+ All Categories
Home > Documents > Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling...

Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling...

Date post: 31-Mar-2021
Category:
Upload: others
View: 3 times
Download: 0 times
Share this document with a friend
137
Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis S. Petersen Submitted in Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy Supervised by Dr. Jake Bromage The Institute of Optics Arts, Sciences and Engineering Edmund A. Hajim School of Engineering and Applied Sciences University of Rochester Rochester, New York 2017
Transcript
Page 1: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

Scaling Infrared Femtosecond Optical Parametric

Oscillators to High Average Powers

by

Travis S. Petersen

Submitted in Partial Fulfillment of the

Requirements for the Degree

Doctor of Philosophy

Supervised by

Dr. Jake Bromage

The Institute of Optics

Arts, Sciences and Engineering

Edmund A. Hajim School of Engineering and Applied Sciences

University of Rochester

Rochester, New York

2017

Page 2: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

ii

Dedicated to my Mom and Dad for all the help they gave

to make me the man I am today.

Page 3: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

iii

Table of Contents

Biographical Sketch .............................................................................................. vii

Acknowledgments.................................................................................................. ix

Abstract ................................................................................................................ xiii

Contributors and Funding Sources.........................................................................xv

List of Figures ..................................................................................................... xvii

1 Introduction .......................................................................................................1

1.1 Brief History of Optical Parametric Oscillators ........................................ 1

1.2 Recent Advancements in Scalable Ultrafast OPO’s ................................. 3

1.3 Applications for Infrared Ultrafast Optical Pulses .................................... 6

1.4 Outline ....................................................................................................... 7

2 Optical Parametric Oscillator Fundamentals .....................................................8

2.1 Fundamental Principles of Nonlinear Optics ............................................ 8

2.1.1 The Nonlinear Optical Susceptibility .................................................. 8

2.1.2 Description of Nonlinear Optical Processes...................................... 10

2.1.3 Coupled-Wave Equations .................................................................. 11

2.1.4 Critical Phase Matching .................................................................... 13

2.2 Characteristics of Optical Parametric Oscillators ................................... 17

2.2.1 Parametric Gain ................................................................................. 17

Page 4: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

iv

2.2.2 Synchronous Pumping and Cavity Length Stabilization................... 19

2.3 Properties of Ultrafast Pulses .................................................................. 20

2.3.1 Definition of Spectral Phase and Dispersion..................................... 20

2.3.2 Sources of Dispersion........................................................................ 22

2.3.3 Self-phase Modulation ...................................................................... 24

2.4 Pulse Characterization Techniques ......................................................... 24

3 Thin-Disk Mode-Locked Pump Laser .............................................................26

3.1 A Review of Thin-Disk Lasers................................................................ 26

3.2 Principles of Thin-Disk Operation .......................................................... 27

3.2.1 Thin-Disk Laser................................................................................. 27

3.2.2 Yb:YAG as a Laser Material ............................................................. 28

3.2.3 Thermal Lensing ............................................................................... 30

3.3 Cavity Design .......................................................................................... 31

3.3.1 Thin-Disk Head ................................................................................. 31

3.3.2 Active Multipass Cavity .................................................................... 34

3.3.3 Cavity Design .................................................................................... 35

3.4 Mode-Locking ......................................................................................... 37

3.4.1 Soliton Formation .............................................................................. 37

3.4.2 Initiation of Mode-Locking ............................................................... 40

3.4.3 Mode-Locking Instabilities ............................................................... 41

3.5 TDML as a Pump Source ........................................................................ 45

3.5.1 Output Power and Stability Regimes ................................................ 45

3.5.2 3.5.1 Pulse Duration .......................................................................... 46

3.5.3 Beam Quality..................................................................................... 48

Page 5: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

v

3.5.4 Temporal Stability ............................................................................. 49

4 Design of the Optical Parametric Oscillator ....................................................51

4.1 Nonlinear crystal ..................................................................................... 51

4.1.1 Availability of Crystals ..................................................................... 51

4.1.2 BiBO Transparency ........................................................................... 55

4.1.3 Noncollinear Geometry ..................................................................... 57

4.1.4 Crystal Coatings ................................................................................ 58

4.2 Cavity Design .......................................................................................... 59

4.2.1 Multipass Cavity Design ................................................................... 59

4.2.2 Folded Cavity Design ........................................................................ 64

4.2.3 ABCD Analysis ................................................................................. 67

5 Characterization of the Ultrafast Optical Parametric Oscillator ......................69

5.1 Pulse Duration ......................................................................................... 69

5.2 Spectral Measurements ........................................................................... 71

5.3 Temporal Stability ................................................................................... 75

5.4 Output Power and Scaling Limitations ................................................... 77

5.5 Beam Quality........................................................................................... 78

6 Thermal Analysis and Limitations of High-Power Operation .........................82

6.1 Motivation of Thermal Analysis ............................................................. 82

6.1.1 2-µm Absorption in BiBO and Total Absorbed Power ..................... 82

6.1.2 Active Cooling of the OPO ............................................................... 83

6.2 Measurement of Thermal Properties ....................................................... 86

6.2.1 Thermal Camera Selection and Image Acquisition........................... 86

Page 6: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

vi

6.2.2 Preparation of Thermal Images for Data Analysis ............................ 87

6.2.3 Final Thermal Images........................................................................ 88

6.3 Impact of Internal Power on the Thermal Gradient ................................ 89

6.4 Thermally Induced Phase-Matching Errors ............................................ 90

6.5 Power Limitations from Thermal Lensing .............................................. 93

7 Outlook and Conclusions.................................................................................97

7.1 Summary ................................................................................................. 97

7.2 Outlook .................................................................................................... 98

7.2.1 OPO Pulse Duration .......................................................................... 99

7.2.2 Pump Laser Power and Energy ......................................................... 99

7.2.3 Timing Jitter Stabilization ............................................................... 100

7.2.4 Thermal Management of Nonlinear Crystals .................................. 102

7.2.5 Cavity Design .................................................................................. 103

7.3 Conclusions ........................................................................................... 105

References ............................................................................................................107

Page 7: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

vii

Biographical Sketch

Travis S. Petersen was born in Sacramento, CA in 1986 and grew up in time split

between Reno, NV, and Puyallup, WA. In 2009, he graduated cum laude from Central

Washington University (CWU) with a Bachelor of Science degree in Physics, a Bachelor

of Arts degree in Philosophy and a minor in Mathematics. While the study of philosophy

was his first passion, a critical understanding of physical science seemed imperative to

comprehend modern philosophical arguments. Physics became the primary focus of his

education and he pursued a variety of research interests.

The first major research Travis performed was with the assistance of the cwU

Science Honors Program under Dr. Andy Piacsek. The propagation of sound from wind

turbines was modeled from acoustical measurements of nearby wind farms to assess the

impact of a proposed wind farm in Ellensburg, WA. Involvement with the local chapter of

the Society of Physics Students (SPS) led to an interesting research opportunity with the

help of Dr. Mike Braunstein to construct a circuit to model chaotic differential equations

and model key parameters for characterizing chaotic behavior. After this research was

completed, a new opportunity with Dr. Mike Jackson became available, which sparked

Travis’s interest in optics. A laser laboratory was constructed from the ground up, which

characterized isotopic methanol using heterodyned CO2 lasers. This research was funded

in part by the cwU Science Honors Program, which led to his passion for laser science.

In 2009, Travis entered the doctoral program at The Institute of Optics at the

University of Rochester. One year later, he joined the Laser Development Group at the

Laboratory for Laser Energetics (LLE) under the supervision of Dr. Jake Bromage to study

high-average-power ultrafast parametric systems with the support of a Horton Fellowship.

He received a Master of Science degree in Optics in 2015.

Page 8: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

viii

Publications

1. T. Petersen, J. Zuegel, and J. Bromage, “Thermal Effects in an Ultrafast BiB3O6

Optical Parametric Oscillator at High Average Powers,” Applied Optics, 2017,

56(24): pp. 6923-6929.

2. T. Petersen, J. Zuegel, and J. Bromage, “High-average-power, 2-μm femtosecond

optical parametric oscillator synchronously pumped by a thin-disk, mode-locked

laser,” Optics Express, 2017, 25(8): pp. 8840–8844.

3. M. Jackson, T. Petersen, and L. R. Zink. “Frequencies and Wavelengths From a

New Far-Infrared Lasing Gas: 13CHD2OH,” IEEE Journal of Quantum Electronics,

2009, 45(7): pp. 830–832.

Conference Proceedings

1. T. Petersen, J. Zuegel, and J. Bromage, “High-Energy Infrared Femtosecond

Optical Parametric Oscillator Synchronously Pumped by a Thin-Disk Laser,” in

Lasers Congress (ASSL, LCL, LAC), pATu1A.6 (2016).

2. T. Petersen and J. Bromage, “Thin-Disk Pumped, High-Average-Power, 2.06-μm

BiB3O6 Femtosecond Optical Parametric Oscillator”, Frontiers in Optics,

pLTh4I.6 (2015).

3. T. Petersen and J. Bromage, “Intracavity Chirped-Pulse Amplification for High-

Energy Ultrafast Optical Parametric Oscillators,” Frontiers in Optics, pFM4G.3

(2012).

Page 9: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

ix

Acknowledgments

This work would have not been possible without the support of many talented

people and the patience of many friends. Therefore, I will use this small space to give a

small dedication of my gratitude to the following.

Many thanks are due to Dr. Jake Bromage, my advisor. Being Jake’s first graduate

student has been an amazing adventure. It is quite an honor to be able to have taken this

journey with such a talented laser physicist. His deep understanding of ultrafast phenomena

and practical laser engineering has taught me so much in the past years that I would not

have been able to learn in many other places. Jake is an extremely driven individual who

is active in the lab, supports the Laser Development Group (not only a group member but

also through his recent promotion as the group leader), and serves as a volunteer firefighter,

all on top of being an academic advisor. The success of this research would not have been

possible without the countless hours of planning and review during our weekly hour-long

meetings. It has truly been a pleasure to work and learn under Jake’s guidance, and I will

always be grateful for this experience.

Humble cheers are due to the members of the Laser Development Group. Rick

Roides, who gave unique insight into laser development techniques with his many years of

experience. Dr. Ildar Begishev, who stayed late on a few occasions so I could work longer

in the lab, provided engineering insight for parametric devices, and babysat my laser from

time to time. Dr. Christophe Dorrer, who shared years of insight on ultrafast pulse

dynamics during group meetings. Bob Cuffney, who helped me find a great deal of

equipment around the lab that I could borrow for my experiments. Dr. Andrey Okishev,

who helped with foreign correspondences and provided key infrared diagnostics. Dr.

Seung-Whan Bahk, who provided a great deal of information and code for the imaging

needs of this study. Finally, Dr. Jon Zuegel, whose tireless work over the years as group

Page 10: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

x

leader, and more recently as Laser Development and Engineering Division Director,

ensured that finances were never a concern and that working with his group was always a

joy.

From a student’s point of view, LLE has many different disciplines under the same

roof, providing great services from various departments that allow one to spend more time

on one’s research. The Purchasing Department, headed by Bill Byrne, has been absolutely

phenomenal in handling my many requisitions and complicated orders. The Publications

Department, led by Jen Taylor, has never missed a deadline and has always gone the extra

mile to make any last-minute changes when I asked. It is amazing what they accomplish in

this building with such a small crew. Cindy Dorfner and the members of the Computer

Support Group helped me solve problems in a fraction of the time of what I could have

done on my own, if at all. The Machine Shop has been a wonderful resource, with both

requisitions for specialty equipment and the availability to make modifications on my own.

Alexei Kozlov gave me a lot of advice on his personal time about laser damage and thermal

mitigation. The Optical Manufacturing Group has been an invaluable resource to have in

the same building. Gary Mitchell taught me how to get optical surfaces cleaner than I

thought possible, cleaned specialty optics and optomechanics for me on his own time, and

showed me that not all isopropyl alcohols are created equal. Alex Maltsev created custom

optical cuts to mirrors at a fraction of the cost of a commercial vendor in half the time.

There are so many more to thank that I could probably fill half the thesis with an adequate

description. LLE truly has been an amazing place to spend my time as a graduate student.

The staff in The Institute of Optics, which includes Kari Brick, Maria Schnitzler,

Gina Kern, Lori Russell, Gayle Thompson, Noelene Votens, kept all the paperwork in order

and sheltered me from the bureaucracy of higher education. Kari deserves special thanks

because she has helped me navigate many complicated issues arising from performing my

research off-campus and questions about the graduate process. Kari and the entire staff

have always been cheerful and willing to help at a moment’s notice.

Page 11: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

xi

The Central Washington University Physics Department should receive more

attention for the service they provide to their undergraduate students. I was not sure I

wanted to pursue physics when I started college, and looking back I know that many

students would not have selected cwU to study physics. However, the opportunities

available to undergraduate students in that department, as well as cwU , are quite

astounding. Being able to perform four years of research alongside my curriculum gave me

skills that I am unsure if I could have obtained elsewhere. The wonderful professors in our

small department maximized the potential of having such small classes with personal

attention and feedback that I felt honored to be a part of the process. I would also like to

thank the Science Honors program, which provided research funds and stipends that

introduced students from every discipline to the work expected from you at a graduate level

while also diversifying a student’s scientific knowledge.

I could not thank Professor Mike Jackson enough, who served as my advisor during

my last year of undergraduate research. I am sure that without the research opportunities

he brought with him, I would not have been accepted into the doctoral program at The

Institute of Optics. It was Professor Jackson who first sparked my interest in optics. The

experimental work with lasers led me to find my true passion in research, which might not

have ever happened if he had not come to cwU. He was truly a gift to the students at cwU

and I will be forever grateful for the impact he had on my life.

Although it has been many years since, Mrs. Julie Brownlee, my high school debate

coach continues to be a source of inspiration to me. Although I did quite well in high

school, I was never particularly fond of it. If it were not for my involvement in debate, it is

questionable if I would have attended college at all. Debate nurtured an interest in arguing

into a passion for philosophical thought and critical reasoning. This likely would not have

appealed to me if it were not for Mrs. Brownlee’s effort, which went well beyond the

expectations as an instructor.

It goes without saying that I could not have made it this far without my wonderful

family. I will always cherish the love and support my mother gave me as a child to help

Page 12: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

xii

me excel at school and nurtured my inquisitive mind. Always a source of love, you have

fueled my life with excitement and optimism that has kept my sanity during these past

crazy years. My father has been a solid foundation of support and encouragement my entire

life. A man of his word and one of the most respectable people I know, I am honored to

have had you guide me through this journey. Karen, your love and support have enriched

my life. It seems crazy that 20 years have gone by but I just cannot imagine how life would

have been without you in it. Ken, you have done so much to make California feel like home

and it is hard to express my appreciation for what you have brought to my life.

Even though we have lived on opposite sides of the country for many years, my

longtime friend Jeston Loux and I have maintained a friendship that has been a vital to my

sanity throughout the college experience. Our paths in life have been about as different as

they get, yet a bond that I will never fully understand has always kept us on the same page.

You have kept me grounded, realizing that there are other things in life besides graduate

school, which at times can be all-consuming. Thanks for always being there, man.

I was lucky to have met my colleague and dear friend Justin Schultz and to have

your friendship throughout graduate school. Although our work was in very different

regions of optics, the ability to discuss research difficulties or blow off steam about

graduate school with you is something for which I am very grateful. Our occasional dinners

and late-night scotches made some of the most difficult parts of the graduate process

bearable, and I am glad to have known someone with such a brilliant mind and kind heart.

Words cannot express how thankful I am to have had my love, Amelia McMullen,

in my life during the most challenging time of my research. Your patience, motivation, and

willingness to listen put me in a debt that I can never fully repay. When staying the course

seemed so difficult, you were there to help carry the weight. I will always treasure what

you brought to my life. Thank you, from the bottom of my heart.

Page 13: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

xiii

Abstract

A demand for ultrafast, highly energetic laser sources in the infrared exists for

various applications such as high-harmonic generation, waveguide inscription, and remote

sensing. These applications require high repetition rates for faster processing speed and

better signal-to-noise ratios. Until recently, laser pulses used in these systems were usually

provided by optical parametric oscillators and amplifiers pumped by Ti:sapphire laser

systems. The output power and spectral bandwidth of this pump source are fundamentally

restricted due to unavoidable limitations caused by the quantum-defect heating and

restricted emission spectrum of the gain medium. Advancements to Yb-based sources have

provided a significant increase in average power and pulse energy available from ultrafast

pumps, thereby increasing the capabilities of parametric devices.

In this thesis, an ultrafast optical parametric oscillator was constructed as a test bed

for investigating the scaling potential of these devices. The main goals of this study were

to utilize recent technological advancements to ultrafast pump sources and test the limits

for obtaining high-energy pulses with high average powers from an optical parametric

oscillator. Various nonlinear crystals and parametric pumping geometries were

investigated to determine how best to use the available power of a home-built pump laser

and the general potential for scalability. To utilize recent developments of similar pump

sources, a long cavity was designed to take advantage of the high-average-powers with

scalable pulse energies. While the designing the cavity, it became apparent that the cavity

length could be done easily without notable change to critical cavity parameters by using

telescopic imaging relays.

It was discovered that limitations to the power scalability of this system were

directly linked to problems with the thermal management of the nonlinear crystal. While

minor changes could have been implemented to reduce the impact of these boundaries,

Page 14: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

xiv

understanding thermal effects in parametric devices will be critical to fully utilizing the

state-of-the-art pump sources. Simple and cost-effective methods were developed for

characterizing the impact of the thermal gradient on the available gain. This analysis could

easily be extended to other systems for predicting and mitigating thermally induced power

scaling limitations.

Page 15: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

xv

Contributors and Funding Sources

This dissertation was supervised by Jake Bromage, Senior Scientist and Group

Leader at the Laboratory for Laser Energetics. The dissertation committee, chaired by

Professor Todd Krauss (Chemistry) consisted of Dr. Jake Bromage (LLE), Professors

Wayne Knox (Optics), Robert Boyd (Optics), and Nicholas Bigelow (Physics).

This research was supported by the Department of Energy National Nuclear

Security Administration under Award Number DE-NA0001944, the University of

Rochester, and the New York State Energy Research and Development Authority. This

report was prepared as an account of work sponsored by an agency of the U.S. Government.

Neither the U.S. Government nor any agency thereof, nor any of their employees, makes

any warranty, express or implied, or assumes any legal liability or responsibility for the

accuracy, completeness, or usefulness of any information, apparatus, product, or process

disclosed, or represents that its use would not infringe privately owned rights. Reference

herein to any specific commercial product, process, or service by trade name, trademark,

manufacturer, or otherwise does not necessarily constitute or imply its endorsement,

recommendation, or favoring by the U.S. Government or any agency thereof. The views

and opinions of authors expressed herein do not necessarily state or reflect those of the

U.S. Government or any agency thereof.

The author’s graduate study was funded by a Horton Fellowship, which has been

an invaluable resource for countless students working at LLE.

This work would have not been possible if not for the numerous contributions from

the Electronics Department at LLE. Numerous home-built electronic devices including

laser locks, noise characterization circuits, and thermoelectric cooling controllers were

constructed that were vital to the operation of the systems used in this study. Wade Bittle

Page 16: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

xvi

was a great resource for general knowledge of general electronics and characterization

equipment.

The work shown here could not have been accomplished without the generous

contributions from TRUMPF GmbH. The pump laser used in this investigation was built

using a TRUMPF thin-disk head obtained through a collaborative agreement. Prior to

receiving this equipment, commercial thin-disk devices for scientific investigations were

not available, and the ability to test the potential of Yb-based pump sources would not have

been possible if it were not for support from TRUMPF, headed by Dr. Dirk Sutter. The

author would like to personally thank Dr. Dominik Bauer. During his graduate studies, he

provided invaluable guidance for optimizing the thin-disk pump laser during long

conference calls and email correspondences. This assistance was provided on his own time

with no direct personal benefit. He went above and beyond to ensure the success of this

laser at the LLE.

Page 17: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

xvii

List of Figures

Figure 1.1: Historical sampling of ultrafast OPO’s ............................................................ 5

Figure 2.1: Noncollinear interaction geometry . ............................................................... 15

Figure 3.1: Schematic of the thin-disk laser head ............................................................ 28

Figure 3.2: Yb:YAG material information ....................................................................... 29

Figure 3.3: Diagram of thin-disk head baseplate .............................................................. 33

Figure 3.4: Alignment of cw-pump spot ........................................................................... 34

Figure 3.5: Diagram of the TDML cavity ......................................................................... 36

Figure 3.6: Example of cw background ............................................................................ 44

Figure 3.7: Efficiency of the TDML. ................................................................................ 45

Figure 3.8: Autocorrelation and spectrum of the TDML .................................................. 47

Figure 3.9 Near-field beam profile of the TDML ............................................................. 49

Figure 4.1: OPO crystal parameters ................................................................................. 54

Figure 4.2: Unpolarized transmission spectrum of BiBO ................................................. 55

Figure 4.3: Measured BiBO crystal transmission ............................................................. 56

Figure 4.4: Noncollinear phase-matching curves ............................................................. 57

Figure 4.5: Multipass OPO cavity schematic ................................................................... 60

Figure 4.6: Mirror reflectivity and GDD for various cavities ........................................... 62

Figure 4.7: Folded OPO cavity schematic ........................................................................ 64

Figure 4.8: Total linear dispersion of the folded OPO cavity ........................................... 66

Figure 4.9: Images of the crystal mount and beam orientation ......................................... 67

Page 18: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

xviii

Figure 5.1: Autocorrelation trace of the OPO ................................................................... 71

Figure 5.2: Diagram of the home-built spectrometer........................................................ 72

Figure 5.3: OPO spectrum ................................................................................................ 74

Figure 5.4: Radio-frequency spectrum of the signal and pump ....................................... 76

Figure 5.5: Output power of the signal and idler .............................................................. 77

Figure 5.6: Beam profiles at different output powers ....................................................... 79

Figure 5.7: Beam-quality measurements and M2 fit ........................................................ 80

Figure 6.1: Schematic of the thermoelectric cooler and heat sink .................................... 84

Figure 6.2: TEC temperature setting and the signal power .............................................. 85

Figure 6.3: Schematic and sample thermal image of crystal face ..................................... 89

Figure 6.4: Change in thermal gradient as a function of SWIR power ............................. 90

Figure 6.5: Thermal images various powers and wavelengths ......................................... 91

Figure 6.6: Changes to cavity dynamics predicted by thermal analysis ........................... 95

Figure 6.7: Signal power and calculated beam spot size versus pump power .................. 96

Page 19: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

1

1 Introduction

1.1 Brief History of Optical Parametric Oscillators

An optical parametric oscillator (OPO) is an optical frequency conversion system

using the frequency-mixing characteristics of a nonlinear crystal, where one or two waves

are resonated inside a cavity during the process of nonlinear frequency conversion. An

OPO is a wavelength-tunable light source with a unique set of advantages including a broad

tuning range, high stability, and a simple structure. Traditionally, OPOs have been of

interest for their ability to generate coherent light at frequencies not available to

conventional lasers.

The development of OPOs has grown with advancements in both nonlinear crystals

and available pump sources. In the early 1960’s, multiple groups predicted the possibility

of parametric gain during a three-wave interaction [1-3]. The first OPO was demonstrated

quickly after the development of the first laser with pulsed operation. In 1965, Giodmaine

and Miller obtained a signal-beam tuning range of 0.97 to 1.15 µm using a lithium niobate

crystal pumped by a Q-switched multimode laser [4]. Soon, Boyd et al. proposed the

possibility of a continuous-wave (cw) OPO to overcome inherent instabilities and produce

narrow linewidths [5]. In 1968 cw OPO’s were successfully obtained, and numerous

studies investigated improving the tuning range, uniform tuning power and narrow

linewidth of OPO’s for numerous applications, most notably spectroscopy [5-7]. For

pulsed OPO’s, numerous improvements to Q-switched pump sources provided

breakthroughs in obtaining pump thresholds with higher peak powers for nanosecond (ns)

pulses without the need for high average powers.

Page 20: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

2

The next large development for OPO’s was extending pulsed operation to the

picosecond (ps) and femtosecond (fs) regimes using ultrafast mode-locked sources. These

devices had significantly higher peak powers than nanosecond (ns) OPO’s but lower pulse

energies. Unlike Q-switched laser–pumped OPO’s, where the ns pulses usually span

several cavity lengths, mode-locked OPO’s can generate pulse durations much shorter than

a cavity length. This allows the resonated pulse to be amplified by an undepleted pump

pulse for each round-trip inside the cavity. The output of the OPO follows the pulse

dynamics of the mode-locked pump laser, so characteristics of the output can be controlled

and designed for a specific application.

First demonstrated in 1972, synchronous pumping of an OPO occurs when the

signal pulse arrives at the nonlinear crystal at the same rate as the pulse laser such that the

𝑐/2𝐿 frequencies of the pump laser and the OPO are identical [8]. After this achievement,

analytic expressions for an OPO pumped by a mode-locked laser were presented for

transient and steady states [9]. The timing synchronization requires identical repetition

rates between the pump laser and the synchronously pumped OPO. After a characteristic

buildup time, the signal reaches steady-state operation and oscillates in the cavity. Initial

synchronously pumped systems were restricted to pump sources using Q-switched mode

locking, which had room for improvement due to unwanted damage caused by the high

intensities from nonuniform pulse trains.

As interest in synchronously pumped systems grew, improvements to numerical

simulations of the parametric process were derived from the master mode-locking equation

and using the split-step Fourier method [10-14]. The first fs OPO was demonstrated by

D.C. Edelstein et al. in 1989 using a KTP crystal synchronously pumped by a colliding-

pulse passively mode-locked dye laser [15]. This analysis was done alongside experimental

demonstrations of synchronously pumped fs OPO’s pumped by the growing popularity of

Kerr-lens mode-locked Ti:sapphire lasers using a variety of phase-matching techniques

[16-20]. Additionally, soliton formation in fs OPO’s has been achieved with appropriate

conditions for dispersion and self-phase modulation [21, 22]. Broad tunability around

Page 21: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

3

degeneracy using diffraction gratings or birefringent filters as frequency-selective devices

to induce singly resonant operation in synchronously pumped OPO’s has also been

demonstrated [23, 24].

Many advancements have been made with free-space fs OPO’s synchronously

pumped by Ti:sapphire lasers in the last 25 years. Ultrafast systems have been developed

from wavelengths between 0.25 and 8.0 µm with a wide variety of goals for individual

investigations. The focus of some investigations has been wide tunability in a single

system. Examples of this are a periodically poled lithium niobate OPO that has near-

continuous tunability from 0.98 to 4.55 µm with pulse durations of 150 fs and a AgGaSe2-

based OPO with tunability from 2 to 8 µm with pulse widths ranging from 400 to 500 fs

across the available wavelengths [20, 25]. When tunability was not a concern, pulse

durations below 100 fs have been achieved at energies between 1 and 10 nJ from the visible

to the mid-infrared [26-34].

Ti:sapphire–pumped ultrafast OPO’s are typically limited to producing sub-watt

average powers due to restrictions on the available power from Ti:sapphire lasers. Using

state-of-the-art 18-W, solid-state green lasers to pump fs Ti:sapphire oscillators, it is now

possible for commercial ultrafast singly resonant (and thus tunable) OPO’s to reach near-

10-nJ energy levels [35]. This pulse energy is likely to represent an upper limit for this

technology because of the increasingly difficult thermal management required in strongly

pumped Ti:sapphire oscillators.

1.2 Recent Advancements in Scalable Ultrafast OPO’s

In the past decade, to surpass the capabilities of the Ti:sapphire laser, development

of ultrafast, high-average-power laser sources has been dominated by three technologies:

Innoslab amplifiers, fiber amplifiers, and thin-disk lasers. Innoslab-based amplifiers have

recently achieved kilowatt (kW) average powers with pulse durations of 615 fs [36].

Similarly, fiber amplifier systems have achieved average powers of 830 W with pulse

durations of 640 fs [37]. Both types of systems are based on a master-oscillator/power-

Page 22: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

4

amplifier scheme, which must employ complex amplifier stages to achieve these powers.

To reduce these complexities, high-average-power femtosecond pulses have been achieved

in a single oscillator using a thin disk as the gain medium.

These ytterbium-based systems have shown great promise in scaling high-average-

power, ultrafast pulses near 1-μm wavelengths. Recently 145-W, 1.1-ps pulses containing

more than 41 J of energy pulses with peak powers of 37 MW have been produced when

operated in an ambient atmosphere [38]. Additionally, 250-W, 210-fs pulses with peak

powers of 38 MW have been achieved in air with use of Kerr lens mode-locking instead of

the conventional semiconductor saturable absorbing mirror techniques [39]. This

technology shows promise of reaching 100-MW peak powers within a few years in addition

to different approaches to scaling techniques [40].

While each of these systems produces high-average-power fs pulses, there still exist

inherent limitations to tunability and wavelength selection due to the relatively narrow

emission spectra of their respective gain medium. The use of these systems as pump

sources for ultrafast OPO’s makes it possible to achieve high-average-powers at

wavelengths across the visible and infrared.

A few groups have already employed these pump sources in ultrafast OPO’s with

significant return in the output power and energy. Using a noncollinear pump geometry in

tandem with a Yb:KLu(WO4)2 thin-disk oscillator, an ultrafast OPO was able to produce

visible light with average powers exceeding 3 W with a repetition rate of 34 MHz and pulse

durations of ~70 fs from a β-barium borate (BBO) crystal [26]. Another system employed

a Yb:fiber laser pump source, capable of producing watt-level 1530-nm pulses from a

MgO:PPLN crystal with energies greater than 70 nJ with pulse durations of 1.5 ps [41].

Modifications to this system by lowering the repetition rate through cavity dumping

produced 650 nJ at 1.52 μm with pulse durations of 228 fs, at sub-watt levels, which are

the most energetic pulses from a synchronously pumped fs OPO to date [42].

Page 23: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

5

To understand the significance of these improvements, it is useful to compare the

performance of Yb-pumped ultrafast OPO’s to ones using a Ti:sapphire pump. A sampling

of synchronously pumped free-space ultrafast OPO’s created in the last 25 years is shown

in Fig. 1.1.

Figure 1.1: Sampling of free-space ultrafast OPO’s created in the last 25 years operated at

different wavelength regions. Wavelength regions represent areas where the OPO primarily

operated; however, in some cases resonated signal and idler waves spanned multiple

regions (e.g., from 650 to 900 nm). The OPO’s represented in the top half of the figure

were pumped by Yb-based pump sources, while those on the bottom were Ti:sapphire

pumped.

The majority of ultrafast OPO’s with pulse durations below 500 fs were pumped with a

Ti:sapphire laser. Since OPO’s were only sampled in this figure, the prevalence of this

pump source is vastly understated, which is understandable considering the Ti:sapphire

laser has truly been the workhorse of the ultrafast field. Despite the presence of nanojoule

(nJ) level pulses across the optical spectrum, the increase in available energy from the new

class of Yb-based pump sources becomes obvious. These systems did change how OPO’s

were fundamentally operated, but were simply able to take advantage of the dramatic

Page 24: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

6

increase in pump pulse energy available from recent technological advancements. The key

question becomes; What are the limitations to this new available power and for what

applications would they be best suited?

1.3 Applications for Infrared Ultrafast Optical Pulses

Recent improvements to OPO’s and the potential for scalability, both in power and

energy, from a tunable ultrafast source open applications previously not addressable with

OPO’s. Illustrated by the examples below, an increasing number of applications require

high-intensity pulses with high average powers in the infrared (IR). In general, ultrafast

OPO’s have the capability to meet the needs of these applications because of their wide

tunability and strong potential for scalable ultrafast laser pulses with superb beam quality.

In high-harmonic generation (HHG) [43], the electron energies for production of

extreme ultraviolet (UV) light scale with pump intensity and 𝜆2, making high-intensity

pulses in the IR an attractive source for HHG. Furthermore, high repetition rates are

beneficial for improved performance in HHG applications including spectroscopy,

microscopy, and frequency metrology [44-46].

Ultrafast sources with high repetition rates reduce the modification threshold and

improve waveguide quality in ultrafast waveguide inscription [47]. However, no suitable

sources exist beyond 1040 nm, despite a real demand. Advancements in the energy

available from ultrafast OPO’s offer an alternative method for achieving high-energy, high-

repetition-rate pulses with the broad spectral selection that is traditionally available from

an OPO.

For remote sensing applications, there is a particular need for high-intensity pulses

with high repetition rates for monitoring atmospheric CO2 and H2O at eye-safe

wavelengths near 2 µm [48-50]. The large bandwidth and spatial coherence available from

these types of OPO’s enable remote sensing over long distances.

Page 25: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

7

1.4 Outline

In this thesis, the author investigates the power-scaling potential of ultrafast OPO’s.

So far, a brief introduction to OPO’s has been given and the capabilities of state-of-the-art

systems has been outlined. The need for increasing the pulse energy of fs oscillators in

high-average-power systems has been described. This thesis will focus on the design and

characterization of one such system. The limitations related to power scaling this OPO will

be discussed and potential improvements for obtaining even higher powers will be

explored.

A description of the components of this thesis are as follows. In Chap. 2 the

fundamental physics behind the relevant processes is presented to the reader. In Chap. 3,

the design and characterization of the scalable thin-disk laser used to pump the OPO are

described. Chapter 4 outlines the design principles used to build the OPO used in this

investigation. Alternative cavity designs and considerations for scaling the cavity to longer

lengths for higher-energy pulses are explored. Chapter 5 shows the results of the OPO

characterization, and explains in detail how these measurements were accomplished.

During the characterization of the OPO it became obvious that full use of the available

pump power was not possible due to difficulties in thermal management of the nonlinear

crystal at high average powers. In Chapt. 6, the methods used to analyze the thermal effects

in the crystal are explored. Using information about the thermal gradient of the crystal

caused by the high-average-power beams, the power limitations of the OPO could be

understood. Finally, Chap. 7 summarizes the findings of this investigation and briefly

discusses possible improvements and future work.

Page 26: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

8

2 Optical Parametric Oscillator

Fundamentals

Starting with the basic principles of nonlinear optics, fundamental concepts that are

critical to the operation and characterization of an ultrafast OPO are explained. The reader

is provided with a brief introduction to nonlinear optics, focusing on χ(2)-related processes.

Phase-matching conditions for these processes are covered with special attention to

critically phase-matched systems. Parametric gain from such devices and the conditions

necessary to achieve synchronous pumping are discussed. Special considerations necessary

for the characterization of subnanosecond pulses such as spectral phase, dispersion, and

self-phase modulation are covered. The chapter concludes with a description of

interferometric autocorrelation as a method for characterizing the temporal shape of

ultrafast pulses experimentally.

2.1 Fundamental Principles of Nonlinear Optics

2.1.1 The Nonlinear Optical Susceptibility

Nonlinear optics is the field of study for phenomena that occur when light of

sufficient intensity can modify the optical properties of a material. To understand how this

occurs, it is useful to consider how the dipole moment per unit volume, or polarization

𝑃(𝑡), of a material depends on the strength of an applied electric field 𝐸(𝑡). For

Page 27: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

9

conventional optical systems, the induced polarization depends linearly on the electric field

strength and is commonly expressed as

𝑃(𝑡) = 𝜖0𝜒(1)𝐸(𝑡), (2.1)

where χ(1) is known as the linear susceptibility and 𝜖0 is the permittivity of free space. In

the presence of intense electrical fields, which is commonly the case in ultrafast systems,

the optical response can often be described by expanding Eq. (2.1) into a power series [51]

as

𝑃(𝑡) = 𝜖0[𝜒(1)𝐸(𝑡) + 𝜒(2)𝐸2(𝑡) + 𝜒(3)𝐸3(𝑡) + ⋯ ], (2.2)

which assumes a lossless and dispersionless medium. The quantities χ(2) and χ(3) are

referred to as the second- and third-order nonlinear optical susceptibilities, respectively. In

general, the nonlinear susceptibilities are complex variables that typically depend on the

frequencies of the applied fields. The real and imaginary parts can generally be represented

by the equation

𝜒(𝑛) = 𝜒(𝑛)′ − 𝑖𝜒(𝑛)′′. (2.3)

The real part of the linear susceptibility is associated with the frequency-dependent linear

refractive index by

𝑛0(𝜔) = √1 + Re{𝜒(1)}, (2.4)

which is also related to linear dispersion and birefringence. The imaginary part of the linear

susceptibility describes the dissipative effect of linear absorption of photon energy by the

medium and the re-emission at longer wavelengths, the latter of which is related to

parametric gain.

The real part of the second-order nonlinear susceptibility is related to parametric

frequency interactions such as three-wave mixing [sum–frequency generation (SFG),

difference–frequency generation (DFG), second-harmonic generation (SHG), and

Page 28: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

10

parametric amplification] and the electro-optic effect. The imaginary part 𝑖𝜒(2)′′ is zero,

which means that 𝜒(2) processes do not store photon energy in the interacting medium.

The real part of the third-order nonlinear susceptibility is connected to four-wave

mixing [e.g., third-harmonic generation (THG)] and the Kerr effect (self-focusing, self-

phase modulation). The imaginary part of the third-order nonlinearity is associated energy

exchange between the photon and the medium, which is responsible for Raman scattering,

Brillouin scattering, and multiphoton absorption.

Since all 𝜒(2) processes involve only phase-related interactions between the optical

fields, with no energy stored in the medium, the nonlinear process becomes particularly

appealing due to the high potential efficiency and relaxed cooling requirements that are

essential for traditional laser gain materials operated at high peak powers. As a result, 𝜒(2)

nonlinear interactions are power scalable, limited only by the damage-threshold (both from

linear and multiphoton absorption) and photorefractive effects [52-54]. The impact of these

factors is dependent on the material properties of the nonlinear crystal used in the

interaction, so the relevance of these limitations is very application specific.

2.1.2 Description of Nonlinear Optical Processes

In the previous section the various nonlinear optical phenomena that can occur from

the second-order nonlinear susceptibility were described. These can be described in terms

of the nonlinear contributions to the polarization described by Eq. (2.2). First, consider a

case in which the optical field incident on a second-order nonlinear optical medium consists

of two frequency components, which can be represented as

𝐸(𝑡) = 𝐸1𝑒−𝑖𝜔1𝑡 + 𝐸2𝑒−𝑖𝜔2𝑡 + 𝑐. 𝑐. . (2.5)

It follows from Eq. (2.2) that the second-order contribution to the nonlinear polarization

can be written as

𝑃(2)(𝑡) = 𝜖0𝜒(2)𝐸(𝑡)2. (2.6)

Page 29: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

11

The nonlinear polarization is then given by

𝑃(2)(𝑡) = 𝜖0𝜒(2)[𝐸12𝑒−2𝑖𝜔1𝑡 + 𝐸2

2𝑒−2𝑖𝜔2𝑡 + 2𝐸1𝐸2𝑒−𝑖(𝜔1+𝜔2)𝑡 + 2𝐸1𝐸2∗𝑒−𝑖(𝜔1−𝜔2)𝑡 +

𝑐. 𝑐. ] + 2𝜖0𝜒(2)[𝐸1𝐸1∗ + 𝐸2𝐸2

∗].

(2.7)

This result can be expressed more conveniently as

𝑃(2)(𝑡) = ∑ 𝑃(𝜔𝑛)𝑒−𝑖𝜔𝑛𝑡𝑛 , (2.8)

where the summation extends over positive and negative frequencies 𝜔𝑛. The complex

amplitudes of the different frequency components of the nonlinear polarization are then

[51]

𝑃(2𝜔1) = 𝜖0𝜒(2)𝐸12 (SHG),

𝑃(2𝜔2) = 𝜖0𝜒(2)𝐸22 (SHG),

𝑃(𝜔1 + 𝜔2) = 2𝜖0𝜒(2)𝐸1𝐸2 (SFG), (2.9)

𝑃(𝜔1 − 𝜔2) = 2𝜖0𝜒(2)𝐸1𝐸2∗ (DFG),

𝑃(0) = 2𝜖0𝜒(2)(𝐸1𝐸1∗ + 𝐸2𝐸2

∗) (OR).

Here, each expression has been labeled by the physical process it represents. The

dependence of SHG, SFG, and DFG on the second-order nonlinear susceptibility 𝜒(2) was

described above. The final expression describes the optical rectification (OR), which is the

inverse process of the electro-optic effect [51].

2.1.3 Coupled-Wave Equations

Having described how the nonlinear polarization can be characterized when a

medium is subject to an intense electric field, we can now apply Maxwell’s equations to

describe the generation of these components to the electromagnetic field. This will show

Page 30: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

12

how the interacting fields become coupled by the nonlinear interaction and what conditions

must be met to make the process efficient. An extension of the standard wave equation is

necessary to address the additional second–order nonlinear polarization and can be stated

as [51]

𝛻2𝑬 −1

𝑐2

𝜕2

𝜕𝑡2𝑬 =

1

𝜖0𝑐2

𝜕2

𝜕𝑡2𝑷. (2.10)

The derivation of the wave equations is cumbersome, even with the limitation of

restricting the analysis to three frequencies, 𝜔𝑝, 𝜔𝑠, 𝜔𝑖, and propagation along a single

direction (collinear), commonly denoted as 𝑧. If the three interacting waves are defined

such that 𝜔𝑖 < 𝜔𝑠 < 𝜔𝑝, they can be referred to as the pump, signal, and idler waves,

respectively, as is common in parametric systems. It is convenient to express the electric

field as a linearly polarized, monochromatic plane wave at frequency 𝜔, progagating in the

𝑧 direction:

𝐸(𝑧, 𝑡) = Re{𝐴(𝑡)exp [𝑖(𝜔𝑡 − 𝑘𝑧)]}, (2.11)

where 𝑘 is given by

𝑘 =2𝜋

𝜆. (2.12)

Within the slowly varying amplitude approximation the following propagation equation

can be derived:

𝑑𝐴

𝑑𝑧= −𝑖

𝜇0𝑐0𝜔

2𝑛𝑃𝑒−𝑖𝑘𝑧, (2.13)

where 𝜇0 is the permeability of free space and 𝑛 is the index of the material. The resulting

coupled-wave equations are expressed as [7, 51]

𝑑𝐴𝑖

𝑑𝑧= −𝑖

𝜔𝑖𝑑eff

𝑛𝑖𝑐0𝐴𝑠

∗𝐴𝑝𝑒−𝑖∆𝑘𝑧, (2.14)

𝑑𝐴𝑠

𝑑𝑧= −𝑖

𝜔𝑠𝑑eff

𝑛𝑠𝑐0𝐴𝑖

∗𝐴𝑝𝑒−𝑖∆𝑘𝑧, (2.15)

Page 31: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

13

𝑑𝐴𝑝

𝑑𝑧= −𝑖

𝜔𝑝𝑑eff

𝑛𝑝𝑐0𝐴𝑖𝐴𝑠𝑒𝑖∆𝑘𝑧, (2.16)

where 𝑑eff is the so-called effective nonlinear optical coefficient, which is dependent on

the propagation duration and the polarization of the three beams, and

∆𝑘 = 𝑘𝑝 − 𝑘𝑠 − 𝑘𝑖, (2.17)

defined as the wave-vector mismatch.

2.1.4 Critical Phase Matching

Efficiency in the nonlinear interaction is maximized when ∆𝑘 = 0. This condition

can be achieved the by careful choice of propagation directions inside a dispersive and

birefringent medium. The phase-matching condition can be recast in the form

𝑛𝑝 =𝑛𝑖𝜔𝑖+𝑛𝑠𝜔𝑠

𝜔𝑝. (2.18)

It is straightforward to show that this condition cannot be fulfilled in bulk isotropic

materials in the usual dispersion region (𝑛𝑖 < 𝑛𝑠 < 𝑛𝑝). In certain birefringent crystals,

phase matching can be accomplished by choosing the polarization direction for the higher-

frequency pump (𝜔𝑝) that gives a lower refractive index. In this case, which is common

for ultrafast parametric devices, negative uniaxial crystals (𝑛𝑒 < 𝑛𝑜) can be used. When

the pump beam is polarized along the extraordinary axis of the crystal, phase matching can

be achieved for a certain combination of the signal and idler wavelengths. If both the signal

and idler beams are polarized perpendicular to the ordinary axis of the crystal, it is referred

to as type I (or 𝑜𝑠 + 𝑜𝑖 → 𝑒𝑝) phase matching. If one of the two is polarized in the same

direction as the pump, it is referred to as type II phase matching. This can be the case when

either the signal (𝑒𝑠 + 𝑜𝑖 → 𝑒𝑝) or the idler (𝑜𝑠 + 𝑒𝑖 → 𝑒𝑝) has the polarization along the

extraordinary axis [55]. This condition can of course be reversed for certain situations using

positive uniaxial crystals (𝑛𝑜 < 𝑛𝑒), where the pump is oriented along the ordinary axis.

Additionally, this procedure can be extended to biaxial crystals, which are typically

Page 32: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

14

operated on one of the three optical planes, which can then be treated as a negative or

positive uniaxial crystal. Operation outside of these planes is possible, although many

biaxial nonlinear crystals are difficult to characterize and the nonlinear interactions are hard

to predict [56].

Usually, the phase-matching condition is achieved by adjusting the angle 𝜃𝑚

between the wave vector of the propagating beam and the extraordinary (optical) axis of

the nonlinear crystal (angular phase matching). Alternatively, the refractive indices can be

changed by adjusting the temperature of the crystal (temperature or noncritically phase

matching). Another method, quasi phase matching, utilizes periodic polling of the material

to achieve phase matching over long interaction lengths with pump, signal, and idler beams

having the same polarization. Critical phase matching was employed in this investigation,

and the methods used to optimize this process for broadband beams require further

explanation.

As an example, we consider a negative uniaxial crystal where type I phase matching

is achieved when [57]

1

𝑛𝑒,𝑝2 (𝜃𝑚)

=sin2(𝜃𝑚)

𝑛𝑒,𝑝2 +

cos2(𝜃𝑚)

𝑛𝑜,𝑝2 , (2.19)

where 𝑛𝑒,𝑝 and 𝑛𝑜,𝑝 are the principle extraordinary and ordinary refractive indices at the

pump wavelength. The phase-matching angle can then be obtained using

𝜃𝑚 = asin [𝑛𝑒,𝑝

𝑛𝑒,𝑝(𝜃𝑚)√

𝑛𝑜,𝑝2 −𝑛𝑜,𝑝

2 (𝜃𝑚)

𝑛𝑜,𝑝2 −𝑛𝑒,𝑝

2 ]. (2.20)

It is worth noting that, in general, the phase-matching angle shows a less-pronounced

wavelength dependence for type I with respect to type II phase matching. The above

analysis assumes a collinear interaction geometry between the monochromatic pump,

signal, and idler beams. The propagation direction is typically selected to satisfy the phase-

matching condition for a given signal wavelength.

Page 33: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

15

This analysis can also be extended for broadband pulses, which is particularly

useful for ultrafast pulses. The coupled wave equations can be extended to include waves

propagating in the nonlinear crystal with different group velocities 𝜐𝑔 = 𝜕𝜔/𝜕𝑘 [58, 59].

The resulting wave equations show how the amplification of ultrafast pulses is significantly

related to the group-velocity mismatch (GVM) between the interacting pulses. The GVM

between the pump and amplified pulses limits the interaction length, where energy can be

efficiency transferred from the pump due to temporal walk-off. Additionally, the GVM

between the signal and idler pulses limits the phase matching bandwidth available and

therefore is the main limitation for creating shorter pulses. It can be shown that if the signal

frequency increases to 𝜔𝑠 + ∆𝜔 (and by energy conservation, the idler frequency decreases

to 𝜔𝑖 − ∆𝜔), then the wave-vector mismatch can be approximated to first order, as [60]

∆𝑘 ≅ −𝜕𝑘𝑠

𝜕𝜔𝑠∆𝜔 +

𝜕𝑘𝑖

𝜕𝜔𝑖∆𝜔 = (

1

𝜐𝑔𝑖−

1

𝜐𝑔𝑠)∆𝜔. (2.21)

For a parametric process using a collinear interaction geometry, the propagation

direction is selected to satisfy the phase-matching condition for a certain signal wavelength.

In this condition, the signal and idler group velocities are fixed and, as a result, the-phase

matching bandwidth. An additional degree of freedom can be introduced by using a

noncollinear geometry, which is illustrated in Fig. 2.1.

Figure 2.1: A schematic of the noncollinear interaction geometry where angles have been

exaggerated for clarity. Not shown here is the angle Ω between the signal and idler beams,

which varies with signal wavelength.

Page 34: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

16

The angle 𝛼 between the pump and signal wave vectors is constant for a broadband

signal beam. The idler is emitted at an angle Ω with respect to the signal and is dependent

on the signal wavelength. This results in an angularly dispersed but broadband idler pulse.

The phase-matching condition can be described as a vector equation, which can be

separated into two parts (parallel and perpendicular to the signal wave vector), and

becomes

𝛥𝑘par = 𝑘𝑝 cos 𝛼 − 𝑘𝑠 − 𝑘𝑖 cos 𝛺 = 0, (2.22)

𝛥𝑘perp = 𝑘𝑝 sin 𝛼 − 𝑘𝑖 sin 𝛺 = 0. (2.23)

Like the collinear analysis, if the signal frequency increases by Δ𝜔, the wave-vector

mismatch along the two directions can be approximated (to first order) as

∆𝑘par ≅ −𝜕𝑘𝑠

𝜕𝜔𝑠∆𝜔 + cos Ω

𝜕𝑘𝑖

𝜕𝜔𝑖∆𝜔 − 𝑘𝑖 sin Ω

𝜕𝛺

𝜕𝜔𝑖𝛥𝜔, (2.24)

∆𝑘perp ≅ − sin Ω𝜕𝑘𝑖

𝜕𝜔𝑖∆𝜔 + 𝑘𝑖 cos Ω

𝜕Ω

𝜕𝜔𝑖𝛥𝜔. (2.25)

To obtain broadband phase matching, both Δ𝑘par and Δ𝑘perp must be zero. With some

simple arithmetic, this condition be shown to be equivalent to

𝜕𝑘𝑖

𝜕𝜔𝑖− cos Ω

𝜕𝑘𝑠

𝜕𝜔𝑠= 0, (2.26)

which can be expressed more simply as

𝜐𝑔𝑠 = 𝜐𝑔𝑖 cos Ω. (2.27)

Equation (2.27) shows that broadband phase matching can be achieved for an angle

Ω so that the group velocity of the signal equals the projection of the group velocity of the

idler along the signal direction. In the collinear case, the signal and idler moving with

different group velocities will get separated over a small interaction length, causing a

reduction in bandwidth. For a noncollinear geometry the two pulses can stay overlapped,

Page 35: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

17

allowing for significantly more broadband amplification. The signal–idler angle Ω is

important for understanding the improvement in performance between collinear and

noncollinear geometries; however, it is much more helpful to know the pump–signal angle

𝛼 for the design and operation of an ultrafast device. Without losing conditions required

for broadband phase matching, the pump–signal angle is given by [60]

𝛼 = arcsin (1−𝜐𝑔𝑠

2 /𝜐𝑔𝑖2

1+2𝜐𝑔𝑠𝑛𝑠𝜆𝑖𝜐𝑔𝑖𝑛𝑖𝜆𝑠

+𝑛𝑠2𝜆𝑖

2/𝑛𝑖2𝜆𝑠

2)

1/2

. (2.28)

2.2 Characteristics of Optical Parametric Oscillators

2.2.1 Parametric Gain

The gain and amplification of the signal and idler fields in the parametric process

can be derived from the coupled-wave equations presented above. As is characteristic of

the system in this study, if pump depletion is neglected (𝐴𝑝 ≅ const.) and assuming the

signal intensity 𝐴𝑠0 and no initial idler beam, Eqs. (2.14)–(2.16) can be solved for the signal

and idler intensities after passing through a nonlinear crystal of length 𝐿 [60]:

𝐼𝑠(𝐿) = 𝐼𝑠0 [1 +𝛤2

𝑔2 sinh2(𝑔𝐿)], (2.29)

𝐼𝑖(𝐿) = 𝐼𝑠0𝜔𝑖

𝜔𝑠

𝛤2

𝑔2 sinh2(𝑔𝐿), (2.30)

where

𝑔 = √Γ2 − (𝛥𝑘

2)

2

, (2.31)

Γ2 =2𝜔𝑖𝜔𝑠𝑑eff

2 𝐼𝑝

𝑛𝑖𝑛𝑠𝑛𝑝𝜖0𝑐03 . (2.32)

Page 36: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

18

In the case of perfect phase matching (Δ𝑘 = 0, 𝑔 = Γ) and in the large gain approximation

(Γ𝐿 ≫ 1), the intensities of the signal and idler simplify to

𝐼𝑠(𝐿) ≅1

4𝐼𝑠0exp (2Γ𝐿), (2.33)

𝐼𝑖(𝐿) ≅𝜔𝑖

4𝜔𝑠𝐼𝑠0exp (2Γ𝐿). (2.34)

It is also important to note that the ratio of signal and idler intensities must be such that an

equal number of signal and idler photons are created. Equations (2.33) and (2.34) allow the

parametric gain to be defined as

𝐺 =𝐼𝑠(𝐿)

𝐼𝑠0=

1

4exp(2Γ𝐿), (2.35)

where the gain grows exponentially with the crystal length and the nonlinear coefficient Γ.

It is worth mentioning that most of the gain control is done through careful consideration

of the parameters in Γ, which is dependent on the pump intensity, signal and idler

wavelengths, the effective nonlinearity of the material, and the refractive indices at the

interacting wavelengths.

This exponential growth of the signal and idler beams is qualitatively different from

the gain seen for other second-order nonlinear processes, such as SFG and SHG. In a strong

pump field, the presence of an existing signal photon stimulates the generation of an

additional signal photon as well as a photon at the idler wavelength. Additionally, due to

the symmetry of the signal and the idler, the amplification of the idler stimulates the

generation of a signal photon. The amplification of the signal field promotes the generation

of the idler field and vice versa, giving rise to exponential growth of both waves.

In the presence of imperfect phase matching, it is useful to create a metric that

expresses the gain as a function of the phase matching and crystal length. In the large gain

approximation, the parametric gain can be shown to be proportional to the phase-matching

condition and the crystal length as [61, 62]

Page 37: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

19

𝐺 ∝ sinc2(∆𝑘∙𝐿

2). (2.36)

As has been shown for critically phase-matched systems, ∆𝑘 is dependent on the angle

between the extraordinary beam and the crystal axis, which allows this expression to

become useful in determining the gain available in a parametric device when the angular

acceptance or crystal length is a concern for efficient operation.

2.2.2 Synchronous Pumping and Cavity Length Stabilization

The development of ultrafast pump sources has greatly increased the use of

synchronous pumping in OPO’s. To achieve optimal synchronous pumping, the optical

path length inside the OPO cavity must equal that of the pump source. For ultrafast systems,

this tolerance can be quite tight (of the order of tens of µm), creating restrictions on the

cavity length not typically present in conventional lasers.

The stability of the pulse duration, operating wavelength, repetition frequency, and

output power is extremely important for utilizing the potential of synchronously pumped

OPO’s. The stabilization of the OPO cavity is an important parameter to control when it is

synchronously pumped. In traditional synchronously pumped lasers, the intrinsic gain

storage allows for a small repetition rate mismatch at the expense of an increase in the pulse

duration [63]. However, a synchronously pumped OPO responds to the mismatch by

utilizing the group-delay dispersion (GDD) in the cavity. The wavelength-dependent gain

of an OPO usually results in a change in the central wavelength by modifying the cavity

length, which subsequently modifies the power of the OPO [64]. As a result, changes in

the pump laser’s repetition rate, or environmental fluctuations (e.g., heating or air flow)

that modify the optical path length for either the pump or the OPO, will result in a myriad

of instabilities in the OPO itself.

A straightforward way to stabilize the power level of any source could be to monitor

the output power and adjust the cavity length to maintain a constant power. In

synchronously pumped lasers, the sensitivity of the pulse repetition rate to the cavity length

Page 38: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

20

can be exploited to provide an error signal by comparing the rate to either an electronic

oscillator or a reference cavity [65-67].

2.3 Properties of Ultrafast Pulses

Ultrafast pulses are continuously reshaped by their temporal and spectral profile by

both linear and nonlinear mechanisms. These can be from the material in which the pulse

propagates as well as dispersion from the surfaces on which the beam is incident inside the

resonator. In some cases, this can be used to optimize the pulse dynamics and counteract

undesirable effects inside the cavity. To understand these effects, a brief description of the

involved process is given in this section.

2.3.1 Definition of Spectral Phase and Dispersion

It is convenient to consider a wave traveling through a medium where the refractive

index does not change as the pulse propagates. Similar to the procedure employed in Sec.

2.1 but neglecting spatial variation, consider that a pulse with a center frequency at 𝜔0 can

be represented by a time-dependent complex envelope function with the real electric field

as

𝐸(𝑡) ∝ Re{𝐴(𝑡)𝑒𝑖𝜔0𝑡}, (2.37)

where 𝐴(𝑡) is the normalized time-dependent electric field amplitude with the

instantaneous optical power |𝐴(𝑡)|2, which is monitored by a conventional detector. A

Fourier transform of the complex field gives the spectral domain analytically as

𝐴(𝜔) =1

2𝜋∫ 𝐴(𝑡)𝑒−𝑖𝜔0𝑡𝑑𝑡

−∞. (2.38)

Since the electric field amplitude has been defined here as a normalized quantity, we can

equivalently obtain the real power spectrum of the pulse with |𝐴(𝜔)|2.

The relationship between the temporal width of a pulse ∆𝜏𝑝 and its spectral

bandwidth ∆𝜈𝑝 is commonly referred to as the time–bandwidth product (TBP). In this

Page 39: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

21

study, both the spectral and temporal widths of a pulse are measured at the full width at

half maximum (FWHM), although other definitions (such as 1/𝑒2) are used. Two

important temporal shapes for ultrafast pulses are the Gaussian pulse (𝑒−2𝑙𝑛2 𝑡2/𝛥𝜏𝑝2) and

the hyperbolic secant pulse [sech2(1.763𝑡/∆𝜏𝑝)]. The minimum TPB of a Gaussian pulse

and hyperbolic secant pulse is 0.441 and 0.315, respectively. Any chirping (either positive

or negative) on the pulse or higher-order dispersion effects will increase the time-

bandwidth product and is typically useful for determining the minimum pulse duration

possible from the amplified spectrum in the parametric process.

Except for a vacuum, any medium introduces dispersion, which arises from the fact

that the refractive index is frequency dependent, implying that the spectral phase

components of a pulse travel at different velocities through the medium, introducing a

phase delay. This can be characterized by first introducing a frequency-dependent phase

term with

𝐴(𝜔)′ = 𝐴(𝜔)𝑒−𝑖𝜑(𝜔). (2.39)

The spectral phase variations corresponding to the first-, second-, and higher-order

dispersions is usually defined via the Taylor expansion of the wave vector 𝑘, giving the

spectral phase change per unit length as a function of the angular frequency 𝜔 relative to

the center frequency of the pulse 𝜔0:

𝜑(𝜔) = 𝜑(𝜔0) + (𝜔 − 𝜔0) (𝜕𝜑

𝜕𝜔) |𝜔0

+1

2(𝜔 − 𝜔0)2 (

𝜕2𝜑

𝜕𝜔2) |𝜔0

+ ⋯ . (2.40)

The zero-order term represents a common phase shift of the carrier wave that has no effect

on the pulse shape and phase distortions. The first-order term introduces the linear change

of the phase in the frequency domain, where the corresponding time domain component is

commonly referred to as “group delay.”

Qualitatively, the higher-order terms of the expansion are a measure of the change

of the group delay with respect to wavelength and have a significant impact pulse dynamics

Page 40: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

22

during propagation. The GDD describes how the delay changes during propagation inside

a dispersive medium and is given by

GDD = (𝜕2𝜑

𝜕𝜔2) |𝜔0 (2.41)

and is typically given in units of fs2.

The third-order term of the expansion is associated with the cubic spectral phase

variations and is known as third-order dispersion (TOD), where

TOD = (𝜕3𝜑

𝜕𝜔3) |𝜔0 (2.42)

and has units of fs3. This process leads to a steepening of one edge of the pulse and a

stretching of the other, which can eventually lead to a separation of the pulse, commonly

referred to as “pulse breakup” [68]. TOD can be the cause of unforeseen pulse defects due

to the difficulty involved in controlling material properties in manufacturing. Optical

coatings that are widely available can reduce GDD or be able to achieve discrete levels of

GDD. However, balancing the needs for an ultrafast system’s high-power or ultra-

broadband reflectivity usually comes at the cost of significant TOD, leading to unwanted

pulse distortions. Controlling the net intracavity dispersion is critical for achieving ultrafast

pulses (especially those in the few-cycle regime). This is normally achieved once the

quadradic and higher-order spectral phase terms in a single round-trip are negligible across

the spectral bandwidth of the pulse. The accumulated dispersion may result from several

sources, some of which will be discussed in the next section.

2.3.2 Sources of Dispersion

Material dispersion is associated with the frequency-dependent refractive index of

a dielectric medium, which, in an OPO, is introduced by the nonlinear crystal. For

propagation in bulk dielectric materials, the dispersion can be determined in terms of the

rate of change of the index with respect to the wavelength. The wavelength-dependent

refractive index is available for many materials through the use of a Sellmeier equation

Page 41: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

23

[69, 70]. Therefore, in practice it is much more useful to define the GDD and TOD in terms

of the refractive index and wavelength, which are given by

GDD =𝜆3𝑧

2𝜋𝑐2

𝑑2𝑛

𝑑𝜆2 (2.43)

and

TOD = −𝜆3𝑧

2𝜋𝑐2 (3𝑑2𝑛

𝑑𝜆2 + 𝜆𝑑3𝑛

𝑑𝜆3). (2.44)

Contrary to material dispersion, geometrical dispersion arises from the wavelength-

dependent path lengths that can be overserved in optical systems that have angular

dispersion, such as prisms or diffraction gratings. The GDD and TOD of these dispersive

elements may be controllable by a careful combination of both the materials and layout

used for dispersion compensation.

The use of a prism to control dispersion (both inside or outside the cavity) requires

careful calculation of the material dispersion introduced by traveling through additional

glass and the relative angles of incidence for the cut of the prism. This can be calculated

using [71, 72]

GDD =𝜕2𝜑

𝜕𝜔2 =𝜆3

𝜋𝑐2𝑟2

𝑑2𝐼

𝑑𝜆2. (2.45)

Another common method to achieve geometrical dispersion is the use of diffraction

gratings, although none were used in this investigation. The GDD available from a pair of

gratings and the relevant properties of such devices is given in [73, 74].

Another alternative that has been widely used in fs lasers is chirped mirrors with

multilayer coatings [75, 76]. The use of chirped mirrors allows one to add discrete amounts

of GDD to the laser with low loss; these mirrors can be specifically designed to compensate

for higher orders of dispersion [77]. While chirped mirrors usually limit the available GDD

to a few fs2, Gires-Tournois interferometer (GTI) mirrors have been used to generate larger

amounts of negative GDD inside a laser resonator [78, 79].

Page 42: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

24

2.3.3 Self-phase Modulation

Self-phase modulation (SPM) occurs when an optical pulse with sufficiently high

intensity changes the refractive index of the material in such a way that the phase of the

pulse is affected. The origin of this effect is directly related to the optical Kerr effect and

can be characterized by a refractive index according to [51, 80]

𝑛 = 𝑛0 + 𝑛2𝐼(𝑡), (2.46)

where 𝐼(𝑡) = 2𝑛0𝜖0𝑐|𝐴(𝑡)|2. It is assumed that the medium is able to respond

instantaneously to the pulse and the length of the medium is sufficiently small that the

optical pulse will not be shaped by the medium. The resulting frequency shift of the pulse

after traveling through a medium with length 𝐿 can be expressed as

𝜑𝑁𝐿(𝑡) = −𝑛2𝐼(𝑡)𝜔0𝐿/𝑐. (2.47)

The result of the time-varying phase component of the wave is that the spectrum will be

modified and, in most cases, broader than the initial pulse. In such cases the center of the

pulse undergoes a nearly linear positive chirp, which can be compensated for with an equal

amount of negative GDD, as is commonly done with systems employing temporal solitons.

2.4 Pulse Characterization Techniques

Temporal properties of ultrafast pulses are not accessible with any direct electrical

measurement technique. With the growing popularity of ultrafast systems in the last few

decades, many techniques have been created to either estimate or precisely acquire the

temporal shape. Ultrafast pulses must be sampled with the use of a second ultrafast pulse

and measured indirectly. One commonly employed estimation is temporal autocorrelation.

The interferometric autocorrelation function is the correlation of the electric field with

itself [81]:

𝑔2(𝜏) =∫ |{𝐸(𝑡)+𝐸(𝑡−𝜏)}2|

2𝑑𝑡

∞−∞

∫ |𝐸2(𝑡)|2𝑑𝑡∞

−∞

, (2.48)

Page 43: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

25

requiring two identical pulses with a variable delay 𝜏 between them. The most common

experimental setup is a classical Michelson interferometer with a beam splitter providing

equal power to the pulse in each arm. The length of one arm of the interferometer is

repetitively changed to induce a delay from −𝜏 to 𝜏. The two pulses are sent to a nonlinear

crystal, where the SHG signal of the interfering pulses is measured, normally using a

photodiode or a photomultiplier tube. If the detector has insufficient bandwidth, the

autocorrelation fringes become smeared into averaged traces known as an intensity

autocorrelation, analytically described as

𝑠(𝜏) = 2∫ 𝐼(𝑡)𝐼(𝑡−𝜏)𝑑𝑡

∞−∞

∫ 𝐼2(𝑡)𝑑𝑡∞

−∞

. (2.49)

The measured autocorrelation signal is always symmetric and requires an assumption of

the pulse shape to calculate the pulse duration. For pulses longer than a few cycles, this

assumption can be normally justified by mode-locking, such as the case with soliton pulses

having a sech2 pulse shape.

Often, the spectrum of the pulse used in the autocorrelation is measured

simultaneously (or separately given the same operating conditions) to determine qualitative

effects of the pulse’s phase. The width of the spectrum can easily be compared the to

calculated width of the pulse. If the TBP has not been minimized for the given pulse shape,

uncompensated GDD and higher-order dispersion can be attributed to the increase in pulse

duration.

Because the spectral phase cannot be determined with this approach, it is impossible

to retrieve the exact pulse shape. For pulses containing complicated spectra and complex

phase profiles, frequency-resolved optical gating (FROG) or spectral phase interferometry

for direct electric-field reconstruction (SPIDER) offer a much more systematic way to

obtain temporal shape of the pulse [82-84]. These techniques can provide the information

necessary for a true characterization of the pulse profile and are particularly useful for few-

cycle ultrafast pulses.

Page 44: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

26

3 Thin-Disk Mode-Locked Pump

Laser

A critical component to a synchronously pumped OPO is the pump laser. This

section provides a description of the home-built thin-disk laser used to pump the ultrafast

OPO. First, a brief history of thin-disk systems is presented. The fundamental properties

of Yb:YAG and physics for operating thin-disk lasers at high powers are discussed.

Modifications to the design of this laser relative to previous system are described, with

special attention to the active multipass cavity used to extend the cavity length and achieve

high-energy pulses at high average powers. Finally, the performance of the laser as a pump

source is characterized.

3.1 A Review of Thin-Disk Lasers

The creation of the thin-disk laser in 1994 gave rise to a reliable source of scalable

ultrafast pulses that has been in strict competition with fiber- and slab- based oscillators

[85]. The common point between these technologies is the cooling geometry of the gain

medium, which is suitable for high-average-power operation and power scaling. Among

these, thin-disk technology is ideally suited for short pulses from the minimization of both

nonlinearities and thermal distortions [86].

Thin-disk mode-locked oscillators and amplifiers have both gained significant

ground in the past few years toward the power scaling of ultrafast pulses [87-90].

Unprecedented high-energy and average-power levels can be reached directly from a

passively mode-locked oscillator without any external amplification. The advantages of

this oscillator over amplifier systems with similar performance are in terms of noise level,

Page 45: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

27

pulse quality, reliably, long-term stability, and complexity, if adequate engineering effort

is applied.

Most of the energy and power scaling of mode-locked thin-disk oscillators has used

Yb:YAG as the gain medium [91]. Many alternative Yb-doped materials are currently

being investigated to improve thin-disk systems by increasing the available gain

bandwidth, which has already been shown to produce pulses with durations less than 200

fs [92-96].

The thin-disk mode-locked laser (TDML) used in this study was based on earlier

work that has since been surpassed due to advances in both systems engineering and

material properties [97, 98]. The TDML was a home-built laser that helped to improve the

expertise needed to build a complicated free-space oscillator as well as serve as a test bed

for the capabilities in power-scaling pump sources. With the recent advancements to thin-

disk technology, it is obvious that these systems can be used as extraordinary pump

sources, well surpassing traditional pump lasers.

3.2 Principles of Thin-Disk Operation

3.2.1 Thin-Disk Laser

Thin-disk lasers can achieve stable operation at extremely high average powers by

using a material thickness significantly smaller than the spot diameter of the pump and

laser mode [85]. The thin-disk is mounted on a heat sink to allow for nearly one-

dimensional heat flow, which reduces the transverse refractive-index gradient. The thin

disk is chemically bonded to the heat sink, which is water cooled with a ceramic–diamond

composite [99]. Compared to other solid-state lasers, the ratio of the cooling surface to the

pumped volume is increased, which is a basic advantage to extracting high average powers

from a small volume [100].

An antireflection (AR) coating is placed on the front of the disk to minimize the cw

pump and laser reflectivity and a high-reflection (HR) coating on the back surface. This

Page 46: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

28

allows the thin disk to act as an end mirror for the pumping geometry and the laser cavity

while being actively cooled.

Since the thickness of the disk is ~100 µm, the absorption efficiency of the pump

light is very low. To get a reasonable efficiency, the unabsorbed cw-pump beam is re-

imaged onto the disk several times, as shown in Fig 3.1. The thin-disk laser head used in

this work was provided by TRUMPF Laser GmbH, which was based on the concept of

pump geometry allowing up to 20 passes of pump light through the thin disk [85].

Figure 3.1: Schematic of the thin-disk laser head. The parabolic mirror allows for 20 passes

of pump light through the thin disk by using several roof prisms (not shown here). Ten

passes are completed before the pump beam is reflected [101]. The front of the disk is AR

coated for 940 nm and 1030 nm and HR coated on the back at the same wavelengths.

3.2.2 Yb:YAG as a Laser Material

The spectroscopic properties of Yb3+-doped laser materials are based on a simple

[Xe]4f116s2 electronic structure, as shown in Fig 3.2(a). Two manifolds, 2F7/2 and its

excited-state counterpart 2F5/2, and their subordinate Stark levels participate in the lasing

process. The two manifolds and the spectral distribution of the cross section are shown in

Fig 3.2(a). Since YAG has a relatively high Debye temperature (~ 550 K) and Stark levels

Page 47: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

29

with a distance ≈1000 cm-1 in between one another, a Boltzmann population distribution

between them will be established at the lattice temperature in less thn 10 ps [102].

Figure 3.2: (a) Energy levels for Yb:YAG [103]; (b) absorption and emission cross section

of a 5%-doped Yb:YAG crystal at room temperature [103].

The pump ground level, which is the lowest Stark level of the lower manifold, is

populated at room temperature with a probability of 88%, while the ground level is given

by the upper three Stark levels of the lower manifold, each being only 200 cm–1 apart [103].

The specific arrangement of the energy levels results in a thermal population of the lasing

ground levels at room temperatures. This results in Yb3+-doped laser materials behaving as

a quasi-three-level system and show ground-state absorption at the laser wavelength. As a

result, the lasing threshold pump power is significantly increased ( > 1.5 kW/cm2 at 300 K

[85]) in comparison to operation at lower temperatures as a result of cooling, which can

reduce the lower laser-level population and improve efficiency.

Quantum defect is one of the greatest sources of heat generation for solid-state laser

sources. A small quantum defect of 9% when pumped at 941 nm is a major advantage of

Yb3+ compared with other four-level systems. The resulting low heat deposition makes

Yb3+ an attractive material for high-power lasers. The ratio of generated heat to absorbed

energy is < 11% for Yb:YAG [85], while Nd:YAG is ~35% when pumped at 808 nm [104].

Page 48: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

30

Small pump-absorption bandwidths require sophisticated pumping concepts, while

large bandwidths allow for conventional high-power pump sources (e.g., semiconductor

diode arrays) without strict requirements on their operating temperature to keep the

emission wavelength constant. Yb3+-doped laser materials can be pumped by many

conventional sources, such as strained layer InGaAs diode lasers in relatively large

wavelength ranges between 0.9 and 1.05 µm as shown by the absorption and emission cross

sections in Fig. 3.2(b).

While pumping at 972 nm [105] would result in a lower quantum defect than

pumping at 941 nm, the absorption bandwidth is significantly lower, requiring more

sophisticated and expensive wavelength-stabilized pump diodes. The absorption peak at

941 nm has a bandwidth roughly 5x larger than 972 nm while maintaining a low quantum

defect. The wide absorption peak at 941 nm allows one to use less-expensive cw-pump

sources that do not require fine control of the pump’s linewidth.

3.2.3 Thermal Lensing

Intense pumping of solid-state media inevitably leads to optical-phase distortions.

For this type of media, the thermally induced stress can result in a different expansion of

the front and back sides of the thin disk and bending of the disk, leading to a thermal lens.

Although the thin disk allows for improved one-dimensional heat flow over conventional

solid-state gain media, the presence of high average powers still results in some phase

distortions that can have a significant impact on laser performance.

For Yb:YAG the bending of the disk dominates for large-mode sizes on the disk

[106], resulting in defocusing. Additionally, the refractive index of the material might

change by the excitation of the Yb3+ ions. The linear polarizability of the excited ion differs

from that of the unexcited ion [107] and for intensively pumped Yb:YAG thin disks, the

index change caused by electronic mechanisms can become even larger than those caused

by thermal heating [108].

Page 49: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

31

The thermal-lens dioptric power caused by the change in index is different when

operating under lasing or non-lasing conditions. When lasing, the dioptric power shows

an approximate linear relationship with the absorbed pump power, but under non-lasing

conditions there is a noticeable roll-off in the observable dioptric power [109]. The index

change and thermal bending lead to a spherical optical phase distortion with opposite signs,

suggesting that there must be a combination of the thickness of the thin disk and pump-

spot diameter for a specific lasing condition, such that thermal lensing vanishes.

One can compensate for spherical optical phase distortions by adjusting the stability

zone of the resonator. However, nonspherical distortions result in surface scattering losses

and higher mode operation. For this investigation, these should be negligible since it has

been shown that nonspherical phase distortions for Yb:YAG thin-disk lasers occur only for

pump powers exceeding 300 W (with optical phase distortions < 100 nm [100]).

Nonspherical phase distortions occur mostly at the edges of the pump spot suggesting that

the mode diameter of the laser is optimized when the mode overlap is 80% of the pump for

fundamental mode operation [101].

3.3 Cavity Design

To utilize the potential of thin-disk systems, a home-built Yb:YAG thin-disk mode-

locked laser (TDML) was constructed. In the following section the design of the TDML

pump is discussed.

3.3.1 Thin-Disk Head

The pump laser was a high-power diode laser module from Jenoptik, model JOLD-

250-CPXF-2P2-938-0.4. This laser can provide up to 250 W at wavelengths near 940 nm.

The power supply was a Xantrex XFR12-220 DC Supply. Operation of the laser was

operated by keeping the voltage at a near constant value of 6.05 V and varying the current

to modify the output power and wavelength.

Page 50: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

32

The diode laser was water cooled with a Thermo Scientific NESLAB RTE-220

stand-alone chiller, capable of providing 4 liters per minute at a temperature range of 8 to

40°C. An interlock was used to protect the laser from overheating and to ensure a constant

wavelength of the pump laser. The output of the laser module was connected to a fiber

coupler, injecting the light into a fiber with a 600-µm core diameter with no alignment or

fine tuning needed.

Power measurements were made for a specific temperature setting, which was used

to tune the wavelength output at optimal power. In all cases the spectrum changed as the

current supplied to the diode was increased. Before an optimal pump power was identified,

spectral measurements were made across the operating powers for different temperatures

to ensure that the correct spectrum would be used after the pump power had been

optimized.

Alignment of the pump laser onto the thin disk required optimization of six separate

degrees of freedom. The first was the proper collimation of the fiber output, which was

tested with low powers (in LED mode) over a 9-ft distance. The next were the x and y

positions of the collimation optics. Both the x and y axes are controlled using a custom

collimation package from TRUMPF. The y axis is controlled with a spring-loaded screw.

An optimum y position was chosen such that the collimation optics were in the same plane

as the thin disk. The x axis was controlled using two set screws on the left and right sides

of the collimation optics. The optimal position was chosen such that the first reflected beam

was located near the center of the thin disk.

It was crucial to ensure the flatness between the parabolic mirror and the

collimation optics. Both the collimation optics and the thin-disk head were fastened to a

custom metal block that was screwed into the optical table, as shown in Fig. 3.3. The

collimation optics allowed for rough pointing and x-axis centering, while fine tuning could

be done with optomechanical controls. The pointing of the thin-disk head was more

restricted. Slight tuning of the pointing of the parabolic mirror and roof mirrors (enclosed

in a single mount shown in red) with respect to the thin-disk mount (shown in purple) could

Page 51: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

33

be done by rotating of the mount. The thin-disk mount was the focal point of the alignment

after making the mount flush with the height change in the plate, shown as the black line.

The 22.5-mm change in height was due to the adaptation of TRUMPF components to fit

on the optical table.

Figure 3.3: Diagram of baseplate created for the thin-disk head and collimation optics. The

collimator is shown in blue. The thin-disk mount and heat sink are shown in purple. The

parabolic mirror and roof prisms are enclosed in a metal housing, shown in red.

This mount had minor degrees of freedom to adjust the flatness between the

parabolic mirror and the collimation mount. The x and y positions of the parabolic mirror

and roof prisms could be used to fine tune the position of the mirror with respect to the

disk. The thin disk could be moved in and out of the focal plane of the parabolic mirror to

achieve the uniform pump spot needed to create a super-Gaussian mode profile.

The entire alignment procedure was necessary to create a super-Gaussian mode

profile for the pump. Figure 3.4 shows the effect of alignment on the cw-pump beam.

Individual passes of the cw-pump beams can easily be seen in Fig. 3.4(a) as bright

ellipsoidal spots. The ellipsoidal shape is due to the projection of the Gaussian beam profile

on a flat surface at an angle. The outer ring is the edge of the thin disk heat sink. The faint

ring shows the thin-disk, which is 1 cm in diameter. The steps above were used iteratively

to optimize the pump spot, which is shown in Fig 3.4(c). When all ten passes of the pump

light were properly aligned, the individual ellipsoidal spots could be combined into a single

super-Gaussian mode profile.

Page 52: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

34

Figure 3.4: (a) Severe misalignment of the pump spot. All ten reflections of the pump spot

can be seen with poor overlap. (b) Improved alignment as the ten spots become closer to

each other. (c) A properly aligned pump spot with a super-Gaussian mode profile.

Fine adjustments to the collimation optics and the parabolic mirror were made at

high powers for optimal alignment. When the laser was operational, the thin disk was

imaged using a short-pass filter to image the green pump fluorescence. This was done to

avoid imaging of the 1030-nm laser output as well as the 940-nm pump light. Lineouts in

the x and y axis were taken from these images and used to optimize the mode profile.

3.3.2 Active Multipass Cavity

The weak small-signal gain (SSG) of a thin-disk laser appears to call for a low-loss

cavity. In many thin-disk oscillators, the intracavity energies exceed the external energies

by a factor of ten or more. This can have a significant effect on the intracavity SPM and

pulse shaping in mode-locked operation.

Normally, to increase the SSG, the doping level or the thickness of the disk must

be increased. However, without losing the advantage of the thin disk in terms of its good

thermal management (and therefore power scaling potential), a cavity configuration using

multiple passes of the laser mode through the gain medium was used. Multiple passes

through the gain media have been used in solid-state laser amplifiers with low gain rods

[110], as well as thin-disk amplifiers [111]. The use of an active multipass cavity (AMC)

compensates the low SSG by achieving many amplifications in a single round-trip and,

consequently, greater output coupling (OC) and lower nonlinearities.

Page 53: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

35

In addition to the large OC rates available through multiple passes, the energy of

the pulse also increases. Multiple passes increase the cavity length while maintaing a

relatively constant average power, leading to an increase in the pulse energy. This is

important to consider for dispersion management and soliton formation, which is discussed

further in Sec. 3.4.1.

The number of passes in the AMC was determined by maximizing the number of

passes without sacrificing cavity stability. A maximum of 12 passes was achieved using

the cavity configuration shown in Fig. 3.5. Power instabilities and mode-locking

difficulties were minimized however when using only six passes through the AMC,

resulting in a repetition rate of 7.08 MHz.

3.3.3 Cavity Design

A schematic of the TDML based on previous thin-disk designs [98], is shown in

Fig. 3.5. All mirrors were provided by Layertec, GmbH, except for the semiconductor

saturable absorber mirror (SESAM), which was provided by BATOP, GmbH. All mirrors

denoted with M were HR with reflectivity > 99.9% from 1020 to 1040 nm with near-zero

dispersion. The GTI mirrors used for this cavity were provided by Layertec, with a

reflectivity >99.5% from 1020 to 1040 nm. GTI2 and GTI7 had a focal length of 1.5 m.

This dispersion of these mirrors is discussed further in section 3.4.1.

All mirrors were 1-in. optics with the exception of M2–M9, which were 2-in

mirrors that allowed for an increased number of passes in the AMC. The pick-off mirror

(M10) was a D-cut mirror with a shaved edge for increased clearance along the beam path

between M8 and GTI7. This also prevented beam clipping when the space between cavity

modes of successive passes in the AMC were very close together.

The focal power of M11, GTI2, and GTI7 was chosen to create a cavity mode size

of 0.95 mm on the thin disk for optimal mode matching with the cw-pump spot size. The

GTI2 and GTI7 mirrors were set at distances 2f from each other to achieve a telescopic

Page 54: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

36

relay. This allowed for an arbitrary number of passes in the AMC without disturbing the

vital cavity parameters, mainly the spot sizes on the thin disk and the SESAM.

Figure 3.5: Diagram of the TDML cavity. The number of passes depicted has been reduced

for simplicity. SESAM: semiconductor saturable absorbing mirror; M: HR mirror; GTI:

Gires-Tournois Interferometer mirror [112].

To achieve a telescopic relay in the AMC the dioptric power of the thin-disk had to

be compensated for by adjusting the distance between M7 and M8. An ABCD matrix

calculation was performed to establish a constraint on the distance dtele between the GTI7,

M8, M9, and GIT2 in the AMC [113]. The relationship between this distance and the focal

power of the thin disk is given by

𝑑tele = 2𝑓tele −𝑓tele

2

𝑓TD, (3.1)

where ftd is the focal length of the thin disk and ftele is the focal length of GTI2/GTI7. The

design of the AMC provides a simple way to increase the number of passes through the

AMC by tilting the pick-off mirror M10 (as shown in Fig. 3.5) toward the outer edge of

GTI7, decreasing the separation between adjacent passes in the cavity. Once a suitable

distance between modes was established, M9 could be moved perpendicular to the M1

beamline to easily change the number of passes.

A thin-film polarizer was used in conjunction with a double pass through a quarter-

wave plate (QWP) as a variable output coupler. The light was s polarized inside the AMC.

Page 55: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

37

The polarizer was Brewster cut to minimize the losses of the incoming light. After a double

pass through a QWP, a portion of the light was transferred to p polarization, which passed

through the polarizer. The output coupler had a transmission of 73%, which was optimized

for maximum power when the optical-to-optical efficiency peaked at 34%. Alignment

sensitivity of this optic was quite critical due to the lack of sufficient humidity control in

the room, which changed the Brewster angle of the s polarized beams and, consequently,

the efficient output coupling of light.

Unless gross realignment was necessary, alignment was always changed by

adjusting the tip and tilt of the M1 and the SESAM mount. Alignment of the cavity had to

be fine tuned at full power for reliable long-term operation.

3.4 Mode-Locking

To generate laser pulses, a number of lasing modes must be coupled together. When

no attempt is made to control the laser spectrum, a laser always runs on a multitude of

different modes with a spacing of 𝛿𝜈 = 𝑐/2𝐿, given by the geometry of the laser resonator.

Only on a very short timescale, given by the inverse of the optical bandwidth ∆𝜈, some or

all the modes can radiate in phase, producing a sharp feature in the laser output described

as a coherence spike. If all the modes of the laser can be made to oscillate in phase, e.g., if

they can be locked together, the laser output becomes temporally well defined, with a

period corresponding to the cavity round-trip time, giving rise to ultrafast laser pulses. A

good general review of mode-locking is given by [114].

3.4.1 Soliton Formation

For the TDML, pulsed operation was possible through temporal soliton formation.

A thorough review of soliton formation and control is given by [51], although a short

summary of the fundamental principles will be given here with their relation to pulse

duration control in the TDML.

Page 56: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

38

Soliton formation is possible by balancing the self-phase modulation (SPM, 𝛾SPM)

of the pulse, the group-delay dispersion (GDD, 𝛽2), the pulse duration (𝜏FWHM), and the

peak power (𝜈peak2 ) of the pulse. This relationship is given by [115-117]

𝜏FWHM = √|𝛽2|

𝛾SPM|𝜈peak|

−1≈ √

1.76∙2∙0.88|𝛽2|

𝐵≈ 1.76 ∙

2∙|𝛽2|

|𝛾SPM|𝐸𝑃. (3.2)

Here, B is the B-integral and the approximation is only valid for a sech2 pulse shape. The

SPM coefficient and total amount of GDD used when applying this formula on the solitons

inside a laser resonator are specified for one complete round-trip, and both quantities must

have the opposite sign.

For an AMC-based laser with large OC, a modified version of the soliton area

theorem must be used to describe the relationship between pulse duration, external pulse

energy, total GDD, and the SPM coefficient. Assuming an exponential increase in the pulse

energy inside the cavity, the soliton area theorem can be approximated by

𝜏FWHM =≈ 1.76 ∙2∙|𝛽2|𝑙𝑛 (1−𝑂𝐶)

|𝛾𝑆𝑃𝑀|𝐸𝑃,ext. (3.3)

Due to the complexity of the mode size in the cavity, the B-integral can be quite difficult

to calculate. However, it must be precisely known to control the intracavity GDD for

control of the pulse duration. To calculate the B-integral, we first look at the change in the

refractive index when the intensity of the beam is very large. This can be expressed

mathematically as

𝑛(𝐼) = 𝑛0 + 𝑛2𝐼, (3.4)

where 𝑛0 is the linear refractive index, 𝑛2 is the nonlinear refractive index, and 𝐼 is the

pulse intensity along the beam axis, which for a Gaussian beam is given by 𝐼 = 2𝑃/𝐴. The

phase change due to the nonlinear refractive index is given by [118]

𝑑𝜑(𝑧) = −2𝜋

𝜆[𝑛0 + 𝑛2𝐼(𝑧)]𝑑𝑧. (3.5)

Page 57: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

39

The round-trip B-integral can be calculated by [101]

𝐵 = ∫ 𝑑𝜑(𝑧) 𝑑𝑧 =2𝜋

𝜆∫ 𝑛2𝐼peak(𝑧)𝑑𝑧 = ∑ 𝛾𝑆𝑃𝑀(𝑧𝑖)𝑃peak(𝑧𝑖)𝑖

2𝐿

0, (3.6)

where L is the length of the cavity. For resonators with low output coupling, 𝐼(𝑧) can be

calculated for one complete round-trip. In the case of a laser using an AMC, the peak power

strongly depends on the position of the pulse with respect to the output coupler and any

other gain or loss element. The calculation of 𝛾SPM can be simplified for telescopic image

relays of Gaussian beams. For a Gaussian beam with radius 𝑤, 𝛾SPM is given by

𝛾𝑆𝑃𝑀 = ∫4𝑛2

𝜆𝑤2(𝑧)𝑑𝑧

∆𝐿

0. (3.7)

For a Gaussian beam over a propagation distance through a telescopic arrangement, we see

that [101]

∫1

𝑤2(𝑧)𝑑𝑧

′4𝑓′

0=

𝜋2

𝜆. (3.8)

Equation (3.7) becomes

𝛾SPM =4𝑛2

𝜆∙ 2 ∙ 𝑁img ∙

𝜋2

𝜆𝑚−1 ≈ 2 ∙ 𝑁img1.4885 ∙ 10−3𝑀𝑊−1. (3.9)

𝑁𝑖𝑚𝑔 is given by the number of passes through the AMC, corresponding to one pass

through a 4f imaging system, plus the number of telescopic arrangements in the remaining

cavity arm and some fraction giving the SPM coefficient for the rest of the cavity. The

factor of 2 accounts for the fact that the SPM coefficient is calculated for one complete

round-trip. Therefore, the SPM coefficient scales linearly with the number of optical

images inside the laser cavity. For our cavity configuration, 𝑁𝑖𝑚𝑔 = 6. Using Eq. (3.9), we

can accurately calculate the B-integral, allowing for adequate control of the pulse duration

with appropriate control of the GDD.

Page 58: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

40

The GDD per mirror was specified at –1300 fs2 at 1030 nm, although the dispersion

varied by ±150 fs2 over the clear aperture of the mirror. Measurements of the spatially

dependent GDD, provided by Layertec, were used to calculate the total GDD of the cavity,

which for the optimal output power of the TDML was –0.125 ps2. Fine control of the GDD

was not possible and the total GDD could be modified only by swapping GTI mirrors for

HR mirrors or by changing the number of passes in the AMC. Since the number of passes,

was optimized to provide the greatest stability with the largest number of passes only

swapping HR mirrors with GTI mirrors could be used to control the GDD, resulting in

limited control of the pulse duration.

3.4.2 Initiation of Mode-Locking

For passive mode-locking of the TDML, an intracavity SESAM was used [119-

121]. Upon the incidence of laser radiation, due to the strong absorption bleaching, caused

mostly by band-filling through Pauli’s exclusion principle, semiconductors show a

nonlinear transmission characteristic, which makes them suitable to be used as saturable

absorbers. Epitaxial growth techniques, such as molecular beam epitaxy and metal-organic

chemical vapor deposition, can be used with different growth parameters to tailor linear

and nonlinear optical properties of semiconductor layer structures. The advanced growth

and processing technology offers the possibility of precise control of the optical

nonlinearity of these devices.

The key design parameters of a SESAM are its saturation behavior, the dynamic

response and the wavelength of operation. The saturation properties of a SESAM can be

characterized after determining the reflectivity R for various incident pulse energy fluences

F. Nonlinear optical absorption and refraction in general are investigated by using z-scan

experiments [122-126], while time-resolved experiments are performed with “pump–

probe” setups [127]. SESAM specific measurement techniques are also reviewed in Haiml

et al. [128].

Page 59: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

41

The key component for efficient initiation of mode-locking is making sure that the

losses for pulsed operation are less than the losses of cw radiation. This is achieved by

having a high modulation depth ΔR. The reflectivity losses for cw operation can be seen

when the reflectivity of the SESAM is considered as a function of pulse fluence. The

reflectivity of a SESAM is given by

𝑅 = 1 − 𝐴𝑛𝑠 −𝛥𝑅∙𝐹sat

𝐹∙𝑙𝑛(1+𝐹

𝐹𝑠𝑎𝑡)

−𝛽∙𝐹∙𝑑

𝜏𝑃, (3.10)

where Ans is the nonsaturable absorption loss, Fsat is the saturation fluence, F is the pulse

fluence, τP is the pulse duration, 𝑑 is the thickness of the absorption layer, and β is the two-

photon absorption coefficient. The modulation depth, which is calculated by taking the

difference between reflectivity of the pulse and cw radiation, must be carefully considered

for the initiation of mode-locking and control of mode-locking instabilities.

When the TDML was constructed, the options for choice parameters of

commercially available SESAM’s were quite limited. In general, modulation depth and

saturation fluence of the SESAM were tied in such a way that mode-locking instabilities

(discussed in the next section) were difficult to eliminate. Furthermore, choices in

relaxation times, which are a critical parameter for the behavior of the SESAM as a slow-

saturable absorber, were limited.

For this experiment, a BATOP SAM-1040-4-5ps-4.0-12.7s-c SESAM was chosen.

The absorbance was 4% with a modulation depth of 2.6% and a relaxation time of 5 ps.

The chip was 4.0 mm square on a 12.7-mm-diameter copper heat sink. This heat sink was

attached to a custom water-cooled kinematic mount to prevent damage and to limit thermal

lensing effects from use at high average powers. The saturation fluence of 40 µJ/cm2 was

chosen to create stable pulsing at maximum output power.

3.4.3 Mode-Locking Instabilities

Passive continuous-wave mode-locking (CWML) is ideally characterized by one

single pulse oscillating inside the laser cavity with the same pulse energy, amplitude and

Page 60: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

42

pulse duration for each round-trip. Although stable operation could be achieved for >8 h

over the course of a few years, it was necessary to overcome instabilities that interfere with

stable pulsing by carefully designing the cavity. These instabilities included Q-switching,

cw background and multiple pulsing, which were affected by the gain material, cavity

design, and choice of SESAM. These instabilities had to be understood and controlled for

the TDML to function adequately as a pump source for the OPO.

When pulse energies are below a certain marginal value, determined by both the

gain medium and the saturable absorber, Q-switched mode-locking (QML) instabilities

will occur as a result of undamped relaxation oscillations [129]. Historically, the saturation

energies of most solid-state lasers were large enough to ensure that saturable absorbers

would always produce QML behavior.

In 1992 the invention of semiconductor saturable absorbers led to saturation

fluences that were able to suppress the QML output of solid-state lasers [130, 131]. The

pulse energy required to overcome QML for a saturable absorber with a relaxation time

much shorter than the laser’s repetition rate and small enough pulse fluences to prevent

effects like two-photon absorption is given by [131]

𝐸𝑃2 > 𝐸𝑠𝑎𝑡,𝑔𝐸𝑠𝑎𝑡,𝑙∆𝑅. (3.11)

Esat,g and Esat,l are the saturation energies for the gain and saturable absorber, respectively.

EP is the intracavity pulse energy and ΔR is the modulation depth of the SESAM. When

QML instabilities occur, the amplitudes of the mode-locked pulses are controlled by an

oscillating envelope function with a frequency lower than the repetition rate of the laser.

The maximum pulse energy can be significantly increased, which can lead to damaged

optical components.

The choice of the SESAM, mode size and intracavity power are the only points of

control to eliminate the QML instability. Due to limitations from the collimator and

wanting to avoid paying for additional collimation optics, 𝐸𝑠𝑎𝑡,𝑔 = 79.2 µW. The spot size

on the SESAM could be decreased to ensure that QML could be overcome at lower

Page 61: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

43

energies. The focal length of M11 in Fig. 3.5 was chosen to be 750 mm to create a 600 µm

spot size on the SESAM, resulting in 𝐸sat,𝑙 = 2.74 mW with a modulation depth of 2.4%.

The resulting pulse energy to overcome QML required an average power of 30.7 W with a

cavity repetition rate of 7.08 MHz, which was well within the capabilities of TDML.

Various contributing factors can give rise to a gain advantage of a cw background.

At the center of the spectral gain distribution, the continuum might have a gain advantage

when compared to the spectrum of the pulse due to the limited gain bandwidth. Caused by

an additional loss provided by the saturable absorber for the cw background and any

additional pulses, these instabilities are suppressed in the case of stable CWML. For soliton

mode-locking, the relaxation time of the saturable absorber might be larger than the pulse

itself. Therefore, the gain window is relatively large, giving rise for a possible amplification

of the continuum.

The cw background did not affect the total average power of the TDML. As shown

in Fig. 3.6, a sharp peak located at 1030.8 nm was clearly visible in the spectrum in the

presence of cw-background. The total power of the cw background could be calculated by

fitting the spectrum to a sech2 pulse shape, subtracting that from the measured spectrum

and comparing the area between the two to give a percentage of their contribution to the

total average power.

Regardless of the cw background’s contribution to the total power, the average

output power of the TDML was unaffected. This was confirmed by simultaneous

measurements of the spectrum and output power, which showed that if the cavity was

CWML, the total power stayed stable within 200 mW.

Page 62: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

44

Figure 3.6: Example of cw background by changing the alignment of the cavity. Output

power remained constant but the spectrum of the TDML changes significantly. As power

shifts into the pulse, the spectrum become wider (allowing for shorter pulses) and

experiences a slight red shift. The green curve represents the best cw background reduction

through alignment only, even though it was completely removed with proper wavelength

calibration of the pump laser, as shown in Fig. 3.10(b).

When strong cw background was present, the pulse duration increased, as would

be expected when pulse energy is being shifted into cw radiation. Since the total average

power of the TDML remained unchanged despite the amount of cw background, the pulse

energy was reduced as the amount of cw light grew. Due to soliton mode-locking, the result

of this energy transfer was an increase in the pulse duration to balance the soliton theorem

as shown in Eq. (3.2). This was tested and verified with autocorrelation measurements at

varying amounts of cw background.

While the impact of each of the factors that contribute to the cw background is not

clearly understood in the TDML, both the alignment of the cavity and the wavelength of

the cw pump beam had a significant impact on the presence of the cw background. The cw

light could be completely suppressed by carefully aligning the cavity and maintaining

proper calibration of the pump wavelength.

Another instability that had to be considered was multiple pulsing. Multiple pulsing

can occur when, due to gain filtering, double pulses (or higher combinations) with a

respectively smaller spectrum experience a gain advantage large enough to overcome

Page 63: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

45

additional losses required for saturating the SESAM [132]. There is a great deal of

information about multiple pulsing in thin-disk oscillators and SESAM-based devices since

it is critical to the initial design of the system [117, 133]. In the TDML, double pulsing was

observed but only when the laser was pushed well beyond maximum efficiency. This

source of instability was not nearly as much of a concern for this study as the two

mentioned above.

3.5 The TDML as a Pump Source

3.5.1 Output Power and Stability Regimes

To maintain safe operation of the TDML, efficient power extraction of the absorbed

pump light on the thin disk was necessary. The efficiency falloff at high powers is well

known in thin-disk oscillators, and maintaining an optical to optical efficiency greater than

30% was required to prevent damage [98]. The efficiency of the TDML is plotted in Fig.

3.7.

Figure 3.7: Efficiency of the TDML as a function of pump power. Pump power was

increased until falloff was noticeable and efficiency was near minimum safety levels. The

different regimes of operation are color coded. CW: continuous-wave operation; QML: Q-

switched mode-locked; CWML: continuous-wave mode-locked (stable); DP: double

pulsing.

Page 64: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

46

The SESAM could not initiate mode-locking until the pulse fluence rose high

enough to create a substantial difference between cw losses and pulsed losses. Continuous-

wave operation is present until the pump power reaches roughly 55-W. As the power

increases, the TDML begins to operate in QML. Because of the risk of damage, the pump

power is rapidly increased during this time to minimize operation in QML. CWML is

achieved for a relatively small region. For stability reasons, the TDML was operated near

the top of the efficiency curve. This removed long-term behavior from sporadically

switching to either QML or double pulsing.

The TDML operated at maximum stable power to ensure a constant pulse duration

as a pump source. The maximum average power used to pump the OPO was 44.8 W, which

was measured after the losses introduced by relay optics and a variable throttle outside the

cavity. A throttle was introduced outside the cavity to change the available power to the

OPO without changing the cavity dynamics of the TDML. The throttle was a half-wave

plate used in combination with a thin-film polarizer that allowed for partial convergence

of the p-polarized output light to be converted into s-polarized light and reflected off the

polarizer. The s-polarized light was sent to a high-power beam dump, while the other beam

was sent to either the diagnostics or the OPO cavity. With careful alignment, the output

power of the TDML could be varied between 99.5% and 0.5% of the available power with

negligible impact on the beam quality and pulse duration. The throttle made it possible to

use a pulse with variable power without impacting the pulse duration, which was of critical

importance for the TDML to operate as an adequate pump source for the OPO.

3.5.2 Pulse Duration

An autocorrelation was performed to measure the pulse duration with a commercial

FR-103MN/LiIO3 autocorrelator from Femtochrome. A typical autocorrelation trace,

averaged over ten acquisitions, is shown in Fig. 3.8(a). The FWHM of the autocorrelation

was 1.5 ps, corresponding to a pulse with of 1.0 ps, assuming a sech2 pulse shape.

Page 65: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

47

Figure 3.8: (a) Autocorrelation trace of the TDML; (b) corresponding spectrum of the

TDML pulse [112].

Pulse widths as high as 2.1 ps were recorded, which depended both on operation at

lower powers and with more GTI mirrors than were shown in Fig. 3.5. Since GTI mirrors

were the dominant form of negative dispersion in the cavity, decreasing the number of GTI

mirrors should also decrease the pulse duration, which is evident from Eq. (3.2). Pulse

duration tests were performed with fewer GTI mirrors; however, mode-locking instabilities

dominated over the full range of available power. Although shorter pulse durations were

desirable as a pump source, 1.0 ps was adequate for this study.

The spectrum corresponding to the autocorrelation measurement is shown in Fig.

3.8(b) with a FWHM of 1.1-nm. The spectrum revealed that the time-bandwidth product

[51] of the pulse 0.316, showing near transform limited operation for a sech2 pulse, which

is to be expected from soliton mode-locking.

The spectrum was typically measured with an Ocean Optics 4000 spectrometer.

The spectrometer used an H6 composite grating with optimal performance for 500 to 1100

nm. The slit width was 5 µm to maximize resolution with the ample power available from

the TDML. For ease of use, a multimode fiber was used to couple light into the

spectrometer; however, this produced heavy fringes during normal use. These fringes could

be eliminated by quickly shaking the fiber during spectral measurements that were

averaged 20x with 50-ms acquisition windows.

Page 66: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

48

The spectrometer was the primary tool to determine if stable mode-locking had

been achieved, as well as optimization of alignment through the removal of cw background.

A higher-resolution optical spectrum analyzer was also used to verify resolution of the

spectral width. Due to the low scanning speeds and the bulk of the instrument, it was

deemed unnecessary for daily use.

Both cw and QML operations showed narrow peaks in the spectrum while cwML

had a broad sech2 shape. This metric was most helpful when determining if the TDML had

stable mode-locking or if adjustments to the alignment were necessary. There was a

negligible difference in the output power in each of these regimes and the spectrum

provided a quick method by which to determine not only if the TDML was CWML but if

cw background had been suppressed.

3.5.3 Beam Quality

High beam quality was necessary for the TDML to perform as an adequate pump

source for the OPO. Multiple beam-quality measurements were performed to ensure that

the beam could be focused inside a crystal and could accurately represent the beam

parameters used to calculate the gain of the OPO. Images of the output beam were used to

qualitatively characterize the beam shape and 𝑀2 measurements were taken to verify the

beam’s Gaussian behavior.

The near-field beam profile had few irregularities and was quite Gaussian, as shown

in Fig. 3.9. The beam profile looks fairly uniform and Gaussian, which is to be expected

from the output of an oscillator. However, the beam is slightly tilted off-axis and elongated

slightly in the x direction. Slight astigmatism of the pump beam was unavoidable due to

off-axis angles of incidence on spherical mirrors inside the AMC. In order to obtain a

more-quantitative analysis of beam quality, an 𝑀2 measurement was performed on the

TDML output [134].

Page 67: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

49

Figure 3.9 Image of the near-field beam profile of the TDML. The image was taken with

an Imaging Source DMK 41BU02 camera with a 1030-nm bandpass filter.

Despite the astigmatism, the measured beam quality was fairly high with a 𝑀2 <

1.4, which is consistent with previous cavities of similar design [98]. The locations of the

focal point on the x and y axes were separated by 30 mm. 𝑀2 measurements were

performed using a knife-edge technique that allowed for beam characterization at high

powers without throttling the pump beam below 5% [134, 135]. The beam size was

measured using an 84/16 method to directly measure the 1/e2 spot size [136]. A 400-mm-

focal-length lens was used to focus the beam, ensuring that the focal spot was large enough

to be measured and the beam could be adequately sampled though focus.

3.5.4 Temporal Stability

The TDML was not locked to an external source, as is typically done for many

mode-locked oscillators to maximize temporal stability [137]. Without external locking,

variations in pulse amplitude and long-term drift were a significant concern. These

characteristics were investigated to determine their impact on the TDML as a pump source.

The variation in pulse amplitude was measured using oscilloscope traces from a

DET10A Thorlabs photodetector. The pulse amplitude measured from the photodetector

has a linear relationship to the peak power of the pulse. Traces were taken with maximum

resolution on a 300-MHz oscilloscope for accuracy in the peak voltages. These traces were

Page 68: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

50

analyzed in MATLAB by gating the temporal window of an individual pulse and

characterizing the entire acquisition, which had over one million pulses. The maximum

value of the pulse amplitude was stored; then a mean and standard deviation was calculated

from these values. The standard deviation was calculated to be less than 0.5% of the peak

amplitude and was consistent for multiple traces taken.

Long-term drift was evident when the pulse train was monitored with a radio-

frequency (rf) spectrum analyzer. Drift was characterized by measuring the repetition rate

over multiple hours of use, showing a typical change of 120 Hz in 4 h of operation,

corresponding to a change in cavity length of ~350 µm. This neglects the rapid change that

occurred in the first 30 min of operation during the warm-up period of the TDML. The

change in cavity length was likely due to heating inside the TDML enclosure, which saw

a typical rise of 2.5°C. The temperature’s effect on optomechanics inside the enclosure

would always cause the cavity to decrease in length and then reset to almost the exact same

length after given sufficient time to cool.

Page 69: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

51

4 Design of the Optical Parametric

Oscillator

The construction of a cavity that could be synchronously pumped by the TDML

just presented required careful design. In this chapter, the most important aspects of the

OPO design are described. Section 4.1 is dedicated to the nonlinear crystal. The reasoning

behind crystal choice is given. Material properties of the chosen crystal are discussed with

particular focus on the linear absorption. The section concludes with the phase-matching

geometry and choice of optical coatings for the crystal. Section 4.2 describes cavities that

were explored during this investigation. Finally, in Sec. 4.2.3, an ABCD model was used

to hypothesize how the cavity design principles could be used to achieve longer cavity

lengths and obtain higher pulse energies.

4.1 Nonlinear crystal

4.1.1 Availability of Crystals

The choice of the nonlinear crystal was a critical design consideration of the OPO.

The key factors for determining a proper crystal were the transparency range, phase-

matching capabilities, effective nonlinearity, and commercial availability.

To get the maximum output power from the OPO, it was desirable to use direct

1030-nm output of the TDML without any frequency conversion. The nonlinear crystal

must be transparent not only for the pump but also for the signal and idler wavelengths.

With a pump wavelength near 1 µm, the signal and idler wavelengths could span between

1250 and 6000 nm [138-140]. Appropriate choice of the signal wavelength was critical

when considering possible absorption of most conventional crystals in this wavelength

range [141]. With high average powers, even if the idler is not being resonated (therefore

Page 70: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

52

having significantly less power than the signal), high absorption at longer wavelengths can

create unwanted heating inside the crystal [142].

If the OPO is operated near degeneracy, the signal and idler wavelengths will be

nearly the same, which limits the transparency range requirements of the crystal.

Additionally, this allows the crystal length to be much longer than normal while

maintaining enough bandwidth to create ultrashort pulses [143]. Operating at exact

degeneracy is beneficial because both the signal and idler will be in phase, creating

additional output power from the combination of both beams. However, it removes the

inherent benefit of tunability from the OPO, restricting the output to a single central

wavelength as well as increased requirements on oscillator stability [5, 144].

Phase matching was a critical design consideration of the OPO. Multiple methods

could be employed to achieve the phase-matching condition required to make the

parametric process efficient, namely critically phase matching, quasi-phase matching, and

noncritical phase matching (sometimes called temperature phase matching). Choosing of

the type of method to employ for this study relied on the ability to power scale these

methods.

Noncritically phase-matched systems are typically not used in high-power ultrafast

systems since these devices are typically employed for their ease of use and tunability in a

single device [55, 145, 146]. Difficulties in power scaling these devices stem from the

limited phase-matching bandwidth and temperature dependence, which is difficult to

control at high average powers. Also, the damage threshold of ovens used to control the

temperature typically have a damage threshold lower than the crystals themselves, making

this option the least attractive of the three.

Recent advances to femtosecond quasi-phase-matched systems have shown

potential for their power scaling capabilities [3, 147]. However, to achieve broadband

phase-matching at high average powers, exotic crystal designs are required, greatly

increasing the cost and complexity of these systems. One notable crystal using this type of

Page 71: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

53

phase matching is periodically polled lithium niobate (PPNL) and doped variations, which

have shown great promise for high-power ultrafast OPO’s [32, 41, 148, 149]. There are

concerns, however, that quasi-phase-matched systems can be applied only to crystals with

limited thickness, which excludes large-aperture devices for high average powers.

Critically phase-matched systems have shown great potential for power scaling,

limited only by the size of the crystal growth [44, 150, 151]. Additionally, techniques used

in different experiments can easily be applied to other systems as long as phase matching

can be achieved in regions with relatively high effective nonlinearities. As such, the search

for crystals for this investigation was limited to ones using critical phase matching. Figure

4.1 summarizes the properties of various crystals that were considered.

Of the crystals considered, bismuth triborate (BiB3O6, BiBO) was the most

promising nonlinear crystal due to the combination of transparency, effective nonlinearity,

damage threshold and maturity of the crystal development. Although KTA and KTP have

been used extensively in parametric devices, the effective nonlinearity was not useful for

near degenerate operation with a 1-µm pump. CSP and AGS were also attractive due to

their high effective nonlinearity and transparency, however, the low damage threshold

made them problematic for power scaling. Commonly used crystals like BBO and LBO

were also considered; however, the transparency of these crystals limited the power scaling

potential due to high absorption.

Although the transparency range of these crystals (noted in Fig. 4.1) seemed useful

for operation near 2 µm, transmission values greater than 95% were desirable for scalability

into higher powers. Typically, when transparency is reported, all wavelengths providing

transparency values greater than 50% are reported. However, these levels of absorption

would result in significant thermal problems at the anticipated high average powers of the

OPO. Exotic crystals, such as BGS, were highly attractive but were not obtainable since

the growth and characterization of these crystals are in their infancy.

Page 72: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

54

Crystal Transparency

Range

(nm)

Usable*

Effective

Nonlinearity

(pm/V)

Damage

Threshold

(J/cm2)

Maturity

AgGaS2

(AGS) 470 to 13,000 12.5 2.3 √√√

BaB2O4

(BBO) 190 to 2250 1.9 13 √√√

BiB3O6

(BiBO) 286 to 2500 2.1 10 √√

BaGa4S7

(BGS) 350 to 10,500 5.4 25 √

CdSiP2

(CSP) 500 to 9000 85 3.0 √√

LiB3O5

(LBO) 160 to 2600 0.6 34 √√√

KTiOAsO4

(KTA) 350 to 4000 0.0 15 √√√

KTiOPO4

(KTP) 350 to 4500 0.0 5.0 √√√

ZnGeP2

(ZGP) 720 to12,000 75 3.0 √√

Figure 4.1: Table of some of the relevant parameters for various critically phase-matched

crystals investigated as a gain medium for the OPO [152-162]. The transparency range

corresponds to transparencies greater than 50% for typical crystal lengths. The maturity

level is ranked from one to three check marks, with three being the highest level of

maturity. *Usable refers to the effective nonlinearity in the region that can be phase

matched for a 1-µm pump and a near-2-µm signal.

Page 73: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

55

BiBO was the best option for a power-scalable ultrafast OPO. It has been used in

numerous OPO’s and has had most of its relevant material properties characterized [156,

163, 164]. The parameters of the crystal relevant to the operation of a scalable ultrafast

OPO are discussed in depth in the following subsections.

4.1.2 BiBO Transparency

The transparency reported from earlier investigations was made using a BiBO

crystal attractive for operation near 2 µm [156]. Figure 4.2 shows the unpolarized

transmission of a 1-mm BiBO crystal measured by Hellwig et al. with reflection losses

from an uncoated crystal. Anticipated absorption from 1950 to 2150 nm was less than 5%.

The sharp drop in transmission near 2250 nm was well outside the anticipated wavelengths

of the OPO.

Figure 4.2: Unpolarized transmission spectrum of BiBO [156].

Preliminary tests on the BiBO crystals used in this study showed unanticipated

signs of heating. To investigate the source of this heating, it was necessary to measure the

transmission of the BiBO crystals that were used. Measurements were attempted at LLE

with a Lambda 900 spectrophotometer from PerkinElmer. However, lack of sensitivity in

the detector along with alignment difficulties made it impossible to obtain a meaningful

Page 74: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

56

transmission spectrum. Transmission data were available for wavelengths up to 3 µm but

this required a switch in the photodetector, resulting in significant noise for wavelengths

exceeding 1300 nm. This resulted in an error of 6% over the wavelengths of interest,

making it difficult to measure fine details in transmission spectrum. The size of the crystal

also made it difficult to make precise measurements. The spectrophotometer was designed

for 12- to 48-mm-diam optics, while the crystal face was a 5 5-mm clear aperture.

Modifications were made to accommodate the crystal size but at an additional sacrifice to

the precision of the results.

The crystal manufacturer (Conex System Technology) was able to provide

transmission measurements from the crystal batch used in this investigation after a special

request was made and their equipment was upgraded. Using the absorption data provided,

the transmission through an 8.4-mm BiBO crystal was calculated, neglecting reflection

losses (shown in Fig. 4.3).

Figure 4.3: Transmission through an 8.4-mm BiBO crystal, neglecting reflection losses for

the pump wavelength (dashed blue line) and the OPO tunability range (dashed red lines)

[165].

At 1030 nm (pump wavelength), negligible absorption was observed, which was

expected from earlier measurements of the transmission. Absorption exceeded

expectations for wavelengths greater than 1600 nm and narrowband absorption peaks are

evident. The sharp drops in transmission suggest that there were likely contaminants

Page 75: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

57

present in the crystal growth process [166, 167]. These contaminants were likely caused

by impurities in the bismuth oxide or water deposition during the vapor deposition process,

which could be prevented at additional manufacturing cost along with care during the

crystal growth process. For this study, both time and cost prohibited investigating

improvements to the crystal. Evidence of thermal effects did arise from this absorption,

which is discussed in more detail in Chap. 6.

4.1.3 Noncollinear Geometry

A small noncollinear angle could be introduced between the pump and signal

beams. The noncollinear geometry increases the available bandwidth in the parametric gain

and prevents the idler beam from resonating, which reduces the complication for a doubly

resonant cavity [60, 168]. The calculated phase-matching curve is shown in Fig. 4.4.

Figure 4.4: Phase-matching calculation for an 8.4-mm BiBO crystal with an internal

noncollinear angle of 0.2° [165].

Despite the 8.4-mm crystal length (which is relatively long compared to typical

crystal lengths of less than 2 mm in OPO’s), the phase-matching bandwidth available in

this system is quite impressive. This is due to near-degenerate operation as well as the

exceptional phase-matching capabilities of BiBO. As shown in Fig. 4.4, the phase-

matching curve is flattest near degeneracy at 2060 nm. The optimal phase-matching angle

was chosen as 8.65°, which gives near-uniform phase matching from 1.95 to 2.17 µm.

Page 76: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

58

Increasing the noncollinear angle also increases the bandwidth to a limit, but there

are multiple reasons why a small noncollinear angle is sufficient. The bandwidth in this

setup is already large enough to produce sub-100-fs pulses. Increasing the noncollinear

angle would have placed additional constraints on the beam dumps needed to collect

reflected beams from the crystal face. Additionally, the idler beam produced would have

greater angular dispersion. The available space for collecting stray beams and appropriate

locations for beam dumps was already limited with a small collinear angle, so the increased

angular dispersion of the idler beam made safe engineering practices increasingly difficult.

Finally, because of the large crystal length, the spatial walk-off between the pump and the

signal was important. As the noncollinear angle was increased, the impact of the spatial

walk-off between the pump and the signal had a negative impact on the gain as the spatial

overlap between the beams decreased [169, 170].

4.1.4 Crystal Coatings

Antireflection (AR) coatings on the crystal were necessary to mitigate unwanted

watt-level stray beams inside the cavity and to prevent unnecessary losses in the cavity. If

left uncoated, the high index (n =1.78) of the material caused Fresnel reflections of roughly

7.5% for pump, signal, and idler beams [171]. Multiple AR coating strategies were

reviewed to minimize the impact of these reflections on the cavity performance.

A multilayer dielectric coating was optimal for minimizing the reflectivity at the

pump, signal, and idler wavelengths. Calculations showed that reflectivity of less than

0.1% could be achieved at 1030 nm and from 2000 to 2200 nm. Coating difficulties were

encountered, however, due to the high-index of the material and the wavelength range of

the OPO. The thickness of the coating had a considerable risk of cracking due to poor

adhesion to the substrate, regardless of the power incident on the crystal [172]. These

cracks give rise to high-order aberrations, affecting the mode quality of the OPO and

reducing the available gain from poor mode overlap between the pump and signal beams.

Furthermore, cracks would also reduce the damage threshold of the coating below the bulk-

damage threshold of the crystal. Due to the anticipated powers of the OPO and the high

Page 77: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

59

repetition rates, it was very likely that a multilayer dielectric–coated BiBO crystal would

quickly cause the OPO to cease lasing.

An acceptable alternative was found by using a single-layer coating. Using a single-

layer SiO2 AR coating could minimize the reflectivity for 2060 nm at a sacrifice of

increased reflectivity at the pump wavelength. By using a single-layer coating, the damage

threshold could be higher than the bulk damage of the crystal. Also, since single-layer

coating deposition does not change significantly for different materials or wavelengths, the

coating could be reliably engineered such that optical defects would depend less on the

coating than the surface quality of the crystal itself [173]. By optimizing the thickness, the

reflectivity was minimized for degenerate operation. This yielded a reflectivity of 8.7% at

1030 nm and less than 2% reflectivity from 1950 to 2200 nm. The decrease in available

pump power was negligible when compared to the uncoated crystal, while the reflectivity

at the signal and idler wavelengths was dramatically improved.

4.2 Cavity Design

4.2.1 Multipass Cavity Design

Multiple cavity designs were considered before the final decision was made. The

first cavity constructed had a cavity length 6× shorter than the pump cavity and consisted

of only four mirrors. This cavity was built to perform initial tests on the crystal quality and

diagnostic performance; however, irregularities in the amplitude of pulses and stability

problems made this design suboptimal compared to a synchronously pumped cavity.

The first synchronously pumped cavity used a multipass design similar to the

TDML, in order to scale the cavity length in unison with the pump. Although the pump

cavity was operated at a repetition rate of 7.08 MHz, the cavity length could be extended

to achieve pulse energies by more than a factor of 2 at the same average power. The

increase in cavity length was achieved by increasing the number of passes in the multipass

section of the cavity, as described in Sec. 3.3.2. The multipass OPO is shown in Fig. 4.5.

Page 78: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

60

Figure 4.5: Multipass OPO. Unlabeled mirrors are high-reflectivity mirrors (either 0° or

45° angles of incidence); DC: dichroic mirrors; OC: output coupler, 5% transmission. All

curved mirrors had a focal length of 1.5-m [174].

The compact cavity design accommodated the pump laser, OPO and diagnostics

for both lasers on the same 4-ft 10-ft optical table. The OPO was designed for collinear

pumping, although it would be straightforward to adjust the cavity alignment slightly to

accommodate a noncollinear pump/signal geometry. The cavity can be separated into two

main parts, the pump section and the multipass cavity, which are shown in the Fig. 4.5 on

the left and right sides, respectively.

In the pump section, dichroic mirrors were used to achieve collinear pumping. The

back of the mirrors were AR coated for 1030 nm to prevent unnecessary losses as well as

heating and ghost reflections from light reflected off the front surface of the mirrors. The

front surfaces of the mirrors were AR coated for 1030-nm light and HR coated for

wavelengths between 1950-2150 nm. Folding mirrors were used to ensure the BiBO crystal

was placed at the focus of the pump section. Folding mirrors were also used to minimize

the angles of incidence (AOI’s) on the curved mirrors to reduce astigmatism. The focal

length and position of these mirrors were chosen to recreate the non-AMC length segment

of the pump laser.

The multipass section worked similarly to the AMC in the pump laser. The largest

difference is the length segment of the dtele section. In the pump laser, this segment served

two purposes: the first was to accommodate the focal power of the thin disk at high-power

operation; the second was for ease of alignment of the angularly multiplexed cavity

extension. The latter worked the same in the OPO cavity; however, the focal spot no longer

Page 79: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

61

occurred on an optic that changed with internal power. A flat mirror was used in the OPO

cavity, which increased the length of dtele. In the final arm of the OPO, space was left for a

prism set to manage the internal dispersion of the cavity. The length of this segment could

be adjusted without changing the critical cavity parameters, making it ideal for fine tuning

the cavity length to achieve synchronous operation. The prism set was not implemented

and the inclusion of dispersion control will be discussed in Chap. 7.

Despite a great deal of effort, this cavity was never able to achieve lasing, which

was primarily due to alignment difficulties in the laser. To achieve lasing, the cavity length

of the OPO must match that of the pump laser within 40 µm as well plus the pump and

cavity mode in the crystal must precisely overlap. Although practice in aligning the

multipass section of the cavity had been thoroughly explored with the pump laser, finding

an appropriate alignment source made alignment through the multipass section exceedingly

difficult. Due to the complicated cavity layout, four separate types of mirror coatings had

to be employed: an output coupler, a 0° broadband HR, a 45° s-polarized broadband HR,

and the dichroic mirrors. Finding an adequate alignment laser that provided enough usable

light for a single round-trip in the cavity proved more difficult than expected. The

reflectivity and GDD of the mirrors used in this investigation are shown in Fig. . The

reflectivity was measured by the Optical Manufacturing Group at LLE using a Lambda

900 spectrophotometer from PerkinElmer, and the GDD was measured using an

experimental Chromatis upgrade from KMLabs.

Although each mirror had its reflectivity optimized at 2060 nm, the performance at

other wavelengths varied significantly. The reflectivity of each mirror is shown in Fig. (a).

It is immediately obvious that the reflectivity of the dichroic mirrors falls off at higher

wavelengths significantly more than the other mirrors. Additionally, there is no obvious

wavelength that would be good to use for an alignment source due to high-reflectivity

bands for different mirrors having poor overlap for wavelengths between 400-1100 nm. In

Fig. (c) the total reflectivity for each cavities designed for this investigation is shown. The

multipass cavity, shown in blue, has no useful source for alignment, with the best

Page 80: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

62

wavelength being 690 nm at a reflectivity of less than 1%. As mentioned above, this made

it extremely difficult to align this complicated cavity design. When one compares the total

reflectivity between the two cavities, the benefits of the alternative design are obvious.

Figure 4.6: (a) Measured reflectivity for all mirrors used in both OPO cavities excluding

the output coupler. (b) The measured GDD of each type of mirror. (c) Total reflectivity

from the mirrors in each OPO cavity. Losses from the output coupler and crystal face are

not considered in this calculation. (d) Total GDD in each OPO cavity, not accounting for

dispersion obtained from the crystal or SPM. In both (a) and (c) the tunability of the OPO

has been highlighted with dashed black lines.

The total reflectivity of the cavity was calculated by taking the reflectivity of each

individual mirror into account for a single round-trip. The impact of the large number of

mirror bounces becomes obvious when comparing the contribution of a single mirror to

the losses seen in one oscillation. In the multipass cavity, there are 76 bounces off various

mirror coatings, whereas in the final design (folded cavity), there are only 24 bounces from

Page 81: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

63

a single coating. In these calculations neither the output coupler, the crystal absorption, nor

the losses from the crystal faces is taken into consideration.

Due to the insensitivity of the spectrophotometer at wavelengths exceeding 1300

nm the reflectivity measurements become quite noisy. This noise is amplified in the total

reflectivity calculation, which is not very precise for wavelength regions near the tunability

range of the OPO. For instance, the folded cavity has less than a total 5% loss from the

mirrors in the cavity during one round-trip, which is nearly uniform from 1950 to 2200 nm.

This value was calculated from manufacture-provide high-precision measurements that

were difficult to use for broadband analysis.

There are multiple benefits to a cavity using only 0° HR mirrors, which is discussed

thoroughly in the next section. Multiple wavelengths are available to align the cavity since

losses at discrete bands are minimal for an individual coating type. Additionally, the

reflectivity around the wavelengths of interest for the OPO are considerably higher. Near

2060 nm the total losses are ~5% less than with the multipass cavity. For wavelengths

greater than 2070 nm the reflectivity is significantly higher, suggesting that the multipass

cavity would have significantly less tunability and potential bandwidth than a cavity using

only 0° HR mirrors.

The presence of significant GDD also varies between the two cavity designs. In

Fig. (b) the GDD from each mirror can be seen. Although the GDD from both the 45° HR

and the 0° HR was fairly uniform, the GDD from the dichroic mirrors was significantly

higher than expected with a fairly large defect at 2140 nm. The mirrors were initially

designed to have near-zero GDD from 2010 to 2110 nm, which was almost achieved with

the exception of the dichroic mirrors. During the design phase of the mirrors, the GDD was

an extremely difficult constraint to control due to the tight requirements on broadband high

reflectivity and damage threshold.

A new cavity was designed to alleviate the alignment problems caused by the

previous cavity using only 0° HR mirrors except for the output coupler. This allowed for

Page 82: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

64

the use of an inexpensive 695-nm cw diode laser as an alignment source that could maintain

high brightness despite the considerable number of mirror bounces. The folded cavity

design also improved the uniformity of the GDD and total reflectivity over operational

wavelengths of the OPO.

4.2.2 Folded Cavity Design

The schematic of the cavity, drawn to scale, is shown in Fig. 4.7. The focusing

mirrors inside the cavity—12.8-mm-diam optics; 25-mm optic adapters—were used in

tandem with clear edge mounts to increase the beam clearance inside the cavity and to

accommodate the compact design shown here. In some cases, the adapters were cut to

achieve very close clearance between the beam and the edge of the mirror. The cavity

length is 21.2-m, but the cavity is folded into a 1.0 0.2-m enclosure. The AOI on the

curved mirrors was limited to less than 2° to reduce astigmatism.

Figure 4.7: Schematic of the 21.2-m synchronously pumped OPO. HR: 1950 to 2250-nm

high reflector; OC: 20% output coupler, 3.0-m radius of curvature (ROC); BiBO: nonlinear

crystal; MP: 1.5-m ROC pump mirror; M1.5: 1.5-m ROC; M2: 2.0-m ROC; M3: 3.0-m ROC.

The pump (blue) and idler (dashed red) are not resonated in the cavity with the signal beam

(solid red) [165].

Two focusing optics down-collimated the pump beam with a slightly diverging

wavefront before arriving at a 1.5-m focusing mirror. The pump beam was focused to a

570-µm radius inside the nonlinear crystal and exited without coming into contact with any

mirrors. The unused pump light was sent to a beam dump, not shown in the schematic. The

angularly dispersed idler beam was sent out near the pump beam and monitored with a

Coherent PM30V1Q thermopile sensor or collected in a beam dump.

Page 83: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

65

The position of the M3 end mirror was adjusted using a micrometer to achieve

synchronous pumping. The cavity length was manually tuned while carefully monitoring

the repetition rate of the pump to maintain long-term use, although automated stabilization

could be used [175]. The size of the cavity mode at the crystal and the stability of the

cavity were controlled by the curvature and position of the end mirrors and the mirrors

closest to the crystal. Although these four mirrors were the critical pieces for obtaining

good stability and the correct mode size, it was difficult to find mirrors that could be used

to create the long cavity length required to achieve synchronous pumping. The cavity

length was extended to 21.2 m using image relays by setting all other mirrors at a distance

of 2f from each other [176]. This technique is explained further in the following section.

As discussed in the previous section, there was no additional dispersion control in

the folded cavity. The linear dispersion was dominated by the coatings on the mirror, but

there was a small contribution from the material dispersion of the BiBO crystal. The total

linear dispersion in the cavity over the tunability of the OPO is shown in Fig. 4.8. The

dispersion is slightly positive near 2 µm, but most of the dispersion is slightly negative.

Near the edge of the OPO tunability the material dispersion is significantly increased,

which is due to the multitude of reflections off the mirrors in the folded cavity. The balance

of the negative linear dispersion with the positive GDD obtained through SPM is not well

understood, due to the long cavity length in air and the significant change in mode size

throughout the cavity. Ideally, uniform dispersion over the tunability of the OPO was

desired so that the linear dispersion could easily offset the SPM, which would create

transform limited pulse durations. As shown in Chap. 5, transform-limited pulses were not

achieved, suggesting that the intracavity dispersion could be improved to optimize the

pulse duration.

Page 84: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

66

Figure 4.8: Total linear dispersion of the folded OPO cavity. GDD values were calculated

from the total combination of all mirrors in the cavity and the material dispersion of BiBO

for an 8.4-mm-long crystal.

The crystal was held in a custom mount designed for high precision control of the

crystal orientation and mitigation of thermal effects. The crystal was actively cooled using

a Peltier module (CP0.8-127-06L, Laird Technologies) to enable stable steady-state

operation at higher powers resulting from absorption of the signal and idler beams. The

unpolished crystal face was placed on a chemically etched indium foil for good thermal

contact with a 6 12 12-mm aluminum block that was temperature controlled by the

thermoelectric cooler, as shown in Fig. 4.9.

The thermoelectric cooler (TEC) was bolted to a modified kinematic mount for

initial coarse alignment and to ensure that the crystal was oriented correctly. The aluminum

blocks were attached to a post for variable height control on the crystal. At the base was a

Newport UTR80S precision steel rotation stage for fine tuning control of the angle between

the pump and crystal axis.

Page 85: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

67

Figure 4.9: Images of the crystal mount and beam orientation. (a) Top-down view of the

crystal with the pump (green), signal (red), and idler (red outline). Angles are exaggerated

for clarity. (b) Side view of the crystal mount, with all beams in plane with the middle of

the crystal. (c) Clear image of the crystal, mount thermistor, and thermoelectric cooler

(TEC).

4.2.3 ABCD Analysis

An ABCD analysis [177] was used to design the cavity. This ensured that (1) the

correct cavity mode size for optimal pump/signal beam overlap could be achieved, (2) the

cavity was stable, and (3) the cavity length could be easily adjusted.

Using an ABCD model, the mode size at the crystal was modified by adjusting the

radius of curvature (ROC) and distances between the cavity mirrors. The beam size at the

crystal was determined by the bulk damage threshold of the BiBO crystal, which had to be

greater than 550 µm. This limitation was due to maximum intensity anticipated in the cavity

and the damage threshold of the crystal, which was reported by the manufacturer to be 1

GW/cm2. Higher intensities may be possible without causing damage to the crystal, but the

damage threshold of BiBO has not been thoroughly investigated for ultrafast pulses. The

beam waist was designed to have a 570-µm radius for both the pump and signal beams,

which ensured adequate gain and a minimal risk of damage to the crystal.

The stability of the cavity was important to increasing the reliability of the laser.

Instabilities could result in pointing fluctuations in the beam, which could cause significant

Page 86: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

68

problems due to the sensitive alignment requirements of a critically phase-matched

oscillator. The cavity stability, determined from the mean of the main diagonal of the

ABCD matrix for the cavity, was calculated to be 0.07 on a scale of –1 to 1, suggesting the

cavity was very stable [113, 177].

The core components that determined the spot size and stability of the OPO were

the separation between the two mirrors closest to the crystal and the curvatures of these

two mirrors and the end mirrors of the cavity. If all other mirrors were set at a distance of

2f from the neighboring mirror, the fundamental cavity parameters did not change. Either

end mirror could be moved over 300 mm, resulting in a 0.1% effect in both the cavity

stability and the beam size. However, movement of over 5 mm of the position between the

inner mirrors and/or the crystal caused a notable change to both the beam size and the

stability. Invariance to the cavity length was an important feature to be able to achieve

synchronous pumping since the pump laser experienced significant drift over the course of

the day, as mentioned above.

The ABCD model showed that additional telescopic relays between these mirrors

had no noticeable effect on the cavity. This shows that additional mirrors can be added to

arbitrarily extend the cavity length as long as they are placed at distances of twice their

focal length from the neighboring mirror. This can be a useful tool for extending the cavity

and taking advantage of the high-energy pulses available from longer pump cavities.

Page 87: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

69

5 Characterization of the Ultrafast

Optical Parametric Oscillator

After the ultrafast OPO was constructed and stable operation was achieved, the

cavity was characterized. The spectrum and pulse duration were measured as well as the

beam quality and output power. The characterization methods and the results of the

measurements will be discussed in depth in this chapter.

5.1 Pulse Duration

One of the most critical measurements of an ultrafast system is the pulse duration.

The most practical method for measuring an ultrafast pulse near 2-µm was to use a

commercial autocorrelator that could make an intensity autocorrelation of the OPO output

[81]. The system used to measure the autocorrelation signal was a modified FR-

103MN/LiIO3 autocorrelator from Femtochrome. The sensor was swapped from the

traditional photomultiplier tube (PMT) to a high sensitivity photodiode. The PMT was

insensitive to wavelengths of less than 800 nm so it could not be used to detect the second-

harmonic output of the LiIO3 crystal for the range of available signal wavelengths. The

photodiode required a separate alignment of the autocorrelator due to the placement of the

detector. Additional internal electronics were also required to switch between operation for

the pump and OPO output. The LiIO3 crystal could be used in both cases to obtain a second-

harmonic beam for the pump and OPO, even though different tuning angles were required.

Although these modifications allowed for autocorrelation measurements of both systems,

significantly more power was required from the OPO. The low conversion efficiency of

LiIO3 for creating SHG from 2 µm meant that more power was required from the OPO to

obtain a usable autocorrelation signal. This problem was exacerbated by the poor

Page 88: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

70

sensitivity of the photodetector compared to the PMT. A minimum of 0.4 W of 2-µm power

was required to obtain a useable autocorrelation trace.

A typical autocorrelation trace is shown in Fig. 5.1. The 455-fs pulse duration of

was obtained assuming a sech2 pulse shape, which was chosen due to the characteristics of

the pump pulse as well as the measured spectrum of the OPO output, discussed in the

following section. The autocorrelation trace is heavily averaged due to the acquisition time

of the autocorrelator and averaging done in the oscilloscope. The autocorrelator used

rotating mirrors to sweep over the length of the two arms of the interferometer. The mirrors

rotated at a rate of 45 Hz, which means that individual traces consisted of over 150,000

pulses. To improve the signal-to-noise, the averaging function of the oscilloscope was used

to calculate the mean of 16 traces. The pulse duration was obtained by performing this

method ten times, which resulted in an average pulse duration of 455 fs with a standard

deviation of 47 fs.

The autocorrelation measurement shown in Fig. 5.1 was taken with an output power

of 0.7 W. No measurable change in the pulse duration was observed between output powers

of 0.5 and 1.0 W. Additionally, autocorrelation measurements with center wavelengths

between 2040 and 2080 nm showed no measurable effect on the pulse duration.

Autocorrelation measurements were not possible at wavelengths outside this range due to

small fluctuations in the peak power that interfered with the shape of the autocorrelation

trace.

Page 89: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

71

Figure 5.1: Autocorrelation trace of the OPO giving a 455-fs pulse duration assuming a

sech2 pulse shape [165].

5.2 Spectral Measurements

Measurements of the pulse spectrum are another fundamental characterization of

ultrafast pulses. Since more-advanced techniques like FROG and SPIDER were not used

for pulse characterization, the shape and width of the pulse were necessary to infer

information about the spectral phase [82, 83].

The spectrum of the OPO was measured using a home-built spectrometer, a

schematic of which is shown in Fig. 5.2. Light from the output of the OPO passed first

through a variable slit. A lens, placed a focal length away from the slit, then collimated the

light before it continued on to two separate uncoated SF10 prisms. In this configuration,

two prisms were necessary to obtain adequate spectral resolution. After the prisms, a 92-

mm-focal-length lens was used to focus the angularly dispersed light onto a Pyrocam III

beam analyzer with a Si window, AR coated for 1.05 to 2.5 µm.

Various components of the spectrometer had to be optimized to ensure that the

resolution of the spectrometer did not interfere with the width of the measured pulse

spectrum [178]. The variable slit was optimized to give the maximum amount of light

without affecting the width of the spectral measurements. The optimal slit size was found

to be 35 µm. The dispersion obtained from two prisms allowed the spectrum to be sampled

by the detector array over the entire tuning range of the OPO. The pixels on the detector

Page 90: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

72

were 70 70 µm with an array size of 12 mm, corresponding to 171 pixels. Tests were

performed with a single prism, but the resulting spectral resolution was too broad for an

accurate measurement of the pulse spectrum. Since the spectrometer was built in-house, a

combination of calibration methods were used to ensure accurate measurements. In

general, this required that the pixels of the detector were mapped to the correct wavelengths

and that the nm/pixel was well understood. Typically, narrowband sources at multiple

wavelengths are used to calibrate spectrometers. However, due to the expense of these

sources and the intensity required for a strong signal on the detector that was used,

alternative methods of calibration were necessary.

Figure 5.2: Diagram of the home-built spectrometer, not drawn to scale. L1: 210-mm

focusing lens; L2: 92-mm focusing lens; prisms: uncoated SF10 prisms with a 60.9° apex

angle.

The first calibration technique focused on characterizing the wavelength range of

the spectrometer. The SHG of the OPO output was first focused through a 1-mm crystal.

The unused 2-µm light was reflected off a dichroic mirror (described in Chap. 4). The

second-harmonic light was injected into a multimode fiber and sent to an Ocean Optics

HR4000+ spectrometer. The center wavelength of the SHG spectrum was used to

characterize the tunability of the OPO. The spectrum collected from the home-built

spectrometer was then calibrated such that the edges were known; then a distribution of the

Page 91: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

73

how many nm/pixel could be calculated. The dispersion of the prisms and the focusing lens

were arranged such that the full tuning range of the OPO could be viewed over the pixel

array of the detector.

An alternative method was used to determine the wavelength range by changing

the cavity length to map the wavelength between the SHG output and the home-built

spectrometer. The SHG and OPO spectrum were measured by the same method just

described. Instead of measuring the edges of the spectrum, however, a map of the spectrum

as a function of cavity length was investigated. The cavity length was changed at a constant

speed in the same direction for both spectra. For the SHG, the central wavelength could be

plotted as a function of the cavity length. When this process was repeated for the home-

built spectrometer, the wavelength as a function of the cavity length was used to calibrate

the discrete wavelength bands for the individual pixels. The result of this calibration

confirmed the previous method, in which the nm/pixel distribution was inferred strictly

from the edges of the OPO tunability. In this case, a nm/pixel map was created as a function

of the change in cavity length and showed good agreement between the two methods. Even

though narrow linewidth sources were not used in calibration, the agreement between these

two measurements gave strong confidence in the accuracy of the spectrometer.

The final calibration ensured that the resolution of the spectrometer did not interfere

with the spectral width of the measured pulses. The resolution was controlled by several

different spectrometer parameters: the slit width, the pixel size on the detector, the focal

power of the optical element before the detector, and the dispersion placed on the beam

from the prism pair [178]. Except for the pixel size (which was fixed due to the detector),

the spectrometer elements were optimized in a trial by error due to interconnectedness of

their performance. The slit width was initially optimized by minimizing the slit so that the

weakest parts of the spectrum had just enough power to be measured by the detector. The

first prism of the prism pair almost introduced enough dispersion on its own, so the second

prism could be used to fine tune the angular dispersion of the spectrum. The second prism

was optimized such that a full tuning range of the spectrum could be viewed on the detector.

Page 92: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

74

To ensure that the resolution in this arrangement did not affect the pulse duration, the

second prism was tuned such that the spectrum from 2140 to 2200 nm could be viewed.

The spectral width between the two configurations was compared and found to be identical,

suggesting that the resolution of the spectrometer could accurately measure the spectrum

of a pulse over the entire tunability of the OPO.

The spectrum of the OPO is shown in Fig. . The spectrum of an individual pulse is

shown in blue. The sech2 pulse shape is evident, supporting the use of the sech2

deconvolution factor in the autocorrelation. The FWHM of the spectrum is nearly constant

over the full tunability of the OPO, which is shown in black. The tunability spectrum of

the OPO is slightly more complex than the figure would suggest.

Figure 5.3: The spectrum of on individual pulse with a bandwidth of ~30 nm is shown in

blue; the spectrum over the tunability of the OPO is shown in black [165].

It would be easiest to understand this curve as the relative spectral power of pulses

across the full tunability of the OPO. The shape of the tunability curve was determined by

compiling the pulse spectrum while tuning the OPO over the entire bandwidth (which was

done by slowly changing the cavity length). The cavity was tuned over 100 µm at a rate of

0.1 µm/s over the full range of cavity lengths that allowed the cavity to lase. The rate was

chosen to be slow enough to avoid power fluctuations that arise during rapid tuning of the

cavity length. The spectrum was sampled at 24 Hz and stored for analysis.

Page 93: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

75

Using this data, a program was able to analyze a few thousand spectrum samples to

compile the relative spectral power over the tunability of the OPO. An empty array was

created for wavelengths between 1950 and 2250 nm. A pulse spectrum was then loaded

and compared to the array. If the amplitude of the spectrum was larger than the array, those

values were replaced. This procedure was iterated for each measured spectrum over the

full tunability of the OPO. As the cavity was tuned, the maximum value was stored,

replacing previous spectral amplitudes when applicable. The amplitudes of the spectrum

over the tunability of the OPO were compared with power measurements and shown to be

in good agreement, as was expected, to the uniform spectral response of the detector in the

area of interest. The result is the dashed black curve shown in Fig. 5.3, which represents

the relative spectral power for a device that is continuously tunable over 200 nm.

Features in the tunability spectrum present a few interesting facts about the

behavior of the OPO. The tunability curve is extremely similar to the transmission

measurements shown in Fig. 4.3. This is not too surprising considering that losses from

linear absorption in the crystal dominated the wavelength-dependent losses in the cavity.

The spectrum at 2150 nm is roughly 40% as powerful as the spectrum at 2060 nm. This

amplitude difference is directly related to the power of the OPO, where the maximum

power can be obtained only in the spectral range between 2040 and 2100 nm. In addition

to power, this area of operation also gave the most spectral and temporal stability. Near

2060 nm (degenerate operation) the center wavelength was very stable and would stay

constant within 5 nm. However, the spectrum near the edges would drift back and forth

over 20 nm, making it difficult to maintain at a constant wavelength.

5.3 Temporal Stability

The 7.08-MHz pulse repetition rate and temporal stability were investigated using

a rf spectrum analyzer connected to a Thorlabs DET100A Si photodetector and DET10D

InGaAs photodetector for the pump and signal, respectively. The free-running pump

showed significant temporal jitter that could be reduced with active stabilization, which is

Page 94: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

76

discussed in Chap. 7. This noise appeared in the rf spectrum of the OPO as well, as was

expected because of the lack of active control (shown in Fig. 5.4).

Figure 5.4: Radio-frequency (rf) spectrum of the signal (blue) and pump (red) [165].

The presence of noise in the rf spectrum for frequencies < 50 kHz can easily be

seen in Fig. 5.4. Most of the features in the noise seen in the pump’s rf spectrum can also

be seen in OPO’s the rf spectrum. The difference in amplitude between the two suggests

that the temporal jitter of the pump was amplified in the OPO, which could be eliminated

by active stabilization of the TDML. Additionally, low-frequency noise is present in the

OPO near 10 kHz, which is not present in the pump. The exact cause of this noise unknown,

but it is not too surprising due to the lack of independent stabilization of the OPO and the

operation of the OPO in air without a full enclosure to protect against air flow. Despite the

presence of this noise, the OPO output had less than 2% rms peak-to-peak power

fluctuations over the course of 4 h, and control of the center wavelength of the OPO to

within 5 nm for operation between 2040 and 2100 nm was possible. Active stabilization of

the pump is discussed in section 7.2.3. Active stabilization of the OPO based on wavelength

stabilization with position-sensitive photodiodes is a commonly employed technology that

could have been used in this system but was outside the scope of this investigation [18,

169, 175].

Page 95: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

77

5.4 Output Power and Scaling Limitations

Signal and idler powers measured simultaneously in this system are shown in Fig.

5.5. The pump power threshold of the OPO was 14.5 W. Reliable performance was

obtained for signal (and idler) powers of 2.5 W (5.2 W) corresponding to 350-nJ (730-nJ)

pulses with a slope efficiency of 14%. At maximum pump power, the OPO was unstable

and the beam became deformed from local heating of the BiBO crystal, which is discussed

thoroughly in Chap. 6. The maximum output powers obtained under these conditions were

3.0 W with 430 nJ pulses and 7.6 W with 1.1-J pulses from the signal and idler,

respectively. The conversion efficiency, calculated by comparing the amplified signal

beam and the measured power of the idler with the total pump power, was estimated to be

25%, decreasing near maximum power to 22%.

Figure 5.5: Output power of the signal (blue) and nonresonated idler (red) [165].

Power measurements were taken with a Coherent PM10V1 thermopile sensor

connected to a LabMax-TOP power meter. Using the built-in features of the power meter,

the output power was measured using data collected at 10 Hz over the course of 120 s. For

measurements with less than 30 W of pump power the standard deviation of the output

power was less than 10% of the mean. At higher powers, however, instabilities began to

cause a significant reduction in performance, with standard deviations to the powers shown

up to 22%.

Page 96: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

78

Thermal stabilization of the laser was of critical importance to obtain the powers

shown in Fig. 5.5. Initially the TEC was set to a temperature of 29.8°C, which ensured the

crystal was properly phase matched at low temperatures. As the laser power was increased,

the TEC was unable to sufficiently cool the crystal, so the temperature setting had to be

reduced to achieve maximum output power. At maximum power the optimal TEC

temperature was 24.3°C. The combination of the decreased temperature setting of the TEC

and the falloff in efficiency suggested that the TEC could not sufficiently cool the crystal

at higher powers. This was the motivation for an investigation of the thermal properties of

the OPO, which is the focus of Chap. 6.

5.5 Beam Quality

Beam-quality measurements of the OPO were important for evaluating the

usefulness of the OPO for various applications and understanding problems in the system.

For pump powers less than 30 W, the output beam had very high quality. At higher powers

the beam showed significant deformation in addition to the instabilities mentioned earlier.

The far-field signal beam profiles of each case are shown in Fig. 5.6. Images of the idler

output could not be obtained due to the high average powers and limited table space.

Qualitatively it was possible to see that the idler beam had high angular dispersion, which

was expected from the noncollinear interaction geometry between the pump and signal

beams [60]. Although the high output powers of the idler would make it of interest for other

applications, the poor beam quality would make the beam much less useful than the signal.

The beam quality shown in Fig. 5.6(a) is representative of the output beam for pump

powers of less than 30 W where the slope efficiency is 14%. The beam is nearly a perfect

Gaussian profile in both the x and y axes, while being slightly larger along the x axis. This

is likely due to a slight astigmatism in the cavity from the off-axis AOI of eight separate

spherical mirrors. At pump powers exceeding 30 W the beam became deformed due to

local heating effects inside the crystal. Local deformations on the bottom-right quadrant of

the beam can easily be seen and the beam becomes slightly elongated in the y axis. The

Page 97: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

79

exact cause of these deformations is not well understood, even though it must be related to

operation near maximum power.

Figure 5.6: Beam profiles at output powers of (a) 0.4 W and (b) 3.0 W. Local heating

effects become noticeable and cannot be mitigated by actively cooling the crystal. As can

be seen in (b), the local heating begins to deform the beam [179].

Beam quality was measured to ensure that the beam shown in Fig. 5.6(a) was indeed

Gaussian. Unlike the TDML, this measurement was not performed with the knife-edge

technique. Initial attempts were made to use this technique but the 10% power fluctuations

made this exceedingly difficult. Furthermore, the time required to perform the knife-edge

measurements is usually of the order of a few hours for a single set of measurements. The

power fluctuations made the knife-edge technique impractical since precise power

measurements within 1% of the maximum power were required to get accurate

measurements of the beam size using the 84/16 method [136]. An additional difficulty was

controlling the long-term drift of the pump laser. Since compensating for drift was done

manually, it proved very difficult to ensure that the OPO’s center wavelength was

unchanged for multiple hours.

To obtain an accurate 𝑀2 measurement, images of the output beam at various

distances from a focusing element were used. Beam quality could be measured by this

technique much quicker than by using the knife-edge technique. A Pyrocam III beam

profiler was used to measure the size of the output beam. The OPO output was passed

Page 98: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

80

through an uncoated 500-mm-focal-length lens, and images of the beam were measured

through focus to obtain adequate sampling of the beam at discreate distances from the lens

[134, 135, 170]. The result of this measurement is shown in Fig. 5.7.

Figure 5.7: Beam-quality measurements along both the x (red) and y (blue) axes with

corresponding M 2 fit [165].

The refresh rate of the beam profiler was 24 Hz, which provided nearly 300,000

pulses per measurement. Electronic triggering for images of individual pulses was not

possible with this repetition rate without significant cost to the detector. On top of the

averaging as a result of the sampling rage, individual images were averaged over 20

acquisitions to account for the power fluctuations over time. The centroid of the beam was

then identified, and lineouts along the x and y axes were taken. Due to the pixel size, these

lineouts were quite rough. To smooth the curves these pixel arrays were averaged for ±1

pixel in both directions. A built-in interpolation function in MATLAB was used to

calculate the 1/e2 full width, half of which was used to calculate the beam radius. This

process was repeated 3×, and the resulting average of the calculated beam radius was used

in a fitting algorithm to calculated the value of 𝑀2 [134]. The fitting routine created a curve

from the measured data to fit the equation

𝑓 = 𝑏√1 + [𝑎(𝑧−𝑐)𝜆

𝜋 𝑏2 ]2

. (5.1)

Page 99: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

81

Here, z is the distance from the lens, c is the location of the beam waist, b is the size

of the beam at focus, and a is the value of the beam quality 𝑀2. The fitting parameters

were confined such that the values must be physical, e.g., a > 1.0. The fitting algorithm

was run to calculate the mean value of the three trials, as well as the minimum and

maximum values to account for error. The minimum and maximum values of the beam size

were used to calculate the standard deviation of the three trials. The resulting 𝑀2 value is

shown in Fig. 5.7, which suggests a maximum 𝑀2 value of 1.10 and 1.12 for the x axis and

y axis, respectively. As expected, the beam quality from the output of the OPO is quite

good when the cavity is well behaved.

Page 100: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

82

6 Thermal Analysis and Limitations

of High-Power Operation

In this chapter, the impact of the high intracavity powers on the nonlinear crystal is

explored. Although strategies were employed to mitigate thermal effects, limitations in the

output power were still present. First, the motivation of the thermal analysis is explored.

The methods that were used to measure the thermal properties of the crystal are then

explained. Next, the impact of the total absorbed power from the signal and idler power on

the thermal gradient in the crystal is discussed. The chapter concludes by investigating the

impact of the thermal effects on the OPO, the most significant of which are the thermally

induced phase matching errors and thermal lensing.

6.1 Motivation of Thermal Analysis

6.1.1 2-µm Absorption in BiBO and Total Absorbed Power

The most important reason why an investigation into the thermal properties of the

nonlinear crystal was necessary was the unexpected absorption near 2 µm, which can be

seen in Fig. 4.3. Sharp drops in transmission due to absorption near the signal and idler

wavelengths make operating with high average powers a difficult endeavor in a critically

phase-matched parametric device.

Since the OPO is operated near degeneracy (which occurs at 2060 nm), both the

signal and idler usually have some component of their spectrum in the sharp absorption

peaks near 2150 nm. Due to the relatively broad spectrum of the signal and idler pulsed,

which is roughly 30 nm, and the tunability of the OPO, it is difficult to quantify the exact

amount of light that is being absorbed by the crystal at any given moment. Even though

the transmission losses are far from ideal, mitigation strategies for dealing with some level

Page 101: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

83

of absorbed laser light will be necessary for any applications where the scaling of average

power is of interest. It will be shown in Sec. 6.3 that there is a direct relationship between

the intracavity power and the presence of thermal effects.

Due to the complex relationship between the signal and the idler and their

respective bandwidths, the most useful metric to relate the average power with the absorbed

light was the short-wave IR (SWIR) power. This value was calculated as the sum of the

amplified and returned signal beam and the exiting idler power. The signal beam, which is

resonated through the OPO cavity, sees the crystal twice during one round-trip: once during

the amplification stage and again when the beam is headed toward the output coupler.

During this time, the idler passes through the crystal only once and is not resonated in the

cavity. With measurement values of the idler’s average power, the output power of the

OPO and the known transmission of the output coupler, a total contribution of near 2 µm,

could be calculated.

The SWIR power is the only contributing factor to the heat dissipated in the crystal.

Although the average power of the pump light is considerable, the absorption is less than

0.5%. With the cavity blocked, thermal measurements of the crystal were made at

increasing pump powers, and even in the absence of active cooling, no measurable change

in the crystal temperature was observed. This suggested that any heating of the crystal was

the result of the near-2-µm light from the signal and idler beams.

6.1.2 Active Cooling of the OPO

As mentioned above, a TEC was used to actively cool the BiBO crystal during

operation. It was necessary to removing the dissipated heat from the crystal was necessary

for high power operation. Images of the setup are shown in Fig. 4.9, and a more detailed

schematic of the device is shown in Fig. 6.1.

Page 102: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

84

Figure 6.1: Schematic of the thermoelectric cooler (TEC) and heat sink for the BiBO crystal

shown from the side (not drawn to scale).

The 5 5 8.4-mm BiBO crystal is held in place by a small plastic clamp that is

mechanically connected to the large-area aluminum plate, which is connected to a

kinematic mount for tip and tilt control of the crystal. The unpolished crystal face was

placed on a thin layer of chemically etched indium foil that ensures good thermal contact

between the BiBO and the TEC. The TEC was a Peltier module (CP0.8-127-06L, Laird

Technologies) and actively transferred heat from the crystal to a 6 12 12-mm aluminum

block. A thermistor was attached to the block to monitor the temperature and was

connected to a home-built controller for the TEC. The current sent to the TEC from the

controller was varied to achieve a stable temperature, which could be set to within 0.1°C

at a range between 10°C to 50°C. The signal beam was placed vertically in the middle of

the crystal to ensure proper clearance of the beam from the crystal mount regardless of

changes to the alignment. The beam was place roughly 1.5 mm from the edge of the crystal

to give additional usable space in the event of damage. Although placing the beam closer

to the cooling surface may have helped with cooling, the size of the beam was a concern

for creating a uniform temperature distribution.

As shown in Fig. 5.5 the falloff in efficiency for pump powers that exceeded 30 W

signified a change in the behavior of the system at high intracavity powers. Thermal

management became a prime concern when the output power was compared with the

Page 103: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

85

optimal temperature setting of the TEC. As stated above, to initiate lasing, the TEC was

set to a temperature of 29.8°C, which was slightly above the 24.2°C room temperature . Of

course, the laser could have been operated with a slightly different angle between the pump

and the crystal axis to initiate lasing, but it was found that this was the optimal method for

achieving maximum power over the full range of available pump powers. The initial

temperature setting allowed for maximum output power at the lasing threshold. As the

pump power was increased, the TEC operating temperature had to be reduced to maximize

the output power, as shown in Fig. 6.2.

Figure 6.2: The relationship between the TEC temperature setting and the signal power

shows how the temperature had to be reduced to obtain optimal power of the OPO output.

When an optimal TEC temperature was found, the laser could achieve stable

operation for pump powers of less than 30 W. This suggested that the bulk temperature of

the crystal was constant and that the TEC could adequately remove the dissipated heat from

the crystal. However, if lasing was stopped for any reason (e.g., cavity length mismatch

between the pump and OPO), the temperature of the TEC had to be reset to achieve lasing

again. This behavior heavily suggested that local heating effects require that the bulk

temperature of the crystal be reduced to achieve adequate phase matching at higher powers.

Page 104: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

86

Furthermore, the falloff in efficiency at higher powers suggested that the TEC was unable

to adequately cool the crystal. This is supported by the optimal temperature setting of the

TEC for output powers exceeding 2.5 W. For the full range of signal powers, the TEC

could maintain the temperature setting of the controller. However, output powers beyond

2.5 W (which is the point where the efficiency falls off significantly) required only a small

decrease in the TEC temperature for optimal output powers. This can be explained for by

local heating effects that could no longer be adequately compensated for by active cooling;

it was not a result of the TEC having insufficient enough power to draw out the total heat

dissipated from the crystal.

Tuning the angle of the crystal relative to the pump beam was a viable, although

problematic, alternative to actively cooling the crystal. Practically, this made it difficult to

handle the high-powered reflections of the pump beam. At maximum power, the reflected

pump beam had an average power of 7.5 W from the reflection of the front and back

surfaces of the crystal. This reflected beam had to be captured by a beam dump, which was

placed near the turning mirror before the 1.5-m ROC pump mirror. As can be seen in Fig.

4.7, the available clearance was quite small, so large adjustments to the crystal angle were

avoided unless absolutely necessary. Angle tuning the crystal was tested in conjunction

with actively cooling the crystal. It was found that angle tuning the crystal could not

improve the OPO performance for powers where the efficiency had begun to decrease.

6.2 Measurement of Thermal Properties

6.2.1 Thermal Camera Selection and Image Acquisition

To investigate suspicions mentioned in the previous section, thermal images were

taken of the crystal while operating the OPO. Images were taken with an HT THT45

compact thermal camera. This device is a relatively inexpensive handheld microbolometer

array that allows one to take detailed thermal images of the crystal. Higher-resolution

cameras were available at a greatly increased cost were found to be unnecessary for the

needs of this investigation.

Page 105: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

87

Obtaining images with enough resolution to capture the fine details of the thermal

gradient was critical for modeling thermally induced effects. The camera had to be placed

as close to the crystal as possible within the minimum focusing distance of the device. The

tight beam clearance and limited space in the plane of the crystal created additional

constraints on the optimal placement of camera. The best location was determined to be

130-mm away from the front face of the crystal (the side from which the pump beam

arrives) slightly above the plane of the crystal. A variable focus lens was available in the

unit and was used to focus the front face of the crystal. The height difference between the

camera and crystal was chosen such that thermal radiation from the surface of the indium

foil was not reflected off the front face of the crystal.

Thermal calibration of the crystal was important for obtaining accurate temperature

measurements. It was found that temperature values were highly dependent on the

emissivity of the material, which was a variable that could be changed both during image

acquisition or in post-processing using the THTLink software package provided with the

hardware. Images were calibrated by heating and cooling the crystal using both the TEC

and a hot plate and then measuring the front face of the crystal to determine the correct

emissivity value that would have the software accurately predict the steady state

temperature of the crystal for different images.

6.2.2 Preparation of Thermal Images for Data Analysis

Importing the data to MATLAB was necessary to adequately analyze the thermal

properties of the crystal. Although the THTLink software was a powerful tool for

investigating the thermal images, it lacked the flexibility needed to extensively investigate

the thermal impact on the OPO cavity. Several steps were necessary to prepare the images

before they could be used to analyze the thermal properties of the crystal.

First, the images taken were saved to a jpg file format that be could analyzed using

THTLink. This program was necessary to pull the metadata from the image, including the

temperature’s relation to the image’s colormap as well as various material information

Page 106: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

88

(e.g., emissivity). All images were uploaded into THTLink, which allowed for the color

scale to be set to a given temperature range. An appropriate scale was chosen after

analyzing a given data set so that fine details were easy to observe. These modified images

were then saved by the program into a separate jpg file as a grayscale image. By knowing

the temperature range of the grayscale, the jpg file could be imported into MATLAB with

a known temperature scale that allowed for further analysis. Due to the exportation of files

in this manner, an interpolation of the thermal data was performed, giving a higher

resolution of the thermal gradients than was possible with the initial images taken.

Before data could be analyzed, it was necessary to rescale and crop the images.

Rescaling was necessary due to the angle of the camera relative to the crystal face. For

thermal images to be more representative of the physical dimensions of the crystal, the

height of each image was changed so that the height of the crystal in the images was equal

to the width. The result produced square images of the crystal face that could be easily

scaled to the physical dimensions provided by the manufacturer.

Although the signal beam never moved relative to the crystal face, there were slight

changes between individual pictures due to the manual image acquisition as mentioned

above. To more easily compare the images, they were cropped so that only the crystal face

was visible. In greyscale, this process was difficult to do, however when the images were

colored, it was much easier to crop. Therefore, duplicates of each image were made for

both color and greyscale. The scaled images were then cropped in color mode and the

cropping dimensions relative to the entire picture were applied to the grayscale images.

The final images were then saved as new files ready to be imported to MATLAB.

6.2.3 Final Thermal Images

It would be useful to first interpret a sample thermal image to understand the key

features before using this data to understand the dynamics of the OPO. An example of the

final thermal images after they were calibrated and uploaded to MATLAB is shown in Fig.

6.3. The heat distribution caused by the signal and idler beam is shown schematically in

Page 107: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

89

Fig. 6.3(a) along with the relative location of the TEC. Although these two beams start in

the same location there is a small spatial separation as they exit the crystal due to the

noncollinear pump/signal interaction. Figure 6.3(b) shows a sample thermal image where

local heating effects from the SWIR power are evident. Preliminary investigation showed

that the surface of the crystal closest to the TEC was much cooler than the temperature

around the beam, reinforcing the suspicion that the local heating effects played a significant

role in the cavity’s behavior.

Figure 6.3: (a) Schematic of the front face of the BiBO crystal. The crystal’s dimensions

are 5 × 5 mm. The locations of the signal and idler beam are shown by the solid and dashed

red circles, respectively. (b) Sample thermal image showing the local heating effects from

the presence of these two beams [179].

6.3 Impact of Internal Power on the Thermal Gradient

As mentioned in Sec. 6.1, at higher powers the impact of the falloff in efficiency

was likely related to the thermal gradient. With a method in place to take thermal images

of the crystal face, it was of interest to measure the thermal gradient as a function of the

SWIR power. Figure 6.4 shows thermal images are shown with the corresponding SWIR

power.

Page 108: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

90

Figure 6.4: Thermal images of the front face of the BiBO crystal during OPO operation

with the corresponding short-wavelength IR (SWIR) power. Each image was taken after

optimizing the thermoelectric cooler (TEC) temperature for maximum output power [179].

At low powers, it is evident that local heating is not a problem. The TEC can control

the temperature of the BiBO crystal and optimize the signal power without creating a

thermal gradient. As the SWIR begins to increase, significant local heating occurs around

the signal and idler beams. At 23.7 W SWIR power, the slope efficiency of the cavity has

not yet fallen off, yet one can see the onset of local heating effects. As the SWIR power

approaches maximum, the thermal gradient becomes more pronounced, which directly

corresponds to the falloff in efficiency at higher powers.

The thermal gradient was of critical importance for understanding the falloff in

efficiency at higher powers. As will be shown in the following sections, the thermal

gradient causes a change in the index of the material, which affects phase matching and

cavity dynamics through thermal lensing.

6.4 Thermally Induced Phase-Matching Errors

Temperature has a significant effect on critically phase-matched systems [142]. The

maximum change observed in the crystal temperature over the available power in the OPO

was 4.8°C. This corresponds to an optimal change in the phase-matching angle of 1.78

mrad at degeneracy, which is significant compared to the angular acceptance of the OPO,

which was calculated to be 3.04 mrad [163]. An additional complication is the spatially

varying index, which causes a complicated gain profile across the beam area from non-

uniform phase matching. To investigate this complication, thermal images were taken at

Page 109: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

91

different output powers while the OPO was running and used to determine the effect of

local heating on the phase matching, which is shown in Fig. 6.5.

In this figure, thermal images were used to calculate a 2-D profile of the index using

the thermo-optic dispersion formula for the BiBO crystal [163]. The index values were

then used calculate the phase matching that is typically done for simple 1-D models. It was

useful to understand the reduction of gain at higher powers by investigating the impact of

the thermal gradient on the phase matching. This was done by comparing the phase

matching for different center wavelengths of the signal beam and how that was affected for

different output powers.

Figure 6.5: Thermal images of the front face of the BiBO crystal during OPO operation

with the corresponding SWIR power. Each image was taken after optimizing the TEC

temperature for maximum output power. The circle shows the 1/e2 diameter of the beam

area [180].

Page 110: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

92

The impact of the thermal profile varies significantly over the tunability of the

OPO. Near threshold, the 4.5 W SWIR power does not produce a strong thermal gradient,

which leads to a uniform phase-matching profile across the tunability of the OPO, as shown

in Figs. 6.5(a), 6.5(d), and 6.5(g). Here the metric of phase matching is the commonly used

sinc2 (∆𝑘𝐿

2) function that directly relates the phase matching to the maximum-available

parametric gain [61]. In Figs. 6.5(b), 6.5(e), and 6.5(h) the incident SWIR power on the

crystal has risen to 23.7 W, where the local heating becomes apparent. Figures 6.5(c),

6.5(f), and 6.5(i) show the most-pronounced thermal gradient, which occurs at maximum

output power. Although higher output powers could be obtained with more pump power,

the increase in the thermal gradient suggests that the slope efficiency would continue to

decline because of the unmitigated thermal effects.

The center wavelength of the signal beam, which can be tuned over a range of

roughly 200 nm, changes the phase matching when a strong thermal gradient is present. As

shown in Figs. 6.5(a)–6.5(c), the phase-matching gradient does not significantly change

despite the increasing thermal gradient in the crystal. The minimum value of the sinc2

function over the beam area is 0.77, suggesting the OPO is well phase matched near the

degenerate wavelength of 2060 nm [56]. This is not too surprising considering how flat the

phase matching curve is near degeneracy, as shown in Fig. 4.4. For wavelengths away from

degeneracy, the phase matching degrades as the SWIR power is increased. This makes

sense since the phase matching curve changes more rapidly as a function of the angle

between the pump and crystal axis, which will shift as the temperature changes. Figures

6.5(g)–6.5(i) show that when the center wavelength is at 2150 nm, the beam is no longer

adequately phase matched at high powers. The most obvious method for resolving this

problem would be to either rotate the crystal to achieve the optimal pump angle or to

operate the TEC at a different temperature to optimize performance at this wavelength.

The phase-matching gradient at wavelengths away from degeneracy suggested that

this effect could not be eliminated entirely. The total change in the sinc2 function over the

cross section of the beam is 0.23, 0.25, and 0.29 for Figs 6.5(c), 6.5(f), and 6.5(i),

Page 111: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

93

respectively. At high powers, the tunability of a BiBO-based OPO will inevitably be

limited as the phase matching becomes unsuitable for broadband operation.

6.5 Power Limitations from Thermal Lensing

In addition to thermally induced phase-matching difficulties, the temperature

gradient of the crystal along the cross section of the beam induces thermal lensing. It is

well known that these types of thermal effects can create a thermal lens through a gradient

index, resulting in spherical aberration of the wavefront, index variation through the

photoelastic effect, and mechanical stress that results in the bulging of the crystal’s end

faces [181-183]. For this investigation, we focus only on the lensing contribution of the

thermally induced gradient index, which was shown to be the most dominant factor for

thermal lensing present in the cavity.

The impact thermal lensing creates two separate problems that interfere with the

gain of the cavity. In this cavity, introducing a thermal lens changes the beam size at the

crystal as well as the wavefront curvature of the beam. A change in the beam radius creates

an imperfect overlap between the pump and signal beams. This can decrease the gain when

the cavity mode (signal beam) gets larger or smaller. In the case where the signal beam is

larger than the pump, the signal beam will have a higher intensity near the center, even for

non-saturated amplification. This will result in either the propagation of higher-order

modes within the cavity (and therefore loss of beam quality) or the loss of efficiency

because of the amplification of higher-order modes that are not resonated in the cavity

[169, 170]. In the case where the signal beam is smaller than the pump, the signal beam is

unable to utilize the total available pump energy, resulting in a lower efficiency than when

the modes are perfectly matched [60].

The wavefront also creates a reduction in the available gain, which is determined

by the angular acceptance of the crystal. When the beam focus shifts away from the crystal,

the wavefront’s radius of curvature (ROC) begins to decrease. As this occurs, the

Page 112: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

94

components of the beam’s spatial frequency distribution that exceed the angular acceptance

of the system will not be adequately phase matched, reducing the system’s efficiency [184].

Normally, both factors are negligible by ensuring the pump and signal beams are

of equal size and at focus inside the crystal. With the introduction of the thermal lens, both

factors can negatively affect the available gain, thereby reducing the system’s efficiency,

which contributes to the falloff at higher powers, as shown in Fig. 5.5. These quantities are

difficult to quantify precisely without extensive modeling of the pump depletion (temporal

and spatial), changes in the cavity mode, and phase-matching errors. Since this extensive

characterization was outside the scope of this investigation, a simpler model was developed

that assumed the thermal gradients, measured at the face of the crystal, were constant along

the length of the crystal. This model was tested by comparing predictions of the cavity

mode size to the measurements made as the power was varied.

The spatially varying refractive index across the beam was calculated from the

thermal gradient measurements using temperature-dependent Sellmeier coefficients for

BiBO [142]. The second-order coefficient of a fourth-order polynomial fit was used to

calculate the refractive power of the thermal lens using [56]

𝑓 ≈ 1/𝑛max𝑘𝑑, (6.1)

where d is the length of the crystal, nmax is the highest component of the index of the crystal

along the polynomial fit, and k is defined by the paraxial approximation of the index:

𝑛 = 𝑛max(1 −𝑘𝑟2

2). (6.2)

In this approximation, higher orders of the polynomial were ignored. The calculation was

performed separately for both the x and y axes of the beam. The impact of the thermal lens

on the beam size and ROC at the crystal was calculated using an ABCD model [177].

Measurements were made to verify the model’s predictions over a range of SWIR

powers. First, the cavity was reconfigured by swapping the positions of M3 and the OC

Page 113: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

95

shown in Fig. 4.7. This made it possible to measure the output beam at a location that was

predicted to be more sensitive to thermal lensing. Figure 6.6(a) compares direct

measurements of the output-beam size (solid lines for x and y axes) and the predictions

from the ABCD model that were derived from thermal measurements (dots). The

agreement suggests that, despite the simplicity of the model, an analysis based on thermal

images can be used to understand the impact of the thermal lens on the cavity.

Figure 6.6: (a) The measured radius (solid line) of the output beam compared against the

SWIR power on the crystal in the alternative cavity configuration for the x and y (blue and

red, respectively) dimensions of the beam. The predicted beam size from the measured

thermal gradient is also shown (dots). (b) The calculated beam size and (c) wavefront ROC

at the nonlinear crystal over the full range of incident power for the x and y dimensions of

the beam (blue and red), respectively [180].

With the ABCD model, thermal images were used to predict the cavity mode size

and the ROC at the crystal as a function of the SWIR power, as shown in Figs. 6.6(b) and

6.6(c), respectively. Figure 6.6(b) shows that the beam size increases as the power of the

cavity rises. The thermal gradient is higher along the y direction, in part because the thermal

conductivity is approximately 60% that of the x direction. The thermal conductivity was

calculated to be 18.4 Wm–1K–1 and 11.8 Wm–1K–1 for the x and y directions, respectively

[164], leading to a stronger thermal lens that produces astigmatism and a non-Gaussian mode

profile at high powers. The ROC, as shown in Fig. 6.6(c), also changes significantly as the

SWIR power is increased. This occurs because the cavity design is not perfectly symmetric

about the crystal; therefore, the introduction of a lensing element causes the position of cavity

beam waist to shift.

Page 114: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

96

Although the loss gain was not quantified for this investigation, there is a

remarkable connection between thermally induced cavity dynamics calculated from the

thermal effects and the loss of efficiency. Figure shows this correlation directly by

comparing the OPO power and calculated spot size against the pump power. Near 25 W of

pump power, the calculated cavity mode begins to change rapidly, occurring at the same

pump power at which the slope efficiency falls off. Independent measurements determined

the spot size and output power, yet the strong correlation between them suggests that this

simple technique of imaging the thermal properties of the crystal can be a strong tool for

predicting the cause of gain reduction in parametric devices.

Figure 6.7: Correlation between the signal power (black) and calculated beam-spot size at

the crystal for different pump powers. The dashed line shows the slope efficiency of 14%

before thermal effects begin to take effect at 30 W of pump power [180].

Page 115: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

97

7 Outlook and Conclusions

7.1 Summary

A cw-pumped Yb:YAG thin-disk laser constructed as a scalable pump source for

an ultrafast OPO was described in Chap. 3. The basic principles of operation were

explained and the process of scaling this technology to higher pulse intensities was

explored. The laser could produce 45 W of average power with pulse durations of 1.0-ps

and pulse energies exceeding 6 µJ. The suitability of this laser as a pump source for a

parametric device was investigated, and the beam quality and stability were shown to be

suitable for the goals of this investigation.

Chapter 4 presented the design of an ultrafast OPO utilizing a high-average-power

thin-disk pump laser. Multiple crystals were investigated as a gain medium for providing

scalable ultrafast pulses in the IR. The reasons for choosing BiBO can be summarized as

an optimization of available gain, technological maturity and transparency range.

Unanticipated absorption near the operational wavelengths of the OPO were observed with

new transparency measurements of commercially available BiBO crystals. The use of a

noncollinear interaction geometry was explained, and optimization of AR coatings for an

oscillator with high average power was explored. Multiple designs of the OPO cavity were

discussed and alignment strategies for extended cavities were considered. An ABCD

analysis showed that the folded cavity design could be extended to longer lengths using

telescopic image relays, ultimately achieving high pulse energies.

Chapter 5 described the characterization and results of the OPO. The pulse duration

and measurements of the spectrum showed that the output pulses were heavily chirped. It

was shown that the pulse duration was nearly constant for different output powers. The

temporal stability of the free-running OPO was investigated and showed that the noise

Page 116: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

98

present in the pump laser was amplified in the OPO. Methods to improve the temporal

stability with automated electronics were discussed. The efficiency and scalability of the

OPO were characterized, showing that there was room for improvement in the efficiency

compared to other BiBO-based OPO’s and that higher powers could have been achieved if

thermal limitations could had been overcome. The beam quality was also measured and, as

expected, was near diffraction limited.

The limitations to the output power were investigated by characterizing the thermal

properties of the nonlinear crystal in Chap. 6. Although exceptionally high average powers

were obtained with the OPO, limitations of the system used to actively cool the crystal

were explored. This was done by mapping the 2-D change in the refractive index as a

function of SWIR power through thermal images of the nonlinear crystal. The impact of

the thermal gradient on both the phase-matching properties at different wavelengths and

the induced thermal lens was explored. Uniform phase matching over the cross section of

the beam was severely limited at wavelengths away from degeneracy at higher powers.

Thermal-lensing effects were shown to shift the beam waist away from the nonlinear

crystal. A strong correlation between changes in the cavity dynamics and the decrease in

efficiency was observed. Although the impact of the thermal gradient on the gain of the

OPO was not quantified, the method employed showed a meaningful connection between

the thermal properties and the falloff in efficiency near the maximum-available pump

power in the OPO.

7.2 Outlook

This section, a brief outlook for potential improvements to the OPO constructed for

this investigation and recent related technological advancements. Some of the most

promising areas for improvement will be discussed, including optimization of the signal

pulse’s duration, improvements to the pump laser’s power/energy, active jitter stabilization

techniques, thermal management for the OPO crystal, and enhancements to the OPO cavity

design.

Page 117: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

99

7.2.1 OPO Pulse Duration

One of the most immediate places for improvement with the current system is the

optimization of pulse duration. External pulse compression is a commonly employed

technique to minimize the duration when the pulse coming out of the cavity is chirped. In

this study, dispersion was designed to be nearly uniform across the tunability of the OPO

but not optimized for near-zero dispersion for the output pulses. External pulse

compression was not employed due to the large tuning bandwidth of the OPO and the

additional costs from broadband pulse compression near 2 µm. From the pulse

measurements shown in Chap. 5, the TBP was measured to be 0.928 (from a minimum of

0.315 for a sech2 pulse) suggesting that the duration of the output pulses could be reduced

to ~170 fs; however, the minimum-achievable pulse duration is not clear. Higher-order

dispersion effects could easily increase the pulse length, and since no measurement of the

spectral phase was made it is difficult to predict the shortest pulse duration that could be

obtained. Improving the pump source is another strategy that could be used to reduce the

pulse duration.

7.2.2 Pump-Laser Power and Energy

There are several ways the pump laser could be upgraded as a pump source for an

ultrafast OPO. Direct improvements to the pulse energy have already been obtained from

a thin-disk laser with a similar design and identical core components. The pulse energy

could be improved by increasing the number of passes in the cavity. An active multipass

cavity using 11 passes (instead of the 6 used here) would nearly double the available pulse

energy without sacrificing the average power of both the pump and the signal [98].

Additional improvements to Yb:YAG thin-disk lasers have been made, including removing

astigmatism with a unique multipass cavity geometry, using KLM to initiate mode-locking

instead of a SESAM for shorter pulses and increasing cw-pump power to achieve higher

average powers for ultrafast pulses [38, 97].

Other thin-disk materials have been investigated to improve the bandwidth

available from a Yb:YAG. Recently, Yb:Lu2O3 was used to achieve pulse durations as

Page 118: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

100

short as 142 fs with an average power of 7 W [185], Yb:LuScO3 achieved sub-100-fs pulses

[186], and Yb:CALGO reached 62 fs with a sub 10 W average powers [187]. Although

the average power of Yb:YAG is still significantly higher, it is reasonable to expect that

the power scalability of these devices will lead to pulses with several tens of microjoules

and hundreds of watts over average power in the near future.

7.2.3 Timing-Jitter Stabilization

The temporal stability and timing jitter could be improved in both the OPO and

TDML. In the TDML the method of improvement is straightforward. Of the options

reviewed, the best solution is to lock the TDML to an external crystal rf signal generator.

A commercially available solution was found through TEM Messtechnik but was not

employed in this test-bed study. The repetition rate of the most-stable multipass

configuration was 7.08 MHz. The TDML could be locked to a crystal oscillator generating

a frequency of 212.5 MHz (30th harmonic of the TDML’s repetition rate). This was an

affordable choice that also reduced the requirements of the locking circuit.

The circuit would control three types of motion to compensate for rapid changes in

the cavity and long-term drift. High-frequency drift could be compensated for with a short

stroke (2-µm), large-bandwidth lead zirconate titanate (PZT) mount attached to the optical

mount of one of the end mirrors. Low-frequency noise could be similarly compensated for

using a PZT with a longer stroke (12-µm). For long term drift, one of the end mirrors could

be mounted to a linear micrometer if the PZTs could not compensate for the change in

cavity length, which was expected due to the hundreds of microns in drift observed during

operation. The three-axis motion control would significantly reduce the phase noise for

nearly all frequencies below 15 kHz, greatly reducing the temporal jitter of the laser.

Due to the potential of misalignment from these controls a slave cavity would be

required to prevent damage to the thin disk. A slave cavity can be placed inside the TDML

by using a variable output coupler in a simple linear cavity arranged so that the loss for that

cavity would be higher than the TDML but would lase if alignment problems from the

Page 119: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

101

main cavity created significant losses. Exceptional care would be required when installing

the timing stability system to avoid the introduction of power fluctuations from minor

alignment changes, which would not be monitored by the locking circuit.

Improving the temporal jitter for the OPO is slightly more complicated. Since the

wavelength of an OPO will change with the cavity length within a certain tolerance, locking

the OPO to an external frequency source is inadequate to ensure stable operation. An

alternative cavity stabilization approach based on wavelength fluctuations was first

demonstrated in a synchronously pumped dye laser [188]. Two photodiodes were used to

monitor the intensity on both sides of a spectrally dispersed beam, and their voltage

difference was used to adjust the cavity length in a simple electronic circuit. Although

limitations of this technology due to errors in finely structured spectra were present, it was

successfully implemented for femtosecond OPO’s to reduce temporal jitter [189].

This method was improved using of a one-dimensional position-sensitive detector

that allowed continuous monitoring of the OPO wavelength, independent of the incident

power level and shape of the pulse spectrum [190]. This technique has been successfully

utilized in ultrafast OPO’s, greatly improving their temporal stability [175]. The

implantation of this technology would be relatively straightforward in a 2-µm OPO,

although careful choice of a detector would be required or a portion of the output would

have to be frequency doubled to use as a signal for the control circuit.

An additional improvement could be a redesign of the OPO’s enclosure. The OPO

was exposed to lab environment without a full enclosure or environmental controls.

Implementing independent temperature and humidity controls would reduce instabilities

caused by sporadic air flow in the cavity and would result in more-reliable, long-term

operation. Another improvement that would be highly desirable is an upgrade to the

thermal management system.

Page 120: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

102

7.2.4 Thermal Management of Nonlinear Crystals

As discussed in Chap. 6, limitations of the system used to actively cool the crystal

led to a loss in potential power. Multiple strategies could be employed to improve the

thermal management in this system, aside from different choices in the nonlinear crystal or

improvements to the manufacturing process. Devising ways to mitigate the types of thermal

effects discussed in the previous chapter will be of great importance as the power of

ultrafast OPO’s is scaled to the maximum-available power of various pump sources.

One strategy would be to place the beams closer to the interface between the TEC

and the nonlinear crystal. This commonly employed technique ensures that the heat

diffusion occurs very close to the location where heat is generated; however, this is

typically done for beams sizes much smaller than were required for this system. For this

investigation, a beam size of 570 µm was used while most OPO’s use a beam radius of less

than 100 µm inside the crystal. As seen in Fig. 6.4(e) the thermal gradient of the crystal

temperature near the cooling interface changes rapidly at a 1-mm distance. It is quite

possible that placing the cavity mode closer to the cooling surface would lead to an

irregular thermal gradient. It is unclear if this new gradient would cause nonuniform (and

thus suboptimal) phase matching over the beam surface or induce aberrations into the

beam.

Another method to improve the thermal management would be to actively cool

every non-optical surface of the crystal (i.e., surround the crystal). In this study, the TEC

drew heat only from the bottom face of the crystal. Looking again at Fig. 6.4(e), near the

maximum available power, a few details in the thermal distribution show the disadvantages

of drawing heat from just one side. Near the cooling surface the temperature of the crystal

is a few degrees cooler than top of the crystal. Neglecting the local heating near the beam,

the temperature across the crystal face is not uniform. If heat was dissipated from every

non-optical surface, it would likely reduce the thermal gradient across the beam. It would

be possible to create a metal housing with indium foil on the four sides of the crystal,

similar to the device used in this study. Additionally, a TEC could be attached to each side

Page 121: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

103

of the housing, which would improve heat extraction in four directions. Although this

would improve the bulk cooling of the nonlinear crystal, it is uncertain how much it would

reduce the effect of local heating. Since this would be relatively easy to implement, it would

be worth investigating.

Other cooling strategies have recently been investigated for other parametric

devices that could possibly be used in this system to achieve higher powers. One promising

technique was the atomic bonding of a material with high thermal conductivity to the

optical surfaces of the nonlinear crystal. This technique was recently employed in an

optical parametric amplifier using a BBO crystal bonded to actively cooled sapphire plates

[191, 192]. Modifications to this approach would be necessary for it to be used in the BiBO

based OPO. To prevent Fresnel reflections between the cooling surface and the nonlinear

crystal, the indices of the two materials must be nearly identical (which is the case for BBO

and sapphire). An alternative material would have to be identified for this technique to be

extended to other nonlinear crystals, especially BiBO where the index was significantly

higher, in this orientation, than BBO.

7.2.5 Cavity Design

Other strategies could be employed that mitigate the result of the thermal effects

rather than prevent them. The calculated effects of the thermal gradient give insight into

the best methods to compensate for thermal lensing and phase-matching errors. Thermal

lensing had a large effect on imperfect pump/signal mode overlap and changes to the ROC

of the signal beam, leading to suboptimal gain. Phase-matching limitations from the

thermal gradient had a significant effect on wavelengths away from degeneracy, suggesting

that improvements to the width of the phase-matching curve would provide the maximum

benefit to the thermal limitations.

Thermally induced phase-matching errors can be mitigated by decreasing the

crystal length or increasing the noncollinear angle between the pump and signal beams.

The 8.4-mm crystal length chosen for this investigation optimized the gain without limiting

Page 122: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

104

the bandwidth necessary for sub-100-fs pulses. Sufficient gain was available with

significantly shorter crystals (< 1 mm) using beams with smaller radii to achieve high

intensities. Due to the high energies in this system, small beam sizes led to a risk of damage.

The damage threshold of BiBO has not been well studied, however, and was based on

reports from other BiBO-based systems and the crystal vendor. Recent reports suggest that

ultrafast pulse intensities of 20 GW/cm2 [193] are possible in BiBO, suggesting that the

crystal length could be significantly reduced to accommodate broader phase matching

using a smaller beam radius.

The results of the thermal analysis highlight multiple strategies available to reduce

the impact of the thermal lensing. The increase in the ROC of the signal beam and the size

of the beam radius indicate that the beam waist was shifting away from the nonlinear

crystal. For this system, the shift of the beam radius was strictly due to the asymmetry of

the cavity. Since the cavity is longer on one side of the crystal than the other, the

introduction of an optical element with variable power causes a shift in the beam waist.

This can be mitigated in two ways: first, the cavity could be redesigned so that it is perfectly

symmetric around the nonlinear crystal. The cavity changes from the thermal lens would

then have a negligible effect on the position of the beam waist. This would require a careful

selection of the mirrors to achieve a completely symmetric design that could also have a

cavity length equal to the pump. A second option would be the incorporation of a variable

telescope. Previous systems with a large thermal load could implement such a device to

compensate for the thermally induced changes. A well-aligned variable telescope would

allowone to make changes during laser operation that would not interfere with the cavity

alignment and would maintain a constant beam waist in the gain medium. Due to the critical

alignment restrictions of this cavity, the former strategy is preferable to the latter, but either

method could be employed.

Page 123: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

105

7.3 Conclusions

In this study, the construction and characterization of a scalable ultrafast OPO have

been presented. A high-average-power Yb:YAG thin-disk laser was constructed, capable

of producing 1.0-ps pulses at 1030 nm with a repetition rate of 7.08 MHz and average

powers exceeding 45 W. This laser was used to pump a high-energy, extended-cavity

femtosecond BiBO OPO. Using a noncollinear interaction geometry, the OPO could

produce 455-fs pulses at 7.08 MHz from 1.99 to 2.20 µm with energies up to 350 nJ with

a stable output. Even higher energies would be obtainable at the cost of efficiency and

stability, although the potential of even more power was available if thermally induced

limitations were overcome.

Despite the limitations of this system, what was accomplished in this investigation

was still quite impressive. The system described here achieved the highest pulse energy to

date (under stable performance) from any watt-class synchronously pumped ultrafast OPO.

These pulse energies belong to a class of OPO’s pumped by cutting-edge laser sources,

which are scalable to powers and pulse energies well beyond what was used in any of these

investigations. There have been no reports of a BiBO-based OPO with pulse energies

higher than reported here. Additionally, to the best of our knowledge, no femtosecond

oscillator operating beyond 2 µm has achieved energies greater than those reported here.

This research is another example of the potential from scalable ultrafast OPO’s that

have been constructed in recent years. Systems built in the past decade showcase the

potential of such devices, proving high-intensity pulses with high average powers from a

simple system, capable of operating over a wide range of wavelengths not traditionally

obtainable from other oscillators. The performance of these OPO’s come without any

significant change in classical design principles, but by strategic utilization of modern

technology to obtain pulse energies previously unavailable to OPO’s. With improvements

to the pulse energy and average power of pump sources for parametric devices, this trend

will continue to grow and open new doors for applications that can be solved by OPO’s.

Page 124: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

106

While the system described in this work represents several landmarks in

performance, it is certainly work that could only be built on the shoulders of giants. So

much of the effort that has come before has made such a great contribution to the scientific

community that the work described here seems like a drop in the pond. Only after countless

hours toiling with the difficulties of designing and building one of these devices was the

author truly able to appreciate the dedication required to become an OPO expert. I can only

hope that the work presented here will be as important to others as previous work has been

to me, and that the results obtained in this research will aid in increasing the potential of

these devices.

Page 125: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

107

References

1. R. Kingston, "Parametric amplification and oscillation at optical frequencies," Proceedings of the

Institute of Radio Engineers 50, 472-& (1962).

2. S. Akhmanov and R. Khokhlov, "Concerning one possibility of amplification of light waves," Sov.

Phys. JETP 16, 252-257 (1963).

3. J. A. Armstrong, N. Bloembergen, J. Ducuing, and P. S. Pershan, "Interactions between Light

Waves in a Nonlinear Dielectric," Physical Review 127, 1918-1939 (1962).

4. J. Giordmaine and R. C. Miller, "Tunable Coherent Parametric Oscillation in LiNb O 3 at Optical

Frequencies," Physical Review Letters 14, 973 (1965).

5. R. Byer, M. Oshman, J. Young, and S. Harris, "Visible CW parametric oscillator," Applied

Physics Letters 13, 109-111 (1968).

6. R. Smith, J. Geusic, H. Levinstein, J. Rubin, S. Singh, and L. Van Uitert, "Continuous optical

parametric oscillation in Ba2NaNb5O15," Applied Physics Letters 12, 308-310 (1968).

7. Y. R. Shen, The Principles of Nonlinear Optics (Wiley, 1984).

8. K. Burneika, M. Ignatavicius, V. Kabelka, A. Piskarskas, and A. Stabinis, "Parametric light

amplification and oscillation in KDP with mode-locked pump," IEEE Journal of Quantum

Electronics 8, 574-574 (1972).

9. M. Becker, D. Kuizenga, D. Phillion, and A. Siegman, "Analytic expressions for ultrashort pulse

generation in mode‐locked optical parametric oscillators," Journal of Applied Physics 45, 3996-

4005 (1974).

10. E. Cheung and J. Liu, "Theory of a synchronously pumped optical parametric oscillator in steady-

state operation," JOSA B 7, 1385-1401 (1990).

11. E. Cheung and J. Liu, "Efficient generation of ultrashort, wavelength-tunable infrared pulses,"

JOSA B 8, 1491-1506 (1991).

12. C. Fallnich, B. Ruffing, T. Herrmann, A. Nebel, R. Beigang, and R. Wallenstein, "Experimental

investigation and numerical simulation of the influence of resonator-length detuning on the output

power, pulse duration and spectral width of a cw mode-locked picosecond optical parametric

oscillator," Applied Physics B: Lasers and Optics 60, 427-436 (1995).

13. R. Laenen and T. Roth, "Numerical study and experimental results for a synchronously pumped

optical parametric oscillator in the saturation regime," JOSA B 14, 454-459 (1997).

Page 126: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

108

14. G. Gale, M. Cavallari, and F. Hache, "Femtosecond visible optical parametric oscillator," JOSA B

15, 702-714 (1998).

15. D. Edelstein, E. Wachman, and C. Tang, "Broadly tunable high repetition rate femtosecond optical

parametric oscillator," Applied Physics Letters 54, 1728-1730 (1989).

16. Q. Fu, G. Mak, and H. Van Driel, "High-power, 62-fs infrared optical parametric oscillator

synchronously pumped by a 76-MHz Ti: sapphire laser," Optics letters 17, 1006-1008 (1992).

17. J. Dudley, D. Reid, M. Ebrahimzadeh, and W. Sibbett, "Characteristics of a noncritically

phasematched Ti: sapphire pumped femtosecond optical parametric oscillator," Optics

communications 104, 419-430 (1994).

18. G. Gale, M. Cavallari, T. Driscoll, and F. Hache, "Sub-20-fs tunable pulses in the visible from an

82-MHz optical parametric oscillator," Optics letters 20, 1562-1564 (1995).

19. K. C. Burr, C. Tang, M. A. Arbore, and M. M. Fejer, "Broadly tunable mid-infrared femtosecond

optical parametric oscillator using all-solid-state-pumped periodically poled lithium niobate,"

Optics letters 22, 1458-1460 (1997).

20. C. McGowan, D. Reid, Z. Penman, M. Ebrahimzadeh, W. Sibbett, and D. Jundt, "Femtosecond

optical parametric oscillator based on periodically poled lithium niobate," JOSA B 15, 694-701

(1998).

21. D. T. Reid, J. M. Dudley, M. Ebrahimzadeh, and W. Sibbett, "Ti: Sapphire Pumped Femtosecond

Optical Parametric Oscillator Exhibiting Soliton Formation," Journal of Modern Optics 41, 1231-

1242 (1994).

22. D. Reid, J. Dudley, M. Ebrahimzadeh, and W. Sibbett, "Soliton formation in a femtosecond

optical parametric oscillator," Optics letters 19, 825-827 (1994).

23. K. Puech, L. Lefort, and D. Hanna, "Broad tuning around degeneracy in a singly resonant

synchronously pumped parametric oscillator by means of a diffraction grating," JOSA B 16, 1533-

1538 (1999).

24. M. V. O’Connor, M. A. Watson, D. P. Shepherd, and D. C. Hanna, "Use of a birefringent filter for

tuning a synchronously pumped optical parametric oscillator," Applied Physics B: Lasers and

Optics 79, 15-23 (2004).

25. S. Marzenell, R. Beigang, and R. Wallenstein, "Synchronously pumped femtosecond optical

parametric oscillator based on AgGaSe2 tunable from 2 μm to 8 μm," Applied Physics B: Lasers

and Optics 69, 423-428 (1999).

26. T. Lang, T. Binhammer, S. Rausch, G. Palmer, M. Emons, M. Schultze, A. Harth, and U.

Morgner, "High power ultra-widely tuneable femtosecond pulses from a non-collinear optical

parametric oscillator (NOPO)," Optics express 20, 912-917 (2012).

27. M. Ghotbi, A. Esteban-Martin, and M. Ebrahim-Zadeh, "Tunable, high-repetition-rate,

femtosecond pulse generation in the ultraviolet," Optics letters 33, 345-347 (2008).

Page 127: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

109

28. A. Esteban-Martin, V. Ramaiah-Badarla, V. Petrov, and M. Ebrahim-Zadeh, "Broadband, rapidly

tunable Ti: sapphire-pumped BiB 3 O 6 femtosecond optical parametric oscillator," Optics letters

36, 1671-1673 (2011).

29. C. McGowan, D. Reid, M. Ebrahimzadeh, and W. Sibbett, "Femtosecond pulses tunable beyond 4

μm from a KTA-based optical parametric oscillator," Optics communications 134, 186-190

(1997).

30. S. McCahon, S. Anson, D.-J. Jang, and T. F. Boggess, "Generation of 3–4-μm femtosecond pulses

from a synchronously pumped, critically phase-matched KTiOPO 4 optical parametric oscillator,"

Optics letters 20, 2309-2311 (1995).

31. C. Cleff, J. Epping, P. Gross, and C. Fallnich, "Femtosecond OPO based on LBO pumped

by a frequency-doubled Yb-fiber laser-amplifier system for CARS spectroscopy," Applied Physics

B 103, 795-800 (2011).

32. C. W. Rudy, A. Marandi, K. A. Ingold, S. J. Wolf, K. L. Vodopyanov, R. L. Byer, L. Yang, P.

Wan, and J. Liu, "Sub-50 fs pulses around 2070 nm from a synchronously-pumped, degenerate

OPO," Optics Express 20, 27589-27595 (2012).

33. D. Reid, M. Ebrahimzadeh, and W. Sibbett, "Efficient femtosecond pulse generation in the visible

in a frequency-doubled optical parametric oscillator based on RbTiOAsO 4," JOSA B 12, 1157-

1163 (1995).

34. M. Jankowski, A. Marandi, C. Phillips, R. Hamerly, K. Ingold, R. Byer, and M. Fejer,

"Femtosecond Temporal Simulton Formation in Optical Parametric Oscillators," in Nonlinear

Optics, (Optical Society of America, 2017), NTu1B. 3.

35. , retrieved https://www.ape-berlin.de/sites/default/files/OPO%20X.pdf.

36. P. Russbueldt, T. Mans, J. Weitenberg, H. Hoffmann, and R. Poprawe, "Compact diode-pumped

1.1 kW Yb: YAG Innoslab femtosecond amplifier," Optics letters 35, 4169-4171 (2010).

37. T. Eidam, S. Hanf, E. Seise, T. V. Andersen, T. Gabler, C. Wirth, T. Schreiber, J. Limpert, and A.

Tünnermann, "Femtosecond fiber CPA system emitting 830 W average output power," Optics

letters 35, 94-96 (2010).

38. D. Bauer, I. Zawischa, D. H. Sutter, A. Killi, and T. Dekorsy, "Mode-locked Yb: YAG thin-disk

oscillator with 41 µJ pulse energy at 145 W average infrared power and high power frequency

conversion," Optics express 20, 9698-9704 (2012).

39. J. Brons, V. Pervak, E. Fedulova, D. Bauer, D. Sutter, V. Kalashnikov, A. Apolonskiy, O. Pronin,

and F. Krausz, "Energy scaling of Kerr-lens mode-locked thin-disk oscillators," Optics letters 39,

6442-6445 (2014).

40. O. H. Heckl, J. Kleinbauer, D. Bauer, S. Weiler, T. Metzger, and D. H. Sutter, "Ultrafast Thin-

Disk Lasers," in Ultrashort Pulse Laser Technology (Springer International Publishing, 2016), pp.

93-115.

41. T. Lamour, L. Kornaszewski, J. Sun, and D. Reid, "Yb: fiber-laser-pumped high-energy

picosecond optical parametric oscillator," Optics express 17, 14229-14234 (2009).

Page 128: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

110

42. T. P. Lamour and D. T. Reid, "650-nJ pulses from a cavity-dumped Yb: fiber-pumped ultrafast

optical parametric oscillator," Optics express 19, 17557-17562 (2011).

43. J. Tate, T. Auguste, H. Muller, P. Salieres, P. Agostini, and L. DiMauro, "Scaling of wave-packet

dynamics in an intense midinfrared field," Physical review letters 98, 013901 (2007).

44. S. Hädrich, J. Rothhardt, M. Krebs, S. Demmler, A. Klenke, A. Tünnermann, and J. Limpert,

"Single-pass high harmonic generation at high repetition rate and photon flux," Journal of Physics

B: Atomic, Molecular and Optical Physics 49, 172002 (2016).

45. T. Südmeyer, S. Marchese, S. Hashimoto, C. Baer, G. Gingras, B. Witzel, and U. Keller,

"Femtosecond laser oscillators for high-field science," Nature Photonics 2, 599-604 (2008).

46. X. Zhou, P. Ranitovic, C. Hogle, J. Eland, H. Kapteyn, and M. Murnane, "Probing and controlling

non-Born-Oppenheimer dynamics in highly excited molecular ions," Nature Physics 8, 232-237

(2012).

47. C. B. Schaffer, A. Brodeur, J. F. García, and E. Mazur, "Micromachining bulk glass by use of

femtosecond laser pulses with nanojoule energy," Optics letters 26, 93-95 (2001).

48. T. M. Taczak and D. K. Killinger, "Development of a tunable, narrow-linewidth, cw 2.066-μm

Ho: YLF laser for remote sensing of atmospheric CO 2 and H 2 O," Applied optics 37, 8460-8476

(1998).

49. K. Scholle, E. Heumann, and G. Huber, "Single mode Tm and Tm, Ho: LuAG lasers for LIDAR

applications," Laser Physics Letters 1, 285 (2004).

50. K. Scholle, S. Lamrini, P. Koopmann, and P. Fuhrberg, "2 µm laser sources and their possible

applications," (2010).

51. R. W. Boyd, Nonlinear optics (Academic press, 2003).

52. J. Feinberg, "Self-pumped, continuous-wave phase conjugator using internal reflection," Optics

letters 7, 486-488 (1982).

53. R. Boyd, Contemporary nonlinear optics (Academic Press, 2012).

54. N. Kukhtarev, V. Markov, and S. Odulov, "Transient energy transfer during hologram formation

in LiNbO3 in external electric field," Optics Communications 23, 338-343 (1977).

55. V. G. Dmitriev, G. G. Gurzadyan, and D. N. Nikogosyan, Handbook of nonlinear optical crystals

(Springer, 2013), Vol. 64.

56. N. Boeuf, D. Branning, I. Chaperot, E. Dauler, S. Gue, G. Jaeger, A. Muller, and A. Migdall,

"Calculating characteristics of noncollinear phase matching in uniaxial and biaxial crystals,"

Optical Engineering 39, 1016-1024 (2000).

57. W. Koechner, Solid-state laser engineering (Springer, 2013), Vol. 1.

58. S. A. Akhmanov and R. Khokhlov, "New developments in non-linear optics," Physics-Uspekhi 11,

394-402 (1968).

Page 129: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

111

59. S. A. Akhmanov, V. A. Vysloukh, and A. S. Chirkin, "Optics of femtosecond laser pulses,"

Moscow Izdatel Nauka (1988).

60. G. Cerullo and S. De Silvestri, "Ultrafast optical parametric amplifiers," Review of Scientific

Instruments 74, 1 (2003).

61. M. M. Fejer, G. Magel, D. H. Jundt, and R. L. Byer, "Quasi-phase-matched second harmonic

generation: tuning and tolerances," IEEE Journal of Quantum Electronics 28, 2631-2654 (1992).

62. E. Lim, M. Fejer, and R. Byer, "Second-harmonic generation of green light in periodically poled

planar lithium niobate waveguide," Electronics Letters 25, 174-175 (1989).

63. A. Scavennec, "Mismatch effects in synchronous pumping of the continuously operated mode-

locked dye laser," Optics Communications 17, 14-17 (1976).

64. J. Sun, B. J. Gale, and D. T. Reid, "Dual-color operation of a femtosecond optical parametric

oscillator exhibiting stable relative carrier-envelope phase-slip frequencies," Optics letters 31,

2021-2023 (2006).

65. M. Rodwell, T. Baer, B. Kolner, K. Weingarten, and D. Bloom, "Reduction of timing fluctuations

in a mode-locked Nd: YAG laser by electronic feedback," Optics letters 11, 638-640 (1986).

66. M. J. Rodwell, D. M. Bloom, and K. J. Weingarten, "Subpicosecond laser timing stabilization,"

IEEE Journal of Quantum Electronics 25, 817-827 (1989).

67. G. Kennedy, R. Grant, W. E. Sleat, and W. Sibbett, "Amplitude noise reduction of a coupled-

cavity mode-locked NaCl: OH− color-center laser by active synchronization," Optics letters 18,

208-210 (1993).

68. P. Beaud, W. Hodel, B. Zysset, and H. Weber, "Ultrashort pulse propagation, pulse breakup, and

fundamental soliton formation in a single-mode optical fiber," IEEE journal of quantum

electronics 23, 1938-1946 (1987).

69. Sellmeier, "Zur Erklärung der abnormen Farbenfolge im Spectrum einiger Substanzen," Annalen

der Physik 219, 272-282 (1871).

70. I. Malitson, "Interspecimen comparison of the refractive index of fused silica*," Josa 55, 1205-

1209 (1965).

71. R. Fork, O. Martinez, and J. Gordon, "Negative dispersion using pairs of prisms," Optics letters 9,

150-152 (1984).

72. R. L. Fork, C. B. Cruz, P. Becker, and C. V. Shank, "Compression of optical pulses to six

femtoseconds by using cubic phase compensation," Optics letters 12, 483-485 (1987).

73. E. Treacy, "Compression of picosecond light pulses," Physics Letters A 28, 34-35 (1968).

74. E. Treacy, "Optical pulse compression with diffraction gratings," IEEE Journal of quantum

Electronics 5, 454-458 (1969).

75. G. Steinmeyer, "Femtosecond dispersion compensation with multilayer coatings: toward the

optical octave," Applied optics 45, 1484-1490 (2006).

Page 130: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

112

76. R. Szipöcs, C. Spielmann, F. Krausz, and K. Ferencz, "Chirped multilayer coatings for broadband

dispersion control in femtosecond lasers," Optics letters 19, 201-203 (1994).

77. A. Tikhonravov and M. Trubetskov, "Modern status and prospects of the development of methods

of designing multilayer optical coatings," Journal of Optical Technology 74, 845-850 (2007).

78. F. Gires and P. Tournois, "Interferometre utilisable pour la compression dimpulsions lumineuses

modulees en frequence," COMPTES RENDUS HEBDOMADAIRES DES SEANCES DE L

ACADEMIE DES SCIENCES 258, 6112-& (1964).

79. I. Sorokina, A. Cassanho, H. Jenssen, R. Szipöcs, E. Sorokin, and E. Wintner, "Prismless

passively mode-locked femtosecond Cr: LiSGaF laser," Optics letters 21, 1165-1167 (1996).

80. R. Stolen and A. Ashkin, "Optical Kerr effect in glass waveguide," Applied Physics Letters 22,

294-296 (1973).

81. J.-C. M. Diels, J. J. Fontaine, I. C. McMichael, and F. Simoni, "Control and measurement of

ultrashort pulse shapes (in amplitude and phase) with femtosecond accuracy," Applied Optics 24,

1270-1282 (1985).

82. D. J. Kane and R. Trebino, "Characterization of arbitrary femtosecond pulses using frequency-

resolved optical gating," IEEE Journal of Quantum Electronics 29, 571-579 (1993).

83. V. Wong and I. A. Walmsley, "Analysis of ultrashort pulse-shape measurement using linear

interferometers," Optics letters 19, 287-289 (1994).

84. C. Dorrer and I. A. Walmsley, "Accuracy criterion for ultrashort pulse characterization techniques:

application to spectral phase interferometry for direct electric field reconstruction," JOSA B 19,

1019-1029 (2002).

85. A. Giesen, H. Hügel, A. Voss, K. Wittig, U. Brauch, and H. Opower, "Scalable concept for diode-

pumped high-power solid-state lasers," Applied Physics B: Lasers and Optics 58, 365-372 (1994).

86. C. J. Saraceno, F. Emaury, C. Schriber, A. Diebold, M. Hoffmann, M. Golling, T. Südmeyer, and

U. Keller, "Toward millijoule-level high-power ultrafast thin-disk oscillators," IEEE Journal of

Selected Topics in Quantum Electronics 21, 106-123 (2015).

87. J.-P. Negel, A. Voss, M. A. Ahmed, D. Bauer, D. Sutter, A. Killi, and T. Graf, "1.1 kW average

output power from a thin-disk multipass amplifier for ultrashort laser pulses," Optics letters 38,

5442-5445 (2013).

88. C. J. Saraceno, F. Emaury, O. H. Heckl, C. R. Baer, M. Hoffmann, C. Schriber, M. Golling, T.

Südmeyer, and U. Keller, "275 W average output power from a femtosecond thin disk oscillator

operated in a vacuum environment," Optics express 20, 23535-23541 (2012).

89. C. Teisset, M. Schultze, R. Bessing, M. Haefner, S. Prinz, D. Sutter, and T. Metzger, "300 W

picosecond thin-disk regenerative amplifier at 10 kHz repetition rate," in Mid-Infrared Coherent

Sources, (Optical Society of America, 2013), JTh5A. 1.

90. C. J. Saraceno, F. Emaury, C. Schriber, M. Hoffmann, M. Golling, T. Südmeyer, and U. Keller,

"Ultrafast thin-disk laser with 80 μJ pulse energy and 242 W of average power," Optics letters 39,

9-12 (2014).

Page 131: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

113

91. P. Lacovara, H. Choi, C. Wang, R. Aggarwal, and T. Fan, "Room-temperature diode-pumped Yb:

YAG laser," Optics letters 16, 1089-1091 (1991).

92. S. Ricaud, A. Jaffres, K. Wentsch, A. Suganuma, B. Viana, P. Loiseau, B. Weichelt, M. Abdou-

Ahmed, A. Voss, and T. Graf, "Femtosecond Yb: CaGdAlO 4 thin-disk oscillator," Optics letters

37, 3984-3986 (2012).

93. K. S. Wentsch, L. Zheng, J. Xu, M. A. Ahmed, and T. Graf, "Passively mode-locked Yb 3+: Sc 2

SiO 5 thin-disk laser," Optics letters 37, 4750-4752 (2012).

94. C. J. Saraceno, O. H. Heckl, C. R. Baer, M. Golling, T. Südmeyer, K. Beil, C. Kränkel, K.

Petermann, G. Huber, and U. Keller, "SESAMs for high-power femtosecond modelocking: power

scaling of an Yb: LuScO 3 thin disk laser to 23 W and 235 fs," Optics express 19, 20288-20300

(2011).

95. T. Südmeyer, C. Kränkel, C. R. E. Baer, O. H. Heckl, C. J. Saraceno, M. Golling, R. Peters, K.

Petermann, G. Huber, and U. Keller, "High-power ultrafast thin disk laser oscillators and their

potential for sub-100-femtosecond pulse generation," Applied Physics B 97, 281-295 (2009).

96. K. Beil, C. J. Saraceno, C. Schriber, F. Emaury, O. H. Heckl, C. R. Baer, M. Golling, T.

Südmeyer, U. Keller, and C. Kränkel, "Yb-doped mixed sesquioxides for ultrashort pulse

generation in the thin disk laser setup," Applied Physics B 113, 13-18 (2013).

97. J. Neuhaus, D. Bauer, J. Zhang, A. Killi, J. Kleinbauer, M. Kumkar, S. Weiler, M. Guina, D. H.

Sutter, and T. Dekorsy, "Subpicosecond thin-disk laser oscillator with pulse energies of up to 25.9

microjoules by use of an active multipass geometry," Optics express 16, 20530-20539 (2008).

98. J. Neuhaus, J. Kleinbauer, A. Killi, S. Weiler, D. H. Sutter, and T. Dekorsy, "Pulse energies

exceeding 13 microjoules from a passively mode-locked Yb: YAG thin-disk oscillator by use of a

self-imaging active multipass geometry," in Conference on Lasers and Electro-Optics, (Optical

Society of America, 2008), CFB1.

99. A. Killi, K. Wallmeroth, T. Vethake, and V. Kuhn, "Cooling disk lasers," (Google Patents, 2016).

100. A. Giesen and J. Speiser, "Fifteen years of work on thin-disk lasers: results and scaling laws,"

IEEE Journal of Selected Topics in Quantum Electronics 13, 598-609 (2007).

101. J. Neuhaus, "Passively mode-locked Yb: YAG thin-disk laser with active multipass geometry,"

(PhD thesis, University Konstanz, 2009).

102. W. F. Krupke, "Ytterbium solid-state lasers. The first decade," IEEE Journal of Selected Topics in

Quantum Electronics 6, 1287-1296 (2000).

103. S. Reuter, Numerische Modellierung und experimentelle Charakterisierung der physikalischen

Eigenschaften von modengekoppelten Femtosekunden-Yb: YAG-und Yb: KGW-lasern (Cuvillier,

2004).

104. T. Y. Fan, "Heat generation in Nd: YAG and Yb: YAG," IEEE Journal of Quantum Electronics

29, 1457-1459 (1993).

Page 132: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

114

105. V. V. Ter-Mikirtychev and V. A. Fromzel, "Directly single-diode-pumped continuous-wave Yb

3+: YAG laser tunable in the 1047–1051-nm wavelength range," Applied optics 39, 4964-4969

(2000).

106. C. Stewen, K. Contag, M. Larionov, A. Giesen, and H. Hugel, "A 1-kW CW thin disc laser," IEEE

Journal of selected topics in quantum electronics 6, 650-657 (2000).

107. R. C. Powell, Physics of solid-state laser materials (Springer Science & Business Media, 1998),

Vol. 1.

108. O. Antipov, D. Bredikhin, O. Eremeykin, A. Savikin, E. Ivakin, and A. Sukhadolau, "Electronic

mechanism for refractive-index changes in intensively pumped Yb: YAG laser crystals," Optics

letters 31, 763-765 (2006).

109. S. Chénais, F. Balembois, F. Druon, G. Lucas-Leclin, and P. Georges, "Thermal Lensing in

Diode-Pumped Ytterbium Lasers-Part II: evaluation of quantum efficiencies and thermo-optic

coefficients," IEEE journal of quantum electronics 40, 1235-1243 (2004).

110. A. M. Scott, G. Cook, and A. P. Davies, "Efficient high-gain laser amplification from a low-gain

amplifier by use of self-imaging multipass geometry," Applied optics 40, 2461-2467 (2001).

111. M. Kumkar, "Laser amplification system," US Pat 6765947(2006).

112. T. Petersen and J. Bromage, "Intracavity Chirped-Pulse Amplification for High-Energy Ultrafast

Optical Parametric Oscillators," in Frontiers in Optics, (Optical Society of America, 2012),

FM4G. 3.

113. H. Kogelnik and T. Li, "Laser beams and resonators," Applied Optics 5, 1550-1567 (1966).

114. P. French, "The generation of ultrashort laser pulses," Reports on Progress in Physics 58, 169

(1995).

115. F. Kartner, I. Jung, and U. Keller, "Soliton mode-locking with saturable absorbers," IEEE Journal

of Selected Topics in Quantum Electronics 2, 540-556 (1996).

116. F. Kärtner and U. Keller, "Stabilization of solitonlike pulses with a slow saturable absorber,"

Optics letters 20, 16-18 (1995).

117. R. Paschotta and U. Keller, "Passive mode locking with slow saturable absorbers," Applied

Physics B: Lasers and Optics 73, 653-662 (2001).

118. O. Martinez, R. Fork, and J. P. Gordon, "Theory of passively mode-locked lasers including self-

phase modulation and group-velocity dispersion," Optics letters 9, 156-158 (1984).

119. U. Keller, K. J. Weingarten, F. X. Kartner, D. Kopf, B. Braun, I. D. Jung, R. Fluck, C. Honninger,

N. Matuschek, and J. A. Der Au, "Semiconductor saturable absorber mirrors (SESAM's) for

femtosecond to nanosecond pulse generation in solid-state lasers," IEEE Journal of selected topics

in QUANTUM ELECTRONICS 2, 435-453 (1996).

120. A. Kost and E. M. Garmire, Nonlinear optics in Semiconductors (Academic press, 1999), Vol. 58.

121. U. Keller, "Recent developments in compact ultrafast lasers," Nature 424, 831-838 (2003).

Page 133: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

115

122. M. Sheik-Bahae, A. A. Said, T.-H. Wei, D. J. Hagan, and E. W. Van Stryland, "Sensitive

measurement of optical nonlinearities using a single beam," IEEE journal of quantum electronics

26, 760-769 (1990).

123. M. Sheik-Bahae, J. Wang, R. DeSalvo, D. Hagan, and E. Van Stryland, "Measurement of

nondegenerate nonlinearities using a two-color Z scan," Optics letters 17, 258-260 (1992).

124. E. Van Stryland, M. Sheik-Bahae, A. Said, and D. Hagan, "Characterization of nonlinear optical

absorption and refraction," Progress in crystal growth and characterization of materials 27, 279-

311 (1993).

125. E. W. Van Stryland and M. Sheik-Bahae, "Z-scan measurements of optical nonlinearities,"

Characterization techniques and tabulations for organic nonlinear materials 18, 655-692 (1998).

126. J. Wang, M. Sheik-Bahae, A. Said, D. J. Hagan, and E. W. Van Stryland, "Time-resolved Z-scan

measurements of optical nonlinearities," JOSA B 11, 1009-1017 (1994).

127. R. Fluck, R. Häring, R. Paschotta, E. Gini, H. Melchior, and U. Keller, "Eyesafe pulsed microchip

laser using semiconductor saturable absorber mirrors," Applied physics letters 72, 3273-3275

(1998).

128. M. Haiml, R. Grange, and U. Keller, "Optical characterization of semiconductor saturable

absorbers," Applied Physics B: Lasers and Optics 79, 331-339 (2004).

129. A. Schlatter, S. C. Zeller, R. Grange, R. Paschotta, and U. Keller, "Pulse-energy dynamics of

passively mode-locked solid-state lasers above the Q-switching threshold," JOSA B 21, 1469-

1478 (2004).

130. U. Keller, D. Miller, G. Boyd, T. Chiu, J. Ferguson, and M. Asom, "Solid-state low-loss

intracavity saturable absorber for Nd: YLF lasers: an antiresonant semiconductor Fabry–Perot

saturable absorber," Optics letters 17, 505-507 (1992).

131. C. Hönninger, R. Paschotta, F. Morier-Genoud, M. Moser, and U. Keller, "Q-switching stability

limits of continuous-wave passive mode locking," JOSA B 16, 46-56 (1999).

132. J. A. Der Au, D. Kopf, F. Morier-Genoud, M. Moser, and U. Keller, "60-fs pulses from a diode-

pumped Nd: glass laser," Optics letters 22, 307-309 (1997).

133. F. Kurtner, J. A. Der Au, and U. Keller, "Mode-locking with slow and fast saturable absorbers-

what's the difference?," IEEE Journal of Selected Topics in Quantum Electronics 4, 159-168

(1998).

134. A. Siegman, "Wow to (Maybe) Measure Laser Beam Quality," (1998).

135. A. E. Siegman, "Defining, measuring, and optimizing laser beam quality," in OE/LASE'93: Optics,

Electro-Optics, & Laser Applications in Science& Engineering, (International Society for Optics

and Photonics, 1993), 2-12.

136. S. Kimura and C. Munakata, "Measurement of a Gaussian laser beam spot size using a boundary

diffraction wave," Applied Optics 27, 84-88 (1988).

Page 134: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

116

137. R. Pantell, "The laser oscillator with an external signal," Proceedings of the IEEE 53, 474-477

(1965).

138. A. Tyazhev, D. Kolker, G. Marchev, V. Badikov, D. Badikov, G. Shevyrdyaeva, V. Panyutin, and

V. Petrov, "Mid-infrared optical parametric oscillation in the wide-bandgap BaGa 4 S 7 nonlinear

crystal," in Lasers and Electro-Optics (CLEO), 2012 Conference on, (IEEE, 2012), 1-2.

139. Y. Peng, X. Wei, W. Wang, and D. Li, "High-power 3.8μm tunable optical parametric oscillator

based on PPMgO:CLN," Optics Communications 283, 4032-4035 (2010).

140. G. Marchev, F. Pirzio, A. Agnesi, G. Reali, V. Petrov, A. Tyazhev, P. G. Schunemann, and K. T.

Zawilski, "1064nm pumped CdSiP2 optical parametric oscillator generating sub-300 ps pulses

near 6.15μm at 1–10 kHz repetition rates," Optics Communications 291, 326-328 (2013).

141. V. Petrov, "Frequency down-conversion of solid-state laser sources to the mid-infrared spectral

range using non-oxide nonlinear crystals," Progress in Quantum Electronics 42, 1-106 (2015).

142. J. Rothhardt, S. Demmler, S. Hädrich, T. Peschel, J. Limpert, and A. Tünnermann, "Thermal

effects in high average power optical parametric amplifiers," Optics letters 38, 763-765 (2013).

143. M. H. Dunn and M. Ebrahimzadeh, "Parametric generation of tunable light from continuous-wave

to femtosecond pulses," Science 286, 1513-1517 (1999).

144. Y.-R. Shen, "The principles of nonlinear optics," New York, Wiley-Interscience, 1984, 575 p.

1(1984).

145. D. Reid, M. Ebrahimzadeh, and W. Sibbett, "Noncritically phase-matched Ti: sapphire-pumped

femtosecond optical parametric oscillator based on RbTiOAsO 4," Optics letters 20, 55-57 (1995).

146. L. R. Marshall and A. Kaz, "Eye-safe output from noncritically phase-matched parametric

oscillators," JOSA B 10, 1730-1736 (1993).

147. L. E. Myers, R. Eckardt, M. Fejer, R. Byer, W. Bosenberg, and J. Pierce, "Quasi-phase-matched

optical parametric oscillators in bulk periodically poled LiNbO 3," JOSA B 12, 2102-2116 (1995).

148. L. Lefort, K. Puech, S. Butterworth, Y. Svirko, and D. Hanna, "Generation of femtosecond pulses

from order-of-magnitude pulse compression in a synchronously pumped optical parametric

oscillator based on periodically poled lithium niobate," Optics letters 24, 28-30 (1999).

149. S. C. Kumar and M. Ebrahim-Zadeh, "High-power, fiber-laser-pumped, picosecond optical

parametric oscillator based on MgO: sPPLT," Optics express 19, 26660-26665 (2011).

150. A. Dubietis, R. Butkus, and A. P. Piskarskas, "Trends in chirped pulse optical parametric

amplification," IEEE Journal of selected topics in quantum electronics 12, 163-172 (2006).

151. J. Rothhardt, S. Demmler, S. Hädrich, J. Limpert, and A. Tünnermann, "Octave-spanning OPCPA

system delivering CEP-stable few-cycle pulses and 22 W of average power at 1 MHz repetition

rate," Optics express 20, 10870-10878 (2012).

152. S. Banerjee, N. Umemura, and K. Kato, "Sellmeier and thermo-optic dispersion formulas for

AgGa (S1-xSex) 2 with application in mid-IR generation," in Lasers and Applications in Science

and Engineering, (International Society for Optics and Photonics, 2006), 610314-610314-610316.

Page 135: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

117

153. E. Cheung, J. Liu, K. Koch, and G. T. Moore, "Measurements of second-order nonlinear optical

coefficients from the spectral brightness of parametric fluorescence," Optics letters 19, 168-170

(1994).

154. D. Armstrong, W. Alford, T. Raymond, and A. Smith, "Absolute measurement of the effective

nonlinearities of KTP and BBO crystals by optical parametric amplification," Applied optics 35,

2032-2040 (1996).

155. D. Eimerl, S. Velsko, L. Davis, and F. Wang, "Progress in nonlinear optical materials for high

power lasers," Progress in Crystal Growth and Characterization of Materials 20, 59-113 (1990).

156. H. Hellwig, J. Liebertz, and L. Bohatý, "Linear optical properties of the monoclinic bismuth

borate BiB3O6," Journal of Applied Physics 88, 240-244 (2000).

157. V. Badikov, D. Badikov, G. Shevyrdyaeva, A. Tyazhev, G. Marchev, V. Panyutin, V. Petrov, and

A. Kwasniewski, "Phase‐matching properties of BaGa4S7 and BaGa4Se7: Wide‐bandgap

nonlinear crystals for the mid‐infrared," physica status solidi (RRL)-Rapid Research Letters 5, 31-

33 (2011).

158. V. Kemlin, P. Brand, B. Boulanger, P. Segonds, P. G. Schunemann, K. T. Zawilski, B. Ménaert,

and J. Debray, "Phase-matching properties and refined Sellmeier equations of the new nonlinear

infrared crystal CdSiP 2," Optics letters 36, 1800-1802 (2011).

159. N. Pylneva, N. Kononova, A. Yurkin, G. Bazarova, and V. Danilov, "Growth and non-linear

optical properties of lithium triborate crystals," Journal of crystal growth 198, 546-550 (1999).

160. L. Cheng, L.-T. Cheng, J. Bierlein, F. Zumsteg, and A. Ballman, "Properties of doped and

undoped crystals of single domain KTiOAsO4," Applied physics letters 62, 346-348 (1993).

161. J. D. Bierlein, "Potassium Titanyl Phosphate (KTP): Properties, Recent Advances and New

Applications," in 1989 Orlando Symposium, (International Society for Optics and Photonics,

1989), 2-12.

162. S. Das, G. C. Bhar, S. Gangopadhyay, and C. Ghosh, "Linear and nonlinear optical properties of

ZnGeP 2 crystal for infrared laser device applications: revisited," Applied optics 42, 4335-4340

(2003).

163. N. Umemura, S. Banerjee, K. Miyata, F. Tanno, and K. Kato, "Thermo-optic dispersion formula

for BiB3O6," in Quantum Electronics and Laser Science Conference, (Optical Society of

America, 2005), JTuC18.

164. R. Riedel, J. Rothhardt, K. Beil, B. Gronloh, A. Klenke, H. Höppner, M. Schulz, U. Teubner, C.

Kränkel, and J. Limpert, "Thermal properties of borate crystals for high power optical parametric

chirped-pulse amplification," Optics express 22, 17607-17619 (2014).

165. T. Petersen, J. Zuegel, and J. Bromage, "High-average-power, 2-μm femtosecond optical

parametric oscillator synchronously pumped by a thin-disk, mode-locked laser," Optics Express

25, 8840-8844 (2017).

166. P. N. Gavryushkin, L. I. Isaenko, A. P. Yelisseyev, V. A. Gets, and O. S. Il’ina, "Growth,

Morphology and Optical Properties of γ-BiB3O6Single Crystals," Crystal Growth & Design 12,

75-78 (2011).

Page 136: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

118

167. P. Wright and B. Cockayne, "The organometallic chemical vapour deposition of ZnS and ZnSe at

atmospheric pressure," Journal of Crystal Growth 59, 148-154 (1982).

168. C. Nabors, S. Yang, T. Day, and R. Byer, "Coherence properties of a doubly resonant monolithic

optical parametric oscillator," JOSA B 7, 815-820 (1990).

169. S. Brosnan and R. Byer, "Optical parametric oscillator threshold and linewidth studies," IEEE

Journal of Quantum Electronics 15, 415-431 (1979).

170. R. Paschotta, "Beam quality deterioration of lasers caused by intracavity beam distortions," Optics

Express 14, 6069-6074 (2006).

171. E. Hecht, "Optics, 4th," International edition, Addison-Wesley, San Francisco 3(2002).

172. N. Bloembergen, "Role of cracks, pores, and absorbing inclusions on laser induced damage

threshold at surfaces of transparent dielectrics," Applied optics 12, 661-664 (1973).

173. J. Dobrowolski, A. Tikhonravov, M. K. Trubetskov, B. T. Sullivan, and P. Verly, "Optimal single-

band normal-incidence antireflection coatings," Applied optics 35, 644-658 (1996).

174. T. Petersen and J. Bromage, "Thin-Disk Pumped, High-Average-Power, 2.06-µm BiB 3 O 6

Femtosecond Optical Parametric Oscillator," in Laser Science, (Optical Society of America,

2015), LTh4I. 6.

175. T. P. Lamour, J. Sun, and D. T. Reid, "Wavelength stabilization of a synchronously pumped

optical parametric oscillator: optimizing proportional-integral control," The Review of scientific

instruments 81, 053101 (2010).

176. V. Shcheslavskiy, V. Yakovlev, and A. Ivanov, "High-energy self-starting femtosecond Cr 4+: Mg

2 SiO 4 oscillator operating at a low repetition rate," Optics letters 26, 1999-2001 (2001).

177. H. Kogelnik, "On the propagation of Gaussian beams of light through lenslike media including

those with a loss or gain variation," Applied Optics 4, 1562-1569 (1965).

178. L. Pedrotti, "SJ Leno M," Pedrotti Leno S. Pedrotti, Frank “Introduction to Optics third edition”

p17-25 (2007).

179. T. Petersen, J. Zuegel, and J. Bromage, "High-Energy Infrared Femtosecond Optical Parametric

Oscillator Synchronously Pumped by a Thin-Disk Laser," in Advanced Solid State Lasers,

(Optical Society of America, 2016), ATu1A. 6.

180. T. Petersen, J. D. Zuegel, and J. Bromage, "Thermal effects in an ultrafast BiB3O6 optical

parametric oscillator at high average powers," Applied Optics 56, 6923-6929 (2017).

181. M. J. Riedl, Optical design fundamentals for infrared systems (SPIE press, 2001), Vol. 48.

182. H. Kogelnik, "Imaging of optical modes—resonators with internal lenses," Bell Labs Technical

Journal 44, 455-494 (1965).

183. W. Koechner, "Thermal lensing in a Nd: YAG laser rod," Applied optics 9, 2548-2553 (1970).

Page 137: Scaling Infrared Femtosecond Optical Parametric Oscillators to … · 2018. 6. 12. · Scaling Infrared Femtosecond Optical Parametric Oscillators to High Average Powers by Travis

119

184. I. Jovanovic, B. J. Comaskey, and D. M. Pennington, "Angular effects and beam quality in optical

parametric amplification," Journal of Applied Physics 90, 4328-4337 (2001).

185. C. J. Saraceno, S. Pekarek, O. H. Heckl, C. R. Baer, C. Schriber, M. Golling, K. Beil, C. Kränkel,

G. Huber, and U. Keller, "Self-referenceable frequency comb from an ultrafast thin disk laser,"

Optics express 20, 9650-9656 (2012).

186. C. J. Saraceno, O. H. Heckl, C. R. Baer, C. Schriber, M. Golling, K. Beil, C. Kränkel, T.

Südmeyer, G. Huber, and U. Keller, "Sub-100 femtosecond pulses from a SESAM modelocked

thin disk laser," Applied Physics B 106, 559-562 (2012).

187. A. Diebold, F. Emaury, C. J. Saraceno, C. Schriber, M. Golling, T. Südmeyer, and U. Keller, "62-

fs pulses from a SESAM modelocked Yb: CALGO thin disk laser," in Advanced Solid State

Lasers, (Optical Society of America, 2013), AF3A. 2.

188. J. Chesnoy and L. Fini, "Stabilization of a femtosecond dye laser synchronously pumped by a

frequency-doubled mode-locked YAG laser," Optics letters 11, 635-637 (1986).

189. M. D. Dawson, T. F. Boggess, A. L. Smirl, and D. Maxson, "Cavity-length detuning effects and

stabilization of a synchronously pumped femtosecond linear dye laser," Optics letters 13, 126-128

(1988).

190. D. Butterworth, S. Girard, and D. Hanna, "A simple technique to achieve active cavity-length

stabilisation in a synchronously pumped optical parametric oscillator," Optics communications

123, 577-582 (1996).

191. C. Rothhardt, J. Rothhardt, A. Klenke, T. Peschel, R. Eberhardt, J. Limpert, and A. Tünnermann,

"BBO-sapphire sandwich structure for frequency conversion of high power lasers," Optical

Materials Express 4, 1092-1103 (2014).

192. J. Rothhardt, C. Rothhardt, M. Müller, A. Klenke, M. Kienel, S. Demmler, T. Elsmann, M.

Rothhardt, J. Limpert, and A. Tünnermann, "100 W average power femtosecond laser at 343 nm,"

Optics letters 41, 1885-1888 (2016).

193. F. Geier, J. Pupeikis, H. Geiseler, N. Ishii, T. Kanai, and J. Itatani, "Generation of Carrier-

Envelope Phase-Stable 8.1-fs Infrared Pulses at 20 kHz in a BiB 3 O 6-based Optical Parametric

Amplifier," in Advanced Solid State Lasers, (Optical Society of America, 2016), ATh1A. 2.


Recommended