

= 1 mW<br />

0<br />

0 1 2 3 4 5 6<br />

Signal Power (mW)<br />

Figure 4.4: Stark shift due to the signal field for several settings <strong>of</strong> coupling power and EIT<br />

optical depth.<br />

less noticeable. In discussing the measurement data we use row 1 <strong>of</strong> table 4.1 to<br />

show all <strong>of</strong> the steps in the experiment. The same experimental procedures and<br />

analysis were also carried out for the other rows <strong>of</strong> table 4.1 although they are<br />

not discussed in the same detail as row 1.<br />

From Fig. 4.3 it is possible to find the Stark shift as a function <strong>of</strong> signal power<br />

by determining the displacement <strong>of</strong> the absorption minimum for the probe. These<br />

measured signal Stark shifts are shown in Fig. 4.4 for all rows <strong>of</strong> Table 4.1.<br />

There is also a Stark shift due to the coupling field interacting with the probe<br />

transitions and vice-versa due to the probe field interacting with the coupling<br />

transitions. This Stark shift is plotted as a function <strong>of</strong> coupling power in Fig.


4.2. CW MEASUREMENTS 102<br />

FWHM (kHz)<br />

Stark Shift (kHz)<br />

80<br />

EIT Linewidth<br />

60 (FWHM)<br />

40<br />

20<br />

0<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Stark shift due to Coupling<br />

and Probe fields only<br />

theory<br />

αL≈1.5<br />

αL≈2.6<br />

αL≈4.5<br />

αL≈16<br />

theory<br />

αL≈1.5<br />

αL≈2.6<br />

αL≈4.5<br />

αL≈16<br />

0 0.5 1 1.5 2 2.5<br />

Coupling Power (mW)<br />

Figure 4.5: Stark shift due to the coupling and probe fields as a function <strong>of</strong> coupling power.<br />

Also, shown is the FWHM <strong>of</strong> the EIT resonance versus coupling power. The signal was <strong>of</strong>f for<br />

all <strong>of</strong> these measurements.<br />

4.5. Also, plotted in Fig. 4.5 is the transparency linewidth versus coupling power.<br />

The linewidth and Stark shift are also slightly dependent on the probe power,<br />

but probe power is sufficiently small that it can be neglected. Figures 4.4 and 4.5<br />

exhibit the expected linear dependence on intensity § .<br />

Using homodyne detection we measured both real and imaginary part <strong>of</strong> the<br />

probe susceptibility. Figure 4.6 shows cw homodyne data corresponding to the<br />

first row <strong>of</strong> table 1. The homodyne signal was measured by first adjusting the<br />

piezo voltage such that φ pzt = 0 and then stepping the the Raman detuning over<br />

§ Note that for large optical depth, ˜αL ≈ 16 the EIT linewidth was slightly wider than for<br />

smaller optical depth and all other parameters the same. This may be due to gain from four-wave<br />

mixing.


4.2. CW MEASUREMENTS 103<br />

2<br />

Homodyne signal (arb. units)<br />

1<br />

0<br />

-1<br />

-2<br />

-3<br />

-30 -20 -10 0 10 20 30 40 50<br />

Raman detuning (kHz)<br />

Figure 4.6: Raw CW homodyne data for coupling power <strong>of</strong> 340 µW. Two sets <strong>of</strong> data are shown;<br />

one for P s = 0 and the other for P s = 4.5 mW. The dotted lines show the homodyne envelope<br />

(i.e. the maximum and minimum values which the homodyne signal can obtain).<br />

the desired range <strong>of</strong> Raman detuning with a dwell time <strong>of</strong> 200 ms at each Raman<br />

detuning.<br />

In Fig. 4.6 the homodyne signal is plotted versus Raman detuning for the first<br />

row <strong>of</strong> Table 4.1 (black lines are in the absences <strong>of</strong> the signal field, and green lines<br />

correspond to a signal power <strong>of</strong> 4.5 mW). The dashed curves show the envelopes for<br />

the maximum and minimum homodyne signals. These envelopes are determined<br />

by measuring the homodyne signal while the phase φ pzt is scanned over several<br />

periods and fitting the measured homodyne current to a sinusoid. This is repeated<br />

at each Raman detuning. In Fig. 4.7 we have used the data from Fig. 4.6 to<br />

calculate the real and imaginary parts <strong>of</strong> the probe susceptibility. The homodyne


4.2. CW MEASUREMENTS 104<br />

3<br />

2<br />

Probe susceptibility (αL/2)<br />

2<br />

1<br />

0<br />

3<br />

2<br />

1<br />

0<br />

-30<br />

Re(χ p<br />

)<br />

Im(χ p<br />

)<br />

δ = 8 kHz<br />

R<br />

δ = 0<br />

R<br />

-20 -10 0 10 20 30 40 50<br />

Raman Detuning (kHz)<br />

1<br />

0<br />

-1<br />

1<br />

0<br />

-1<br />

Probe susceptibility (∆φ in Rad.)<br />

Figure 4.7: Measured absorption αL and phase shift ∆φ derived from the homodyne data in<br />

Fig. 4.6. In addition to obtaining refraction information we are able to confirm the absorption<br />

measurements shown in Fig. 4.3.<br />

envelop gives the absorptive (imaginary) part <strong>of</strong> the probe susceptibility, and the<br />

refractive (real) part <strong>of</strong> the probe susceptibility is extracted using Eq. (4.1).<br />

Using homodyne detection to measure the absorption has a couple <strong>of</strong> advantages<br />

over measuring the transmitted intensity directly.<br />

First, it increases the<br />

dynamic range because we are measuring the field amplitude E P directly instead<br />

<strong>of</strong> measuring the probe intensity I P<br />

∝ |E P | 2 . Without this increased dynamic<br />

range it would not have been possible to measure EIT line shapes with optical<br />

depths <strong>of</strong> ˜α P = 16.4. Also, when we scan the piezo-controlled phase φ pv we obtain<br />

a sinusoidal signal that can be used for lock-in amplification to extract the signal<br />

out <strong>of</strong> the noise.


4.3. TRANSIENTS MEASUREMENTS 105<br />

Homodyne Signal (arb. units)<br />

1<br />

0<br />

-1<br />

Homodyne<br />

Envelope<br />

Signal<br />

0 100 200 500 600 700 800<br />

Time ( µ sec.)<br />

Figure 4.8: Homodyne measurements as a function <strong>of</strong> time (P s = 4.5 mW, P c = 340 µW). The<br />

signal was turned on at t = 0 and then turned <strong>of</strong>f again at t = 500 µs. Similar to Fig. 4.6, the<br />

envelope shows the maximum and minimum values obtainable by the homodyne signal.<br />

4.3 Transients Measurements<br />

Homodyne detection was also used to measure the transient response <strong>of</strong> the EIT<br />

fields to the fast turn-on and turn-<strong>of</strong>f <strong>of</strong> the signal field. In this case, the frequency<br />

difference between the probe and coupling field was optimized for peak probe<br />

transmission with the signal field <strong>of</strong>f. The signal was then turned on abruptly and<br />

left on for approximately 1 ms, and then the signal is turned <strong>of</strong>f abruptly. Figure<br />

4.8 shows the transient homodyne signal and homodyne envelope when the peak<br />

signal power was 4.5 mW and all other parameters are given by the first row in<br />

Table 4.1. The square dashed curve shows the shape <strong>of</strong> the signal pulse. From<br />

these measurements, the absorptive and refractive transients <strong>of</strong> the probe field can


4.3. TRANSIENTS MEASUREMENTS 106<br />

Phase shift (rad.)<br />

Absorption ( αL/2)<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

1.5<br />

1<br />

0.5<br />

0<br />

0 50 100 150<br />

Time ( µ sec.)<br />

Re(χ p ) and Im(χ p )<br />

0 10 20<br />

Raman Detuning (kHz)<br />

30<br />

P s =4.5 mW<br />

P s =2.5 mW<br />

P s =1.5 mW<br />

P s =0.8 mW<br />

P s =0.4 mW<br />

P s =0.2 mW<br />

500 550 600 650 700<br />

Figure 4.9: Change in refraction and absorption as a function <strong>of</strong> time for different values <strong>of</strong> peak<br />

signal power. Both refraction and absorption are measured relative to their respective values<br />

when the signal is <strong>of</strong>f.<br />

be extracted. Figure 4.9 shows the change in probe absorption and probe phase<br />

for different choices <strong>of</strong> peak signal power. These curves were extracted from the<br />

transient homodyne measurements identically to the process used to obtain Fig.<br />

4.7. The inset in Fig. 4.9 shows the cw absorption and dispersion curves from<br />

Fig. 4.7 in the absence <strong>of</strong> a signal . The vertical lines in the inset show the Stark<br />

shifts corresponding to several signal powers.<br />

This inset is intended as a tool<br />

to help visualize the change in the real and imaginary parts <strong>of</strong> the susceptibility<br />

corresponding to a given signal power and Stark shift.<br />

There are several characteristics <strong>of</strong> these transients worth noting. First, the


4.3. TRANSIENTS MEASUREMENTS 107<br />

steady-state changes in absorption and refraction correspond well with the crossings<br />

in the inset where vertical Stark shift lines cross the susceptibility. For example,<br />

the change in refraction reaches a maximum near P s = 2.5 mW, and for<br />

signal powers greater than P s = 2.5 mW the steady-state change in refraction<br />

decreases. Also, for small signal powers (e.g. P S = 0.2 mW and P S = 0.4 mW)<br />

the steady state change in refraction is larger than the change in absorption.<br />

This corresponds with the fact that near Raman resonance the real part <strong>of</strong> the<br />

susceptibility is linear in Raman detuning while the imaginary part is quadratic.<br />

However, we will see that there is also a linear change in the absorption due to<br />

the signal field’s contribution to decoherence (i.e. Γ(Ω S ) = Γ(0) + 2Ω 2 S γ ⊥/∆ 2 S ).<br />

Finally, for small signal powers (i.e. in the linear regime with δ R ≪ R) the rise<br />

time is a constant, but in the nonlinear regime δ R ∼ R corresponding to large<br />

signal powers the rise times become slightly shorter and the oscillatory nature <strong>of</strong><br />

the transients becomes more noticeable.<br />

Additional insight comes from considering the dynamics <strong>of</strong> the GSC. In Fig.<br />

4.10 we have used the inverse mapping <strong>of</strong> Eq. (2.44) to obtain the measured<br />

GSC transients. The real part <strong>of</strong> the GSC is plotted versus the imaginary part <strong>of</strong><br />

the GSC parametrically as a function <strong>of</strong> time. The rise time dynamics look very<br />

similar to transient trajectory shown in the theory plot in Fig. 2.7. The fall time<br />

dynamics are essentially straight lines back towards the dark state GSC. In Sec.<br />

4.4 we return to this plot and compare it to the theory from chapter 3.


4.3. TRANSIENTS MEASUREMENTS 108<br />

0.5<br />

0.4<br />

<strong>Dynamics</strong> <strong>of</strong> Ground State Coherence<br />

Rise <strong>Dynamics</strong><br />

Im<br />

ρ<br />

( 21 (-)<br />

/ρ 21<br />

)<br />

0.3<br />

0.2<br />

0.1<br />

P s =4.5 mW<br />

0<br />

P s =2.5 mW<br />

P s =1.5 mW Fall <strong>Dynamics</strong><br />

P s =0.4 mW<br />

-0.1<br />

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8<br />

ρ 21<br />

( )<br />

Re (-)<br />

/ρ 21<br />

Figure 4.10: Trajectories <strong>of</strong> the ground state coherence GSC in the complex plane. GSC was<br />

derived using Eq. 4.1 and the refraction and absorption transients in Fig. 4.9. Steady state<br />

GSC values (black line) and transient data (non-black textured lines) are both shown. The<br />

curved/spiraling lines show the rise time dynamics (δ R ≠ 0), while the straighter lines show the<br />

fall time dynamics (δ R = 0).


4.3. TRANSIENTS MEASUREMENTS 109<br />

200<br />

100<br />

Rise time ( µ sec.)<br />

50<br />

20<br />

10<br />

5<br />

2<br />

theory τ (Ref..)<br />

rise<br />

theory τ (Amp.)<br />

rise<br />

meas. τ (Ref.)<br />

rise<br />

meas. τ (Ref.)<br />

fall<br />

meas. τ (Amp.)<br />

rise<br />

meas. τ (Amp.)<br />

fall<br />

0.5 1 2 5 10 20<br />

Stark shift (kHz)<br />

Figure 4.11: Rise times <strong>of</strong> the refraction and absorption resulting from the EIT <strong>Kerr</strong> nonlinearity<br />

(P c = 340). Squares and triangles show the measured refractive rise and fall times. Diamonds<br />

and circles show the measured absorptive rise and fall times, and solid curves are theory for the<br />

rise times. The values ˜αL = 3.75 and R + Γ = 2π kHz are used to obtain the theory curve fit.<br />

Figure 4.11 shows measured 1/e rise and fall times for the transient measurements<br />

presented in Fig. 4.9 corresponding the the first row <strong>of</strong> Table 4.1. The<br />

solid lines are theoretical rise times from Fig. 3.3 modified to account for optical<br />

thickness [i.e. the refractive rise times are multiplied by (1 + ˜αL/4) and the<br />

absorptive rise times are multiplied by 1 + ˜αLc 1 /(2 + 2c 1 )]. The fall times are<br />

essentially independent <strong>of</strong> the size <strong>of</strong> the Stark shift as expected because δ R = 0<br />

always when the signal field is <strong>of</strong>f. The theory curves agree very well with the<br />

measured rise times, except for the absorptive rise and fall times for small Stark<br />

shifts. For small Stark shifts the change in absorption is very small (see Fig. 4.9)<br />

making this discrepancy <strong>of</strong> little practical importance.


4.3. TRANSIENTS MEASUREMENTS 110<br />

100<br />

(a)<br />

(b)<br />

Rise & fall times (µsec.)<br />

10<br />

2<br />

100<br />

10<br />

2<br />

αL ~ = 1.7<br />

R+Γ = 12 kHz<br />

~ αL = 4.3<br />

R+Γ = 12 kHz<br />

1 10<br />

(c)<br />

~ αL = 2.6<br />

R+Γ = 12 kHz<br />

~ αL = 4.6<br />

R+Γ = 20 kHz<br />

1 10<br />

Raman detuning (kHz)<br />

(d)<br />

Figure 4.12: Rise and fall time measurements for the conditions (a) T = 48 ◦ C and P c = 1 mW,<br />

(b) T = 52 ◦ C and P c = 1 mW, (c) T = 58 ◦ C and P c = 1 mW, and (d) T = 58 ◦ C and<br />

P c = 1.9 mW. The definitions for measured data and curve fits are the same as in Fig. 4.11,<br />

and the parameters used in the curve-fits are given in each figure.<br />

Figure 4.12 shows the rise time plots for the data sets corresponding to the<br />

second through fifth rows <strong>of</strong> Table 4.1. For each data set in Fig. 4.12, we include<br />

the EIT half width at half maximum HWHM=R + Γ, and EIT depth which were<br />

used to calculate the theory curves.<br />

Finally, Fig. 4.13 shows the measured and theoretical rise times in the linear<br />

regime as a function <strong>of</strong> optical thickness. The measured refractive rise and fall<br />

times fit the theory curve well. For small optical depths (i.e. αL < 6) the<br />

absorptive rise time agrees with theory. However, the absorptive fall times differ<br />

from the theory, and both absorptive rise and fall times disagree with theory for<br />

large optical depths. It is not clear why these differences exist. However, these


4.4. DISCUSSION OF MEASUREMENTS 111<br />

100<br />

Rise Time (µs.)<br />

80<br />

60<br />

40<br />

20<br />

Ref. rise (meas.)<br />

Ref. rise (theory)<br />

Ref. fall (meas.)<br />

Ref. fall (theory)<br />

Abs. rise (meas.)<br />

Abs. rise (theory)<br />

Abs. fall (meas.)<br />

Abs. fall (theory)<br />

0<br />

0 2 4 6 8 10 12 14 16<br />

<strong>Optical</strong> Depth (αL)<br />

Figure 4.13: Rise and fall times for the refractive and absorptive parts <strong>of</strong> the EIT <strong>Kerr</strong> nonlinearity<br />

versus optical depth. The Stark shift is small (δ R ≪ (R + Γ)) for all measurements and<br />

theory.<br />

discrepancies only effect the absorptive rise time and they are not important for<br />

applications based on the refractive EIT <strong>Kerr</strong> nonlinearity.<br />

4.4 Discussion <strong>of</strong> Measurements<br />

The measured EIT <strong>Kerr</strong> transients show good agreement with the theory derived<br />

in chapters 2 and 3. However, there are a few small discrepancies between theory<br />

and experiment, which have rather simple explanations. They mostly arise from<br />

either approximations in the theory or the fact that in the experimental quantities<br />

are averaged over spatially varying intensities (i.e. the experiment uses gaussian<br />

beams rather the plane waves assumed in the theory).<br />

Most <strong>of</strong> the approximations made in the theory are well justified, but our


4.4. DISCUSSION OF MEASUREMENTS 112<br />

Im<br />

ρ<br />

( 21 (-)<br />

/ρ 21<br />

)<br />

Measured GSC <strong>Dynamics</strong><br />

0.6<br />

0.4<br />

0.2<br />

0<br />

(a)<br />

0 0.2 0.4 0.6 0.8<br />

ρ<br />

( 21<br />

)<br />

Re (-)<br />

/ρ 21<br />

Im<br />

ρ<br />

( 21 (-)<br />

/ρ 21<br />

)<br />

Theoretical GSC <strong>Dynamics</strong><br />

0.6<br />

0.4<br />

0.2<br />

0<br />

(b)<br />

0 0.2 0.4 0.6 0.8<br />

ρ<br />

( 21<br />

)<br />

Re (-)<br />

/ρ 21<br />

Figure 4.14: (a) measured transient (solid line with dots) and steady state (solid line) GSC and<br />

(b) transient (dashed line and steady state (solid line) GSC calculated using the theory from<br />

Chapter 3. Both plots show the normalized real part <strong>of</strong> the GSC plotted versus the imaginary<br />

part <strong>of</strong> the GSC. The GSC’s are plotted parametrically as a function <strong>of</strong> time (transient data),<br />

or as a function <strong>of</strong> Raman detuning (steady state data). The transient data show GSC starting<br />

out in steady state for zero Raman detuning, and evolving toward steady state when the Raman<br />

detuning is δ R ≈ 1.2(R + Γ). After which the Raman detuning is sharply switched back to zero,<br />

and the GSC again evolves toward steady state.<br />

choice <strong>of</strong> neglecting the increased decoherence created by the signal field [i.e.<br />

Γ(Ω S ) ≈ Γ(0) + Ω 2 Sγ/4∆ 2 S] (4.3)<br />

leads to some rather noticeable differences between theory and experiment . First,<br />

in Fig. 4.3 it is obvious that increased decoherence at large signal powers makes<br />

the transparency resonance broader and shallower.<br />

The effects <strong>of</strong> the signal dependent decoherence are also obvious in Fig. 4.10,<br />

Neglecting the signal dependent decoherence is a good approximation because it doesn’t<br />

change the predictions fundamentally and it makes the theory much simpler. However, this<br />

omission results in noticeable but small inaccuracy when the approximate theory is compared<br />

to the exact theory or experiment.


4.4. DISCUSSION OF MEASUREMENTS 113<br />

where this decoherence makes the steady state GSC non-circular. Figure 4.14<br />

compares the measured GSC (a) to the idealized theory (b).<br />

By ignoring the<br />

signal dependent decoherence, the theory predicts that the steady state GSC<br />

[Fig. 4.14(b) solid line] makes a circle. The measurements show a deformed<br />

circle that comes to a cusp near Raman resonance. An exact theory including the<br />

signal-dependent decoherence agrees with these measurements.<br />

The transients<br />

curves for theory and experiment are also slightly different due to the omission <strong>of</strong><br />

signal dependent decoherence in the theory. In spite <strong>of</strong> these small differences the<br />

measured and theory curves look remarkably similar.<br />

The EIT <strong>Kerr</strong> theory assumes plane waves, but the experiment uses approximately<br />

Gaussian beams. Many <strong>of</strong> the EIT <strong>Kerr</strong> predictions for plane waves must<br />

be modified when we average over spatially varying field intensities. For example,<br />

when plane waves are considered the EIT rise time has a very simple relation to<br />

the EIT linewidth–i.e<br />

τ Ref = 1 + ˜αL/4<br />

HW HM , (4.4)<br />

where the EIT half width at half maximum is HWHM = R + Γ. However, when<br />

we average the rise dynamics become<br />

∫ ∫<br />

Re[χ p (t)] ≈<br />

{ [<br />

dxdyRe[χ SS<br />

p (x, y)] 1 − exp − 1 + ˜αL/4 ]}<br />

, (4.5)<br />

R(x, y) + Γ


4.4. DISCUSSION OF MEASUREMENTS 114<br />

and EIT line shape becomes<br />

Im[χ p (δ R )] ≈ α p<br />

∫ ∫<br />

δ 2 R<br />

dxdy<br />

δR 2 + R2 (x, y) . (4.6)<br />

These modification require that an extra multiplicative factor f is added to Eq.<br />

(4.4), such that<br />

τ Ref = f 1 + ˜αL/4<br />

HW HM . (4.7)<br />

Calculating the factor f is further complicated by the fact that spatial diffusion<br />

coarse grains the spatial averaging. The characteristic diffusion length is l D (x, y) =<br />

√¯vl/R(x, y), where ¯v is mean velocity, l is the mean free path length between<br />

Rb 87 –N 2 collisions. Most atoms in a characteristic volume ‖ encounter the same<br />

average field intensity during an optical pumping period regardless <strong>of</strong> variations<br />

in the field intensities such that<br />

∫ ∫<br />

R(x, y) ≈<br />

dx ′ dy ′ Ω2 (x ′ , y ′ )<br />

γ eff<br />

exp [ (x − x ′ ) 2 + (y + y ′ ) 2] /2l D 2 (x, y). (4.8)<br />

A couple <strong>of</strong> simple observations allow us to calculate f for our experimental<br />

conditions.<br />

In the experiment we have described transit-time broadening the<br />

dominant source <strong>of</strong> decoherence Γ, which means that the diffusion length is related<br />

to the beam waist by R/Γ ∼ (w p /l D ) 2 (the probe beam waist w p is the smallest<br />

‖ The characteristic volume is defined simply as a cube with sides given by the characteristic<br />

diffusion length.


4.4. DISCUSSION OF MEASUREMENTS 115<br />

beam waist). Thus, when R ∼ Γ, the diffusion length is approximately equal to<br />

the beam diameter and f ≈ 1. By comparing the measured rise times with the<br />

measured linewidths, we are able to find f empirically. For coupling powers <strong>of</strong><br />

P c = 0.34 mW, P c = 1 mW, and P c = 1.9 mW, we find that f ≈ 1, f ≈ 1.3,<br />

f ≈ 1.4 respectively. This agrees with the fact that as the optical pumping rate<br />

increases, the diffusion length decreases and the value <strong>of</strong> f should monotonically<br />

approaches its maximum value. The maximum value <strong>of</strong> f ≈ 1.5 can be calculated<br />

using spatial averaging while assuming an infinitesimal diffusion length ∗∗ .<br />

Finally, regardless <strong>of</strong> our efforts to decrease the decoherence Γ and increase the<br />

optical pumping rate R, the best transparency we could achieve is ˜αL ≈ 5/6αL.<br />

We believe that the maximum transparency was limited by the fact that 4 out <strong>of</strong><br />

5 <strong>of</strong> the dark states are only partially dark states. For a ratio <strong>of</strong> Ω 2 P /Ω2 c = 0.2, the<br />

partially dark states have a transparency <strong>of</strong> about ˜αL ≈ 7/8αL. In the experiment<br />

there is about a 2 kHz decoherence rate due to limited dwell time for the atoms<br />

in the laser beams.<br />

∗∗ In order that calculate f = 1.5 we had to make a few educated assumptions about the<br />

overlap <strong>of</strong> the probe and coupling beams.


116<br />

Chapter 5<br />

EIT <strong>Kerr</strong> nonlinearity with slow<br />

signal group velocity<br />

In optically thick media (αL ≫ 1), the dynamics <strong>of</strong> the EIT <strong>Kerr</strong> effect depends<br />

on the relative group velocities <strong>of</strong> the probe and signal fields. In Chapter 3,<br />

when the EIT <strong>Kerr</strong> dynamics and the SNR for QND measurement were derived<br />

for optically thick media, it was assumed that the signal group velocity v g,s was<br />

close to c and that the EIT medium was short relative to the duration <strong>of</strong> the<br />

signal pulse (i.e. L ≪ v g,s τ sig ). In this chapter, we discuss how the group velocity<br />

<strong>of</strong> the signal affects the EIT <strong>Kerr</strong> dynamics and the SNR for QND measurements.<br />

The most important effect we discuss is that the maximum SNR and XPM are<br />

achieved when the signal group velocity is slowed till it equals the slow-light group<br />

velocity <strong>of</strong> the probe.<br />

Although matching the probe and signal group velocities is a necessary condition<br />

for achieving the maximum SNR and XPM, it is not a sufficient condition.<br />

To illustrate this point, consider the case when the probe and coupling fields are


5.1. SLOW LIGHT EIT KERR NONLINEARITY 117<br />

matched pulses. In this case, rather than slowing the signal’s group velocity, the<br />

probe’s group velocity is increased to approximately c; thus achieving matched<br />

group velocities. Although the group velocities are matched, the desired increase<br />

in phase shift and SNR is not achieved because the signal group velocity is not<br />

matched to the EIT <strong>Kerr</strong> dynamics.<br />

To achieve the effects we discuss in this<br />

chapter, the signal group velocity must be slowed to the slow-light group velocity<br />

<strong>of</strong> the probe.<br />

In Sec. 5.1, we generalize the EIT <strong>Kerr</strong> dynamics theory from Chapter 3 to<br />

account for the group velocities. In Secs. 5.2 and 5.3, we present the details <strong>of</strong><br />

experimental observations for the EIT <strong>Kerr</strong> dynamics when the group velocity <strong>of</strong><br />

the signal is slow.<br />

5.1 Slow Light EIT <strong>Kerr</strong> nonlinearity<br />

Our analysis <strong>of</strong> the EIT <strong>Kerr</strong> effect and dynamics in Chapter 3 led us to the<br />

conclusion that under the most ideal conditions, QND measurements using the<br />

EIT <strong>Kerr</strong> nonlinearity could only achieve a SNR <strong>of</strong> less than one per photon and a<br />

phase shift <strong>of</strong> a few hundredths <strong>of</strong> a radian per photon in the signal field. However,<br />

in Chapter 3 we did not consider the possibility <strong>of</strong> changing the group velocity for<br />

the signal.<br />

We recall from Chapter 3 that the QND signal is the integrated charge<br />

〈Q〉/¯n s = 2q ele η¯n p C 1 / √ κ 2 , (5.1)


5.1. SLOW LIGHT EIT KERR NONLINEARITY 118<br />

where ¯n p ∝ T int is the expectation value for the number <strong>of</strong> probe photons in the<br />

integration time T int , C 1 = ¯φ/¯n s ∝ 1/T int is the XPM per signal photon and ¯φ is<br />

the average XPM over the integration time. If the integration time is sufficiently<br />

long, the signal will be independent <strong>of</strong> the EIT <strong>Kerr</strong> dynamics.<br />

However, the<br />

signal to noise ratio<br />

SNR ≡<br />

〈Q〉<br />

√<br />

〈(∆Q)2 〉 ≈ 2C 1¯n s<br />

√ η¯np ∝ 1/ √ T int , (5.2)<br />

is not independent <strong>of</strong> the dynamics. In Chapter 3, we assumed the impulse response<br />

function h(t) for the XPM had a width proportional to the refractive EIT<br />

<strong>Kerr</strong> rise-time τ eit = (1 + ˜αL/2)/(R + Γ), which is true when signal pulse propagates<br />

with group velocity v g,s ≈ c. Recall that the impulse response function was<br />

defined such that the time dependent phase modulation is<br />

φ(L, t) ∝<br />

∫ t<br />

0<br />

dt ′ h(t ′ )I s (0, t − t ′ ), (5.3)<br />

where I s (z, t) is the signal intensity. However, when the probe and signal group<br />

velocities are similar, the impulse response function and the required integration<br />

time can change significantly in optically thick media.<br />

To calculate the XPM impulse response function, we consider the linearized


5.1. SLOW LIGHT EIT KERR NONLINEARITY 119<br />

equation <strong>of</strong> motion for the GSC ρ 12 = ρ ∗ 21<br />

∂<br />

(<br />

)<br />

∂t ρ 12 = Rρ (−)<br />

12 − (iδ R (Ω S ) + R + Γ)ρ 12 , (5.4)<br />

and the Maxwell wave equations for the fields in the SVEA are<br />

( )<br />

∂<br />

∂t + v ∂<br />

g,s Ω S = 0, (5.5)<br />

∂z<br />

and<br />

( ∂<br />

∂t + c ∂ ∂z<br />

)<br />

Ω P ≈ iω 0<br />

2 (χ pΩ P )<br />

≈ −˜αc ( Ω P − ρ 12 Ω 2 /Ω ∗ C)<br />

. (5.6)<br />

In the last line <strong>of</strong> Eq. (5.6) we have used the fact that<br />

χ P ≈ i N|µ 32| 2<br />

ɛ 0 γ<br />

(<br />

1 − ρ12 Ω 2 /Ω ∗ CΩ P<br />

)<br />

. (5.7)<br />

Obtaining a general analytic solution for these equations is difficult, but they<br />

are straightforward to simulate numerically. Figure 5.1 shows several simulations<br />

<strong>of</strong> Eqs. (5.4)-(5.6) for different signal group velocities. The signal intensity (top)<br />

and probe phase (bottom) are plotted as a function <strong>of</strong> the time they exit the<br />

medium. The group velocity <strong>of</strong> the probe and all other parameters except v g,s


5.1. SLOW LIGHT EIT KERR NONLINEARITY 120<br />

Normalized Signal<br />

Intensity (arb. units)<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

v g,s = 25v g,p<br />

v g,s = 2v g,p<br />

v g,s = v g,p<br />

v g,s = v /2 g,p<br />

Probe Phase (rad.)<br />

0<br />

-0.2<br />

-0.4<br />

-0.6<br />

-0.8<br />

0 L/v<br />

g,p<br />

Time (arbitrary units)<br />

2L/v g,p<br />

Figure 5.1: Numerical simulations <strong>of</strong> the EIT <strong>Kerr</strong> dynamics for various ratios <strong>of</strong> probe and<br />

signal group velocities. The signal intensity and probe phase are both plotted as a function <strong>of</strong><br />

the time they exit the EIT <strong>Kerr</strong> medium.<br />

were held constant. The optical depth was ˜αL = 270, and the width <strong>of</strong> the signal<br />

pulse was F W HM = 10R.<br />

When the group velocities are matched, the phase pulse is approximately the<br />

same shape as the signal intensity, and the the largest peak phase modulation has<br />

is obtained. When the group velocities are not matched, the edges <strong>of</strong> the phase<br />

pulse correspond to times t 1 = t 0 + L/v g,p and t 2 = t 0 + L/v g,s where t 0 is the<br />

time that the signal peak enters the medium. Thus, the peak height <strong>of</strong> the phase<br />

pulse is roughly proportional to ∣ ∣ t1 − t 2 −1 ∣<br />

= ∣vg,p − v g,s L ∣ ∣ ∣ , for ∣t1 − t 2 ≥ τsig .<br />

This result can be derived analytical by further simplifying Eqs. (5.4)-(5.6) under<br />

the assumptions <strong>of</strong> the linear regime (δ R ≪ 1), a weak probe field (Ω C ≫ Ω P ),


5.1. SLOW LIGHT EIT KERR NONLINEARITY 121<br />

no decoherence (Γ = 0), and Raman resonance in the absence <strong>of</strong> the signal field<br />

[δ R (Ω s = 0) = 0]. Also, for simplicity we assume that at the medium entrance,<br />

the fields Ω P (0, t) and Ω C (0, t) are real positive constants. These assumptions allow<br />

us to write Ω P (z, t) ≈ exp(iφ(z, t))Ω P (0, t) and ρ 21 (z, t) ≈ exp(iθ(z, t))|ρ (−)<br />

21 |,<br />

where |φ(z, t)| ≪ 1 and |θ(z, t)| ≪ 1 such that Eqs. (5.4)-(5.6) simplify to<br />

∂<br />

∂t θ ≈ R (φ − θ)) − δ R, (5.8)<br />

δ R (z, t) = Ω2 s(0, t − z/v g,s )<br />

4∆ s<br />

(5.9)<br />

and<br />

( ∂<br />

∂t + c ∂ )<br />

φ ≈ − ˜αc (φ − θ) . (5.10)<br />

∂z 2<br />

These equations can be thought <strong>of</strong> as describing two separate processes; the<br />

generation <strong>of</strong> the probe phase and the propagation <strong>of</strong> the probe phase. By analyzing<br />

these two processes individually and then recombining them, it is straightforward<br />

to calculate an optically thick impulse response function h thick (t).<br />

First, let us consider the propagation <strong>of</strong> the probe phase modulation in the<br />

absence <strong>of</strong> the signal field. With δ R = 0 Eq. (5.8) can be integrated exactly to<br />

obtain<br />

(<br />

θ(z, t) ≈ R φ − 1 ∂φ<br />

R ∂t + 2 )<br />

∂ 2 φ<br />

R 2 ∂t + . . . . (5.11)<br />

2


5.1. SLOW LIGHT EIT KERR NONLINEARITY 122<br />

Plugging Eq. (5.11) into Eq. (5.10) and simplifying gives to first order<br />

(<br />

∂<br />

n g,p<br />

∂t + c ∂ )<br />

φ = 0, (5.12)<br />

∂z<br />

where the group index for the probe is n g,p = 1 + c˜α/2R and the probe group<br />

velocity is v g,p = c/n ∗ g,p . By leaving out higher order terms we are neglecting<br />

absorption and higher order dispersion terms. To understand the broadening <strong>of</strong> a<br />

phase pulse we consider that the index <strong>of</strong> refraction for a Lorentzian EIT resonance<br />

is<br />

√<br />

1 + χp ≈ 1 + χ p /2 = 1 + αc (<br />

− δ )<br />

R<br />

2ω p R + δ3 R<br />

+ i αc δR<br />

2<br />

R 3 2ω p R + . . . , (5.13)<br />

2<br />

and the different orders <strong>of</strong> dispersion † are given by<br />

β j = ∂j<br />

ω n(ω). (5.14)<br />

∂ωj Agrawal [138] has shown that the pulse broadening due to dispersion alone is<br />

∗ It is interesting to note that nominally the group and phase velocities have reversed roles.<br />

The phase modulation propagates at the group velocity, while the probe energy propagates at<br />

the phase velocity. Typically we think <strong>of</strong> the energy propagating at the group velocity. This<br />

reversal <strong>of</strong> roles is in name only and the situation we describe is consistent with what is known<br />

<strong>of</strong> the group and phase velocities. However, this nominally role reversal illustrates the fact that<br />

one must be explicit and careful when discussing the velocity <strong>of</strong> light in EIT systems.<br />

† Technically these dispersion terms apply to amplitude pulses and not phase pulses. However,<br />

our assumption <strong>of</strong> 1 ≫ φ ensures that on average ∂ 2 φ/∂t j ≫ (∂φ/∂t) j for i ≥ 2, which means<br />

that to a good approximation these expressions also apply to the dispersion <strong>of</strong> a phase pulse.


5.1. SLOW LIGHT EIT KERR NONLINEARITY 123<br />

given by<br />

[ ( 2<br />

∆t(z)<br />

∆t(0) = β2 z<br />

1 +<br />

+<br />

2∆t (0)) 1 ( ) ] 2 1/2<br />

β3 z<br />

, (5.15)<br />

2 2 4∆t 3 (0)<br />

where the pulse width ∆t is the temporal variance <strong>of</strong> the pulse. Similarly, the<br />

broadening due to absorption alone is<br />

[<br />

] 1/2<br />

∆t(z)<br />

∆t(0) ≈ αz<br />

1 +<br />

, (5.16)<br />

2∆t 2 (0)R 2<br />

where both Eqs. (5.15) and (5.16) assume Gaussian beams. In the limit <strong>of</strong><br />

large optical depth the dispersive pulse broadening can be minimized by choosing<br />

∆t(0) = (3˜αL/4) 1/3 /R, which gives a dispersive broadened pulse width <strong>of</strong><br />

∆t(z) = ∆t(0) √ 3/2. For this initial pulse width, absorptive broadening will be<br />

the dominant form <strong>of</strong> broadening for large optical thicknesses, and the final pulse<br />

width will be ∆t(L) ≈ √˜αL/R. This is a lower bound for the pulse width after<br />

propagating through a Lorentzian EIT medium regardless <strong>of</strong> the initial pulse<br />

shape if the phase modulation.<br />

When the Stark shift from the signal field is included in Eq. (5.8), it acts as<br />

a source term for the phase modulation φ. To understand the generation <strong>of</strong> the<br />

phase pulse, it is simplest to think about an optically thin medium αL ≈ 1. In<br />

an optically thin medium we can ignore the probe phase φ in Eq. (5.8), and the


5.1. SLOW LIGHT EIT KERR NONLINEARITY 124<br />

medium phase θ is calculated by integrating Eq. (5.8) to obtain<br />

θ(z, t) ≈<br />

∫ ∞<br />

0<br />

δ R (z, t − τ) exp(−Rτ)dτ. (5.17)<br />

We are primarily interested in the case when v g,s ∼ v g,p such that v g,s ≪ c, and<br />

this allows us to ignore the time derivative in Eq. (5.10). Thus, Eq. (5.10) can<br />

be integrated in space to obtain<br />

∫ z<br />

φ 0 (z, t) ≈ ˜α θ(z ′ , t) exp [−˜α(z − z ′ )/2] dz ′ (5.18)<br />

2 0<br />

≈ ˜α ∫ z ∫ ∞<br />

e−˜αz/2 dz ′ e˜αz′ /2<br />

dτe −Rτ δ R (z ′ , t − τ)<br />

2<br />

≈<br />

∫ ∞<br />

−∞<br />

0<br />

0<br />

dτh(z, τ)δ R (t − τ, 0),<br />

where the impulse response function <strong>of</strong> the medium is<br />

⎧⎪<br />

h(z, t) = 1 − exp [−˜αz 0/2(1 + v g,p /v g,s )] ⎨ exp[− ˜α v g,p<br />

2 v g,s<br />

(z 0 − z)]<br />

1 + v g,p /v g,s ⎪ ⎩ exp [ − ˜α (z − z 2 0) ]<br />

0 < z ≤ z 0 , L<br />

,<br />

z 0 < z ≤ L<br />

(5.19)<br />

and where z 0 = tv g,s . Equation (5.19) is approximately equal to Eq. (3.38) in the<br />

limit v g,s ≫ v g,p .<br />

To generalize our results to the case <strong>of</strong> optically thick media, we rewrite the


5.1. SLOW LIGHT EIT KERR NONLINEARITY 125<br />

wave equation for the probe phase [i.e. Eq. (5.20)] to include a source term:<br />

(<br />

∂<br />

n g,p<br />

∂t + c ∂ )<br />

φ = P φ , (5.20)<br />

∂z<br />

where P φ (z, t) ∝ ∫ dz ′ h m (z ′ , t)δ R (z − z ′ , t) is the source term analogous to the<br />

polarization <strong>of</strong> the medium in a wave equation for the electric field. Also,<br />

h m (z, t) = 4˜α exp (−˜α|v g,st − z|/2) (5.21)<br />

is the optically thin impulse response function when the group velocities are<br />

matched (i.e. v g,s = v g,p ). Using Eqs. (5.9) and (5.21) and making the change<br />

<strong>of</strong> variables ξ = z and τ = t − z/v g,p allows us to integrate Eq. (5.20) exactly to<br />

obtain<br />

φ(L, t) = ω pχ (3) ∫<br />

pssL<br />

c 2 ɛ 0<br />

dt ′ h thick (L, t ′ )I s (0, t − t ′ ) (5.22)<br />

where<br />

h thick (L, t) ≈<br />

∫ L−t vg,p<br />

vg,s<br />

vg,p L−t v g,p<br />

dξ Rv g,p exp ( − ˜α 2 |ξ|)<br />

2(v g,p − v g,s )L . (5.23)<br />

It is interesting to note that in the limit<br />

lim h thick (L, t) = h m (L, t). (5.24)<br />

v g,s →v g,p<br />

The integration time must be as large as the final pulse width [i.e.<br />

T int ≈


5.2. EXPERIMENTAL APPARATUS 126<br />

(2 + √˜αL)/(R + Γ)], which results in a SNR <strong>of</strong><br />

SNR<br />

n s<br />

≈<br />

2γ √ 2˜αLη<br />

∆ S k 0<br />

√<br />

A(1 + κ2 ))<br />

√<br />

τEIT<br />

∆t(L)<br />

(5.25)<br />

≈<br />

2γ √ 2˜αLη<br />

∆ S k 0<br />

√<br />

A(1 + κ2 )) (˜αL)1/4 /2.<br />

5.2 Experimental apparatus<br />

To observe the EIT <strong>Kerr</strong> dynamics described in Sec. 5.1 and in Chapter 3, we used<br />

the experimental system shown in Fig. 5.2. This is essentially the same experimental<br />

apparatus used in Chapter 4, except the probe-coupling EIT is resonant<br />

with Rb 87 instead <strong>of</strong> Rb 85 and we have added a signal-coupling field to form an<br />

EIT pair with the signal field. The four laser frequencies used in the experiment<br />

are the probe, coupling, signal, and signal-coupling fields. Figure 5.3 shows these<br />

frequencies relative to the D 1 transitions <strong>of</strong> Rb 87 and Rb 85 .<br />

The probe and coupling beams form an EIT pair for Rb 87 , and they are close to<br />

Raman resonance for the Rb 87 (F = 2 ↔ F ′ = 2 ↔ F = 1) two-photon transition<br />

as shown in Fig. 5.3(a). In addition to Raman resonance condition, EIT also<br />

requires strong phase correlations between the probe and coupling fields, which<br />

we achieve by deriving both probe and coupling beam from the same laser. The<br />

laser is tuned to the coupling frequency, and part <strong>of</strong> the beam is picked <strong>of</strong>f and sent<br />

through a 6.8 GHz electro-optic modulator (EOM) to create the probe frequency.


5.2. EXPERIMENTAL APPARATUS 127<br />

Homodyne<br />

Detection<br />

PD<br />

PD<br />

50/50<br />

BS<br />

Probe<br />

50/50<br />

BS<br />

Signal-<br />

Coupling<br />

Computer<br />

DAQ<br />

Spectral<br />

filter<br />

PBS<br />

Magnetic<br />

Shielding<br />

PBS<br />

Signal<br />

Coupling<br />

PZT<br />

APD<br />

Rb Cell<br />

Rb Cell<br />

Probe<br />

Signal-<br />

Coupling<br />

Coupling<br />

Signal<br />

Figure 5.2: Diagram <strong>of</strong> the experiment. At the first 50/50 beam splitter (BS) the probe field<br />

is divided into two arms <strong>of</strong> a Mach-Zehnder interferometer. In one arm <strong>of</strong> the interferometer,<br />

the probe interacts with the coupling, signal, and signal-coupling fields in a rubidium cell.<br />

These fields are then filtered using polarization (PBS), spatial (pick-<strong>of</strong>f mirror), and spectral<br />

filters before the probe field is interfered on a 50/50 BS with the reference arm <strong>of</strong> the Mach-<br />

Zehnder interferometer. Balanced homodyne detection is recorded via a computer controlled<br />

data acquisition (DAQ) system to measure changes in the probe phase. (PD=photo diodes,<br />

APD= avalanche photo diode, BS=beam splitter, PBS=polarizing beam splitter, PZT= piezoelectric<br />

actuated mirror)


5.2. EXPERIMENTAL APPARATUS 128<br />

After the EOM, the probe field is obtained by frequency filtering ‡ to select only<br />

the lower side band <strong>of</strong> the frequency modulated field. We found that the best<br />

contrast for EIT <strong>of</strong> the probe is achieved when the probe beam diameter is small<br />

compared to the coupling beam diameter, and when coupling and probe beams<br />

are perfectly parallel to within experimental accuracy (i.e. slightly better than<br />

1 mrad). Also, EIT is best when the coupling and probe fields have orthogonal<br />

linear polarizations. The beam areas and their overlap are shown in Fig. 5.4.<br />

The signal and signal-coupling beams create EIT in the Rb 85 atoms, and are<br />

detuned from single-photon resonance as shown in Fig. 5.3(a). They are also<br />

derived from a single laser to ensure strong phase correlations. The signal-coupling<br />

frequency is obtained by upshifting the laser frequency by 3.2 GHz using a double<br />

passed 1.6 GHz acousto-optic modulator (AOM). The signal frequency is obtained<br />

by upshifting the laser frequency using a 200 MHz AOM. This AOM is also used<br />

to pulse the signal field with a minimum rise-time <strong>of</strong> 30 nsec. Both lasers used in<br />

the experiment were Newfocus Vortex lasers. Similar to the probe and coupling<br />

beams, the signal and signal-coupling also have orthogonal linear polarizations,<br />

and have parallel propagation directions to within experimental accuracy. Also,<br />

the signal beam was smaller than the signal-coupling to ensure good contrast for<br />

the signal EIT (see Fig. 5.4).<br />

The signal and probe converge in the cell at an angle <strong>of</strong> 7 mrad and overlap<br />

‡ The frequency filter is a temperature controlled solid etalon with finesse <strong>of</strong> 100 and free<br />

spectral range <strong>of</strong> 40 GHz.


5.2. EXPERIMENTAL APPARATUS 129<br />

F’=2<br />

F’=1<br />

(a)<br />

F’=2,3<br />

Rb 85<br />

Rb 87 Coupling<br />

Probe<br />

F=2<br />

Signal<br />

F=3<br />

Signal-<br />

Coupling<br />

F=1<br />

F=2<br />

Transmission<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

Rb 87<br />

Rb 85<br />

Probe<br />

Coupl.<br />

Signal<br />

Signal-<br />

Coupl.<br />

F=2<br />

(b)<br />

F=3<br />

F=2<br />

-5 -4 -3 -2 -1 0 1 2 3 4 5<br />

Frequency - 377,107.46 (GHz)<br />

F=1<br />

Figure 5.3: (a) Level diagram and (b) laser frequencies relative to the Rb 85 and Rb 87 D 1 Doppler<br />

valleys.


5.2. EXPERIMENTAL APPARATUS 130<br />

1<br />

2<br />

Beam Circumfrance ( 1/e )<br />

Sig.Coupl .<br />

Signal<br />

Probe<br />

Coupl.<br />

Position (mm)<br />

0<br />

-1<br />

-2<br />

-2 -1 0 1 2<br />

Position (mm)<br />

Figure 5.4: 1/e 2 intensity beam radius for the probe, coupling, signal, and signal-coupling<br />

beams. The probe and signal beams were smaller than the coupling and signal-coupling beams<br />

respectively in order to ensure good EIT for the entire probe and coupling beams. The probe<br />

beam is also smaller than the signal beam in order to obtain a uniform Stark shift.<br />

through the rubidium cell. The vapor cell contains both rubidium isotopes in their<br />

natural abundance (72% Rb 85 and 28% Rb 87 ). It also contains 20 Torr <strong>of</strong> nitrogen<br />

(N 2 ) buffer gas. The rubidium cell was a 30 mm long cylinder with a diameter<br />

<strong>of</strong> 25.4 mm, and was magnetically shielded. For the experiments discussed here,<br />

the optical depth on resonance with the Rb 87 (F = 2 → F ′ = 1) transition (i.e.<br />

the probe transition) is αL ≈ 7. For the Rb 85 (F = 3 → F ′ = {2, 3}) transition,<br />

(i.e. the signal transition) the on-resonance optical depth is αL ≈ 40. The large<br />

optical depth for Rb 85 transitions is why the signal and signal-coupling fields are<br />

not on resonance [see Fig. 5.3(b)].


5.3. SLOW SIGNAL EIT KERR MEASUREMENTS 131<br />

Probe Absorption (αL/2)<br />

3<br />

2<br />

1<br />

signal<br />

no sig.<br />

coupl. sig.<br />

coupl. no sig.<br />

0<br />

-150 -100 -50 0 50 100 150<br />

Probe Raman Detuning (kHz)<br />

Figure 5.5: Probe absorption as a function <strong>of</strong> Raman detuning.<br />

5.3 Slow signal EIT <strong>Kerr</strong> measurements<br />

Figure 5.5 shows the probe absorption as a function <strong>of</strong> Raman detuning for four<br />

different cases: 1) both signal and signal-coupling <strong>of</strong>f, 2) signal on and signalcoupling<br />

<strong>of</strong>f, 3) signal <strong>of</strong>f and signal-coupling on, and 4) both signal and signalcoupling<br />

on. When the signal-coupling is <strong>of</strong>f, the group velocity <strong>of</strong> the signal is<br />

approximately c, and when the signal-coupling is on, the signal group velocity<br />

is approximately c × 10 −4 . Ideally, slowing the signal group velocity would be<br />

the only effect <strong>of</strong> the signal-coupling field, but Fig. 5.5 shows that the signalcoupling<br />

field also has the side effects <strong>of</strong> making the probe EIT become shallower


5.3. SLOW SIGNAL EIT KERR MEASUREMENTS 132<br />

Signal Absorption ( αl/2 )<br />

and Refraction ( ω [n-1]/c )<br />

s<br />

1<br />

0.5<br />

0<br />

αL/2<br />

ω (n-1)/c<br />

p<br />

-0.5<br />

-200 -100 0 100 200<br />

Signal Raman Detuning (kHz)<br />

Figure 5.6: Measured signal absorption αL as a function <strong>of</strong> Raman detuning. Also, plotted is<br />

the dispersion curve calculated as the Hilbert transform <strong>of</strong> the signal absorption.<br />

and Stark shifting the Raman resonance for the probe and coupling fields. The<br />

shape <strong>of</strong> the probe EIT resonance is qualitatively the same with and without the<br />

signal coupling field making qualitative comparisons possible. However, a direct<br />

quantitative comparison is difficult due to the signal-coupling’s effects on the line<br />

shape.<br />

In addition to the absorption curves shown in Fig. 5.5, we used homodyne<br />

detection to measure the probe phase as a function <strong>of</strong> Raman detuning (not<br />

shown). In the absences <strong>of</strong> the signal-coupling the dispersion was measured to<br />

have a slope <strong>of</strong> about 5×10 −2 rad/kHz and with signal-coupling on the slope was<br />

3×10 −2 rad/kHz. From these values we infer a probe group velocity <strong>of</strong> about<br />

4-6 km/s depending on whether the signal-coupling is on or <strong>of</strong>f.


5.3. SLOW SIGNAL EIT KERR MEASUREMENTS 133<br />

Figure 5.6 shows the signal absorption (blue) as a function <strong>of</strong> the Raman<br />

detuning for the signal and signal-coupling frequencies. The dispersion curve (red)<br />

was calculated by taking the Hilbert transform <strong>of</strong> the absorption curve. The EIT<br />

for the signal resulted in a signal group delay <strong>of</strong> about 1.5 µsec with about 80%<br />

transmission <strong>of</strong> the signal energy. For a 30 mm long vapor cell, this group delay<br />

implies a signal group velocity <strong>of</strong> 20 km/s. When we tried to obtain larger signal<br />

group delays by tuning the signal frequency closer to resonance EIT for both the<br />

signal and probe deteriorated quickly with only a marginal increase in the signal<br />

group delay. We believe this was due to radiation trapping in Rb 85 resulting from<br />

the combination <strong>of</strong> large optical density for Rb 85 and the larger beam diameters<br />

for the signal and signal-coupling beams.<br />

Thus, in the experiment, the signal<br />

group velocity was limited to the regime v g,s ∼ 5v g,p and the probe optical depth<br />

was limited to ˜αL ∼ 3 − 6.<br />

For our transient measurements <strong>of</strong> the EIT <strong>Kerr</strong> nonlinearity, we optimized the<br />

Raman detunings for both EIT pairs to obtain maximum transmission and then<br />

pulsed the signal field. Figure 5.7 shows the measured probe phase in the absence<br />

<strong>of</strong> a signal-coupling field [shown in Fig. 5.7(a)]. Also, shown in Fig. 5.7(b) is the<br />

signal intensity at the output <strong>of</strong> the rubidium cell. Time t = 0 is determined as<br />

the expectation time for the signal pulse to leave the vapor cell. For long signal<br />

pulses the probe phase has essentially the same pulse shape as the signal intensity–<br />

only slightly wider and delayed. For short signal pulses, the leading edge <strong>of</strong> the


5.3. SLOW SIGNAL EIT KERR MEASUREMENTS 134<br />

Phase Shift<br />

(rad.)<br />

0.2<br />

0.1<br />

(a)<br />

Signal Intensity<br />

(arb. units)<br />

0<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

-60<br />

-40 -20 0 20 40 60 80 100<br />

time (µsec.)<br />

(b)<br />

Figure 5.7: (a) measured EIT <strong>Kerr</strong> phase pulses and (b) measured signal intensity for different<br />

signal widths.<br />

phase pulse is approximately the integral <strong>of</strong> the signal intensity, and the falling<br />

edge <strong>of</strong> the phase pulse is a decaying exponential as predicted by Eq. 3.39. The<br />

probe phase data is also slightly rounded and delayed due to the transimpedance<br />

amplifier used after the homodyne detection. The transimpedance amplifier has<br />

a measured propagation delay <strong>of</strong> 2 µs and a rise-time <strong>of</strong> 5 µs.<br />

Deconvolving the effects <strong>of</strong> the transimpedance amplifier, we found that the<br />

1/e fall-time <strong>of</strong> the shortest phase pulses is about 13 µs, which is consistent with<br />

the observations from Chapter 4, where τ eit = f(1 + αL/4)/HW HM and where<br />

f ≈ 1.4, αL = 6, and R = 2π × 45 kHz .<br />

We repeated the measurements shown in Fig.<br />

5.7 with the signal-coupling


5.3. SLOW SIGNAL EIT KERR MEASUREMENTS 135<br />

30<br />

w/o Sig.-Coupl.<br />

Phase Shift (mrad.)<br />

20<br />

10<br />

with Sig.-Coupl.<br />

0<br />

-10 0 10 20 30 40 50 60 70<br />

Time (µsec.)<br />

Figure 5.8: Measured (dots) and theory (solid) for the EIT <strong>Kerr</strong> phase pulses plotted versus<br />

time (with t = 0 defined as the expectation time for the signal pulse leaving the vapor cell).<br />

field on. The effects <strong>of</strong> the slow signal group velocity are most noticeable for short<br />

signal pulses. Figure 5.8 shows the probe phase as a function <strong>of</strong> time when the<br />

signal-coupling is either on or <strong>of</strong>f (i.e. the signal group velocity is v g,s ≈ 2×10 4 m/s<br />

or v g,s ≈ c). The signal pulse width before the cell is τ sig = 7 µs, and t = 0 is the<br />

expectation time for the signal pulse leaving the cell. The dots are the measured<br />

probe phase and the solid lines are numerical simulations <strong>of</strong> Eqs. (5.4-5.6), where<br />

the simulations used the measured parameters from the experiment–e.g. optical<br />

depths, EIT <strong>Kerr</strong> rise-time, signal absorption etc. The good agreement between<br />

measurements and the numerical simulations give us confidence in the theory.<br />

There are two types <strong>of</strong> effects due to the signal-coupling: effects due to the<br />

slow signal group velocity, and effects due to changes in the probe EIT line shape.


5.3. SLOW SIGNAL EIT KERR MEASUREMENTS 136<br />

Probe Phase (mrad.)<br />

20<br />

15<br />

10<br />

5<br />

v g,s = 0.5vg,p<br />

v g,s = v g,p<br />

v g,s = 5vg,p<br />

v = 20v<br />

g,s<br />

g,p<br />

0<br />

-20 -10 0 10 20 30 40<br />

Time (µsec.)<br />

Figure 5.9: Numerical simulations <strong>of</strong> EIT <strong>Kerr</strong> phase pulse for different signal group velocities.<br />

For these simulation ˜αL = 3, τ s ig=7 µs, τ eit = 1/R = 15 µs.<br />

The slower signal group velocity causes the peak phase to advance in time. Also,<br />

for slow enough signal group velocities, the rising slope <strong>of</strong> the probe phase will<br />

become shallower and the falling slope will become slightly steeper. The signalcoupling<br />

field also has the side effect <strong>of</strong> changing to the EIT line shape for the<br />

probe EIT. In Fig.<br />

5.8 these side effects make the phase pulse shorter with a<br />

slightly longer fall-time.<br />

Although experimental limitations prevented matching the signal and probe<br />

group velocities, using numerical simulations it is possible to explore the parameter<br />

space for matched group velocities. Figure 5.9 shows the probe phase for numerical<br />

simulations using the same parameters as the experiment when the signal-coupling


5.3. SLOW SIGNAL EIT KERR MEASUREMENTS 137<br />

is on (i.e. τ eit =15 µs, αL = 3, τ sig =7 µs, v g,p = 6 km/s etc.) with different group<br />

velocities for the signal pulse. Phase pulses are plotted for four different signal<br />

group velocities. The v g,s = 20v g,p and v g,s = 5v g,p are qualitatively similar to<br />

the measured probe phase in Fig. 5.7. For slower group velocities with v g,s =<<br />

5v g,p , one can see the rising slope <strong>of</strong> the phase becomes shallower and the phase<br />

pulse becomes more symmetric about its peak. For this relatively small optical<br />

thickness, matching the group velocities does not increase the peak phase shift or<br />

decrease the optimal integration time as it would in an optically thicker sample.<br />

In order to achieve the benefits <strong>of</strong> slow group velocities, the EIT medium<br />

must be optically thick. For matched group velocities in optically thick media<br />

we expect an integration time <strong>of</strong> T match<br />

int<br />

≈ (2 + √˜αL)/R. While for fast signal<br />

group velocities (i.e. v g,s ≈ c), the integration time is T fast<br />

int<br />

≈ (1 + ˜αL/4)/R. For<br />

˜αL < 10 the integration time is shorter for a fast signal group velocity than for<br />

matched group velocities.<br />

Figure 5.10 shows the pulse-width <strong>of</strong> the phase plotted as a function <strong>of</strong> the<br />

ratio v g,s /v g,p , when the medium’s optical thickness is ˜αL = 3, ˜αL = 6, and<br />

˜αL = 12. For the optical depth used in the experiment (i.e. ˜αL = 3), there is<br />

no narrowing <strong>of</strong> the phase pulse FWMH when group velocities become matched § .<br />

However, for large optical thicknesses, the pulse-width decreases when the group<br />

velocities are matched. However, even for ˜αL = 12 the decrease in the pulse-width<br />

§ This is just a restatement <strong>of</strong> what is already obvious from Fig. 5.7


5.4. DISCUSSION 138<br />

35<br />

FWHM <strong>of</strong> Probe phase pulse<br />

30<br />

25<br />

20<br />

15<br />

αL = 3<br />

αL = 6<br />

αL = 12<br />

10<br />

0.5 1 2 5 10 20<br />

Normalized group velocity <strong>of</strong> signal (v /v g,p)<br />

g,s<br />

Figure 5.10: Phase pulse width (FWHM) plotted versus the ratio between the group velocities<br />

for the probe and signal fields. Numerical simulations were performed with three optical depths<br />

˜αL for the EIT. The signal pulse width was 7 µs and the optically thin EIT rise-time was<br />

τ eit = 1/R = 15 µs.<br />

is still less than a factor <strong>of</strong> two. Thus, the optical thickness <strong>of</strong> the medium must<br />

be very large before matched group velocities significantly improves the SNR for<br />

QND measurements.<br />

5.4 Discussion<br />

For QND measurement using the EIT <strong>Kerr</strong> nonlinearity, it is theoretically possible<br />

to improve the SNR by matching the group velocities for the probe and signal<br />

fields. However, the advantages <strong>of</strong>fered by matched group velocities only become<br />

significant for very large optical thicknesses (i.e. αL > 30).<br />

We have observed some <strong>of</strong> the effects <strong>of</strong> nearly matched group velocities using


5.4. DISCUSSION 139<br />

EIT in two isotopes <strong>of</strong> rubidium. However, we have not been able to experimentally<br />

obtain the large optical depths and slow signal group velocities required to<br />

see decreased pulse widths for the probe phase modulation on the probe and large<br />

improvements in the SNR for QND measurements.<br />

It may be possible to improve on this experiment using a different distribution<br />

<strong>of</strong> rubidium isotopes and/or a longer rubidium cell in order to decrease the<br />

radiation trapping. Also, using cold atoms would eliminate the problems associated<br />

with overlapping Doppler valleys. However, using cold atoms would require<br />

trapping both isotopes <strong>of</strong> rubidium simultaneously. The benefits achieved by a<br />

better matched group velocity experiment probably would not justify the additional<br />

work.


140<br />

Chapter 6<br />

Conclusions and perspective<br />

As mentioned in the introduction, EIT enhanced <strong>Kerr</strong> nonlinearities may enable<br />

many exciting and challenging few-photon applications such as QND measurements<br />

with single photon resolution. However, some <strong>of</strong> the early optimism regarding<br />

the EIT <strong>Kerr</strong> effect should be tempered. Without resorting to optical<br />

cavities and/or slow group velocities for the signal field, the EIT <strong>Kerr</strong> effect can<br />

only achieve a maximum cross-phase modulation (XPM) <strong>of</strong> a few hundredths <strong>of</strong> a<br />

radian per photon under ideal conditions ∗ . Also under ideal conditions, the signal<br />

to noise ration (SNR) for QND detection <strong>of</strong> a single photon is less than one for<br />

the EIT <strong>Kerr</strong> effect † .<br />

The reasons for these limitations is the slow rise-time <strong>of</strong> the EIT <strong>Kerr</strong> effect. In<br />

∗ For the ideal conditions we assume a single velocity class <strong>of</strong> atoms with the atomic linewidth<br />

equal to the natural linewidth, the dipole moment for the signal field is for the D 2 line <strong>of</strong> an<br />

alkali atom. We also assume no decoherence, the fields are focused down to the diffraction limit,<br />

and the N system originally proposed by Schmidt and Imamoğlu [1].<br />

† This result assumes a constraint <strong>of</strong> at least 90% transmission for the signal field


141<br />

the linear EIT <strong>Kerr</strong> regime (i.e. δ R ≪ R) the rise-time for XPM is proportional to<br />

the depth <strong>of</strong> the EIT ˜αL and the optical pumping time t R ≡ 1/R [i.e. τ ref = (1 +<br />

˜αL/4)/(R + Γ) ≈ ˜αL/4R, where it is assumed that the medium is optically thick<br />

(i.e. ˜αL ≫ 1) and the EIT condition R ≫ Γ is met]. For cw applications <strong>of</strong> the<br />

EIT <strong>Kerr</strong> effect it is theoretically possible to make the nonlinear susceptibility (i.e.<br />

χ (3)<br />

pss ∝ ˜αL/R) arbitrary large by either increasing the optical depth <strong>of</strong> the medium<br />

and/or decreasing the optical pumping rate. However for pulsed applications such<br />

as the few-photon applications that interest us, the figure <strong>of</strong> merit is the ratio<br />

between the nonlinear susceptibility and the rise-time, and this figure <strong>of</strong> merit<br />

cannot be increased by changing either the optical depth or the optical pumping<br />

rate.<br />

The EIT <strong>Kerr</strong> nonlinearity is still one <strong>of</strong> the best, if not the best, candidate for<br />

realizing few-photon applications, but researchers must be clever in their methods<br />

<strong>of</strong> utilizing the EIT <strong>Kerr</strong> effect. We have investigated using a slow group velocity<br />

for the signal field, and found that theoretically this method could be used to<br />

achieve arbitrarily large XPM and SNR for QND measurements. However, experimentally<br />

matching the signal and probe group velocities is challenging due to complications<br />

in creating EIT simultaneously for two nearly resonant atomic species.<br />

Also, matched group velocity experiments require very large optical depths before<br />

significant improvements in XPM and SNR can be observed. These large optical<br />

depths also pose significant experimental problems.


142<br />

Recently there have been a couple <strong>of</strong> proposals to create matched signal and<br />

probe group velocities using a single atomic species [58, 119]. These eliminate<br />

some <strong>of</strong> the problems we encountered in our two atomic species experiment, but<br />

they still require extremely large optical thicknesses to achieve their objectives.<br />

Also, they introduce new experimental problems by demanding precise control <strong>of</strong><br />

magnetic fields.<br />

There are other proposals using EIT and slow light to achieve strong <strong>Kerr</strong><br />

nonlinearities. For example, Friedler et al. have proposed using the long range<br />

dipole-dipole interactions between two slow-light pulses, where the slow light is<br />

due to EIT in a ladder system [83]. The large dipole moments required for this<br />

interaction result from the fact that the highest energy level in the ladder level<br />

system is a Rydberg state with large quantum number n.<br />

Some <strong>of</strong> the best results in XPM and QND measurements have used cavity<br />

QED [139, 140]. In fact, QND measurements <strong>of</strong> single photons have very recently<br />

been realized using microwave photons and super-conducting cavities [140]. However,<br />

in these experiments the photon can only be coupled into the cavity via the<br />

spontaneous emission from an excited atom travelling through the cavity. The<br />

photon remains trapped in the cavity until it is either absorbed by a probe atom<br />

traversing the cavity (more likely), or the photon leaks out <strong>of</strong> the cavity (less<br />

likely). This type <strong>of</strong> QND measurement is not useful for measuring the number<br />

<strong>of</strong> photons in a travelling wave.


143<br />

Figure 6.1: Conceptual diagram for an EIT <strong>Kerr</strong> experiment using cold atoms and an optical<br />

cavity.<br />

The best QND measurements <strong>of</strong> photon number for optical fields have also used<br />

optical cavities, but not in the strong-coupling regime <strong>of</strong> Cavity QED [133, 134].<br />

In these experiments a conventional <strong>Kerr</strong> nonlinearity was used with an optical<br />

cavity to increase the intensity <strong>of</strong> the signal field interacting with the atoms ‡ .<br />

In EIT <strong>Kerr</strong> experiments an optical cavity could also be used to increase the<br />

XPM. This could be accomplished in an experiment like the one shown in Fig.<br />

6.1. In Fig. 6.1 two red detuned laser create a dipole trap for a cold atom cloud.<br />

The EIT fields (i.e. probe and coupling) enter from the top, and the signal field<br />

passes through a confocal Fabry-Perot etalon with finesse § Q. The atomic cloud<br />

is at the center <strong>of</strong> the optical cavity. The signal intensity inside the cavity is Q<br />

‡ This conventional <strong>Kerr</strong> nonlinearity is simply a three-level ladder system, and is exactly<br />

identical to the system Schmidt and Imamoğlu used to compare with the EIT <strong>Kerr</strong> nonlinearity<br />

and demonstrate 10 orders <strong>of</strong> magnitude increase in the size <strong>of</strong> the <strong>Kerr</strong> nonlinearity.<br />

§ The symbol Q is used to signify the quality factor <strong>of</strong> the cavity, which is the same thing a<br />

finesse.


144<br />

times larger than outside the cavity resulting in a XPM that is Q times larger.<br />

For QND measurements the results from chapter 3 for the rubidium D 1 line can<br />

be generalized to account for a cavity, such that Eq. (3.43) becomes<br />

SNR<br />

n s<br />

≈<br />

√ √<br />

πη 1 Qγ ˜αL<br />

√ , (6.1)<br />

2 k 0 A ∆ s 1 + κ 1<br />

where it is assumed that T int = τ eit . The nondemolition condition also must be<br />

modified such that ∆/γ ≫ √ Q˜αL/(1 + κ 1 ). Unlike the no-cavity scenario considered<br />

in chapter 3, the SNR can theoretically be increased indefinitely without<br />

violating the nondemolition condition by simultaneously increasing Q and ∆ while<br />

keeping the ratio ∆/ √ Q constant.<br />

Using realistic parameters for the experiment shown in Fig. 6.1 it should be<br />

possible to realize a QND measurement capable <strong>of</strong> resolving individual number<br />

states. Consider the case <strong>of</strong> a cold atomic cloud <strong>of</strong> Rb 87 , with EIT on the D 1<br />

line and the signal field on the D 2 line. Assuming the parameters are A = 8π 2 λ 2 0,<br />

κ 1 = 5, η = 0.7, Q = 10 4 , ˜αL = αL = 4 , and ∆ s ≈ 600γ, then the signal<br />

absorption is about 15% and the SNR for a single photon <strong>of</strong> about eight. Thus, it<br />

is realistically possible to achieve single photon QND measurements. To achieve<br />

this same result by matching group velocities would require an optical depth <strong>of</strong><br />

˜αL ≈ 10 9 with perfect transparency for the probe field, which is not realistic.<br />

We are assuming that the atomic cloud has a length <strong>of</strong> 0.5 mm with a diameter <strong>of</strong> 0.1mm<br />

and about 10 5 atoms. The 5 to 1 aspect ratio <strong>of</strong> the atom cloud keeps the absorption <strong>of</strong> the<br />

probe field to a minimum while keeping αL large for the signal field.


145<br />

Finally, we return to the question <strong>of</strong> using the EIT <strong>Kerr</strong> nonlinearity as an<br />

optical two-qubit quantum logic gate. As we have already mentioned in chapter 1,<br />

there are several architectures for optical quantum computing that are based on<br />

<strong>Kerr</strong> nonlinearities and/or QND measurements [111–116, 141]. However, Shapiro<br />

has shown theoretically that optical gates based on <strong>Kerr</strong> nonlinearities have poor<br />

fidelity [120, 121]. This is in contrast to other researchers who have used a completely<br />

quantum analysis to predict good fidelity using the EIT <strong>Kerr</strong> effect in a<br />

quantum phase gate [118, 142].<br />

It is interesting to contrast the conclusions <strong>of</strong> Refs. [120] and [118]. In Ref.<br />

[118] the conclusion is, that their “study shows that the implementation <strong>of</strong> efficient<br />

EIT-based nonlinear two-qubit gates for traveling single photons is possible” [118].<br />

While Ref. [120], which is titled Single-photon <strong>Kerr</strong> nonlinearities do not help<br />

quantum computation, comes to the conclusion that “Unfortunately, the phase<br />

noise that is associated with the causal, noninstantaneous nature <strong>of</strong> the response<br />

function precludes high-fidelity operation” <strong>of</strong> EIT-based nonlinear two-qubit gates<br />

[120]. Both references use a completely quantum analysis, and both account for<br />

the dynamics <strong>of</strong> the <strong>Kerr</strong> nonlinearity, although their methods are significantly<br />

different. Ref. [120] derives general analytic expressions for <strong>Kerr</strong> nonlinearities<br />

in general, and Ref.<br />

[118] uses numerical simulations <strong>of</strong> the EIT <strong>Kerr</strong> effect<br />

specifically.<br />

The crux <strong>of</strong> the problem lies with understanding the dynamics <strong>of</strong> the <strong>Kerr</strong>


146<br />

nonlinearity and noise in the <strong>Kerr</strong> nonlinearity due to vacuum fluctuations. In<br />

this work we have discussed the dynamics <strong>of</strong> the EIT <strong>Kerr</strong> nonlinearity at length,<br />

but we have not discussed the vacuum fluctuations because our treatment is semiclassical.<br />

It is well known that the quantum vacuum AC Stark shifts the energy levels;<br />

this energy shift is typically referred to as the Lamb shift. Thus, the noise degrading<br />

the fidelity <strong>of</strong> quantum logic gates based on the EIT <strong>Kerr</strong> nonlinearity arises<br />

from fluctuations in the Raman detuning due to the Lamb shift. To our knowledge<br />

the affects <strong>of</strong> the Lamb shift noise on EIT have not been studied, but such<br />

a study should answer questions raised about the fidelity <strong>of</strong> EIT <strong>Kerr</strong> quantum<br />

gates and their utility for quantum computing and QIP. This noise may also be<br />

significant for other EIT <strong>Kerr</strong> applications.


147<br />

Bibliography<br />

[1] H. Schmidt and A. Imamoǧlu, “Giant <strong>Kerr</strong> nonlinearities obtained by electromagnetically<br />

induced transparency,” Opt. Lett. 21, 1936 (1996).<br />

[2] G. Alzetta, A. Gozzini, and G. Orriols, “An experimental method for the<br />

observations <strong>of</strong> R. F. transitions and laser beam tansitions in oreinted Na<br />

vapors,” Nuovo Cimento B 36, 5 (1976).<br />

[3] G. Alzetta, L. Moi, and G. Orriols, “Nonabsorption hyperfine resonances in<br />

a sodium vapor irradiated by a multimode dyelaser,” Nuovo Cimento B 52,<br />

5209 (1979).<br />

[4] H. R. Gray and C. R. Stround, “Coherent trapping <strong>of</strong> atomic populations,”<br />

Opt. Lett. 3, 218 (1978).<br />

[5] S. E. Harris, J. E. Field, and A. Imamoğlu, “Nonlinear optical processes<br />

using electromagnetically induced transparency,” Phys. Rev. Lett. 64, 1107<br />

(1990).<br />

[6] E. Arimondo, in Progress in Optics, edited by E. Wolf (Elsevier Science,<br />

Amsterdam, 1996), Vol. 35, p. 275.<br />

[7] O. A. Kocharovskaya and Y. I. Khanin, “Coherent amplification <strong>of</strong> an ultrashortpulse<br />

in a three-level medium without a population inversion,” JETP<br />

Lett. 48, 630 (1988).<br />

[8] M. O. Scully, S.-Y. Zhu, and A. Gavrielides, “Degenerate quantum-beat<br />

laser: Lasing without inversion and inversion without lasing,” Phys. Rev.<br />

Lett. 62, 2813 (1989).<br />

[9] S. E. Harris, “Lasers without inversion: Interferences <strong>of</strong> lifetime broadened<br />

resonances,” Phys. Rev. Lett. 62, 1033 (1989).<br />

[10] E. S. Fry et al., “Atomic coherence effects within the sodium D 1 line: Lasing<br />

without inversion via population trapping,” Phys. Rev. Lett. 70, 3235<br />

(1993).


BIBLIOGRAPHY 148<br />

[11] Y. Zhu, “Lasing without inversion in a V-type system: Transient and steadystate<br />

analysis,” Phys. Rev. A 53, 2742 (1995).<br />

[12] P. Hemmer, S. Ezekiel, and J. C.C. Leiby, “Stabilization <strong>of</strong> a microwave<br />

oscillator using a resonance raman transition in a sodium beam,” Opt. Lett.<br />

8, 440 (1983).<br />

[13] M. Scully and M. Fleischhauer, “High-sensitivity magnetometer based on<br />

index-enhanced media,” Phys. Rev. Lett. 69, 1360 (1992).<br />

[14] M. Fleischhauer and M. Scully, “Quantum sensitivity limits <strong>of</strong> an optical<br />

magnetometer based on atomic phase coherence,” Phys. Rev. A 49, 1973<br />

(1994).<br />

[15] D. Budker, V. Yashchuk, and M. Zolotorev, “Nonlinear magneto-optic effects<br />

with ultranarrow widths,” Phys. Rev. Lett. 81, 5788 (1998).<br />

[16] M. Linberg and R. Binder, “Dark states in coherent semiconductor spectroscopy,”<br />

Phys. Rev. Lett. 75, 1403 (1995).<br />

[17] R.Wynands and A. Nagel, “Precision spectroscopy with coherent dark<br />

states,” Appl. Phys. B 68, 1 (1999).<br />

[18] F. Levi, A. Godone, J. Vanier, S. Micalizio, and G. Modugno, “Line-shape<br />

<strong>of</strong> dark line and maser emission pr<strong>of</strong>ile in CPT,” Eur. Phys. J. D 12, 53<br />

(2000).<br />

[19] S. Brandt, A. Nagel, R. Wynands, and D. Meschede, “Buffer-gas-induced<br />

linewidth reduction <strong>of</strong> coherent dark resonances to below 50 Hz,” Phys. Rev.<br />

A 56, R1063 (1997).<br />

[20] M. Erhard and H. Helm, “Buffer-gas effects on dark resonances: Theory and<br />

experiment,” Phys. Rev. A 63, 043813 (2001).<br />

[21] D. Budker, D. F. Kimball, S. M. Rochester, and V. V. Yashchuk, “Nonlinear<br />

magneto-optics and reduced group velocity <strong>of</strong> light in atomic vapor with slow<br />

ground state relaxation,” Phys. Rev. Lett. 83, 1767 (1999).<br />

[22] L. Hau, S. Harris, Z. Dutton, and C. Behroozi, “Light speed reduction to<br />

17 metres per second in an ultracold atomic gas,” Nature 397, 594 (1999).<br />

[23] M. Kash et al., “Ultraslow group velocity and enhanced nonlinear optical<br />

effects in a coherently driven hot atomic gas,” Phys. Rev. Lett. 82, 5229<br />

(1999).


BIBLIOGRAPHY 149<br />

[24] C. Liu, Z. Dutton, C. H. Behroozi, and L. V. Hau, “Observation <strong>of</strong> coherent<br />

optical information storage in an atomic medium using halted light pulses,”<br />

Nature 409, 490 (2001).<br />

[25] D. Phillips, M. Fleischhauer, A. Mair, R. Walsworth, and M. D. Lukin,<br />

“Storage <strong>of</strong> light in atomic vapor,” Phys. Rev. Lett. 86, 783 (2001).<br />

[26] O. Kocharovskaya, Y. Rostovtsev, and M. O. Scully, “Stopping light via hot<br />

atoms,” Phys. Rev. Lett. 86, 628 (2001).<br />

[27] M. J. S. J. J. Longdell, E. Fravel and N. B. Manson, “Stopped light with<br />

storage times greater than one second using electromagnetically induced<br />

transparency in a solid,” Phys. Rev. Lett 95, 063601 (2004).<br />

[28] M. D. Eisaman, A. Andre, F. Massou, M. Fleischhauer, A. S. Zibrov,<br />

and M. D. Lukin, “<strong>Electromagnetically</strong> induced transparency with tunable<br />

single-photon pulses,” Nature 438, 837 (2005).<br />

[29] P. R. Hemmer, D. P. Katz, J. Donoghue, M. Cronin-Golomb, M. S. Shahriar,<br />

and P. Kumar, “Efficient low-intensity optical phase conjugation based on<br />

coherent population trapping in sodium,” Opt. Lett. 20, 982 (1995).<br />

[30] M. Jain, H. Xia, G. Y. Yin, A. J. Merriam, and S. E. Harris, “Efficient<br />

nonlinear frequency conversion with maximal atomic coherence,” Phys. Rev.<br />

Lett. 77, 4326 (1996).<br />

[31] Y.-Q. Li and M. Xiao, “Enhancement <strong>of</strong> nondegenerate four-wave mixing<br />

based on electromagnetically induced transparency in rubidium atoms,”<br />

Opt. Lett. 21, 1064 (1996).<br />

[32] M. D. Lukin, M. Fleischhauer, A. S. Zibrov, H. G. Robinson, and V. L.<br />

Velichansky, “Spectroscopy in dense coherent media: Line narrowing and<br />

interference effects,” Phys. Rev. Lett. 79, 2959 (1997).<br />

[33] G. Zhang, K. Hakuta, and B. Stoicheff, “Nonlinear optical generation using<br />

electromagnetically induced transparency in atomic hydrogen,” Phys. Rev.<br />

Lett. 71, 3099 (1993).<br />

[34] M. Jain, G. Yin, J. Field, and S. Harris, “Observation <strong>of</strong> electromagnetically<br />

induced phase matching,” Opt. Lett. 18, 998 (1993).<br />

[35] C. Dorman and J. Marangos, “Symmetric photon-photon coupling by atoms<br />

with zeeman-split sublevels,” Phys. Rev. A 58, 4121 (1998).<br />

[36] M. T. Johnsson and M. Fleischhauer, “Quantum theory <strong>of</strong> resonantly enhanced<br />

four-wave mixing: Mean-field and exact numerical solutions,” Phys.<br />

Rev. A 66, 043808 (2002).


BIBLIOGRAPHY 150<br />

[37] L. M. Duann, M. D. Lukin, J. I. Cirac, and P. Zoller, “Long-distance quantum<br />

communication with atomic ensembles and linear optics,” Nature 414,<br />

413 (2001).<br />

[38] D. A. Braje, V. Balić, S. Goda, G. Yin, and S. E. Harris, “Generation <strong>of</strong><br />

paired photons with controllable waveforms,” Phys. Rev. Lett. 93, 183601<br />

(2004).<br />

[39] V. Balić, D. A. Braje, P. Kolchin, G. Yin, and S. E. Harris, “Generation <strong>of</strong><br />

paired photons with controllable waveforms,” Phys. Rev. Lett. 94, 183601<br />

(2005).<br />

[40] P. Kolchin, S. Du, C. Belthangady, G. Yin, and S. E. Harris, “Generation<br />

<strong>of</strong> narrow-bandwidth paired photons: Use <strong>of</strong> a single driving laser,” Phys.<br />

Rev. Lett. 97, 113602 (2006).<br />

[41] C. F. McCormick, V. Boyer, E. Arimondo, and P. D. Lett, “Strong relative<br />

intensity squeezing by four-wave mixing in rubidium vapor,” Opt. Lett. 32,<br />

178 (2007).<br />

[42] M. V. Pack, R. M. Camacho, and J. C. Howell, “Transients <strong>of</strong> the<br />

electromagnetically-induced-transparency-enhanced refractive <strong>Kerr</strong> nonlinearity:<br />

Theory,” Phys. Rev. A 74, 013812 (2006).<br />

[43] S. E. Harris and Y. Yamamoto, “Photon switching by quantum interference,”<br />

Phys. Rev. Lett. 81, 3611 (1998).<br />

[44] N. Mulchan, D. G. Ducreay, R. Pina, M. Yan, and Y. Zhu, “Nonlinear<br />

excitation by quantum interference in a doppler-broadened rubidium atomic<br />

system,” J. Opt. Soc. Am. B 17, 820 (2000).<br />

[45] D. A. Braje, V. Balić, G. Y. Yin, and S. E. Harris, “Low-light-level nonlinear<br />

optics with slow light,” Phys. Rev. A 68, 041801(R) (2003).<br />

[46] H. Wang, D. Goorskey, and M. Xiao, “Controlling light by light with threelevel<br />

atoms inside an optical cavity,” Opt. Lett. 27, 1354 (2002).<br />

[47] H. Kang and Y. Zhu, “Observation <strong>of</strong> large <strong>Kerr</strong> nonlinearity at low light<br />

intensities,” Phys. Rev. Lett. 91, 093601 (2003).<br />

[48] H. Chang, Y. Du, J. Yao, C. Xie, and H. Wang, “Observation <strong>of</strong> crossphase<br />

shift in hot atoms with quantum coherence,” Europhys. Lett. 65, 485<br />

(2004).<br />

[49] M. Yan, E. G. Rickey, and Y. Zhu, “Observation <strong>of</strong> absoprtive photon<br />

switching by quantum interference,” Phys. Rev. A 64, 041801(R) (2001).


BIBLIOGRAPHY 151<br />

[50] Y.-F. Chen, G.-C. Pan, and I. A. Yu, “Transient behaviors <strong>of</strong> photon switching<br />

by quantum interference,” Phys. Rev. A 69, 063801 (2004).<br />

[51] S. J. Park, H. Cho, T. Y. Kwon, and H. S. Lee, “Transient cohernce oscillation<br />

induced by a detuned raman field in a rubidum Λ system,” Phys. Rev.<br />

A 69, 023806 (2004).<br />

[52] A. D. Greentree et al., “Intensity-dependent dispersion under conditions <strong>of</strong><br />

electromagnetically induced transparency in coherently prepared multistate<br />

atoms,” Phys. Rev. A 67, 023818 (2003).<br />

[53] A. B. Matsko, I. Novikova, G. R. Welch, and M. S. Zubairy, “Enhancement<br />

<strong>of</strong> <strong>Kerr</strong> nonlinearity by multiphoton coherence,” Opt. Lett. 28, 96 (2003).<br />

[54] S. Rebic, D. Vitali, and C. Ottaviani, “Quantum theory <strong>of</strong> a polarization<br />

phase gate in an atomics tripod configuration,” Opt. and Spect. 99, 264<br />

(2005).<br />

[55] D. Petrosyan and Y. P. Malakyan, “Magneto-optical rotation and crossphase<br />

modulation via coherently driven four-level atoms in a tripod configuration,”<br />

Phys. Rev. A 70, 023822 (2004).<br />

[56] S. E. Harris and L. V. Hau, “Nonlinear optics at low light levels,” Phys.<br />

Rev. Lett. 82, 4611 (1999).<br />

[57] M. D. Lukin and A. Imamoglu, “Nonlinear optics and quantum entanglement<br />

<strong>of</strong> ultraslow single photons,” Phys. Rev. Lett. 84, 1419 (2000).<br />

[58] D. Petrosyan and G. Kurizki, “Symmetric photon-photon coupling by atoms<br />

with zeeman-split sublevels,” Phys. Rev. A 65, 033833 (2002).<br />

[59] W. Hanle, “über magnetische beeinflussung der polarisation der resonanzfluoreszenz,”<br />

Z. Phys. 30, 93 (1924).<br />

[60] U. Fano, “Effects <strong>of</strong> configuration interaction on intensities and phase<br />

shifts,” Phys. Rev. 124, 1866 (1961).<br />

[61] S. H. Autler and C. H. Townes, “Stark effect in rapidly varying fields,” Phys.<br />

Rev. 100, 703 (1955).<br />

[62] M. Fleischhauer, “<strong>Optical</strong> pumping in dense atomic media: Limitations due<br />

to reabsorption and spontaneously emitted photons,” Europhys. Lett. 45,<br />

659 (1999).<br />

[63] M. Fleischhauer, “<strong>Electromagnetically</strong> induced transparency and coherentstate<br />

preparation in optically thick media,” Opt. Exp. 4, 107 (1999).


BIBLIOGRAPHY 152<br />

[64] A. B. Matsko, I. Novikova, M. O. Scully, and G. R. Welsh, “Radiation<br />

trapping in coherent media,” Phys. Rev. Lett. 87, 133601 (2001).<br />

[65] A. B. Matsko, I. Novikova, and G. R. Welsh, “Radiation trapping under<br />

conditions <strong>of</strong> electromagnetically induced transparency,” J. Mod. Opt. 49,<br />

367 .<br />

[66] K. Bergmann, H. Theuer, and B. W. Shore, “Coherent population transfer<br />

among quantum states <strong>of</strong> atoms and molecules,” Rev. Mod. Phys. 70, 1003<br />

(1998).<br />

[67] N. V. Vitanov, T. Halfmann, B. W. Shore, and K. Bergman, “Laser-induced<br />

population transfer by adiabatic passage techniques,” Annu. Rev. Phys.<br />

Chem. 52, 763 (2001).<br />

[68] S. Rebić, S. M. Tan, A. S. Parkins, and D. F. Walls, “Large <strong>Kerr</strong> nonlinearity<br />

with a single atom,” J. Opt. B: Quant. Semiclass. Opt 1, 490 (1999).<br />

[69] T. Opatrny and D.-G. Welsch, “Coupled cavities for enhancing the crossphase-modulation<br />

in electromagnetically induced transparency,” Phys. Rev.<br />

A 75, 023805 (2001).<br />

[70] S. Rebić, An Investigation <strong>of</strong> Giant <strong>Kerr</strong> Nonlinearity, Ph.D. thesis, University<br />

<strong>of</strong> Auckland, 2002.<br />

[71] S. Harris, “Normal modes fro electromagnetically induced transparency,”<br />

Phys. Rev. Lett. 72, 52 (1994).<br />

[72] S. Harris and Z.-F. Lou, “Preparation energy for electromagnetically induced<br />

transparency,” Phys. Rev. A 52, R928 (1995).<br />

[73] J. H. Eberly and V. Kozlov, “Wave equation for dark coherence in three-level<br />

media,” Phys. Rev. Lett. 88, 243604 (2002).<br />

[74] M. Konopnicki and J. Eberly, “Simultaneous propagation <strong>of</strong> short differentwavelength<br />

optical pulses,” Phys. Rev. A 24, 2567 (1981).<br />

[75] M. Konopnicki, P. Drummond, and J. Eberly, “Theory <strong>of</strong> lossless propagation<br />

<strong>of</strong> simultaneous different-wavelength optical pulses,” Opt. Commun.<br />

36, 313 (1981).<br />

[76] F. Hioe and R. Grobe, “Matched optical solitary waves for three- and fivelevel<br />

systems,” Phys. Rev. Lett. 73, 2559 (1994).<br />

[77] R. Grobe, F. Hioe, and J. Eberly, “Formation <strong>of</strong> shape-preserving pulses<br />

in a nonlinear adiabatically integrable system,” Phys. Rev. Lett. 73, 3183<br />

(1995).


BIBLIOGRAPHY 153<br />

[78] E. Cerboneschi and E. Arimondo, “<strong>Transparency</strong> and dressing for optical<br />

pulse pairs through a double-Λ absorbing medium,” Phys. Rev. A 52, R1823<br />

(1995).<br />

[79] Kasapi, M. Jain, G. Yin, and S. Harris, “<strong>Electromagnetically</strong> induced transparency<br />

propagation dyanimcs,” Phys. Rev. Lett. 74, 2447 (1995).<br />

[80] M. Fleischhauer and A. S. Manka, “Propagation <strong>of</strong> laser pulses and coherent<br />

population transfer in dissipative three-level systems: An adiabatic dressedstate<br />

picture,” Phys. Rev. A 54, 794 (1996).<br />

[81] H. Eleuch and R. Bennaceur, “An optical soliton pair among absorbing<br />

three-level atoms,” J. Opt. A 5, 528 (2003).<br />

[82] A. V. Rybin, I. P. Vadeiko, and A. R. Bishop, “Theory <strong>of</strong> slow-light solitons,”<br />

Phys. Rev. E 72, 026613 (2005).<br />

[83] I. Friedler, G. Kurizki, O. Cohen, and M. Segev, “Spatial Thirring-type<br />

solitons via electromagnetically induced transparency,” Opt. Lett 30, 3374<br />

(2005).<br />

[84] Y.-F. Chen, G.-C. Pan, and I. A. Yu, “Frequency-modulation-induced transient<br />

oscillation in the spectra <strong>of</strong> electromagnetically induced transparency,”<br />

J. Opt. Soc. Am. B 21, 1647 (2004).<br />

[85] A. Godone, S. Micalizio, and F. Levi, “Rabi resonances in the l excitation<br />

scheme,” Phys. Rev. A 66, 063807 (2002).<br />

[86] K. Harada, K. Motomura, T. Koshimizu, H. Ueno, and M. Mitsunaga, “Coherent<br />

Raman beats from dark states,” J. Opt. Soc. Am. B 22, 1105 (2005).<br />

[87] A. Mair and M. D. Lukin, “Phase coherence and control <strong>of</strong> stored photonic<br />

information,” Phys. Rev. A 65, 031802(R) (2002).<br />

[88] P. Valente, H. Failache, and A. Lezama, “Comparative study <strong>of</strong> the transient<br />

evolution <strong>of</strong> Hanle electromagnetically induced transparency and absorption<br />

resonances,” Phys. Rev. A 65, 023814 (2001).<br />

[89] A. S. Zibrov and A. B. Matsko, “<strong>Optical</strong> ramsey fringes induced by zeeman<br />

coherence,” Phys. Rev. A 64, 013814 (2001).<br />

[90] Y. Wu and X. Yang, “<strong>Electromagnetically</strong> induced transparency in V-, Λ-,<br />

and cascade-type schemes beyond steady-state analysis,” Phys. Rev. A 71,<br />

053806 (2005).


BIBLIOGRAPHY 154<br />

[91] E. A. Korsunsky, N. Leinfellner, A. Huss, S. Baluschev, and L. Windholz,<br />

“Phase-dependent electromagnetically induced transparency,” Phys. Rev. A<br />

59, 2302 (1999).<br />

[92] W. Maichen, F. Renzoni, I. Mazets, E. Korsunsky, and L. Windholz, “Transient<br />

coherent population trapping in a closed loop interaction scheme,”<br />

Phys. Rev. A 53, 3444 (1996).<br />

[93] D. Budker, W. Gawlik, D. F. Kimball, S. M. Rochester, V. V. Yashchuks,<br />

and A. Weis, “Resonant nonlinear magneto-optical effects in atoms,” Rev.<br />

Mod. Phys. 74, 1153 (2002).<br />

[94] S. E. Harris, “<strong>Electromagnetically</strong> induced transparency,” Physics Today<br />

50, 36 (1997).<br />

[95] J. P. Marangos, “Topical review electromagnetically induced transparency,”<br />

J. Mod. Opt. 45, 471 (1998).<br />

[96] M. Fleischhauer, A. Imamoǧlu, and J. P. Marangos, “<strong>Electromagnetically</strong><br />

induced transparency: Optics in coherent media,” Rev. Mod. Phys. 77, 633<br />

(2005).<br />

[97] M. D. Lukin, P. Hemmer, and M. O. Scully, in Adv. At., Mol., Opt. Phys.,<br />

edited by B. Bederson and H. Walther (Elsevier Inc., San Diego, 2000),<br />

Vol. 42, p. 347.<br />

[98] N. V. Vitanov, M. Fleischhauer, B. W. Shore, and K. Bergmann, in Adv.<br />

At., Mol., Opt. Phys., edited by B. Bederson and H. Walther (Elsevier Inc.,<br />

San Diego, 2001), Vol. 46, p. 55.<br />

[99] A. B.Matsko, O. Kocharovskaya, Y. Rostovtsev, A. S. Z. G. R. Welch, and<br />

M. O. Scully, in Adv. At., Mol., Opt. Phys., edited by B. Bederson and H.<br />

Walther (Elsevier Inc., San Diego, 2001), Vol. 46, p. 191.<br />

[100] M. Lukin, “Trapping and manipulating photon states in atomic ensembles,”<br />

Rev. Mod. Phys. 75, 457 (2003).<br />

[101] J. Howell and J. Yeazell, “Nondestructive single-photon trigger,” Phys. Rev.<br />

A 62, 032311 (2000).<br />

[102] K. Mattle, H. Weinfurter, P. G. Kwiat, and A. Zeilinger, “Dense coding in<br />

experimental quantum communication,” Phys. Rev. Lett, 76, 4656 (1996).<br />

[103] I. L. Chuang and Y. Yamamoto, “Quantum bit regeneration,” Phys. Rev.<br />

Lett. 76, 4281 (1996).


BIBLIOGRAPHY 155<br />

[104] G. D’Ariano, L. Maccone, M. G. A. Paris, and M. F. Sacchi, “<strong>Optical</strong> fockstate<br />

synthesizer,” Phys. Rev. A 61, 053817 (2000).<br />

[105] J. C. Howell, A. Lamas-Linares, and D. Bouwmeester, “Experimental violation<br />

<strong>of</strong> a spin-1 Bell inequality using maximally entangled four-photon<br />

states,” Phys. Rev. Lett. 88, 030401 (2002).<br />

[106] G. Weihs, T. Jennewein, C. Simon, H. Weinfurter, and A. Zeilinger, “Violation<br />

<strong>of</strong> Bell’s inequality under strict einstein locality conditions,” Phys.<br />

Rev. Lett. 81, 5039 (1998).<br />

[107] M. Rowe et al., “Experimental violation <strong>of</strong> a Bell’s inequality with effiecient<br />

detection,” .<br />

[108] A. N. Boto, P. Kok, D. S. Abrams, S. L. Braunstein, C. P. Williams, and<br />

J. P. Dowling, “Quantum interferometric optical lithography: Exploiting entanglement<br />

to beat the diffraction limit,” Phys. Rev. Lett. 85, 2733 (2000).<br />

[109] E. Knill, R. Laflamme, and G. Milburn, “A scheme for efficient quantum<br />

computation with linear optics,” Nature 409, 46 (2001).<br />

[110] P. Kok, W. J. Munro, K. Nemoto, T. C. Ralph, J. P. Dowling, and G. J.<br />

Milburn, “Linear optical quantum computing with photonic qubits,” Rev.<br />

Mod. Phys. 79, 135 (2007).<br />

[111] G. J. Milburn, “Quantum optical fredkin gate,” Phys. Rev. Lett. 62, 2124<br />

(1989).<br />

[112] I. L. Chuang and Y. Yamamoto, “Simple quantum computer,” Phys. Rev.<br />

A 52, 3489 (1995).<br />

[113] J. C. Howell and J. A. Yeazell, “Quantum computation through entangling<br />

single photons in multipath interferometers,” Phys. Rev. Lett. 85, 198<br />

(2000).<br />

[114] J. C. Howell and J. A. Yeazell, “Nondestructive single-photon trigger,” Phys.<br />

Rev. A 62, 032311 (2000).<br />

[115] G. M. DAriano, C. Macchiavello, and L. Maccone, “Quantum computations<br />

with polarized photons,” Fortschr. Phys. 48, 573 (2000).<br />

[116] K. Nemoto and W. J. Munro, “Nearly deterministic linear optical controlled-<br />

NOT gate,” Phys. Rev. Lett. 93, 250502 (2004).<br />

[117] W. J. Munro, K. Nemoto, and T. P. Spiller, “Weak nonlinearities: A new<br />

route to optical quantum compution,” New J. Phys. 7, 137 (2005).


BIBLIOGRAPHY 156<br />

[118] C. Ottaviani, S. Rebic, D. Vitali, and P. Tombesi, “Quantum phase-gate<br />

operation based on nonlinear optics: Full quantum analysis,” Phys. Rev. A<br />

73, 010301(R) (2006).<br />

[119] Z.-B. Wang, K.-P. Marzlin, and B. C. Sanders, “Large cross-phase modulation<br />

between slow copropagating weak pulses in 87Rb,” Phys. Rev. Lett.<br />

97, 063901 (2006).<br />

[120] J. H. Shapiro, “Single-photon kerr nonlinearities do not help quantum computation,”<br />

Phys. Rev. A 73, 062305 (2006).<br />

[121] J. H. Shapiro and M. Razavi, “Continuous-time cross-phase modulation and<br />

quantum computation,” New J. Phys. 9, 16 (2007).<br />

[122] H. Schmidt and A. Imamoglu, “High-speed properties <strong>of</strong> a phase-modulation<br />

scheme based on electromagnetically induced transparency,” Opt. Lett. 23,<br />

107 (1998).<br />

[123] J.-Q. Shen, Z.-C. Ruan, and S. He, “Influence <strong>of</strong> the singal light on the<br />

transient optical properties <strong>of</strong> a four-level EIT medium,” Phys. Lett. A<br />

330, 487 (2004).<br />

[124] S. Alam, Lasers without inversion and electromagnetically induced transparency<br />

(SPIE, Washington, 1999).<br />

[125] A. Javan, O. Kocharovshaya, H. Lee, and M. O. Scully, “Narrowing <strong>of</strong><br />

EIT resonance in a doppler broadened medium,” Phys. Rev. Lett. 81, 3611<br />

(2001).<br />

[126] M. O. Scully and M. S. Zubairy, Qunatum Optics (Cambridge University<br />

Press, Cambridge, 1997).<br />

[127] R. S. Bennink, Frequency Conversion <strong>of</strong> <strong>Optical</strong> Signals Using Coherently<br />

Prepared Media, Ph.D. thesis, University <strong>of</strong> Rochester, Instute <strong>of</strong> Optics,<br />

2004.<br />

[128] J. Vanier and A. Audoin, The quantum physics <strong>of</strong> atomic frequency starndarads<br />

(Adam Hilger, Philadelphia, 1989).<br />

[129] M. D. Rotondaro and G. P. Perram, “Collisional broadening and shift <strong>of</strong> the<br />

rubidium D 1 and D 2 lines (5 2 s 1/2 → 5 2 p 1/2 , 5 2 p 3/2 ) by rare gases H 2 , D 2 ,<br />

N 2 , CH 4 and CF 4 ,” J. Quant. Spectrosc. Radiat. Transfer 57, 497 (1997).<br />

[130] J. Vanier, J.-F. Simard, and J.-S. Boulanger, “Relaxation and frequency<br />

shifts in the ground state <strong>of</strong> Rb 85 ,” Phys. Rev. A 9, 1031 (1974).


BIBLIOGRAPHY 157<br />

[131] D. Budker and D. F. Kimball, Atomic physics: an exploration through problems<br />

and solutions (Oxford University Press, Oxford, 2004).<br />

[132] F. A. Franz and C. Volk, “Spin relaxation <strong>of</strong> rubidium atoms in sudden<br />

and quasimolecular collisions with light-noble-gas atoms,” Phys. Rev. A 14,<br />

1711 (1976).<br />

[133] P. Grangier, J.-F. Roch, and G. Roger, “Observation <strong>of</strong> backaction-evading<br />

measurement <strong>of</strong> an optical intensity in a three-level atomic nonlinear system,”<br />

Phys. Rev. Lett. 66, 1418 (1991).<br />

[134] J. P. Poizat and P. Grangier, “experimental realization <strong>of</strong> a quantum optical<br />

tap,” Phys. Rev. Lett. 70, 271 (1993).<br />

[135] J.-P. Poizat, J.-F. Roch, and P. Grangier, “Characterization <strong>of</strong> quantum<br />

nondemolition measurements in optics,” Ann. Phys. Fr. 19, 265 (1994).<br />

[136] H. A. Haus, Electromagnetic Noise and Quantum <strong>Optical</strong> Measurement (Spinger<br />

Verlag, Berlin, 2000).<br />

[137] H. A. Bachor, A Guide to Experiments in Qunatum Optics (Willey, Cambridge,<br />

1999).<br />

[138] G. P. Agarwal, Nonlinear Fiber Optics; Third Edition (Academic Press, San<br />

Diego, 2001).<br />

[139] Q. A. Turchette, C. J. Hood, W. Lange, H. Mabuchi, and H. J. Kimble,<br />

“Measurement <strong>of</strong> conditional phase shifts for quantum logic,” Phys. Rev.<br />

Lett. 75, 4710 (1995).<br />

[140] S. Gleyzes et al., “Quantum jumps <strong>of</strong> light recording the birth and death <strong>of</strong><br />

a photon in a cavity,” Nature 446, 297 (2007).<br />

[141] K. Nemoto and W. J. Munro, “Universal quantum computation on the<br />

power <strong>of</strong> quantum non-demolition measurements,” Phys. Lett. A 334, 104<br />

(2005).<br />

[142] S. Rebic, C. Ottaviani, G. D. Giuseppe, D. Vitali, and P. Tombesi, “Assessment<br />

<strong>of</strong> a quantum phase-gate operation based on nonlinear optics,” Phys.<br />

Rev. A 74, 032301 (2006).<br />

[143] D. Steck, “Rubidium 87 D line data,” , unpublished, available on-line at<br />

http://steck.us/alkalidata, (unpublished).<br />

[144] D. M. Brink and G. R. Satchler, Angular Momentum (Oxford University<br />

Press, Oxford, 1962).


158<br />

Appendix A<br />

Appendix 1: EIT on the<br />

rubidium D1 line<br />

In a complicated level system like the D 1 lines <strong>of</strong> rubidium, EIT can be significantly<br />

more complicated than EIT in three-level systems. The purpose <strong>of</strong> this<br />

appendix is to supply some <strong>of</strong> the missing details for Chapter 3 by discussing<br />

EIT on the Rb 85 and Rb 87 D 1 lines. This appendix can also be looked at as a<br />

template for understanding EIT in other complicated level structures. We restrict<br />

our discussion to the case where the probe and coupling fields have orthogonal<br />

linear polarizations.<br />

A.1 Choice <strong>of</strong> Polarizations<br />

First, there are several reasons for our choice <strong>of</strong> orthogonal linear polarizations.<br />

The primary reason for our choice <strong>of</strong> polarization is that through experimentation<br />

with various polarizations we found that orthogonal linear polarizations for the<br />

probe and coupling fields provided the best contrast for EIT (by EIT contrast


A.1. CHOICE OF POLARIZATIONS 159<br />

we mean that the ratio between absorption on and <strong>of</strong>f Raman resonance was the<br />

greatest). Following our empirical observations, we come up with Table A.1 to<br />

explain the reasons we should have anticipated good EIT contrast for orthogonal<br />

linear polarizations.<br />

Perhaps the most important criteria good EIT contrast is whether a particular<br />

choice <strong>of</strong> polarizations results in parasitic dark states (DS). By parasitic DS we<br />

mean a DS that is not a coherent superposition <strong>of</strong> both ground hyperfine levels.<br />

Such a DS would be dark regardless <strong>of</strong> Raman detuning between the probe a<br />

coupling fields resulting in poor EIT contrast.<br />

The second most important condition is the number <strong>of</strong> completely dark states.<br />

Although this criteria is somewhat invalidated by the existence <strong>of</strong> partial dark<br />

states, it still seems to be a good indicator <strong>of</strong> how good the transparency can be.<br />

For optically thick media four-wave mixing can become important. Unfortunately,<br />

orthogonal linear polarizations do create four wave mixing, and when the<br />

frequency filter was removed from our experiment we did detect a signal due to<br />

four-wave mixing. These four wave mixing product was an order <strong>of</strong> magnitude<br />

smaller than the probe, and were mostly ignored in the experiment.<br />

Finally, for the purposes <strong>of</strong> combining and separating the probe and coupling<br />

fields in the experiment, it is convenient that they have orthogonal polarizations.<br />

From Table A.1 it is evident that apart from four wave mixing, orthogonal


A.2. TRANSITION MATRIX ELEMENTS 160<br />

linear polarizations have all the desired features, whereas the other choices <strong>of</strong><br />

polarization do not.<br />

A.2 Transition matrix elements<br />

For convenience and completeness we give the transition matrix elements here for<br />

the D 1 lines <strong>of</strong> Rb 85 . Much <strong>of</strong> this information can be found in numerous references<br />

[127, 128, 143]. Thus, in order to not duplicate information unnecessarily<br />

this section will be brief. If there is missing information, consult one <strong>of</strong> the many<br />

references on single electron atoms and rubidium.<br />

First, the Wigner-Eckart theorem states that radial and angular parts <strong>of</strong> the<br />

dipole moment expectation value can be factored such that<br />

〈n, F, m F |er q |n ′ , F ′ , m ′ F 〉 = 〈n, F ||er||n ′ , F ′ 〉〈F, m F |F ′ , 1, m ′ F , q〉,<br />

where<br />

⎧<br />

〈n, F ||er||n ′ , F ′ 〉 = 〈J||er||J ′ 〉(−1) √ ⎪⎨ J J ′ 1<br />

F ′ +1+J+I<br />

(2F ′ + 1)(2J + 1)<br />

⎪⎩ F ′ F I<br />

⎫<br />

⎪⎬<br />

(A.1) ,<br />

⎪⎭<br />

and where we have used the Wigner 6-j notation (i.e.<br />

the curl bracket term).<br />

The angular momenta are defined as: F ≡ J + I and J ≡ L + S, where I is the<br />

angular momentum <strong>of</strong> the nucleus, S is the spin angular momentum <strong>of</strong> the electron


A.3. EIT IN RUBIDIUM 85 161<br />

and L is the orbital angular momentum <strong>of</strong> the electron. The prime denotes as<br />

excited state, and we have used the short hand <strong>of</strong> dropping the n and n ′ . The<br />

Clebsch-Gordon coefficients are given by<br />

⎜<br />

⎝<br />

〈F, m F |F ′ , 1, m ′ F , q〉 = (−1) F ′ −1+m F<br />

√<br />

(2F + 1)<br />

⎛<br />

⎞<br />

F ′ 1 F<br />

⎟<br />

⎠ ,<br />

m ′ F q −m F<br />

(A.2)<br />

where we have used the Wigner 3-j notation.<br />

There are interesting properties<br />

and symmetries <strong>of</strong> these expression that are explained adequately elsewhere [127,<br />

128, 143, 144]. To add to the confussion, there are a couple <strong>of</strong> conventions for<br />

calculating the dipole matrix elements; we follow the more common convention<br />

(i.e. less Russian) <strong>of</strong> Brink and Satchler [144]. For Rb 85 the dipole matrix elements<br />

for the D 1 line are given in Table A.2<br />

A.3 EIT in rubidium 85<br />

By choosing the quantization axis parallel to the coupling polarization (i.e. ɛ c =<br />

z = ɛ 0 , where z is the unit vector for the quantization axis) we only need to<br />

consider the dipole allowed transitions shown in Fig. A.1. For concreteness we<br />

also choose the probe polarization in the y direction such that ɛ p = y = i(ɛ −1 +<br />

ɛ +1 )/ √ 2.<br />

When only one excited hyperfine level is considered at a time there are five dark<br />

states for the F ′ = 2 hyperfine level and five dark states for the F ′ = 3 hyperfine


A.3. EIT IN RUBIDIUM 85 162<br />

(a)<br />

F’=3<br />

F’=2<br />

2<br />

1<br />

1<br />

2<br />

27<br />

3<br />

3<br />

27<br />

5<br />

8<br />

1<br />

8<br />

5<br />

27<br />

27<br />

3<br />

27<br />

27<br />

2<br />

−1<br />

−1<br />

2<br />

−<br />

−<br />

27<br />

3<br />

3<br />

27<br />

m =-3 F<br />

m F<br />

=-2 m F<br />

=-1 m F<br />

=0 m F<br />

=1 m F<br />

=2 m F<br />

=3<br />

F=3<br />

F=2<br />

(b)<br />

F’=3<br />

F’=2<br />

15<br />

27<br />

10<br />

27<br />

6<br />

27<br />

1<br />

3<br />

1<br />

27<br />

−1<br />

4<br />

−1<br />

1<br />

4<br />

1<br />

−<br />

0<br />

3<br />

27<br />

27<br />

27<br />

27<br />

3<br />

1<br />

27<br />

1<br />

3<br />

6<br />

27<br />

10<br />

27<br />

15<br />

27<br />

m F =-3 m F =-2 m F =-1 m F<br />

=0<br />

m F =1 m F =2<br />

m F =3<br />

F=3<br />

F=2<br />

Figure A.1: Dipole matrix elements for the transitions used in EIT with orthogonal linear<br />

polarizations.


A.3. EIT IN RUBIDIUM 85 163<br />

level. For Doppler broadened atomic vapors in a buffer gas it is important that<br />

the dark state be simultaneously dark for both excited hyperfine levels because<br />

the Doppler broadened linewidth is greater than the hyperfine splitting. To find<br />

a completely dark state (i.e. a state that is simultaneously dark for both excited<br />

hyperfine levels) we must find a superposition <strong>of</strong> the five F ′ = 3 dark states that<br />

is also a superposition <strong>of</strong> the F ′ = 2 dark states. It is helpful to divide each set<br />

<strong>of</strong> dark states further into two sets; dark states involving only A) the | F =3<br />

m=−2〉, | 3 0〉,<br />

| 3 2〉, | 2 −1〉, and | 2 1〉 ground states, and B) dark states involving only the | F =3<br />

m=−3〉, | 3 −1〉,<br />

| 3 1〉, | 3 3〉, | 2 −2〉, | 2 0〉, and | 2 2〉 ground states.<br />

Transitions for dark states involving subset A <strong>of</strong> ground states are shown in<br />

Fig. A.2. The three dark states corresponding to subset A are<br />

|M〉 3A = 2 ( | −1〉 2 − | +1〉 ) 2 + a √ 5 ( | −2〉 3 + | 3<br />

(A.3)<br />

8 + 10α 2<br />

+2〉 )<br />

|L〉 2A =<br />

|C〉 3A = | 3 0〉<br />

(<br />

1<br />

−1〉 + a √<br />

6<br />

| 3 0〉 − 1<br />

√<br />

1 + 11α2 /30<br />

| 2<br />

√<br />

5<br />

| −2〉<br />

3<br />

)<br />

(A.4)<br />

(A.5)<br />

and<br />

(<br />

| +1〉 2 + a − √ 1<br />

6<br />

| 3 0〉 + 1<br />

|R〉 2A √<br />

1 + 11α2 /30<br />

√<br />

5<br />

| +2〉<br />

3<br />

)<br />

(A.6)


A.3. EIT IN RUBIDIUM 85 164<br />

(a)<br />

−<br />

4<br />

27<br />

10<br />

27<br />

1<br />

3<br />

1<br />

3<br />

10<br />

27<br />

4<br />

27<br />

(b)<br />

5<br />

27<br />

2<br />

27<br />

1<br />

3<br />

−1<br />

3<br />

(c)<br />

1<br />

3<br />

1<br />

3<br />

5<br />

27<br />

−<br />

2<br />

27<br />

Figure A.2: Dipole matrix elements for dark states in Rb 85 D 1 line that involve the | F =3<br />

m=−2〉, | 3 0〉,<br />

| 3 2〉, | 2 −1〉, and | 2 1〉 ground states.


A.3. EIT IN RUBIDIUM 85 165<br />

(a)<br />

(d)<br />

15<br />

27<br />

6<br />

27<br />

1<br />

27<br />

1<br />

3<br />

2<br />

27<br />

−1<br />

3<br />

1<br />

3<br />

1<br />

27<br />

6<br />

27<br />

15<br />

27<br />

−<br />

2<br />

27<br />

−1<br />

3<br />

(b)<br />

(e)<br />

15<br />

27<br />

−1<br />

3<br />

−1<br />

27<br />

8<br />

27<br />

1<br />

27<br />

−<br />

2<br />

27<br />

(c)<br />

(f)<br />

1<br />

27<br />

2<br />

27<br />

1<br />

27<br />

1<br />

3<br />

8<br />

27<br />

15<br />

27<br />

Figure A.3: Dipole matrix elements for dark states in Rb 85 D 1 line that involve the | F =3<br />

m=−3〉,<br />

| 3 −1〉, | 3 1〉, | 3 3〉, | 2 −2〉, | 2 0〉, and | 2 2〉) ground states.<br />

where a ≡ iΩ p /Ω c , and the subscript means hyperfine level F = 3 or F = 2 and<br />

the subset <strong>of</strong> ground states is subset A. There is no superposition <strong>of</strong> these dark<br />

states that is simultaneously dark for transitions to both excited hyperfine levels.<br />

There is a superposition <strong>of</strong> dark states associated with subset B <strong>of</strong> the ground<br />

states (i.e. magnetic sublevels | F =3<br />

m=−3〉, | 3 −1〉, | 3 1〉, | 3 3〉, | 2 −2〉, | 2 0〉, and | 2 2〉). The dark


A.3. EIT IN RUBIDIUM 85 166<br />

states for subset B shown in Fig. A.3. These dark states are<br />

|M〉 2B = 1 √<br />

3<br />

2 |2 0〉 +<br />

8<br />

(<br />

|<br />

2<br />

−2 〉 + | 2 2〉 )<br />

(A.7)<br />

|L〉 2B =<br />

(√ )<br />

8|<br />

2<br />

−2 〉 + a| −1〉<br />

3 / √ 8 + a 2 . (A.8)<br />

and<br />

|R〉 2B =<br />

(√<br />

8|<br />

2<br />

2 〉 − a| 3 1〉)<br />

/ √ 8 + a 2 , (A.9)<br />

|M〉 3B =<br />

√ (<br />

6 |<br />

2<br />

−2〉 + | 2 2〉 ) − | 2 0〉 + a √ 5 ( | −3〉 3 − | 3 3〉 )<br />

√<br />

13 + 10a<br />

2<br />

|L〉 3B =<br />

(√<br />

6|<br />

2<br />

−2 〉 + a √ 5| −3〉 3 + a √ )<br />

3| −1〉<br />

3 / √ 6 + 8a 2 (A.10)<br />

and<br />

(√<br />

|R〉 3B = 6|<br />

2<br />

2 〉 − a √ 5| 3 3〉 − a √ )<br />

3| 3 1〉 / √ 6 + 8a 2 . (A.11)<br />

The completely dark state is given by<br />

|DD〉 =<br />

√<br />

2<br />

√<br />

13 + 10a2 |M〉 3B + 3 √ 3 + 4a 2 (|L〉 3B + |R〉 3B )<br />

2 √ 20 + 23a 2 . (A.12)


A.4. IMPORTANT PARAMETERS FOR RUBIDIUM 167<br />

4<br />

3<br />

2<br />

F=1<br />

5P 3<br />

2<br />

120.960(20)<br />

63.420(31)<br />

29.260(29)<br />

85<br />

Rb<br />

5P 1<br />

2<br />

F=3<br />

F=2 361.936(14)<br />

5S 1<br />

2<br />

F=3<br />

F=2 3035.732440(10)<br />

Figure A.4: Hyperfine splittings, in MHz, <strong>of</strong> the 5S 1/2 5P 1/2 and 5P 3/2 energy levels in rubidium<br />

85.<br />

A.4 important parameters for rubidium<br />

There are several parameters <strong>of</strong> rubidium that we have used in our calculations but<br />

are not part <strong>of</strong> the main text. We collect those parameters here for easy reference.<br />

The lifetimes for the D 1 (D 2 ) line is 27.70±0.04 ns (26.24±0.04 ns). These are<br />

the same for both isotopes <strong>of</strong> rubidium. From these lifetimes the dipole moments<br />

for the D 1 (D 2 ) transition is calculated to be 〈J||er||J ′ 〉=2.534(3)×10 −29 Cm<br />

(〈J||er||J ′ 〉=3.584(4)×10 −29 Cm). The mass and natural abundance <strong>of</strong> rubidium<br />

85 is 84.911800 amu and 72.17%. While the mass and natural abundance <strong>of</strong><br />

rubidium 87 is 86.909184 amu and 27.84%. The hyperfine splittings for the Rb 85<br />

levels 5S 1/2 5P 1/2 5P 3/2 are shown in Fig. A.4. The hyperfine splittings for for<br />

the Rb 87 levels 5S 1/2 5P 1/2 5P 3/2 are shown in Fig. A.5.


A.5. EIT IN RUBIDIUM 87 168<br />

3<br />

2<br />

1<br />

F=0<br />

5P 3 2<br />

266.650(9)<br />

156.947(7)<br />

72.218(4)<br />

87<br />

Rb<br />

5P 1<br />

2<br />

F=2<br />

F=1 816.656(30)<br />

5S 12<br />

F=2<br />

F=1 6834.682610904 29(9)<br />

Figure A.5: Hyperfine splittings, in MHz, <strong>of</strong> the 5S 1/2 5P 1/2 and 5P 3/2 energy levels in rubidium<br />

87.<br />

A.5 EIT in rubidium 87<br />

EIT in Rb 87 is slightly different from Rb 85 because in Rb 87 the hyperfine splitting<br />

for 5P 1/2 is 816 MHz which is larger than the Doppler broadening. Thus, the dark<br />

states do not have to be simultaneously dark for both excited hyperfine levels<br />

Tables A.4 and A.5.<br />

With the choice <strong>of</strong> orthogonal linear polarizations leads to the EIT transitions<br />

shown in Fig. A.6. Again we can divide the dark states into two subsets. Subset<br />

A has four dark states and involves magnetic sublevels | 1 −1〉, | 1 1〉, | 2 −2〉, | 2 0〉, and<br />

| 2 2〉. Subset B has two dark states includes magnetic sublevels | 1 0〉, | 2 −1〉, and | 2 1〉.<br />

Fig.<br />

A.7 shows the transitions involved in those dark states which require a<br />

superposition <strong>of</strong> multiple states (| 2 0〉 is also a dark state for the F ′ = 2 hyperfine<br />

level). There three dark states for the F ′ = 2 hyperfine level, and there are three<br />

dark states associated with the F ′ = 1 hyperfine level.


A.5. EIT IN RUBIDIUM 87 169<br />

(a)<br />

F’=2<br />

F’=1<br />

1<br />

12<br />

1<br />

12<br />

1<br />

2<br />

−1<br />

12<br />

1<br />

3<br />

m F<br />

=-2 m F<br />

=-1 m F<br />

=0 m F<br />

=1 m F<br />

=2<br />

−1<br />

12<br />

1<br />

2<br />

F=2<br />

F=1<br />

(b)<br />

F’=2<br />

1<br />

2<br />

1<br />

2<br />

1<br />

12<br />

F’=1<br />

1<br />

−<br />

3<br />

−1<br />

12<br />

1<br />

12<br />

m F<br />

=-2 m F<br />

=-1 m F<br />

=0 m F<br />

=1 m F<br />

=2<br />

0<br />

1<br />

2<br />

1<br />

12<br />

1<br />

2<br />

1<br />

3<br />

F=2<br />

F=1<br />

Figure A.6: Dipole matrix elements for the transitions used in EIT with orthogonal linear<br />

polarizations in rubidium 87.


A.5. EIT IN RUBIDIUM 87 170<br />

(a)<br />

(d)<br />

1<br />

1<br />

1<br />

1<br />

1<br />

−<br />

3<br />

2<br />

12<br />

12<br />

2<br />

1<br />

3<br />

−1<br />

12<br />

1<br />

2<br />

1<br />

2<br />

1<br />

12<br />

(b)<br />

(e)<br />

−1<br />

12<br />

1<br />

3<br />

1<br />

2<br />

1<br />

12<br />

−1<br />

12<br />

1<br />

2<br />

(c)<br />

1<br />

3<br />

1<br />

12<br />

Figure A.7: Dipole matrix elements for the dark states on Rb 87 D 1 . Not shown is the dark<br />

state | 2 0〉 because it is an incoherent dark state (i.e. it doesn’t require a superposition <strong>of</strong> ground<br />

states.)


A.5. EIT IN RUBIDIUM 87 171<br />

Table A.1: Comparison <strong>of</strong> EIT for several choices <strong>of</strong> polarization<br />

Comparison <strong>of</strong> polarizations for EIT on the D1 line <strong>of</strong> Rubidium 85.<br />

Lin⊥Lin Lin‖Lin Cir⊥Cir Cir‖Cir<br />

# <strong>of</strong> Dark States (DS) 1 0 0 5<br />

Parasitic DS? No No Yes Yes<br />

4-Wave Mixing Yes Yes Yes No<br />

Probe⊥Coupling? Yes No Yes No<br />

Table A.2: Dipole matrix elements for Rb 85 D 1 line with F = 2<br />

Dipole matrix elements for rubidium D 1 transitions from the F = 2 ground hyperfine<br />

level. All coefficients must be multiplied by 〈J = 1/2||er||J ′ = 1/2〉/ √ 27.<br />

F=2<br />

m F<br />

√<br />

= −2 m F<br />

√<br />

= −1 m F<br />

√<br />

= 0 m F<br />

√<br />

= 1 m F = 2<br />

σ + 2 3 3 2 —<br />

F ′ = 2 π −2 −1 0 1 2<br />

σ − ) — − √ 2 − √ 3 − √ 3 − √ 2<br />

σ + −1 − √ 3 − √ 6 − √ 10 − √ √ √ √ √ √<br />

15<br />

F ′ = 3 π 5 8 9 8 5<br />

σ 1 − √ 15 − √ 10 − √ 6 − √ 3 −1<br />

Table A.3: Dipole matrix elements for Rb 85 D 1 line with F = 3<br />

Clebsch-Gordon coefficients for rubidium D 1 transitions from the F = 3 ground<br />

hyperfine level. All coefficients are divided by 〈J = 1/2||er||J ′ = 1/2〉/ √ 27.<br />

F=3<br />

m<br />

√<br />

= −3 m<br />

√<br />

= −2 m<br />

√<br />

= −1 m<br />

√<br />

= 0 m = 1 m = 2 m = 3<br />

σ + 15<br />

√ 10<br />

√ 6<br />

√ 3<br />

√ 1<br />

√ — —<br />

F ′ = 2 π — 5 8<br />

√ 9<br />

√ 8<br />

√ 5<br />

√ —<br />

σ −<br />

√<br />

—<br />

√<br />

—<br />

√<br />

1<br />

√ 3<br />

√ 6<br />

√ 10 15<br />

σ + 3 5 6 6 5 3 —<br />

F ′ = 3 π -3 -2 -1 0 1 2 3<br />

σ − — − √ 3 − √ 5 − √ 6 − √ 6 − √ 5 − √ 3


A.5. EIT IN RUBIDIUM 87 172<br />

Table A.4: Dipole matrix elements for Rb 87 D 1 line with F = 1<br />

Dipole matrix elements for the Rb 87 5S 1/2 → 5P 1/2 transitions from the ground<br />

hyperfine level F = 1. To obtain the dipole moment each element must be multiplied<br />

by 〈J = 1/2||er||J ′ = 1/2〉/ √ 12.<br />

F=2<br />

m F = −1 m F = 0 m F = 1<br />

σ + -1 -1 —<br />

F ′ = 1 π 1 0 -1<br />

σ − ) — 1 1<br />

σ + −1 − √ 3 − √ √ √<br />

6<br />

F ′ = 2 π 3 2 3<br />

σ 1 − √ 6 − √ 3 −1<br />

Table A.5: Dipole matrix elements for Rb 87 D 1 line with F = 2<br />

Dipole matrix elements for the Rb 87 5S 1/2 → 5P 1/2 transitions from the ground<br />

hyperfine level F = 2. To obtain the dipole moment each element must be multiplied<br />

by 〈J = 1/2||er||J ′ = 1/2〉/ √ 12.<br />

F=2<br />

m F<br />

√<br />

= −2 m F<br />

√<br />

= −1 m F = 0 m F = 1 m F = 2<br />

σ + 6<br />

√ 3 1<br />

√ — —<br />

F ′ = 1 π — 3 2<br />

√ 3<br />

√ —<br />

σ −<br />

√<br />

—<br />

√<br />

—<br />

√<br />

1<br />

√ 3 6<br />

σ + 2 3 3 2 —<br />

F ′ = 2 π -2 -1 0 1 2<br />

σ − — − √ 2 − √ 3 − √ 3 − √ 2

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!