Frequency Comb
Frequency Comb
Frequency Comb
Over the last few years, there has been a convergence between the fields
of ultrafast science, nonlinear optics, optical frequency metrology, and
precision laser spectroscopy. These fields have been developing largely
independently since the birth of the laser, reaching remarkable levels of
performance. On the ultrafast frontier, pulses of only a few cycles long have
been produced, while in optical spectroscopy, the precision and resolution
have reached one part in 1014. Although these two achievements appear to be
completely disconnected, advances in nonlinear optics provided the essential
link between them. The resulting convergence has enabled unprecedented
advances in the control of the electric field of the pulses produced by
femtosecond mode-locked lasers. The corresponding spectrum consists of a
comb of sharp spectral lines with well-defined frequencies. These new
techniques and capabilities are generally known as “femtosecond comb
technology.” They have had dramatic impact on the diverse fields of
precision measurement and extreme nonlinear optical physics.
The historical background for these developments is provided in the
Foreword by two of the pioneers of laser spectroscopy, John Hall and
Theodor Hänsch. Indeed the developments described in this book were
foreshadowed by Hänsch’s early work in the 1970s when he used
picosecond pulses to demonstrate the connection between the time and
frequency domains in laser spectroscopy. This work complemented the
advances in precision laser stabilization developed by Hall. The parallel
efforts on mode-locked lasers by Charles Shank, Erich Ippen, and others laid
the groundwork for the development in the 1990s by Wilson Sibbett of Kerr-
lens mode locking, the instantaneous nature of which yields sub-10 fs pulses
directly from laser oscillators that correspond to strong phase-locking of the
comb components across a broad optical spectrum. The synergy between
precision spectroscopy and ultrafast lasers was catalyzed by the development
of novel optical fiber with high nonlinearity and controlled dispersion.
In Chapter 1 we provide an introductory description of mode-locked
lasers, the connection between time and frequency descriptions of their
ii Preface
output, the physical origins of the electric field dynamics, and an overview
of applications of femtosecond comb technology. Chapter 2 by Ippen,
Kärtner and Cundiff discusses the development of ultrashort lasers,
particularly focusing on how to achieve an octave-spanning spectrum and
pulse dynamics that are relevant to the stability and control of the comb.
Chapter 3 by Bartels describes in detail high-repetition-rate ring oscillators
for precision frequency metrology. Chapter 4 by Gaeta and Windeler
provides in-depth discussions relating to the physics of bandwidth generation
and the underlying noise process during pulse propagation through
microstructure fibers. Certain aspects of comb dynamics and stability are
presented in Chapters 5 by Steinmeyer and Keller. An attractive approach
presented in Chapter 6 by Kobayashi makes use of optical parametric
generation to produce high-peak-power, femtosecond pulses in the IR
spectral domain. A detailed review of the traditional harmonic-based
frequency chain is provided in Chapter 8 by Schnatz, Stenger, Lipphardt,
Haverkamp, and Weiss, while the new epoch of absolute optical frequency
measurement using femtosecond comb technology is reviewed in Chapter 7
by Udem, Zimmermann, Holzwarth, Fischer, Kolachevsky, and Hänsch.
Chapter 9 by Diddams, Ye, and Hollberg provides an account of the current
state-of-the-art performance and characterization of femtosecond comb
systems used for optical frequency measurement, synthesis, and optical
atomic clocks. Chapter 10 by Baltuška, Paulus, Lindner, Kienberger, and
Krausz provides a thorough discussion of the generation of the high-intensity
pulses needed to access the regime of extreme nonlinear optics and a review
of the results obtained for above-threshold ionization. Control of high-
harmonic generation is addressed in Chapter 11 by Gibson, Christov,
Murnane and Kapteyn. Stabilization of mode-locked lasers and their
applications to ultrasensitive sensors are discussed in Chapter 12 by Diels,
Jones, and Arissian.
The rapid progress during the last 5–6 years has been breathtaking and
has made a tremendous impact on both science and technology. We foresee
an undiminished potential for similar advances in the near future. We hope
that the readers of this book will share our enthusiasm and benefit from the
material presented in this book.
The editors thank all of the chapter authors for their contributions. The
efforts of Julie Phillips and Lynn Hogan in the JILA Scientific Reports
Office are also gratefully acknowledged.
Boulder
September, 2004 Jun Ye and Steven T. Cundiff
CONTENTS
Preface.......................................................................................................i
Contents...................................................................................................iii
Introduction ..........................................................................................12
1. Time- and Frequency-Domain Pictures of a Mode-Locked
Laser ..........................................................................................14
1.1 Introduction to mode-locked lasers......................................14
1.2 Frequency spectrum of a mode-locked laser ........................15
1.3 Determining absolute optical frequencies with octave-
spanning spectra..................................................................17
1.4 Femtosecond optical-frequency comb generator ..................19
1.5 Time- and frequency-domain characterizations of f0 ............22
2. Precision Optical Frequency Metrology Using
Femtosecond-Optical-Frequency Combs.....................................24
2.1 Measurement of absolute optical frequency .........................24
2.2 Optical atomic clocks..........................................................27
2.3 Optical frequency synthesizer..............................................29
3. Atomic and Molecular Spectroscopy...........................................30
3.1 Precise, simultaneous determination of global atomic
structure and transition dynamics ........................................30
3.2 I2 hyperfine interactions, optical frequency standards,
and clocks ...........................................................................33
4. Carrier-Envelope Phase Coherence and Time-Domain
Applications ...............................................................................38
4.1 Timing synchronization of mode-locked lasers....................39
4.2 Phase lock between separate mode-locked lasers .................40
4.3 Extending phase-coherent femtosecond combs to the
mid-IR spectral region.........................................................41
4.4 Femtosecond lasers and external optical cavities .................42
iv Contents
Author addresses...................................................................................355
Index ....................................................................................................359
Foreword
HISTORY OF OPTICAL COMB DEVELOPMENT
The factors that limit the shortness of the generated laser pulses arise
from two issues: (1) finite gain-bandwidth product (which is not a problem
for Ti:sapphire) and (2) intracavity dispersion. A short pulse can be viewed
as the superposition of many cw phase-locked modes, all of which oscillate
at their own cavity-defined frequencies. For the pulse train to be stable in
time, the modes must have a common frequency separation. Because of
dispersion in the sapphire, intracavity air, and mirror coatings, these cavity
frequencies are generally not exactly evenly spaced. This is particularly true
as the spectral bandwidth dramatically increases. So even with the ~30%
bandwidth of Ti:sapphire, further shortening of sub-100 fs pulses proved
difficult until Asaki et al. [12] employed a sufficiently general analysis of the
pulse laser cavity. This model included the index of refraction characteristics
of the intracavity dispersion-compensating prisms, the resulting color
influence on the refraction direction, associated cavity path lengths, and
some modeling of the air and laser crystal dispersion. Space-time focusing of
the light bullet in the laser crystal was another important consideration.
In the early 1990s, when most of researchers were working feverishly to
shorten pulse widths, some dreamers began to think of pulses so short that
their Fourier representation would span from radio frequencies up into the
visible domain. This idea seemed like science fiction to one of the authors
(JLH) and a likely possibility to the other (TWH). Of course, spectral self-
broadening was well known. By focusing powerful amplified pulses into
water or some solids, one could basically generate a white-light continuum.
At elevated pulse energy levels, one expected serious disruption of the
calmness of the intermolecular bonds; consequently, one would not expect to
find a phase-stable repetition of the generated white light. Perhaps a glass
sample could melt and recrystallize at a 100 MHz rate, emitting a similar
thermal spectrum on every heating cycle. However, to form a coherent
optical comb in the forward direction, the timing would need to be stable to
~1 radian — at the visible frequency! Few believed that this thermal process
would be stable at the 0.3 fs level needed. Rather, it seemed clear that a more
gentle process would be required, in which somewhat less-powerful laser
pulses would strongly distort some atomic wave functions but not disrupt
chemical bonds. Atomic frequencies are so high that when the pulse is gone,
calmness can return; the next pulse would be able to generate just the same
effect on the system. In this case, the phase-coherence of the source pulses
could insure that the nonlinearly generated frequencies would be mutually
coherent pulse-to-pulse.
Researchers in both Garching and Boulder set out to learn about this
subject. In Europe, an amplified pulse was split into two parts and focused
onto two separate spots in a CaF2 crystal. The white light produced in each
spot in the CaF2 plate interfered with each other in a geometry that allowed
HISTORY OF OPTICAL COMB DEVELOPMENT 5
been in a nearby group in the same part of Bell Labs. By some unknown
means, the JILA team came up with a sample of “Magic Fiber” to test by
November 1999. The Garching group teamed up with the powerful fiber
group of P. St. J. Russell at the University of Bath (UK), which had been
working with both microstructure and tapered fibers. New results began
immediately rolling in [21, 22], and the publication competition began in
earnest! (Since then, many alternatives to the Magic Fiber approach have
emerged, including supercontinuum generation in tapered fibers [21],
octave-spanning laser oscillators without the use of external fibers, mode-
locked fiber lasers with some highly nonlinear ordinary fiber for spectral
broadening, and schemes incorporating difference-frequency generation to
determine the carrier-envelope-offset frequency with combs spanning less
than an octave.)
At this point, the Boulder team had a significant advantage. They had
already worked on laser stabilization and optical frequency standards for
many years because of JILA’s affiliation with the Boulder campus of the
National Institute of Standards and Technology (NIST). Indeed, the authors
of this foreword first met in Novosibirsk in 1969 at a conference organized
by the late Veniamin Chebotayev on the topic of stabilized lasers. There,
JLH presented his progress with a methane-stabilized He-Ne laser.
While other researchers were busy improving cw and pulsed lasers,
national metrology and standards laboratories around the world had been
trying to verify the frequencies of their “as-maintained” national wavelength
standards. The first such measurements occurred in Boulder and led, in
1972, to the measurement of the frequency and wavelength of the methane
standard. This measurement, in turn, led to a new and definitive value for the
speed of light. Other national laboratories joined in, and a long discussion
ensued about the philosophical and practical issues associated with
calculating meters from the frequency of light rather than simply adopting
new wavelength standards as they became available. Within 10 years,
national laboratories in Canada, the United Kingdom, Japan, Germany, and
Gaithersburg, Maryland, had confirmed parts of the Boulder work and the
frequency of the He-Ne iodine-stabilized red laser had been determined. In
1983, the meter was redefined based on the speed of light.
The reproducibility of most of the reference lasers developed during this
era was so good that their imperfections had little practical consequence for
length metrology. Still, the optical frequency standards business continued to
develop as researchers sought better designs and new reference transitions.
More importantly, each nation wanted to confirm its own standards at the
highest level.
Both NIST and PTB built very good systems to measure the calcium
intercombination transition at 657 nm, for example. In addition, the PTB
8 Foreword
common 10 MHz rf reference and comparing comb lines near 350 THz, they
verified agreement within a few parts in 1016; the precision of the experiment
was probably limited by Doppler shifts due to air pressure changes or thermal
expansion of the optical tables.
The Garching group also made additional accuracy tests on the comb-
spacing uniformity using one stage of a frequency-interval divider [14]. The
interval between comb lines near the edges of the “white” spectrum could be
divided to find the “center” by two methods. Or, an edge line could be
combined with one near the middle to seek a dispersive effect. No problems
could be found. To really press to the testing limits, comb vs comb tests —
using different comb frequencies, materials, or whatever we think could be
important — are probably necessary. By late 2004, tests still had discovered no
problems, and the accuracy in the context of frequency combs had reached
10-21 [27].
As we begin the next era in optical comb research, with each group
measuring specific frequencies with femtosecond comb techniques, the
realization of the cesium standard frequency is likely to be the first weak
point. With a day’s averaging, the GPS system can help us know our local
frequency standard’s average performance, but it takes about a day to deliver
an accuracy ~1 x 10-14; achieving this accuracy requires us to know the
timing comparisons separately with each of the satellites used in the test. To
test optical frequency combs against the highest traditional standard, modern
fountain cesium clocks are usually available either by fiber link [4] or by
physical transport.
Femtosecond combs are now ready to accurately measure any desired
frequency such as those of some isolated hydrogen atoms at rest in a field-
free vacuum, single Hg+ or Yb+ ions in an ion trap, or a million cold Sr
atoms trapped in an optical lattice/trap. In addition, frequency comb
techniques are having an impact on ultrafast physics. By making it possible
to produce few-cycle pulses with a stabilized carrier-envelope phase, these
tools are leading to the discovery of novel phenomena in nonlinear light-
matter interactions. For example, by 2004 amplified phase-stabilized pulses
had been used to produce controlled bursts of soft x-rays with time durations
in the attosecond range [28].
So where do we go from here?
In contrast to the digital security of frequency measurement, in the final
accuracy-defining step, it usually comes out that spectroscopic line-shape
issues are what limit our results. How well do we know the connection
between the center of the observed resonance line and the desired physical
quantity? Resonance frequencies can be shifted by fields, laser intensity,
Doppler shifts, and ... Careful treatment of such issues is part of what makes
our field so fun. And, of course, many interesting measurements can be
10 Foreword
*Use of product name for technical information only and does not constitute an
endorsement by NIST.
REFERENCES
[1] J. L. Hall, IEEE J. Sel. Top. Quantum Electron. 6, 1136-1144 (2000).
[2] R. K. Shelton, L. S. Ma, H. C. Kapteyn, M. M. Murnane, J. L. Hall, and J. Ye,
Science 293, 1286-1289 (2001).
[3] E. O. Potma, D. J. Jones, J. X. Cheng, X. S. Xie, and J. Ye, Opt. Lett. 27,
1168-1170 (2002).
[4] J. Ye, J. L. Peng, R. J. Jones, K. W. Holman, J. L. Hall, D. J. Jones, S. A.
Diddams, J. Kitching, S. Bize, J. C. Bergquist, L. W. Hollberg, L. Robertsson,
and L. S. Ma, J. Opt. Soc. Am. B 20, 1459-1467 (2003).
[5] A. I. Ferguson, J. N. Eckstein, and T. W. Hänsch, Appl. Phys. Lett. 49, 5389-
5391 (1978).
[6] J. N. Eckstein, A. I. Ferguson, and T. W. Hänsch, Phys. Rev. Lett. 40, 847-850
(1978).
[7] H. R. Telle, D. Meschede, and T. W. Hansch, Opt. Lett. 15, 532-534 (1990).
[8] M. Kourogi, K. Nakagawa, and M. Ohtsu, IEEE J. Quantum Electron. 29,
2693-2701 (1993).
[9] A. Huber, T. Udem, B. Gross, J. Reichert, M. Kourogi, K. Pachucki, M.
Weitz, and T. W. Hänsch, Phys. Rev. Lett. 80, 468-471 (1998).
[10] J. L. Hall, L. S. Ma, M. Taubman, B. Tiemann, F. L. Hong, O. Pfister, and J.
Ye, IEEE Trans. Instrum. Meas. 48, 583-586 (1999).
[11] D. E. Spence, P. N. Kean, and W. Sibbett, Opt. Lett. 16, 42-44 (1991).
[12] M. T. Asaki, C. P. Huang, D. Garvey, J. P. Zhou, H. C. Kapteyn, and M. M.
Murnane, Opt. Lett. 18, 977-979 (1993).
[13] M. Bellini and T. W. Hänsch, Opt. Lett. 25, 1049-1151 (2000).
[14] T. Udem, J. Reichert, R. Holzwarth, and T. W. Hänsch, Opt. Lett. 24, 881-883
(1999).
[15] M. Niering, R. Holzwarth, J. Reichert, P. Pokasov, T. Udem, M. Weitz, T. W.
Hänsch, P. Lemonde, G. Santarelli, M. Abgrall, P. Laurent, C. Salomon, and
A. Clairon, Phys. Rev. Lett. 84, 5496-5499 (2000).
HISTORY OF OPTICAL COMB DEVELOPMENT 11
Abstract: Recently there has been a remarkable synergy between the technology of
precision laser stabilization and mode-locked ultrafast lasers. This has resulted
in control of the frequency spectrum, which consists of a regular “comb” of
sharp lines, produced by mode-locked lasers. Such a controlled mode-locked
laser is a “femtosecond optical frequency comb generator.” For a sufficiently
broad comb, it is possible to determine the absolute frequencies of all of the
comb lines. This ability has revolutionized optical frequency metrology and
synthesis. It has also served as the basis for the recent demonstrations of
atomic clocks that utilize an optical frequency transition. In addition, it is
having an impact on time-domain applications, including phase-sensitive
extreme nonlinear optics and pulse manipulation and control. In this chapter,
we first review the frequency-domain description of a mode-locked laser and
the connection between the pulse phase and the frequency spectrum to provide
a basis for understanding how the absolute frequencies can be determined and
controlled. Using this understanding, applications in optical frequency
metrology, optical atomic clocks, and precision spectroscopy are discussed.
Next, we discuss applications of the carrier-envelope phase coherence in time-
domain experiments. This chapter serves as a broad introduction and summary
for all subsequent chapters that present detailed discussions of specific topics.
that results in a higher net gain for short pulses as compared to continuous
wave (cw) operation. This mechanism can be either an active element or
implemented passively with saturable absorption (real or effective). Passive
mode locking yields the shortest pulses because, up to a limit, the self-
adjusting mechanism becomes more effective as the pulse shortens [18].
Real saturable absorption occurs in a material with a finite number of
absorbers, for example, a dye or semiconductor. The shortness of the pulses
is limited by the finite lifetime of the excited state. Effective saturable
absorption typically uses the nonlinear index of refraction of some material
together with spatial effects or interference to produce higher net gain for
shorter pulses. The ultimate limit on minimum pulse duration in such a
mode-locked laser is due to the interplay between the mode-locking
mechanism, group-velocity dispersion (GVD), and net gain bandwidth [18].
Chapter 2 discusses the development of ultrashort lasers, particularly
focusing on how to achieve an octave-spanning spectrum and pulse
dynamics that are relevant to the stability and control of the comb. Further
aspects of the comb dynamics and stability are presented in Chapters 5, 7, 9,
10, and 12.
Currently, the generation of ultrashort optical pulses is dominated by the
Kerr-lens–mode-locked Ti:sapphire (KLM Ti:sapphire) laser because of its
excellent performance and relative simplicity. Kerr-lens mode locking is
based on a combination of self-focusing in the Ti:sapphire crystal and an
aperture that selects the spatial mode corresponding to the presence of self-
focusing. The Ti:sapphire crystal is pumped by green light from either an
Ar+-ion laser or a diode-pumped–solid-state (DPSS) laser, which provides
far superior performance in terms of laser stability and noise. The
Ti:sapphire crystal provides gain and serves as the nonlinear material for
mode locking. Prisms or dispersion-compensating mirrors compensate the
GVD in the gain crystal [19]. Chapter 3 describes in detail a high repetition-
rate ring oscillator system. Since the discovery of KLM [20], the pulse width
obtained directly from mode-locked lasers has been shortened by
approximately an order of magnitude by first optimizing the intracavity
dispersion [21] and then using dispersion-compensating mirrors [1], yielding
pulses that are less than 6 fs in duration, i.e., less than two optical cycles.
Recently, output spectra that span an octave (factor of two in frequency)
have been obtained directly from a mode-locked laser [22], which is an
important accomplishment for phase stabilization.
Et τ
T = 1/fr
Iν
bFrequency Domain
fr
f0 1/τ
0 νn = nfr+f0 ν
Figure 1-1. Summary of the time-frequency correspondence for a pulse train with evolving
carrier-envelope phase.
frequency as the comb line on the high-frequency side of the spectrum with
index 2n. Measuring the heterodyne beat between these two families of
optical comb lines yields a difference,
2ν n − ν 2n = 2(nf r + f 0 ) − (2nf r + f 0 ) = f 0 , which is just the offset frequency.
Thus an octave-spanning spectrum enables a direct measurement of f0.
However, an octave-spanning spectrum is not required; it just represents the
simplest approach. We designate this scheme, shown in Figure 1-2(a), as
“self-referencing” since it uses only the output of the mode-locked laser.
(a) f0
I(ν) fr
ν
0
f0
(b) νstandard 2νstandard
f0
fr
ν
0
fbeat1 fbeat2
f0 = fbeat2 – 2fbeat1
Figure 1-2. Two equivalent schemes for the measurement of f0 using an octave-spanning
optical frequency comb. In the self-referencing approach, shown in (a), frequency doubling
and comparison are accomplished with the comb itself. In the second approach shown in (b),
the fundamental frequency (νstandard) and its second harmonic of a cw optical standard are
used to determine f0. These two basic schemes are employed for absolute optical frequency
measurement and implementation of optical atomic clocks.
spectrum. Then the second harmonic of the cw laser will be positioned close
to the comb line 2n. Measurement of the heterodyne beat between the cw-
laser frequency, νs, and the comb line n gives f beat1 = ν s − (nf r + f 0 ) and
between the second harmonic of the cw laser and comb line 2n gives
f beat 2 = 2ν s − (2nf r + f 0 ) . Mixing the beats with appropriate weighting
factors gives f beat 2 − 2 f beat1 = 2ν s − (2nf r + f 0 ) − (2ν s − 2(nf r + f 0 )) = f 0 ,
which represents the second detection scheme shown in Figure 1-2(b).
Another interesting fact is that by mixing the two beat signals, one
establishes a direct link between the optical and rf frequencies (νs and fr) as
in f beat 2 − f beat1 = 2ν s − (2nf r + f 0 ) − ν s + (nf r + f 0 ) = ν s − nf r .
techniques. Experimentally, it has been found that even if the power at the
octave-spanning points is 40 dB below the peak, it is still possible to observe
strong ν-to-2ν heterodyne beats.
Since Ti:sapphire, which has the broadest gain bandwidth of all known
laser media, does not support an octave-spanning spectrum, additional
spectral content must be generated. This is accomplished by using self-phase
modulation, which is based on a temporal variation in the index of refraction
because of a combination of a short optical pulse and an intensity-dependent
index of refraction [29]. The broadening can be achieved outside the laser
cavity by using optical fiber or internally by creating coincident secondary
time and space foci [22]. The latter technique requires carefully designed
mirrors [30] and is challenging to implement; therefore the former is
significantly more common. (For more detailed discussions, see Chapter 2.)
Recent results have shown that additional spectral bandwidth can be
obtained by minor changes in the cavity configuration of a high-repetition-
rate laser, although this technique has not yet yielded sufficient intensity at
the octave points for observation of ν-to-2ν beats. (See Reference [31] and
related discussions in Chapter 3.)
The amount of spectral broadening that can be obtained in ordinary
optical fiber is limited, primarily because temporal spreading of the pulse,
due to GVD in the fiber, reduces the peak intensity. Using a low-repetition-
rate laser (to raise the pulse energy) an octave-spanning spectrum has been
obtained with ordinary fiber [7]. The discovery [16] that microstructure fiber
can have zero group-velocity dispersion within the emission spectrum of a
Ti:sapphire laser eliminated this difficulty and led to rapid progress in the
field of femtosecond-optical-frequency combs by allowing broadband-
continuum generation with only nanojoule-pulse energies.
Microstructure fiber uses air holes surrounding a fused silica core to
obtain the index-of-refraction contrast needed for waveguiding. This method
results in a much larger index contrast than can be obtained using doping.
The large index contrast has two consequences: (1) the ability to generate a
zero in the GVD at visible or near-infrared wavelengths and (2) the
possibility of using a much smaller core size. The first consequence means
that the pulse does not spread temporally and hence maintains its high peak
power. In addition, it results in phase matching between the generated
spectral components. The second outcome greatly increases the light
intensity in the core, thereby enhancing nonlinear effects. Chapter 4 provides
in-depth discussions relating to the physics of bandwidth generation and the
underlying noise process during pulse propagation through these relatively
novel microstructure fibers.
We have now introduced all of the concepts and components needed to
construct a femtosecond-optical-frequency comb generator that produces
1. INTRODUCTION 21
output
microstructure
fiber
KLM Ti:sapphire Laser ν−2ν interferometer
AOM
Pump
Laser Mod.
· n/16 SHX
fr
Micro-
wave servo counter
Clock
f0 photo-
counter
servo diode
symmetric, and the shift of the fringes from the peak of the envelope is due
to the pulse-to-pulse phase shift ∆φce. By stabilizing f0 at different frequency
offsets, differing values of ∆φce are obtained, as discussed in Section 1.2.
Another technique, known as spectral interferometry, can also be employed
to determine the values of ∆φce and the related f0 if the interfering pulses are
separated by at least one cavity round-trip time. This technique also yields
additional spectral-phase information. For single-shot experiments, a
variation of the spectral-interferometry technique that involves interference
between the fundamental and its second harmonic spectra, which overlap if
the fundamental optical bandwidth exceeds one octave, becomes the most
effective approach to measuring ∆φce [36].
Although cross-correlation measurements can demonstrate some degree
of phase coherence, they are actually quite insensitive to phase fluctuation
because they measure the change between pulses that are separated by only a
few round-trip times because of the practical size limitation of the
interferometer. Instead, frequency-domain measurements of the frequency-
noise spectrum of f0 are much more sensitive since they can monitor the
frequency/phase evolution of f0 over much longer time intervals. Typically, f0
is determined from the self-referencing technique discussed earlier, from
which one can make a close examination of the frequency- and phase-noise
power spectral density associated with f0.
Given a measurement of the frequency-noise power spectral density of
f0, sνf 0 , the accumulated root-mean-square fluctuations of φce are given by
∞
sν ( f )df for an observation time τobs [37]. Once a
1 f0
∆φ ce
rms
= 2 ∫
τ obs f2
1 2πτ obs
between the two signal sources with a frequency counter recording the
resultant beat-frequency fluctuations over a period of time. There has been
good progress in the process of extracting high-stability rf-output signals
from an optical clock at a level that approaches the optical stability. Further
work along this line is continuing. We also still face some technical
challenges in making an optical clock a reliable scientific device. Technical
developments necessary for an advanced optical atomic clock include: (1)
highly accurate, cold-atom optical frequency standards and portable, high-
stability optical frequency standards (see Chapters 8, 9); (2) development of
ultrastable optical local oscillators suitable for the most demanding
spectroscopic tasks (see Chapter 9); (3) stabilization and control of wide-
bandwidth optical combs, including exploration of novel generation and
detection techniques and approaches to reduce noise (see Chapters 5, 7, 9);
(4) reliable, stable, and compact ultrafast laser technology for practical
implementation of optical clocks (see Chapters 2, 3, 4, 6); (5) development
and use of femtosecond combs for intercomparison of optical frequency
standards and the cesium primary standard (see Chapters 7, 8, 9); and (6)
development of frequency- and time-transfer methods over extended fiber
optic links that can support the next generation of atomic frequency
standards (see Chapter 9).
F” =
1
5D5/2 2
3 5D δSD
5D3/2 4
3
Φ-coherent 2
comb 1
0 De 6P
νPD = ca
yt
m fr + f0 o5
F’ = S
fr 3
2
5P3/2 1
0
5P1/2 5P
2 δSP
1
νSP =
f0 F= n fr + f0
…
2
5S1/2 1 5S
∆φce 2∆φce
E(t)
ωc
Cold Rb
t
τ = 1/fr ∆φce = 2π (f0/fr)
Figure 1-4. Top: Schematic of the relevant energy levels of the 87Rb atom and the frequency-
domain perspective of the atom-light interaction. Bottom: Time-domain picture showing a
sequence of mode-locked pulses, with the relevant interaction parameters in fr and ∆φce. The
inset at right shows the relevant “three-level” model used for construction of the Bloch
equations to solve for population-transfer dynamics. An example for the on-resonance,
stepwise transition is shown with relevant detunings at δSP and δSD for the pair of comb modes
that make the dominant contribution to the transition probability amplitude.
1. INTRODUCTION 33
νstandard
(a)
fr f0
Transfer stability
~ 10-16
< 10 nW per comb
f
0
(b) 2P
1/2 + 2P3/2
B
E - energy
2P + 2P3/2
1 3/2
u
X
R
R - internuclear distance
Figure 1-5. (a) Random access and precise stabilization of a cw laser (νcw) using the
frequency reference grid provided by the femtosecond comb stabilized by optical (νstandard)
and rf standards. (b) The ground state and the first excited state of I2 with their associated
dissociation limits. The transition linewidth narrows when the excited state approaches the
dissociation limit.
after ν' = 60. For the sake of figure clarity, the eqQB data for ν' > 60 are not
shown. Another important observation is that for levels of ν' = 57–59, all
hyperfine parameters except for CB bear abnormal J-dependences due to
perturbations from a 1g state through accidental rotational resonances.
Combining data from this work and the literature [67], investigations of
the hyperfine spectra now cover the majority of the vibrational levels (3 ≤ v'
≤ 82) in the B state. Therefore, it is now possible and useful to explore the
global trend of these hyperfine parameters in the B state. Suppressing the
rotational dependence, hyperfine parameters as functions of pure vibrational
energy E(v') are found to increase rapidly when molecules approach the
dissociation limit, which is a result of the increasingly strong perturbations
from other high-lying electronic states sharing the same dissociation limit
with the B state. While the variation of CB is smooth over the whole range,
eqQB , dB , and δB all have local irregularities at three positions: (1) ν' = 5
where the B":1u state crosses nearby, (2) around ν' = 57–59 (see discussions
above), and from ν' = 76–78, because of the same 1g state.
To examine these hyperfine parameters in terms of the internuclear
separation R, the vibrational average of the hyperfine parameters is removed
by inverting the expression O (v ′, J ′) = v ′J ′ O(R ) v ′J ′ , where O(v',J') denotes
one of the four hyperfine parameters. eqQB , CB , dB , and δB against R-
centroid are evaluated from v′J ′ R v′J ′ (with v ′J ′ properly normalized).
Consistent with CB’s smooth variation, the interpolation function CB (R) has
small residual errors (within ±0.03, relative) for the entire range from v' = 3–
70. In contrast, the large residual errors in the interpolation of eqQB , dB, and
δB for v' ≥ 56 reflect abnormal variations observed around v' = 57 and 59,
restricting a reliable interpolation only to levels of v' < 56. In the region of R
< 5 Å, valuable information can be readily extracted from eqQB to assist the
investigation of I2’s electronic structure. Unlike the other three hyperfine
parameters whose major parts originate from perturbations at nearly all
possible values of R, a significant part of eqQB is due to the interaction
between the nuclear quadrupole moment Q and the local electric-field
gradient q(R) generated by the surrounding charge distribution of a largely
B-state character. Thus, for R < 5 Å, where perturbations from other
electronic states are negligible, the vibration-removed interpolation function
eqQB (R), coupled with a priori information on q(R), can be used to
determine the I2 nuclear quadrupole moment or serve as a benchmark for
molecular ab initio calculations of the electronic structure at various values
of R.
Precision measurements on B–X hyperfine spectra provide an alternative,
and yet effective, way to investigate the potential energy curves (PECs)
sharing the same dissociation limit with the B state and the associated
1. INTRODUCTION 37
42 (a) 70 (b)
2 69 65
43 63
-560
61
eqQ B (MHz)
60
45 59
CB (MHz)
1 57
-564
47 55
50 6 53
49 5
51 51
-568 4 49
53 50 47
3 45
57 55
43
-572 2
59 60 42
1 70 70
8 69 (c) (d)
69 65
7 63
6
59 1 61
65 60 60
5 63
59 57
-dB (MHz)
δB (MHz)
4 55
61 57 55
3 53
53 51
51
0.1
49
2 50 49 50
47 47
45 45
43 0.01 42
0.1
42 42
8
3 3
0 5 10x10 0 5 10x10
J' (J' +1) J' (J' +1)
Figure 1-6. Rovibrational dependence of the B-state hyperfine parameters: (a) eqQB , (b) CB,
(c) dB , and (d) δB. Graphs (b), (c), and (d) are semilog plots and the vertical scale of (c) has
been inverted. Each solid line is a fit for J-dependence for each vibrational level (ν' indicated
in the figure). Experimental data in squares and open circles show abnormal variations of
eqQB , dB , and δB around ν' = 57 and 59.
38 Chapter 1
achieve a desired timing offset between the two pulse trains. The high
stability loop can then be activated to achieve the ultimate level of
synchronization at the preset value of timing offset. For example, in the all-
electronic implementation of laser synchronization, we use two phase-locked
loops (PLLs). The low-resolution PLL compares and locks the fundamental
repetition frequencies (100 MHz) of the lasers. The second, high-resolution,
PLL compares the phase of the 140th harmonic of the two repetition
frequencies at 14 GHz. A transition of control from the first PLL to the
second PLL is sufficiently smooth to allow synchronization at the
femtosecond level for any timing offsets within the entire dynamic range,
e.g., one pulse period of 10 ns [82]. The synchronization lock can be
maintained for several hours.
The ability to synchronize a passively mode-locked laser to an external
reference or to a second laser has many applications. Previous work in
electronic synchronization of two mode-locked Ti:sapphire lasers
demonstrated timing jitter of, at best, a few hundred femtoseconds.
Therefore the present level of synchronization would make it possible to take
full advantage of this time resolution for applications such as high-power
sum- and difference-frequency mixing [83], novel pulse generation and
shaping [9], new generations of laser/accelerator-based light sources, or
experiments requiring synchronized laser light and x-rays or electron beams
from synchrotrons [84]. Indeed, accurate timing of high-intensity fields is
essential for several important schemes in quantum coherent control and
extreme nonlinear optics such as efficient x-ray generation. Two recent
applications that have been developed in our laboratories include tunable,
subpicosecond pulse generation in the IR [85] and coherent anti–Stokes-
Raman scattering (CARS) microscopy with two tightly synchronized
picosecond lasers [69]. The flexibility and general applicability of the two-
laser-synchronization approach are clearly demonstrated in the
straightforward generation of programmable light sources for these
applications.
The combination of ultrashort pulse trains and optical cavities will open
doors for a variety of exciting experiments. Such experiments will require an
understanding of intricate pulse-cavity interactions and the development of
techniques to efficiently couple ultrashort pulses into a high-finesse optical
1. INTRODUCTION 43
cavity and coherently store them in it. An immediate impact will be on the
precision stabilization of ultrafast lasers (see more detailed discussions in
Chapter 12). Similar to the state-of-the-art stabilization of cw lasers, a
cavity-stabilized ultrafast laser is expected to demonstrate superior short-
term stability of both the pulse-repetition frequency and the carrier-envelope
phase [90]. The improved stability is beneficial in particular for time-domain
applications where the signal-processing bandwidth is necessarily large.
Another attractive application lies in broadband and ultrasensitive
spectroscopy. The use of high-finesse cavities has played a decisive role in
enhancing sensitivity and precision in atomic and molecular spectroscopy.
We expect a dramatic advancement in the efficiency of intracavity
spectroscopy by exploiting the application of ultrashort pulses. In other
words, high detection sensitivity is achievable uniformly across the broad
spectrum of the pulse. Applying cavity-stabilization techniques to
femtosecond lasers, the comb structure of the probe laser can be precisely
matched to the resonance modes of an empty cavity, allowing an efficient
energy coupling for a spectroscopic probe. Molecular samples located inside
the high-finesse cavity will have a strong impact on the dispersive properties
of the cavity. In fact, it is this dispersion-related–cavity-pulling effect that
will aid our sensitive detection process when we analyze the light
transmitted through the cavity. Preliminary data on spectrally resolved, time-
domain–ring-down measurement for intracavity loss over the entire
femtosecond-laser bandwidth are already quite promising.
To develop sources for ultrafast nonlinear spectroscopy, a properly
designed, dispersion-compensated cavity housing a nonlinear crystal will
provide efficient nonlinear optical frequency conversion of ultrashort optical
pulses at spectral regions where no active gain medium exists. Furthermore,
by simultaneously locking two independent mode-locked lasers to the same
optical cavity, efficient SFG and/or DFG can be produced over a large range
of wavelengths. Similarly, a passive cavity can be used to explore coherent
superposition of ultrashort pulses, with cavity stabilization providing the
means to phase coherently superpose a collection of successive pulses from a
mode-locked laser. The coherently enhanced pulse stored in the cavity can
be switched out using a cavity-dumping element (such as a Bragg cell),
resulting in a single phase-coherent, amplified pulse. The use of a passive
cavity also offers the unique ability to effectively amplify pulses in spectral
regions where no suitable gain medium exists such as for IR pulses from
difference-frequency mixing or the ultraviolet light from harmonic
generation. Unlike actively dumped laser systems, the pulse energy is not
limited by the saturation of a gain medium or the requirement for a saturable
absorber for mode locking. Instead, the linear response of the passive cavity
allows the pulse energy to build up inside the cavity until limited by cavity
44 Chapter 1
Mode-locked laser
2
F
Cavity enhancement: N = 4tin τroundtrip = k T or (1/k) T; k: integer
2π
1.2
(a) enhanced (112 x buildup)
FROG trace input spectrum (b)
Amplitude (a.u.)
0.2
0.0
0.0
Figure 1-7. Top: Illustration of the principle of coherent-pulse amplification with the aid of an
optical cavity, showing the matching of the pulse-repetition period with the cavity round-trip
time. The intracavity pulse is switched out when sufficient energy is built up in the cavity.
The build-up factor is given ideally by the cavity finesse, with appropriate intracavity
dispersion compensation. Bottom: Coherent evolution of a 50 fs pulse inside the cavity. The
left panel displays a FROG measurement of the amplified pulse switched out of the passive
cavity shown at FWHM of 49 fs, slightly wider than the input pulse width of 47 fs. The right
panel presents a comparison of the input and output pulse spectra, showing no significant
distortion to the input pulse spectrum after the intracavity power is built up by a factor of 112.
Figure 1-8. Conceptual diagram of quantum interference between one- and two-photon
absorption in a direct-gap semiconductor. The two interfering absorption pathways are driven
by the spectral wings of a single octave-spanning pulse. An imbalance in the otherwise
symmetric carrier-population distribution in momentum space (represented by ovals) occurs
due to interference and results in a net current. The resulting photocurrent is sensitive to φce.
10 dB
Figure 1-9. Spectrum of the measured quantum interference signal in a semiconductor. The
dotted line is the background (no light on the sample).
5. SUMMARY
ACKNOWLEDGEMENTS
REFERENCES
[1] U. Morgner, F. X. Kärtner, S. H. Cho, Y. Chen, H. A. Haus, J. G. Fujimoto,
E. P. Ippen, V. Scheuer, G. Angelow, and T. Tschudi, Opt. Lett. 24, 411-
413 (1999); D. H. Sutter, G. Steinmeyer, L. Gallmann, N. Matuschek, F.
Morier-Genoud, U. Keller, V. Scheuer, G. Angelow, and T. Tschudi, Opt.
Lett. 24, 631-633 (1999).
[2] D. J. Jones, S. A. Diddams, J. K. Ranka, A. Stentz, R. S. Windeler, J. L.
Hall, and S. T. Cundiff, Science 288, 635-639 (2000).
[3] S. A. Diddams, D. J. Jones, J. Ye, S. T. Cundiff, J. L. Hall, J. K. Ranka, R.
S. Windeler, R. Holzwarth, T. Udem, and T. W. Hänsch, Phys. Rev. Lett.
84, 5102-5105 (2000); R. Holzwarth, T. Udem, T. W. Hänsch, J. C.
Knight, W. J. Wadsworth, and P. S. J. Russell, Phys. Rev. Lett. 85, 2264-
2267 (2000); S. T. Cundiff, J. Ye, and J. L. Hall, Rev. Sci. Instrum. 72,
3746-3771 (2001).
[4] J. Stenger, T. Binnewies, G. Wilpers, F. Riehle, H. R. Telle, J. K. Ranka, R.
S. Windeler, and A. J. Stentz, Phys. Rev. A 63, 021802 (2001).
[5] T. Udem, S. A. Diddams, K. R. Vogel, C. W. Oates, E. A. Curtis, W. D.
Lee, W. M. Itano, R. E. Drullinger, J. C. Bergquist, and L. Hollberg, Phys.
Rev. Lett. 86, 4996-4999 (2001).
[6] S. A. Diddams, T. Udem, J. C. Bergquist, E. A. Curtis, R. E. Drullinger, L.
Hollberg, W. M. Itano, W. D. Lee, C. W. Oates, K. R. Vogel, and D. J.
Wineland, Science 293, 825-828 (2001); J. Ye, L. S. Ma, and J. L. Hall,
Phys. Rev. Lett. 87, 270801 (2001).
[7] A. Apolonski, A. Poppe, G. Tempea, C. Spielmann, T. Udem, R.
Holzwarth, T. W. Hänsch, and F. Krausz, Phys. Rev. Lett. 85, 740-743
(2000).
[8] T. Brabec and F. Krausz, Rev. Mod. Phys. 72, 545-591 (2000).
[9] R. K. Shelton, L. S. Ma, H. C. Kapteyn, M. M. Murnane, J. L. Hall, and J.
Ye, Science 293, 1286-1289 (2001).
50 Chapter 1
1. INTRODUCTION
10ps
Nd:glass
Nd:YLF
Nd:YAG
1ps
SOLID-STATE
CW Dye SYSTEMS
Color
Center
100fs
CPM Dye Er:fiber
DYE LASER Cr:YAG
Nd:fiber
TECHNOLOGY
Cr:forsterite
10fs
Cr:LiS(G)AF
w/Compression Ti:sapphire
led to the generation of 5 fs pulses containing less than two optical cycles.
The wings of the spectrum of these pulses extend over more than one octave,
an achievement that has important implications for the control of the electric
field waveform of the pulse. It had been recognized earlier by Xu et al.[12]
that for few-cycle pulses, the maximum electric field in a pulse
quantitatively depends on the phase relationship between the carrier wave
and the maximum of the envelope (see Figure 2-1).
Nonlinear optical processes, which depend on powers of the electric field
amplitude, must therefore also depend on the carrier-envelope phase for very
short pulses. This dependence is visualized in Figure 2-2, which shows
sketches of spectra resulting from instantaneous χ (2) and χ (3) processes
produced by a chirp-free pulse with a rectangular-shaped spectrum spanning
an octave. For a one-octave pulse, the spectra of the second-harmonic and
optical-rectification terms overlap with the fundamental spectrum.
Interference in the overlap regions will be constructive or destructive
depending on the carrier-envelope phase. Use of this interference for carrier-
envelope phase control is called the ν-to-2ν self-heterodyne technique. If
higher-order processes are employed, such as a χ (3) nonlinearity, there are
similar interference terms between the χ (2) and χ (3) terms. Thus a phase-
dependent interference is already possible even if the pulse spans only 2/3 of
an octave. All of these interferences have been used to detect and control the
carrier-envelope phase.
Fundamental
eiφ
O
OR SHG χ(2)
e2iφO
eiφ O
SPM
THG χ(3)
e3iφO
0 ω0 2ω0 3ω0 ω
Figure 2-2. Top: Fundamental spectrum of an octave-spanning pulse centered at ω0; Below:
Spectra generated by instantaneous second- and third-order nonlinearities due to optical
rectification (OR), second-harmonic generation (SHG), self-phase modulation (SPM), and
third-harmonic generation (THG).
2. PULSE DYNAMICS
∂ ∂2
a = −iD (z ) 2 a + iδ (z ) a a ,
2
(1)
∂z ∂t
that the ordering of the pulse-shaping elements within the cavity has a major
effect on the pulse formation [18]. The discrete action of linear dispersion in
the arms of the laser resonator and the discrete, but simultaneous, action of
positive SPM and positive GDD in the laser crystal cannot be neglected any
longer. The importance of the strong variations in dispersion was first
discovered in a fiber laser and called stretched-pulse mode locking [19]. The
positive dispersion in the Er-doped fiber section of a fiber ring laser was
balanced by a negative-dispersion passive fiber. The pulse circulating in the
ring was stretched and compressed by as much as a factor of 20 in one round
trip. One consequence of this behavior was a dramatic decrease of the
nonlinearity and thus increased stability against the SPM-induced
instabilities. Also, the formation of Kelly sidebands [20] occurring for large
nonlinear phase shifts per round trip in conventional soliton lasers was
greatly suppressed. The energy of the output pulses could be increased a
hundredfold [21]. These important consequences can explain the high power
densities that can be achieved in sub-10 fs KLM lasers without undergoing
catastrophic self-focusing in the laser crystal.
a) Silver
Double-Chirped Mirror I
Mirror 1mm BaF2
Laser crystal:
2mm Ti:Al 2O3
PUMP
L = 20 cm
Double-Chirped Mirror II
OC 2
BaF2 - wedges
Base Length = 30cm for 82 MHz Laser
b)
D D
- D, SPM -
2 2
lasers on the spectral shape of the laser pulses. The dispersion distribution
shown in Figure 2-3(b) suggests an analogy with pulse propagation along a
dispersion-managed fiber transmission link [4]. A system with sufficient
variation of dispersion can support waves that have been called nonlinear
Bloch waves [22]. One can show that the Kerr nonlinearity produces a self-
consistent nonlinear scattering potential that permits formation of a periodic
solution with a simple phase factor in a system with zero net dispersion. It
has been shown further that nonlinear propagation along dispersion-managed
fiber near zero net GDD possesses a narrower spectrum in the segment of
positive dispersion than in the segment of negative dispersion [4, 22]. Thus,
the effect of negative dispersion is greater than that of the positive
dispersion, and the pulse sees an effective net negative dispersion. This
effective negative dispersion can balance the Kerr-induced phase, leading to
steady-state pulses at zero net dispersion. This is true even when there is no
nonlinearity in the negative dispersion segment. The pulses are analogous to
solitons in that they are self-consistent solutions of the Hamiltonian
(lossless) problem as are conventional solitons. But, they are not secant
hyperbolic in shape. Figure 2-4 shows a numerical simulation of a self-
consistent solution of the Hamiltonian pulse-propagation problem in a linear
medium of negative dispersion and subsequent propagation in a nonlinear
medium of positive dispersion with positive self-phase modulation following
Equation (1).
The dispersion map D(z) is shown as an inset in Figure 2-5. The
dispersion coefficient D(z) and the nonlinear coefficient δ(z) are defined per
unit length. In Figure 2-4, the steady-state intensity profiles are shown at the
center of the negative dispersion segment over 1000 round trips. In addition,
we can include in the model the saturable gain, Lorentzian gain filtering, and
KLM modeled by a fast saturable absorber. Figure 2-5 shows the behavior in
one period (one round trip through the resonator) including these effects.
( )
The response of the absorber is q(a ) = q0 1 + a PA , with q0 = 0.01/mm
2
Figure 2-4. Simulation of the Hamiltonian problem. Temporal (a) and spectral (b) intensity
profiles at the center of the negatively dispersive segment are shown for successive round
trips. The total extent in 1000 round trips. D = D (± ) = ±60 fs/mm for the segment of crystal
length L = 2 mm, τ FWHM = 5.5 fs, δ = 0 for D < 0, δ = 1 (MW-mm) for D > 0 .
2. FEMTOSECOND LASER DEVELOPMENT 61
Figure 2-5. Pulse shaping in one round trip. The negative segment has no nonlinearity.
After the pulse leaves the crystal, it experiences GVD in the arms of the
laser resonator. The positively chirped pulse is compressed to its transform
limit at the end of each arm, where output couplers can be placed.
Propagation back towards the crystal imposes a negative chirp, generating
the time-reversed solution of the nonlinear Schrödinger Equation (1).
Therefore, subsequent propagation in the nonlinear crystal compresses the
pulse spectrally and temporally to its initial shape in the center of the crystal.
The spectrum is narrower in the crystal than in the negative-dispersion
sections, because it is negatively prechirped before it enters the SPM section.
Spectral spreading occurs again only after the pulse has been compressed.
This result further explains that in a laser with a linear cavity, for which the
negative dispersion is located in only one arm of the laser resonator (i.e., in
the prism pair and no use of chirped mirrors), the spectrum is widest in the
arm that contains the negative dispersion [18]. In a laser with a linear cavity,
for which the negative dispersion is equally distributed in both arms of the
cavity, the pulse runs through the dispersion map twice per round trip. The
pulse is short at each end of the cavity, and, most importantly, the pulses are
identical in all passes through the crystal; thus they can exploit the full KLM
action twice per round trip [23]. For this reason, a symmetric dispersion
distribution may lead to an effective saturable absorption that is twice as
strong as an asymmetric dispersion distribution and that can produce
substantially shorter pulses. Furthermore, the dispersion swing between the
negative and positive dispersion sections is only half as large, which allows
62 Chapter 2
Figure 2-6. Energy of the pulse in the lossless dispersion-managed system with stretching
S = LD τ FWHM
2
or, for a fixed crystal length L and pulse width as parameters;
D = 60 fs2/mm for Ti:sapphire at 800 mm.
(or pulse width with fixed crystal length L). One can see that for a pulse
width longer than 8 fs with crystal length L = 2 mm, no solution of finite
energy exists in the dispersion-managed system for zero or positive net
dispersion. Pulses of durations longer than 8 fs require net negative
dispersion. Hence, one can reach the ultrashort operation at zero net
dispersion only by first providing the system with negative dispersion. At the
same energy, one can form a shorter pulse by reducing the net dispersion,
provided that the 8 fs threshold has been passed. For a fixed-dispersion
swing ± D, the stretching increases quadratically with the spectral width or
the inverse pulse width. Long pulses with no stretching have a sech shape.
For stretching ratios of 3–10, the pulses are Gaussian shaped. For even larger
stretching ratios, the pulse spectra become increasingly more flat topped, as
shown in Figure 2-4.
These dynamics explain why there are steady-state pulses even at zero
and slightly positive linear intracavity dispersion as confirmed in the
experiment. These pulses would, of course, be unstable in the presence of
2. FEMTOSECOND LASER DEVELOPMENT 63
gain filtering due to the finite gain bandwidth of Ti:sapphire and, therefore,
there is a need for strong-saturable-absorber action to keep the pulse stable
against continuum radiation exploiting the peak of the gain. This saturable-
absorber action is due to the Kerr-lens effect and can be optimized by
various techniques [23, 24], which shall not be repeated here. Also, the
spectral shape of the output pulses is greatly distorted because of the finite
bandwidth of the output coupler, which can greatly enhance the wings of the
laser spectrum, as we will see in the next section.
3. OCTAVE-SPANNING LASERS
crystal. This damage takes some time to occur but can limit the duration of
experiments. In addition, the output spatial mode exhibits strong frequency
dependence, which can also impact experiments.
Intracavity continuum generation has been demonstrated in a Ti:sapphire
laser that was operated in the self-Q-switching regime [28]. However, in this
regime, the spectrum is unstable and therefore not useful for femtosecond
comb applications. As described in Chapter 3, intracavity continuum
generation has also been observed in high-repetition-rate lasers [29], but
there was insufficient power at the octave points to implement ν-to-2ν
stabilization. A 2ν-to-3ν scheme had to be used [30].
Finally, we would like to comment on the definition of “octave
spanning.” The span of a spectrum is often taken as the width at some power
below the peak (often full-width at half-maximum or perhaps even at the 10
dB points). However, for femtosecond comb applications, there is a good
operational definition of octave spanning, namely that it is possible to obtain
ν-to-2ν beats. A slightly stronger version of this is that the beats are
sufficiently strong to be used to stabilize the offset frequency of the laser.
This criterion can be met even when the intensity at the octave points is as
much as 40 dB below the peak.
3.1.1 Design
OC
EM
P2 P1
PUMP
M3
M1 M2
BBO
filter
(580 nm)
polarizer fo locking
PMT electronics
*
This information is given for technical purposes only and does not represent an endorsement
on the part of the National Institute of Standards and Technology.
66 Chapter 2
101515), and a protected silver mirror were tested and yielded octave-
spanning spectra. Also, two OC’s (Spectra Physics G034-007 and CVI Laser
PR2 — 2% transmission @ 800nm) resulted in an octave of bandwidth. We
have not been able to reproduce the octave spectrum using fused silica
prisms. Apparently, in this case, the continuum generation does not result
from meticulous dispersion compensation. This is not surprising since the
continuum is created on a single pass through the crystal. As a result, careful
dispersion compensation may only be necessary over the bandwidth
indicated by the OC reflectivity and not the entire spectrum.
The position of the curved mirrors is critical for producing an octave-
spanning spectrum. Typically, KLM is obtained by displacing one curved
mirror away from the optimum mirror position for cw operation. Here, the
mirror furthest from the pump is translated inward to the point that the laser
is on the edge of stability for cw operation. This arrangement increases the
discrimination between mode-locked and cw operation and requires the
formation of a strong Kerr-lens to correct for the misalignment with respect
to the pump. To optimize the bandwidth, both curved mirrors are then
translated in the same direction away from the pump, just before the point
where the laser begins Q-switching.
ν-to-2ν beats
fr
30 dB
Frequency
Figure 2-8. Rf spectrum showing a repetition rate peak and ν-to-2ν heterodyne beats.
0.30
5
10 0.25
0.20
0
10
0.15
-5 locked 0.10
10
0.05
0.00
-3 -2 -1 0 1 2 3 4 5
10 10 10 10 10 10 10 10 10
Frequency (Hz)
Figure 2-9. Power spectral density of phase noise (left axis) and accumulated rms phase
fluctuations (right axis).
away from the resonant Gaussian TEM00 mode and its corresponding 1/λ
diffraction limit. G
587nm 612nm 664nm 690nm
4 10
-1
Spot Size (mm)
sagittal plane
tangential plane -2
3 10
-3
10
2
-4
10
1 10
-5
Figure 2-11. Measured sagittal and tangential spot sizes as a function of wavelength (squares,
left axis). Line shows laser spectrum (right axis).
OC Transmission (%)
60
6
5 40
4
20
3
2
0
600 700 800 900 1000 1100
Wavelength (nm)
Figure 2-12. Waist size at output coupler as a function of wavelength for sagittal and
tangential planes (squares, left axis). Output coupler transmission is line (right axis).
Beam Divergence (mrad)
10
sagittal plane
tangential plane
8
Figure 2-13. Measurement of beam divergence (squares) and predicted divergence based on
waist size (lines).
The mode profiles and the sudden decrease in waist in the extremes of
the spectrum still need to be explained. For resonant modes, the cavity
geometry determines the waist size and divergence. The beam characteristics
of the nonresonant modes, however, can be directly influenced by the Kerr
lens in the crystal and by diffraction and aperturing of optics without the
self-correcting effects of the cavity. The Kerr lensing is considerable since
intracavity continuum generation, for this laser, requires significant
realignment away from ideal cw operation. This realignment results in a
highly asymmetric cw beam (tangential waist/sagittal waist = 1.56) and a
large discrimination in average output power between cw and mode-locked
operation (4:1). The asymmetry implies that the nonlinear correction for the
cw asymmetry may result in the formation of an asymmetric Kerr lens,
which may explain the observed asymmetry in the wings of the spectrum.
This mirror positioning also results in a different initial cw condition than for
70 Chapter 2
3.2.1 Design
operation, the average power is 90 mW. The output power is limited by low
output coupling and scattering losses in the OC coating and the silver mirror.
Presumably, further optimization of output coupling and elimination of the
silver mirror would increase laser efficiency.
a) b)
PUMP 1.0 100
ϕ 0.8 80
Figure 2-14. (a) Schematic diagram of octave-spanning prismless laser. Gray and black
mirrors are type I and II double-chirped mirrors (DCMs), respectively. The BaF2 wedges are
used for fine-tuning of the dispersion. (b) Reflectivity (left scale) of the type I DCMs with
pump window shown as thick solid line. The group-delay design is given by the thick, dash-
dotted line. The individual group delays (right scale) of type I and II DCMs are shown as thin
lines and the average as a dashed line (almost identical to the design goal over the wavelength
range of interest from 650–1200 nm). The measured group delay, using white-light
interferometry, is shown as the thick solid line from 600–1100 nm. Beyond 1100 nm, the
sensitivity of the Si detector limited the measurement.
In one round trip of the laser pulse through the cavity, the total number of
bounces on DCMs (12 bounces) generates the precise negative dispersion
required to compensate for positive second- and higher-order dispersion
caused by the laser crystal, the air path in the cavity, and the BaF2 wedge
pairs used to fine-tune the intracavity dispersion.
We used BaF2 for dispersion compensation because it has the lowest ratio
of third- to second-order dispersion in the wavelength range from 600–1200
nm, and the slope of the dispersion of BaF2 is nearly identical to that of air.
These features make it possible to scale the cavity length and repetition rate
without changing the overall intracavity dispersion. To achieve mode-locked
operation, it is necessary to reduce the amount of BaF2 inside the laser cavity
(by withdrawing one of the wedges).
The broadest spectrum can be achieved by optimizing the insertion of the
BaF2 wedge, so that the laser operates very close to the zero average-
dispersion point. In this case, the spectral width of the laser is critically
dependent on the dispersion balance. With the prismless lasers, adjusting the
dispersion does not significantly change the cavity alignment. Figures 2-
15(a) and (b) show the spectrum under broadband operation of an 80 MHz
and a 150 MHz laser, respectively. The octave is already reached at a
72 Chapter 2
spectral density about 25 dB below the average power level. The same plots
also show the corresponding OC transmission curves. The detailed shape of
this transmission curve is a determining factor in the width of the output
spectrum, since it significantly enhances the spectral wings. A roll-off of the
output coupler matched to the intracavity spectrum is necessary to generate
the octave-spanning spectrum.
a)
Power Spectrum (dB)
OC Transmission (%)
80
-10
-20 60
-30 40
-40 20
-50
600 800 1000 1200
Wavelength (nm)
Figure 2-15. (a) Measured output spectra for lasers with 80 MHz repetition rate and (b) 150
MHz repetition rate. In both cases, the octave is reached at approximately 25 dB below the
average power level. The average mode-locked power is 90 mW for both lasers. Also shown
(dashed lines) are the respective output coupler (OC) transmission curves for the ZnSe/MgF2
Bragg-stack (a) and a broadband SiO2/TiO2 stack (b). Both have about 1% transmission at
800 nm.
following the PMT, which is 8MHz, for more than 10 hours. On the time
scale of seconds, the beat can jitter by about 100 kHz. The f0 lock only
breaks because of a slow drift of the beat note out of the filter bandwidth of
the 8 MHz filter, which is not yet controlled.
(a)
SM
AOM
OC
Loop
filter 580 nm
DM
580 nm Filter
Digital
Phase detector
PMT
DM
λ/2 plate BBO
1160 nm
LO
(b)
RF Spectrum [dB]
-40
-80
6
0 20 40 60 80x10
Frequency [Hz]
Figure 2-16. (a) Setup for f0 detection and lock. (b) Measured carrier-envelope beat signal
from the 80 MHz laser.
Figure 2-17. Spectral densities of the phase-error signal at the output of the digital phase
detector for the f0 lock (solid line) and free running f0 (dashed line).
5. SUMMARY
ACKNOWLEDGEMENTS
REFERENCES
[1] D. E. Spence, P. N. Kean, and W. Sibbett, Opt. Lett. 16, 42-44 (1991).
[2] D. K. Negus, L. Spinelli, N. Goldblatt, and G. Feugnet, in Advanced Solid-
State Lasers (OSA, 1991), Vol. 10; F. Salin, J. Squier, and M. Pichè, Opt.
Lett. 16, 1674-1676 (1991); M. Pichè, Opt. Commun. 86, 156-160 (1991).
[3] H. A. Haus, IEEE J. Quantum Electron. 11, 736-746 (1975).
[4] J. H. B. Nijhof, N. J. Doran, W. Forysiak, and F. M. Knox, Electron. Lett.
33, 1726-1727 (1997).
[5] K. Tamura, H. A. Haus, and E. P. Ippen, Electron. Lett. 28, 2226-2228
(1992).
[6] Y. Chen, F. X. Kärtner, U. Morgner, S. H. Cho, H. A. Haus, E. P. Ippen,
and J. G. Fujimoto, J. Opt. Soc. Am. B 16, 1999-2004 (1999).
[7] J. P. Zhou, G. Taft, C. P. Huang, M. M. Murnane, H. C. Kapteyn, and I. P.
Christov, Opt. Lett. 19, 1149-1151 (1994).
[8] A. Stingl, C. Spielmann, F. Krausz, and R. Szipocs, Opt. Lett. 19, 204-206
(1994).
[9] D. H. Sutter, G. Steinmeyer, L. Gallmann, N. Matuschek, F. Morier-
Genoud, U. Keller, V. Scheuer, G. Angelow, and T. Tschudi, Opt. Lett. 24,
631-633 (1999).
[10] U. Morgner, F. X. Kärtner, S. H. Cho, Y. Chen, H. A. Haus, J. G. Fujimoto,
E. P. Ippen, V. Scheuer, G. Angelow, and T. Tschudi, Opt. Lett. 24, 411-
413 (1999).
[11] R. Ell, U. Morgner, F. X. Kärtner, J. G. Fujimoto, E. P. Ippen, V. Scheuer,
G. Angelow, T. Tschudi, M. J. Lederer, A. Boiko, and B. Luther-Davies,
Opt. Lett. 26, 373-375 (2001).
[12] L. Xu, C. Spielmann, A. Poppe, T. Brabec, F. Krausz, and T. W. Hänsch,
Opt. Lett. 21, 2008-2010 (1996).
[13] A. Baltuška, T. Fuji, and T. Kobayashi, Phys. Rev. Lett. 88, art. no.-133901
(2002).
[14] L. Xu, C. Spielmann, F. Krausz, and R. Szipocs, Opt. Lett. 21, 1259-1261
(1996).
[15] I. P. Christov, H. C. Kapteyn, M. M. Murnane, C. P. Huang, and J. P. Zhou,
Opt. Lett. 20, 309-311 (1995).
[16] I. P. Christov, V. D. Stoev, M. M. Murnane, and H. C. Kapteyn, Opt. Lett.
21, 1493-1495 (1996).
[17] I. P. Christov and V. D. Stoev, J. Opt. Soc. Am. B 15, 1960-1966 (1998).
[18] C. Spielmann, P. F. Curley, T. Brabec, and F. Krausz, IEEE J. Quantum
Electron. 30, 1100-1114 (1994).
[19] K. Tamura, E. P. Ippen, H. A. Haus, and L. E. Nelson, Opt. Lett. 18, 1080-
1082 (1993).
2. FEMTOSECOND LASER DEVELOPMENT 77
Albrecht Bartels
Time and Frequency Division, National Institute of Standards and Technology
1. INTRODUCTION
Figure 3-1. Illustration of the advantages for frequency metrology of the frequency comb
emitted by a 1 GHz femtosecond laser (left) compared to that of a 100 MHz laser (right). The
length of the two sets of comb teeth (solid vertical lines) scales with the power per mode. The
lengths of the gray horizontal bars represent the necessary resolution for premeasurement of
the frequency to be measured (fT, indicated by the dashed line) to determine the index k of the
mode against which fT is beat.
vg
fr = . (1)
L
Thus, increasing the repetition rate means shrinking the cavity length. The
realization of gigahertz repetition rates for Ti:sapphire femtosecond lasers
has been straightforward. A few simple specific design criteria now exist.
They can be discussed in the context of a steady-state duration τ of a pulse
circulating in a cavity in which only SPM, provided by the nonlinear
refractive index n2 ≈ 3×10-16m2/W of the Ti:sapphire gain crystal, and
negative group-delay dispersion (GDD), designated as D2, shape the pulse
(see, for example, Reference [11]). This limit applies in good approximation
to all Kerr-lens–mode-locked femtosecond lasers as long as higher-order
dispersion is negligible:
(
τ = 4 ln 1 + 2 ) dn E
D2 λ w02
, (2)
2 P
where λ is the carrier wavelength, w0 is the waist radius inside the laser
crystal, d the length of the crystal, and EP the pulse energy.
When shrinking a 100 MHz cavity to yield a 1 GHz repetition rate, the
pulse energy E P is reduced by a factor of 10 (assuming constant average
power). This results in a steady-state pulse duration 10 times longer. The
combination of lower pulse energy and greater pulse duration reduces the
peak by a factor of 100. This alone does not present a fundamental problem
for achieving higher repetition rates. However, in practice Kerr-lens mode
locking (KLM) is required to maintain stable pulsed operation over cw
operation. A Kerr lens is induced by a high-intensity Gaussian beam profile
in the Ti:sapphire crystal via its nonlinear refractive index. To achieve KLM,
the cavity is arranged in a way that the Kerr lens modulates the net gain or
loss of the cavity for a pulsed beam. Either it reduces the net cavity losses by
increasing the transmission through a hard aperture (e.g., a slit) at an
appropriate position in the resonator (hard-aperture KLM), or it increases the
net gain by increasing the overlap with the finite gain volume inside the
crystal (soft-aperture KLM). KLM gets more efficient with higher peak
intensity. (High pulse energies and short pulse widths in the cavity are
essential.) Thus the reduction of peak power due to the increased repetition
rate must be compensated for. This is accomplished by reducing the amount
of output coupling with respect to a 100 MHz oscillator (from ≈10% to ≈1%)
and shrinking the waist radius w0 inside the gain crystal by tighter focusing.
The second major concern is the achievement of negative net-GDD in the
cavity. The standard method of employing a four-prism sequence (or a
82 Chapter 3
Figure 3-2. (a) Four-mirror cavity with the gain crystal enclosed by focusing mirrors M1 and
M2. The flat mirror M3 and the output coupler (OC) form the collimated part of the cavity.
The pump light enters the crystal through lens L and M1. (b) Equivalent to (a) but with extra
folding using flat mirrors M4 and M5.
The ring oscillators shown in Figure 3-2 are the basis for all lasers
presented here. A ring configuration has been chosen for geometrical
reasons. With a given number of components, it allows a higher repetition
rate than a linear cavity. Also, ring cavities are less sensitive to back-
reflections. In the following sections, the cavities will be discussed in detail
along with actually realized lasers.
As with all Kerr-lens–mode-locked femtosecond lasers, these oscillators
need an initial perturbation, such as tapping a mirror mount, to generate an
intracavity power fluctuation that builds up to a stable circulating
femtosecond pulse. As opposed to cw operation, which is bidirectional,
mode-locked operation is unidirectional and starts in a random direction
[14].
3. GIGAHERTZ FEMTOSECOND LASERS 83
1.5
1.0
0.5
0.0
0 2 4 6 8 10
pump power (W)
Figure 3-3. Laser output power as a function of incident pump power for the four-mirror
(triangles) and six-mirror (squares) lasers described in the text. The solid lines are linear fits
to the datasets.
Figure 3-4. (a) Interferometric autocorrelation trace of the pulses from the four-mirror laser at
5.5 W pump power. Inset: Pulse duration as a function of intracavity pulse energy (squares)
and theoretical values according to Equation (2). (b) Corresponding output spectrum. Inset:
Series of output spectra of a continuously tunable version of the six-mirror cavity [16].
through the beam profile with a fitted Gaussian, again with no noticeable
deviation from ideal. A measurement of the M2-factor has been performed,
yielding values of 1.1 for both tangential and saggital planes. (See Reference
[17] for a definition of M2.) The excellent beam quality is attributed to the
fact that spatially dispersive elements are absent in the cavity.
With the four-mirror cavity, repetition rates between 1 GHz and 3.5 GHz
were attained. The upper limit was set by geometrical constraints, i.e., it has
not been possible to move the components closer together. It is anticipated
that even higher repetition rates of up to 10 GHz are feasible by minimizing
mechanical components and further reducing the focal lengths of M1 and
M2. The disadvantage of this cavity is that the short-gain crystal results in a
relatively low output power.
Figure 3-5. (a) Mode profile of a four-mirror laser at the focus of a 200 mm lens. (b) Cross
section (squares) through the profile shown in (a) and fit with a Gaussian (solid line).
The two oscillators described in this section are widely used frequency
comb generators for optical frequency metrology. They are commonly
operated between 0.8 GHz and 1 GHz. This range has proven to be the best
compromise between high-repetition rate and sufficient pulse energy to
generate an octave-spanning spectrum in common microstructure fibers.
(See e.g., Reference [18] for a typical octave-spanning spectrum from a 1
GHz four-mirror laser and a microstructure fiber.)
The six-mirror cavity has also been realized in a wavelength-tunable
modification at a repetition rate of 1 GHz [16]. A half-Brewster prism was
used to introduce spatial dispersion into the resonator with the effect that
horizontally tilting a cavity mirror allowed continuous tunability from 733 to
850 nm with pulse durations of ≈40 fs. A series of output spectra is shown in
the inset of Figure 3-4(b).
It is a logical step to extend the concept that has successfully been used
with Ti:sapphire to gain media that operate at longer wavelengths. One
reason is to obtain access to the telecommunication bands and thus be able to
calibrate and characterize system components, such as wavelength
references or wavelength-division multiplexing filters, in that region. A
second goal is to generate a mode-locked spectrum that is adjacent to and
that can be phase coherently combined with the output spectrum of a
Ti:sapphire laser (specifically the one described in the next section) to
generate an extraordinarily broad “superfrequency comb.” The gain material
that is best suited for both tasks is Cr:forsterite, which has a gain spectrum
extending from wavelengths of ≈1150 to 1350 nm.
A femtosecond laser with Cr:forsterite was realized at a repetition rate of
433 MHz in the six-mirror configuration [see Figure 3-2(b)] [19]. The 10
mm gain crystal with an absorption coefficient of α=1.1 cm-1 at 1075 nm is
significantly longer than in the Ti:sapphire lasers to allow for sufficient
pump light absorption. The crystal is pumped with 10 W from an Yb:glass
fiber laser emitting at 1075 nm through a lens with fL = 40 mm. The crystal is
cooled to 0° C to increase the gain and to reduce problems with thermal
lensing. (Cr:forsterite has a thermal conductivity ≈5 times lower than
Ti:sapphire.) The focal lengths of mirrors M1 and M2 are 25 mm; they are
longer than what has been used for the Ti:sapphire lasers to better match the
confocal length of the cavity mode to the length of the gain crystal. With
chirped mirrors at positions M1–M3, a GTI mirror at M4, a low-dispersive
high reflector at M5, and the GDD of the gain crystal of 185 fs2 [20], the net
GDD is ≈ -260 fs2. A 1.5% transmission output coupler at 1280 nm is used.
3. GIGAHERTZ FEMTOSECOND LASERS 87
1
10
0
10
-1
10
-2
10
-3
10
900 1200 1500 1800 2100
wavelength (nm)
Figure 3-6. Output spectrum of the Cr:forsterite laser (solid line) and spectrum after
broadening in a highly nonlinear Ge-doped silica fiber (HNLF) (dashed line).
The output from this laser was successfully broadened via SPM in highly
nonlinear Ge-doped silica fiber (HNLF) to form a frequency comb that spans
more than one octave [19, 21], as shown in Figure 3-6. This frequency comb
has been referenced to the National Institute of Standards and Technology’s
(NIST’s) Ca optical frequency standard and a cesium atomic clock. Its utility
for optical frequency measurements was demonstrated by Corwin et al. [22]
who used the broadened Cr:forsterite comb for the characterization of
telecommunication-band frequency standards.
The creation of a phase-coherent superfrequency comb, ranging from
wavelengths of 570 to 1450 nm, was demonstrated by actively linking the
output of the Cr:forsterite laser, described above, and a broadband
Ti:sapphire laser described in the next section [23].
88 Chapter 3
Figure 3-7. Left panel: Net cavity group-delay dispersion (GDD) of the broadband laser.
Right panel: Reflectance of the chirped mirrors (solid line) and the output coupler (dashed
line).
The output spectrum of the laser is shown in Figure 3-8. It spans from
570 to 1100 nm at a power level of 1 nW per 1 GHz frequency mode. (1 nW
per mode is the power typically required to measure a beat note against a
milliwatt-scale cw laser with sufficient signal-to-noise ratio for a frequency
measurement.) The spectrum has pronounced maxima at 670 nm, 845 nm,
900 nm, and 927 nm. The peak around 670 nm can be isolated by a double
reflection off a pair of 657 nm high reflectors, leading to the spectrum shown
in the inset of Figure 3-8(b). It contains 450 mW of average power.
The duration of the output pulses has only been roughly characterized by
using an intensity autocorrelation measurement. Assuming a temporal sech2
shape, it is 11 fs when the nonlinear crystal is angle-tuned to optimally phase
match the central part of the spectrum. The peak around 670 nm yields
bandwidth-limited pulses of 33 fs after isolation from the longer-wavelength
part of the spectrum.
3. GIGAHERTZ FEMTOSECOND LASERS 89
Figure 3-8. (a) Output spectrum of the broadband laser on a logarithmic scale (solid line). At
a level of 1 nW per 1 GHz mode, it extends from 570 to 1100 nm. The dashed line shows a
simulated spectrum, offset by 20 dB for clarity. (b) Output spectrum of the broadband laser on
a linear scale. Inset: spectrum after reflection off a pair of 657 nm high reflectors.
approach has been taken with great success by Ell et al. [24]. Their carefully
engineered intracavity dispersion is capable of generating spectra exceeding
one octave with 5 fs pulses at a repetition rate of 64 MHz. In our case,
however, because of the narrow bandwidth of the chirped mirrors, both with
respect to their dispersive and reflective properties, this limit is inapplicable.
Alternatively, one can allow higher-order dispersion in the resonator but, at
the same time, ensure that the leading and trailing edges of the temporally
spreading spectrum are sufficiently suppressed, such that only a short pulse
remains stable in the cavity. This effect can be attained by employing an
effective fast saturable absorber. In the case of our Ti:sapphire lasers, this
effective saturable absorber is provided by a soft aperture KLM (see Section
2.1). While this effect is actually a self-gain modulation, it can theoretically
be treated as an equivalent fast saturable absorber. These effects are more
generally referred to as self-amplitude modulation (SAM).
There are strong indications that the broadband laser described above
operates in the limit of an increased SAM. The strongest support for this idea
is the experimental observation that the mode-locked output power and the
continuous output power of the laser differ by more than an order of
magnitude. This difference indicates that the Kerr-lens-induced effective
saturable absorber has a saturable absorption of approximately 30%.
Theoretical calculations of the change in beam waist diameter inside the gain
medium (i.e., the soft-aperture Kerr-lens effect) show that replacing flat
mirror M3 with a slightly convex mirror (in our case with a radius of
curvature of 1 m) can increase the SAM of pulses circulating in the ring
cavity [25]. While the theory applied to calculate this effect is valid only at
power levels of about a factor of four below the power at which our
oscillator operates, it still has provided a clue to understanding this new type
of femtosecond oscillator.
A full theory explaining why the convex mirror leads to a stronger
effective saturable absorber does not yet exist. However, modeling the
resonator with a split-step Fourier simulation allows us to understand why
this laser functions. The model includes the Ti:sapphire gain spectrum and
dispersion, all reflective amplitude and phase properties of the cavity
mirrors, SPM, and a fast saturable absorber (of the same type as used by
Chen et al. [26]). The dashed curve in Figure 3-9 shows a simulation of a
cavity with the same mirrors as the broadband laser cavity but with a shorter
crystal (d = 1.5 mm) and a saturable absorption of qsat. = 2%. Its net GDD at
800 nm is -45 fs2. It represents a standard laser with a relatively narrow
spectrum of 45 nm bandwidth. An attempt to get shorter pulses from this
laser by increasing the net cavity GDD to ≈ -20 fs2 by using a longer crystal
(d = 2.0 mm) fails because the simulation does not yield a stable solution
anymore. Higher-order dispersion spreads the seed pulse and randomizes its
3. GIGAHERTZ FEMTOSECOND LASERS 91
1
10
0
10
-1
10
-2 d=1.5 mm, qsat.=0.02
10 d=2.0 mm, qsat.=0.3
-3
10
600 700 800 900 1000 1100
wavelength (nm)
Figure 3-9. Modeled output spectra of a standard laser (d = 1.5 mm, qsat.= 0.02) and a
broadband laser (d = 2.0 mm, qsat. = 0.3) using a split-step Fourier simulation.
has been SPM in microstructure fibers. However, coupling light into the core
of fibers with a diameter less than 2 µ m is a difficult task. As SPM is a third-
order nonlinearity, slight changes in coupling efficiency can result in large
changes of the attained spectrum and usually limit the time over which a
measurement of f0 can be maintained with sufficient signal-to-noise ratio.
Apart from that, microstructure fibers also limit the amount of useful
average power achievable for the continuum spectrum because the nonlinear
broadening process inherently amplifies both technical and shot noise
present on the input light. (References 10 and 27 provide an overview on
noise amplification during supercontinuum generation in microstructure
fibers.)
The broadband laser allows one to circumvent these problems with
microstructure fiber because its broad spectral coverage allows for a direct
measurement of its offset frequency. Although the output spectrum has
wavelength components that are one octave apart, the power contained in
them is too low to facilitate a measurement of f0 using the standard ν-to-2ν
method. Therefore, we employ a 2ν-to-3ν method: A portion around 930 nm
is frequency tripled and beat against the second harmonic of a portion
around 620 nm to yield a signal at f0. Details of this setup can be found in
Ramond et al. [28].
Although the 2ν-to-3ν method is more complicated than the standard
method, the broadband laser has an unprecedented long-term stability as a
frequency measurement tool. First, it routinely stays mode locked for periods
of many days (>1 week demonstrated) without noticeable changes to its
output power or spectrum. Second, the f0 signal usually has a signal-to-noise
ratio of 25–30 dB in the 300 kHz bandwidth and does not degrade over time
because of the absence of microstructure fiber in the measurement apparatus.
In a test of its long-term stability, we have operated the laser in an optical
clock configuration. Here, the repetition rate fr of the broadband laser is
phase locked to a cavity-stabilized single-frequency laser diode at frequency
fLD = 456 THz that represents the optical frequency standard that would be
used in an optical clock. Specifically, a beat between the laser diode and the
neighboring component of the frequency comb with frequency fb has been
phase locked to a synthesizer at 600 MHz by feedback to the femtosecond
laser’s cavity length via a mirror mounted onto a piezoelectric transducer.
(All synthesizers were referenced to the NIST primary standard.)
Additionally, f0 was phase locked to a synthesizer at 100 MHz by feedback
to the pump power via an acousto-optic modulator. It can be shown that fr
now derives its frequency and instability entirely from the laser diode or, in
the case of an optical clock, from the optical frequency standard and is
effectively the microwave output of the optical clock [23]. We then counted
f0, fb and fr simultaneously with frequency counters at 10 s gate time. The
3. GIGAHERTZ FEMTOSECOND LASERS 93
offset of the counter readings from their preset values (f0 = 100 MHz, fb =
600 MHz, set by the synthesizers; fr = 998,092,449.54 Hz, defined by fLD and
the choices of f0 and fb) over a period of 21 hours are displayed in Figure 3-
10. The data set shows uninterrupted and hands-off operation of the f0 phase
lock for 21 hours with counter-resolution-limited deviations on the millihertz
scale until the system was turned off. The phase lock on fb operated similarly
for ≈14 h with only one detected feedback-loop error (cycle slip) at ≈4 h, i.e.,
≈10 h of continuously phase-locked data were obtained. The failure at ≈14 h
was likely caused by the failure of the stabilization of fLD to the Fabry-Perot
cavity that prevented the feedback loop from tracking the diode laser
frequency rather than failure of the femtosecond laser itself. As fr represents
a measurement of the laser diode frequency (fLD = f0 + nLD fr + fb, where nLD is
the mode number of the frequency comb component against which fLD is
beat), the time record of the offset of fr from 998,092,449.54 Hz can be
multiplied with nLD = 456,857 to give a record of the temporal drift of the
Fabry-Perot cavity on the right axis of the graph in Figure 3-10. The cycle
slip in the fb feedback loop does not appear in the fr record because the effect
of the 120 mHz excursion in fb results in an error of 260 nHz in fr, which is
below our measurement limit.
These results show that the broadband laser facilitates a phase-coherent,
cycle-slip-free link between an optical oscillator at 456 THz and the 1 GHz
laser repetition rate for 10 h. In other words, we have the ability to count
1.6×1019 optical cycles at 456 THz without ever losing track of the
oscillation. It has also been shown that fr and f0 of the broadband laser can be
continuously phase locked to a synthesizer for more than 48 hours.
4. CONCLUSION
Ti:sapphire ring oscillators at 1 GHz repetition rate are currently the best
available frequency comb generators for optical frequency metrology. They
are compact, have a conveniently large comb spacing, and yield 100 times
more power per mode than conventional 100 MHz lasers. Their simple ring-
laser architecture employing negative dispersion mirrors for GDD control
allows repetition rates between 300 MHz and 3.5 GHz with Ti:sapphire. The
concept is extendable to other gain media, as has been demonstrated with
Cr:forsterite at 433 MHz repetition rate.
94 Chapter 3
5
a)
f0 - 100 M Hz
(m Hz)
0
-5
b)
f b - 600 M Hz
100
(m Hz)
50
0
f R - 998,092,449.54 Hz
30 0
0.6 c)
f LD Drift (kHz)
0.4 20 0
(Hz)
0.2 10 0
0.0 0
-0.2 -100
0 2 4 6 8 10 12 14 16 18 20
Time (h)
Figure 3-10. (a) A record of consecutive counter readings at 10 s gate time for the phase-
locked f0. The preset value of 100 MHz has been subtracted. (b) The simultaneously measured
offset of fb from 600 MHz. In this plot, one cycle slip of the phase-locked loop controlling fb
has occurred. (c) A time record of the offset of the laser repetition rate from 998,092,449.54
Hz (left scale) that can be translated into a drift of fLD (right scale).
The most critical development has been the 1 GHz broadband Ti:sapphire
laser that eliminates troublesome microstructure fibers from optical-
frequency-measurement setups. This laser has enabled the construction of an
optical clockwork with an unprecedented continuous operation time in
excess of 48 hours. This laser might also become interesting in basic
research as a source of femtosecond pulses in the range from 620 to 690 nm.
ACKNOWLEDGEMENTS
REFERENCES
[1] D. E. Spence, P. N. Kean, and W. Sibbett, Opt. Lett. 16, 42-44 (1991).
[2] M. Ramaswamypaye and J. G. Fujimoto, Opt. Lett. 19, 1756-1758 (1994).
[3] A. Stingl, C. Spielmann, R. Szipocs, and F. Krausz, in Conference on
Lasers and Electrooptics (Opt. Soc. Am., 1996), p. 66
[4] T. Udem, J. Reichert, R. Holzwarth, and T. W. Hänsch, Phys. Rev. Lett. 82,
3568-3571 (1999).
[5] D. J. Jones, S. A. Diddams, J. K. Ranka, A. Stentz, R. S. Windeler, J. L.
Hall, and S. T. Cundiff, Science 288, 635-639 (2000).
[6] S. A. Diddams, D. J. Jones, J. Ye, S. T. Cundiff, J. L. Hall, J. K. Ranka, R.
S. Windeler, R. Holzwarth, T. Udem, and T. W. Hänsch, Phys. Rev. Lett.
84, 5102-5105 (2000); M. Niering, R. Holzwarth, J. Reichert, P. Pokasov,
T. Udem, M. Weitz, T. W. Hänsch, P. Lemonde, G. Santarelli, M. Abgrall,
P. Laurent, C. Salomon, and A. Clairon, Phys. Rev. Lett. 84, 5496-5499
(2000); S. A. Diddams, T. Udem, J. C. Bergquist, E. A. Curtis, R. E.
Drullinger, L. Hollberg, W. M. Itano, W. D. Lee, C. W. Oates, K. R. Vogel,
and D. J. Wineland, Science 293, 825-828 (2001); J. Stenger, T. Binnewies,
G. Wilpers, F. Riehle, H. R. Telle, J. K. Ranka, R. S. Windeler, and A. J.
Stentz, Phys. Rev. A 63, 021802 (2001); T. Udem, S. A. Diddams, K. R.
Vogel, C. W. Oates, E. A. Curtis, W. D. Lee, W. M. Itano, R. E. Drullinger,
J. C. Bergquist, and L. Hollberg, Phys. Rev. Lett. 86, 4996-4999 (2001); G.
D. Rovera, F. Ducos, J. J. Zondy, O. Acef, J. P. Wallerand, J. C. Knight,
and P. S. Russell, Meas. Sci. Techn. 13, 918-922 (2002).
[7] J. Reichert, R. Holzwarth, T. Udem, and T. W. Hänsch, Opt. Commun.
172, 59-68 (1999).
[8] J. K. Ranka, R. S. Windeler, and A. J. Stentz, Opt. Lett. 25, 25-27 (2000); J.
C. Knight, T. A. Birks, P. S. Russell, and D. M. Atkin, Optics Letters 21,
1547-1549 (1996).
[9] L. Hollberg, C. W. Oates, E. A. Curtis, E. N. Ivanov, S. A. Diddams, T.
Udem, H. G. Robinson, J. C. Bergquist, R. J. Rafac, W. M. Itano, R. E.
Drullinger, and D. J. Wineland, IEEE J. Quantum Electron. 37, 1502-1513
(2001).
[10] K. L. Corwin, N. R. Newbury, J. M. Dudley, S. Coen, S. A. Diddams, K.
Weber, and R. S. Windeler, Phys. Rev. Lett. 90, 113904 (2003).
[11] T. Brabec, C. Spielmann, and F. Krausz, Opt. Lett. 17, 748-750 (1992).
[12] R. L. Fork, O. E. Martinez, and J. P. Gordon, Opt. Lett. 9, 150-152 (1984).
[13] F. Gires and C. R. Tournois, Science 258, 6112 (1964); R. Szipocs, K.
Ferencz, C. Spielmann, and F. Krausz, Opt. Lett. 19, 201-203 (1994).
[14] A. Bartels, T. Dekorsy, and H. Kurz, Opt. Lett. 24, 996-998 (1999).
[15] H. W. Kogelnik, C. V. Shank, A. Dienes, and E. P. Ippen, IEEE J.
Quantum Electron. 8, 373 (1972).
[16] A. Bartels, T. Dekorsy, and H. Kurz, in Conference on Lasers and Electro-
Optics (OSA Technical Digest, 2000), p. CMF3
96 Chapter 3
Abstract: Microstructure fibers have played a key role in the production of coherent
frequency combs that span more than an octave of bandwidth. In this chapter,
we review the fabrication process for such fibers and their linear and nonlinear
optical properties. We also discuss the underlying physical processes that give
rise to supercontinuum generation.
1. INTRODUCTION
Figure 4-1. Left: Drawing of a preform fashioned from a close-pack arrangement of tubes and
rods. Right: A scanning-electron-microscope (SEM) micrograph showing the fiber after the
preform is drawn.
replace or coexist with stacking are extrusion and casting. These processes
have the advantage of not requiring precision tubing and are ideal for
polymer, sol-gel slurry, or low melting-point glass. The main advantage of
extrusion and casting is that complicated structures can be fabricated in
which the position, size, and shape of the air regions are independent of one
another. These methods may become more common as complex air
structures are needed to make advanced microstructure fibers.
Another method consists of drilling holes in a traditional preform or rod.
Drilling is well understood and is used for other specialty fibers. The
advantage of this method is that it is easy to put variously sized holes in any
position in a preform, including doped regions. The disadvantages are that
the holes cannot be drilled very deep compared to a standard preform length;
the distance between holes may be limited due to cracking; and the fiber may
experience high loss due to surface roughness of the holes and impurities
incorporated during drilling.
Figure 4-2. SEM photographs of several types of microstructure fiber. (a) High-nonlinearity
fiber. (b) Air-clad fiber. (c) Endlessly single-mode fiber [4]. (d) Band-gap fiber [6].
Air-clad fibers consist of large air holes closely packed around an inner
cladding containing a doped core [Figure 4-2(b)]. These fibers have inner
claddings with very high numerical apertures and are ideal for generating
efficient, high-power cladding pump amplifiers and lasers.
At the other extreme, fibers with small air holes have been designed [4]
such that the cladding index changes with wavelength to create fibers that
are endlessly single moded [Figure 4-2(c)]. These fibers can have very large
core diameters and therefore can propagate high powers at low intensities.
Fibers with precise, periodically spaced index regions in the cladding can
create band-gap guidance, allowing light to be guided in a core with an index
lower than the cladding [Figure 4-2(d)]. Guidance in an air core has been
demonstrated [5] and has the potential to have an attenuation over an order
of magnitude lower than the best traditional transmission fiber [6].
102 Chapter 4
2
neff = f + (1− f ) nglass
2
, (1)
where f is the air-filling fraction of the cladding region and nglass is the index
of the glass that comprises the structure. The solutions for the modes and the
dispersion using this simple step-index model can be obtained analytically,
and the dispersion for the fundamental mode agrees very well with the
predictions of the full-vector model [7]. It is apparent by calculating the
effective V-number [8] that even for small cores and reasonably large air-
filling fractions (f > 0.5), the fiber is multimode (i.e., V > 2.405).
Nevertheless, since the difference in the propagation constant between the
fundamental mode and the higher-order modes is large, a strong perturbation
is required to couple energy from the fundamental mode to the higher-order
modes. Thus if the fundamental mode is initially excited, the energy does not
easily couple to the higher-order modes [9]. In addition, the large index
contrast also leads to a strong confinement of the fundamental mode with
little energy lying outside the core. This results in a large effective
nonlinearity that allows for strong nonlinear interactions with relatively
modest pulse energies such as those from a femtosecond Ti:sapphire
oscillator.
4. MICROSTRUCTURE FIBER AND WHITE-LIGHT GENERATION 103
Figure 4-3. Plots of the group-velocity-dispersion parameter D for (a) various values of the
diameters d and an air-filling fraction f = 1 and (b) various values of f for a core diameter d =
1.7. The dashed line represents the dispersion curve for bulk-fused-silica glass.
5. SUPERCONTINUUM GENERATION
∂u
∑ i ∂
Lds
= −i sgn( β 2 ) + i 1+ p , (2)
∂ζ n=2 L
( n)
ω 0τ p ∂τ nl
ds
τ
Lds
pnl = (1 − f )|u| + f
2
∫ dτ g (τ − τ ') | u |2 u , (3)
Lnl −∞
where Lnl = (c/ω0 n2I0) is the nonlinear length, I0 = n0c|A0|2/2π is the peak
input intensity, f is the fractional contribution of Raman scattering to the
nonlinear refractive index, and g(τ) is the Raman-response function [12].
The presence of the operator 1+ i∂/ω0τp∂τ in the nonlinear polarization term
accounts for self-steepening effects and allows for the modeling of the
propagation of pulses with spectral widths comparable to the central
frequency ω0.
Figure 4-4. Theoretically predicted supercontinuum spectra and temporal profiles for z = 0.02
cm (a,b) and z = 1.2 cm (c,d).
For the case in which picosecond and nanosecond pulses are used, four-
wave mixing and Raman scattering can produce a broadband
supercontinuum in long fiber lengths [15]. Tailoring the dispersion of the
4. MICROSTRUCTURE FIBER AND WHITE-LIGHT GENERATION 107
fiber can result in control of the continuum such that certain spectral parts of
the continuum are suppressed and others are enhanced [16]. The generated
supercontinuum has also been found to be highly dependent on polarization
because of the birefringent nature of microstructure fibers [17].
Figure 4-5. (a) Output spectrum for an input peak power P = 16 kW and the propagation
distance 1.5 cm. (b) Same as (a), but with 0.1% higher peak power. (c) High resolution
window of the spectra in (a) (solid line) and (b) (dotted line).
4. MICROSTRUCTURE FIBER AND WHITE-LIGHT GENERATION 109
6. CONCLUSIONS
ACKNOWLEDGEMENTS
REFERENCES
[1] J. C. Knight, T. A. Birks, P. S. Russell, and D. M. Atkin, Opt. Lett. 21,
1547-1549 (1997).
[2] J. K. Ranka, R. S. Windeler, and A. J. Stentz, Opt. Lett. 25, 25-27 (2000).
[3] I. Hartl, X. D. Li, C. Chudoba, R. K. Ghanta, T. H. Ko, J. G. Fujimoto, J.
K. Ranka, and R. S. Windeler, Opt. Lett. 26, 608-610 (2001).
[4] T. A. Birks, J. C. Knight, and P. S. Russell, Opt. Lett. 22, 961-963 (1997).
[5] J. C. Knight, J. Broeng, T. A. Birks, and P. S. J. Russel, Science 282, 1476-
1478 (1998).
[6] C. M. Smith, N. Venkataraman, M. T. Gallagher, D. Muller, J. A. West, N.
F. Borrelli, D. C. Allan, and K. W. Koch, Nature 424, 657-659 (2003).
[7] R. D. Meade, A. M. Rappe, K. D. Brommer, J. D. Joannopoulos, and O. L.
Alerhand, Phys. Rev. B 48, 8434-8437 (1993).
[8] G. P. Agrawal, Nonlinear Optics (Academic Press, San Diego, 2001).
[9] J. K. Ranka, R. S. Windeler, and A. J. Stentz, Opt. Lett. 25, 796-798
(2000).
[10] J. C. Knight, J. Arriaga, T. A. Birks, A. Ortigosa-Blanch, W. J. Wadsworth,
and P. S. Russell, IEEE Photonics Technol. Lett 12, 807-809 (2000); D.
Ouzounov, D. Homoelle, W. Zipfel, W. W. Webb, A. L. Gaeta, J. A. West,
J. C. Fajardo, and K. W. Koch, Opt. Commun. 192, 219-223 (2001).
[11] T. Brabec and F. Krausz, Phys. Rev. Lett. 78, 3282-3285 (1997).
[12] R. H. Stolen, J. P. Gordon, W. J. Tomlinson, and H. A. Haus, J. Opt. Soc.
Am. B 6, 1159-1166 (1989).
[13] A. L. Gaeta, Opt. Lett. 27, 924-926 (2002).
[14] A. V. Husakou and J. Herrmann, Phys. Rev. Lett. 8720, art. no.-203901
(2001); J. Herrmann, U. Griebner, N. Zhavoronkov, A. Husakou, D. Nickel,
J. C. Knight, W. J. Wadsworth, P. S. J. Russell, and G. Korn, Phys. Rev.
Lett. 88, art. no.-173901 (2002).
[15] J. M. Dudley, L. Provino, N. Grossard, H. Maillotte, R. S. Windeler, B. J.
Eggleton, and S. Coen, J. Opt. Soc. Am. B 19, 765-771 (2002); S. Coen, A.
H. L. Chau, R. Leonhardt, J. D. Harvey, J. C. Knight, W. J. Wadsworth,
and P. S. J. Russell, J. Opt. Soc. Am. B 19, 753-764 (2002).
[16] K. M. Hilligsoe, T. V. Andersen, H. N. Paulsen, C. K. Nielsen, K. Molmer,
S. Keiding, R. Kristiansen, K. P. Hansen, and J. J. Larsen, Opt. Expr. 12,
1045-1054 (2004).
[17] Z. M. Zhu and T. G. Brown, Opt. Expr. 12, 791-796 (2004).
[18] X. Gu, L. Xu, M. Kimmel, E. Zeek, P. O'Shea, A. P. Shreenath, R. Trebino,
and R. S. Windeler, Opt. Lett. 27, 1174-1176 (2002).
[19] J. M. Dudley and S. Coen, Opt. Lett. 27, 1180-1182 (2002); X. Gu, M.
Kimmel, A. P. Shreenath, R. Trebino, J. M. Dudley, S. Coen, and R. S.
Windeler, Opt. Expr. 11, 2697-2703 (2003).
[20] K. L. Corwin, N. R. Newbury, J. M. Dudley, S. Coen, S. A. Diddams, K.
Weber, and R. S. Windeler, Phys. Rev. Lett. 90, 113904 (2003); J. N.
4. MICROSTRUCTURE FIBER AND WHITE-LIGHT GENERATION 111
1. INTRODUCTION
Ultrashort pulse generation reached pulse durations of a few
femtoseconds around the turn of the century [1]. Some of the shortest pulses
have a width of only two optical cycles [2-4]. When pulse durations
approach this regime, the commonly used approach of the slowly varying
envelope approximation (SVEA) starts to fail. Nonlinear optical effects are
then expected to depend not only on the envelope structure of the pulses, but
also on the structure of the electric field itself, including its relative phase to
5. OPTICAL COMB DYNAMICS AND STABILIZATION 113
the envelope, which has been referred to as the absolute phase. In 2001 and
2003, the first experimental evidence for the failure of the SVEA was
reported [5].
In this chapter, we describe how to monitor the relative phase between
the envelope and the carrier of an optical pulse train from a mode-locked
optical oscillator. The chapter is organized as follows: First, we introduce the
definition of the carrier-envelope-offset (CEO) phase and frequency. We
show how both entities are connected with intracavity dispersion. Methods
to measure the CEO frequency are discussed, with an emphasis on the most
wide spread ν-to-2ν scheme. Careful characterizations of the CEO-phase-
noise spectra are used to isolate the physical mechanisms behind the
excessive fluctuations of the CEO phase of a free-running oscillator. This
insight allows building femtosecond lasers with an increased passive
stability of the CEO and forms the basis for subsequent stabilization of the
CEO frequency. Finally, we discuss how to optimize control of the CEO
phase and how to push residual phase jitter into the attosecond range.
φce
TR = 1/fr
Figure 5-1. The electric field E (t ) of two subsequent pulses from a mode-locked laser (solid
line). The envelope ± A(t ) is shown as dashed lines. The electric-field patterns of the pulses
experience a pulse-to-pulse phase shift ∆ φ ce according to Equation (1).
2π L ω 2 L dn( z )
∆φce = ∫ n g ( z ) − n( z ) d z mod 2π = ∫ d ω d z mod 2π .
(1)
λ 0
c 0
Here L is the length of the dispersive material. For the case of a linear cavity,
L takes the role of twice the cavity length, and the carrier-envelope offset
(CEO) phase ∆φce is the change of the phase φ ce per round trip:
∆φce (t ) = φ ce (t ) − φ ce (t − TR ) . (2)
The CEO phase ∆φ ce must not be confused with the phase φ ce , which is
typically defined such that a pulse with φ ce = 0 has the largest possible value
of the electric field [6]. Some authors have referred to φ ce as the absolute
phase. An example for such a pulse is shown as the left pulse in Figure 5-1.
∆φ ce , however, is defined as the difference of the absolute phase of two
subsequent pulses. It is useful to introduce the CEO frequency [7]
∆φce
f0 = fr , (3)
2π
where fr equals the inverse round-trip time 1/TR of the cavity and f0 is time
dependent unless the intracavity dispersion and the cavity length are
5. OPTICAL COMB DYNAMICS AND STABILIZATION 115
fr
f0
Figure 5-2. Equidistant frequency comb of a mode-locked laser. The comb lines are spaced
by the repetition rate fr and exhibit a nonvanishing offset frequency f0 at zero frequency unless
the electric-field pattern exactly reproduces from pulse to pulse (compare to the time domain
picture in Figure 5-1).
ν i = f 0 + if r . (4)
This leaves only two degrees of freedom for the dynamics of the frequency
comb, translation via f0 and breathing via fr, as illustrated in Figure 5-3. Any
kind of perturbation of the cavity, e.g., by a thermal change of the refractive
index of the laser crystal, will typically affect both the repetition rate and the
CEO frequency.
breathing
translation
Figure 5-3. Comb dynamics [10]. The frequencies inside the comb structure are determined
by two parameters, f0 and fr . This gives rise to a translational degree of freedom and a
breathing mode. Noise contributions will induce a characteristic linear combination of both
degrees of freedom.
−1
∂f 0 ∂f r
f x = f0 + fr , (5)
∂X ∂X
where X could be any physical parameter of the cavity, e.g., its length or the
temperature of the laser crystal. Let us illustrate the concept of a fixed point
by choosing X as the cavity length. Cavity length fluctuations only affect the
repetition rate fr of the laser but leave the per-round-trip phase shift ∆φce
between envelope and carrier unchanged. Inserting Equation (3) into
Equation (5) yields f x = 2 f 0 , i.e., a value very close to zero frequency. A
complementary example would be an effect that causes only a change of the
cavity group delay but leaves the phase delay unchanged. This could be
achieved, e.g., by tilting a mirror in an intracavity prism sequence with the
pivot point adjusted to the center frequency of the mode-locked spectrum
5. OPTICAL COMB DYNAMICS AND STABILIZATION 117
[12]. Retarding the group by one cycle relative to the phase changes f0 by
one free spectral range (i.e., the repetition rate), whereas the repetition rate
itself only changes by a very small amount. In this case, one calculates that fx
equals the carrier frequency of the pulse. Most environmental contributions
to comb dynamics, such as thermal or nonlinear changes of the intracavity
refractive indices, have an f x located between zero frequency and the
carrier frequency [11]. This means that they neither add a pure contribution
to the group delay nor do they only affect the phase delay of the group.
Measuring the fixed frequency f x can help to pinpoint the source of
dominant frequency comb dynamics.
Nν m1 − Mν m2 = ( N − M ) f 0 , (6)
which requires that Nm1 = Mm2 . Equation (6) is the key to any measurement
of the carrier-envelope offset and was used in the first experimental
demonstrations by Jones et al. [14] and Apolonski et al. [15] for the case of
118 Chapter 5
ν2 2ν1
f0
ν1 = f0 + n fr 2ν1 = 2f0 + 2n fr
ν2 = f0 + 2n fr
Figure 5-4. Scheme for measuring the CEO-frequency of a laser comb for the case of
heterodyning the fundamental and the second harmonic, i.e., N = 1 and M = 2 in Equation (6).
Graphically, this scheme mirrors the origin at f = ν1, transferring the f0 beat from dc into a
region with nonvanishing spectral content.
-10
CEO-beats
intermode beat
Power density (dBc)
-20
spurious
-30
-40
-50
-60
0 20 40 60 80 100 120
Frequency (MHz)
Figure 5-5. Typical rf spectrum of the CEO beat note signal. This signal was measured at a
Ti:sapphire laser heterodyning the fundamental and the second-harmonic-generation (SHG)
signal from a continuum generated in a microstructure fiber [16]. The CEO beat is located at
35 MHz with a signal-to-noise ratio of >45 dB in a 100 kHz bandwidth. Its mirror frequency
is also visible at 65 MHz. The laser has a 100 MHz repetition rate. Some spurious
contributions have been generated by nonlinear electronic mixing processes in the detector
circuitry.
SHG
530 nm
530 nm CEO
Figure 5-6. Schematic drawing of a practical implementation of Equation (6) and Figure 5-4.
The laser is spectrally broadened to more than an optical octave using continuum generation
in a microstructure fiber. Two wavelengths forming one octave are separated. The long-
wavelength component is frequency doubled and heterodyned with the fundamental signal.
The beat note contains an rf component at the CEO frequency. Specific wavelengths shown
are meant as example values.
ν a 3ν b 2 3
= = . (7)
2ν b 2ν b 4
This way, one now has available two phase-locked oscillators at only half-
an-octave spectral separation. Locking one to the comb and phase
comparison of the other allow extraction of the CEO frequency of the comb.
σ f0
σ φce = . (8)
f
2
σ f0
fr 2 fr 2
δf 0(f low ) = 2 ∫ σ 2f 0 d f , δφ ce(f low ) = 2 ∫
f
df . (9)
flow flow
The integration spans from a lower frequency flow given by the inverse
measurement time to an upper bound that is ideally half the repetition rate
itself. Typically, one can only carry the integration to a few tens or hundreds
of kilohertz, which is normally considered sufficient as the noise rolls off
very rapidly at high frequencies. The integrated noise densities δφ ce ( f low )
and δf 0 ( f low ) can be interpreted as rms widths of the fluctuation range of
phase or frequency, respectively.
inspect the integrated phase noise starting at high frequencies and locating
the point where the δφce approaches unity. In Figure 5-8, one can clearly see
that the integrated phase noise already reaches one radian at several kHz
Fourier frequency. Around 1 kHz, the noise grows dramatically, reaching
values of several hundred to thousands of radians at 100 Hz offset frequency.
In summary, this means that severe phase-noise contributions accumulate
within approximately 1 ms of measurement time. For a successful phase lock
of the CEO frequency, however, it is mandatory to keep residual jitters
below δφce ≈ 0.3 rad [22]. This requirement is only set by the cycle-slip-free
functioning of the phase-locked loop, whereas applications may demand an
even tighter locking to smaller jitter values. These considerations make it
clear that for any meaningful stabilization of the CEO frequency, servo
bandwidths of 10 kHz or more are required, which makes the use of acousto-
optic or electro-optic controls preferable to a mechanical adjustment of
intracavity dispersion.
4
10 without prisms
with prisms
Freq. noise density (Hz/÷Hz)
3
10
2
10
phase locked
1
10
0
10
0 1 2 3 4 5
10 10 10 10 10 10
Figure 5-7. Frequency noise density of different Ti:sapphire lasers [16, 21]. The top trace
shows a measurement for a laser with intracavity prisms; the middle trace is measured for a
prismless variant of the same laser. The bottom trace shows how the measured frequency
noise density drops farther when a phase lock to a reference oscillator is activated. This
measurement has to be interpreted as a noise floor, as it is limited by the stability of the local
oscillator in the measurement.
Moreover, Figures 5-7 and 5-8 contain measurements for both oscillators
with and without prisms for intracavity dispersion compensation. In these
measurements, an identical pump laser was used for the prismless and the
prism setup. It becomes evident that prism-based femtosecond oscillators are
about ten times as noisy as prismless lasers. This can only be explained by
an additional physical mechanism present because of the intracavity prisms.
5. OPTICAL COMB DYNAMICS AND STABILIZATION 123
Figure 5-8. Integrated phase-noise density of different Ti:sapphire lasers [16, 21]. The top and
middle traces have been computed from the data displayed in Figure 5-7 using Equations (8)
and (9). The bottom trace is based on a direct phase comparison of a stabilized oscillator and
an rf reference using an rf lock-in amplifier.
∂ ∂ω ∂n ∂n ∂L ∂2n
∆φ ce = 2ω c c L + ω c2 + ω c2 L. (10)
∂X ∂X ∂ω ∂ω ∂X ∂ω∂X
of angles inside and outside the laser crystal [21]. Beam-pointing effects are
held responsible for an approximately tenfold increase of CEO-phase noise
of prism laser cavities as compared to prismless variants.
The third term in Equation (10) contains contributions to CEO-phase
noise via intensity-induced changes of the refractive index [25]. Nonlinear
refraction is well known as the all-optical Kerr effect [26], but according to
Equation (10), only the dispersion of the Kerr effect affects changes of the
CEO phase. The issue of dispersion of the Kerr effect has been addressed by
[27, 28]. According to Sheik-Bahae et al. [27], the main contribution to the
first-order dispersion of a dielectric medium well below half the band edge
stems from a Kramers-Kronig term induced by two-photon absorption. As
per their example, for sapphire at 800 nm, one calculates
∂ 2 n / ∂ω ∂I ≈ 10 −36 s m2/W rad. Inserting values for typical Ti:sapphire laser
cavities [4], one computes a theoretical estimate of ∂f 0 / ∂I =5×10-9 HzW/m2,
which agrees well with the lowest experimentally observed values of
∂f 0 / ∂I . Again, these low APC coefficients can only be reached in the
absence of geometrical effects and spectral shifting.
8
Coupling coeff. (10 Hz m /W)
measurement
2
6
-8
theoretical estimate
0
1 2 3 4
10 10 10 10
Modulation frequency (Hz)
Figure 5-9. Transfer function of laser intensity noise into fluctuations of the CEO frequency
[16]. These measurements were done in a prismless laser in the absence of spectral shifting.
The measurements reflect Kerr contributions to the APC and thermally induced amplitude-to-
phase conversion (APC) effects at low frequencies.
HzW/m2. Spectral shifting does not seem to play a role in these experiments,
and geometrical effects are also not a concern because of a prismless cavity.
At lower frequencies, additional contributions from thermally induced
changes of the refractive index increase the APC effect. For modulation
frequencies of about 10 kHz or more, the coupling dynamics appear to be
restricted to a purely electronic-refractive nonlinearity.
From the experimental observations, some guidelines can be given on
how to keep APC effects to a minimum. The first recommendation is to use
a prismless cavity, which is also strongly supported by the data in Figures 5-
7 and 5-8. In prismless cavities, beam pointing does not translate into CEO-
phase noise [21]. Spectral shifting is the other APC effect that can be
avoided by suitable design of the laser. For a stable position of the laser
spectrum, a broad mode-locked bandwidth of more than 50 nanometers and
a high pulse energy appear to be favorable conditions [23]. If geometric
effects and spectral shifting can be avoided, the APC effects are restricted to
nonlinear refractive mechanisms, both Kerr-type and an additional thermally
induced mechanism at low Fourier frequencies. Values on the order of
∂f 0 / ∂I =10-8 HzW/m2 or less are indicative of a dominance of nonlinear
refraction in the APC dynamics.
(w )
GD
differential
mode
Figure 5-10. Control of intracavity first-order dispersion by tilting of an end mirror after a
dispersive delay line. Translation of the mirror parallel to the optical axis acts on both group
and phase delay; tilting changes the difference between them. Provided a choice of the correct
pivot point has been made, tilting only affects the CEO-frequency.
Mirror tilting is not an option when a prismless setup is used. Then the
method of choice is modulation of the pump power either with an acousto-
optic modulator [16] or with an electro-optic device [20]. As the required
pump-power modulation is on the order of 10-3, it is typically very easy to
reach bandwidths of several tens to hundreds of kHz. Pump-power
modulation relies on the APC mechanisms discussed in the previous section
and is currently the most widespread mechanism for CEO-frequency control.
broadening of the laser spectra. A first attempt to directly stabilize the CEO
frequency of an octave-spanning laser was reported by Morgner et al. [3].
However, because this laser only spanned the octave at about -40 dBc, the
authors achieved a CEO beat note that was considered minimum for a robust
stabilization. The authors thus proposed a 2ν-to-3ν scheme. This scheme has
been carried out recently with resulting small residual timing jitters and
excellent long-term stability [31]. Still, a direct stabilization is more
challenging than stabilization based on additional spectral broadening.
Another important issue is the setup of the locking electronics. For any
meaningful application, a phase lock to an rf-reference source is required. A
phase lock can be as simple as that depicted in Figure 5-11, which consists
of a double-balanced mixer and some means to adjust the servo loop gain.
The gain has to be optimized for a sufficient phase margin of the loop to
prevent self-oscillation of the servo circuit.
APD Bandpass-filter
Ref. oscillator
AM
AOM Driver
Pump Beam
Ti:Sa
Figure 5-11. Simple phase-lock circuit used for stabilization of the CEO frequency [16]. The
avalanche photo detector (APD) measures the beat note signal [Equation (6), Figure 5-5].
Suitable bandpass filtering isolates the beat note and suppresses mirror frequencies and
spurious contributions. After mixing the signal with the reference oscillator, the mixing
product is directly fed back via an acousto-optic modulator. The servo loop gain has to be
adjusted for sufficient phase margin.
The simple circuit of Figure 5-11 has only a very limited capture range
and may not be able to avoid cycle slips in the presence of strong CEO-
phase noise. An alternative is usage of a phase detector with enhanced
capture range [32]. Such a circuit is based on an electronic counter and can
boost the phase capture range to tens or hundreds of π. This strategy comes
at the price of decreased sensitivity that will ultimately limit the overall
performance of the lock. The general recommendation is to reduce noise
mechanisms as far as possible by enhancing the passive stability of the laser.
Capture range enhancement should only be used as a last resort and then
moderately; otherwise, extra noise of the stabilized laser will result.
5. OPTICAL COMB DYNAMICS AND STABILIZATION 129
Some of the best results in terms of residual phase noise were achieved
with the simple double-balanced mixer (Reference [16], Figure 5-11). The
measured data is also shown in Figures 5-7 and 5-8. From the data in Figure
5-8, one can conclude that the residual phase jitter in these measurements
was only about 20 mrad in a 10 kHz to 0.01 Hz interval. This corresponds to
residual timing jitters of only 10 as associated with the CEO phase.
1 (intracavity)
σ φ(extracavity)
(f) = σ (f). (11)
ce
f r f0
Figure 5-12. Differential phase-noise spectrum of two independent measurements of the CEO
phase.
Several ways have been suggested for overcoming these residual effects.
Rather than using the laser pulses directly, one can use the white-light
continuum pulses for any kind of CEO-sensitive application. Using the
spectrum directly from an octave-spanning oscillator would also rule out
APC effects in the microstructure fiber. Still, both these solutions suffer
from interferometer drift. One way to strongly reduce drift effects is a
common-path interferometer, as suggested by Kakehata et al. [34] for single-
shot CEO-phase measurements.
6. SUMMARY
We have discussed fluctuations of the CEO phase in oscillators and
explained their origin. The CEO phase turns out to be a very sensitive
parameter that is easily influenced by nearly all laser and environmental
parameters. Many of these contributions either give rise to slow drift effects
or can be easily shielded by enclosing the laser. From the perspective of
stabilizing the CEO frequency of the laser, however, amplitude-to-phase
noise conversion turns out to be a much more significant problem that
cannot totally be avoided. Several mechanisms contribute to the conversion
of laser amplitude noise into CEO frequency fluctuations. Again, some of
these contributions, such as spectral shifting or beam pointing, can be
avoided, or at least reduced, by construction and choice of favorable
operating conditions. With all these measures in place, it is possible to
provide a tight lock to an external rf reference, with resulting residual timing
jitters between carrier and envelope of the laser of only a few attoseconds.
Even with such superior performance, one has to be careful to avoid intrinsic
sources of CEO phase noise in the measurement set-up itself. Amplitude-to-
phase conversion also takes place in external continuum generation, which is
often used to broaden the laser spectrum to an optical octave. The ν-to-2ν
5. OPTICAL COMB DYNAMICS AND STABILIZATION 131
interferometer is another weak point and can give rise to a small residual
drift. These residual effects are relatively weak and do not appear to corrupt
most applications of frequency combs demonstrated to date. Nevertheless,
they can be avoided. An improved control of frequency comb parameters
offers even higher precision in metrology applications and opens up novel
applications in extreme nonlinear optics. Understanding the dynamics of the
comb is the key to further progress in these areas.
REFERENCES
[1] G. Steinmeyer, D. H. Sutter, L. Gallmann, N. Matuschek, and U. Keller,
Science 286, 1507-1512 (1999); U. Keller, Nature 424, 831-838 (2003).
[2] M. Nisoli, S. Desilvestri, O. Svelto, R. Szipocs, K. Ferencz, C. Spielmann,
S. Sartania, and F. Krausz, Opt. Lett. 22, 522-524 (1997); A. Baltuška, M.
S. Pshenichnikov, and D. A. Wiersma, Opt. Lett. 23, 1474-1476 (1998); A.
Baltuška, T. Fuji, and T. Kobayashi, Opt. Lett. 27, 1241-1243 (2002); R.
Ell, U. Morgner, F. X. Kärtner, J. G. Fujimoto, E. P. Ippen, V. Scheuer, G.
Angelow, T. Tschudi, M. J. Lederer, A. Boiko, and B. Luther-Davies, Opt.
Lett. 26, 373-375 (2001).
[3] U. Morgner, R. Ell, G. Metzler, T. R. Schibli, F. X. Kärtner, J. G. Fujimoto,
H. A. Haus, and E. P. Ippen, Phys. Rev. Lett. 86, 5462-5465 (2001).
[4] D. H. Sutter, G. Steinmeyer, L. Gallmann, N. Matuschek, F. Morier-
Genoud, U. Keller, V. Scheuer, G. Angelow, and T. Tschudi, Opt. Lett. 24,
631-633 (1999).
[5] G. G. Paulus, F. Grasbon, H. Walther, P. Villoresi, M. Nisoli, S. Stagira, E.
Priori, and S. De Silvestri, Nature 414, 182-184 (2001); G. G. Paulus, F.
Lindner, H. Walther, A. Baltuška, E. Goulielmakis, M. Lezius, and F.
Krausz, Phys. Rev. Lett. 91, 253004 (2003).
[6] H. R. Telle, G. Steinmeyer, A. E. Dunlop, J. Stenger, D. H. Sutter, and U.
Keller, Appl. Phys. B 69, 327-332 (1999).
[7] T. Brabec and F. Krausz, Rev. Mod. Phys. 72, 545-591 (2000).
[8] F. W. Helbing, G. Steinmeyer, and U. Keller, IEEE J. Sel. Top. Quantum
Electron. 9, 1030-1040 (2003).
[9] T. Udem, J. Reichert, R. Holzwarth, and T. W. Hänsch, Opt. Lett. 24, 881-
883 (1999).
[10] T. F. Albrecht, K. Bott, T. Meier, A. Schulze, M. Koch, S. T. Cundiff, J.
Feldmann, W. Stolz, P. Thomas, S. W. Koch, and E. O. Göbel, Phys. Rev.
B 54, 4436-4439 (1996).
[11] N. Haverkamp, B. Lipphardt, J. Stenger, H. R. Telle, C. Fallnich, and H.
Hundertmark, In Conference On Ultrafast Phenomena, Vancouver, BC,
2002), P. Me31-31.
[12] K. F. Kwong, D. Yankelevich, K. C. Chu, J. P. Heritage, and A. Dienes,
Opt. Lett. 18, 558-560 (1993).
[13] L. Xu, C. Spielmann, A. Poppe, T. Brabec, F. Krausz, and T. W. Hänsch,
Opt. Lett. 21, 2008-2010 (1996).
[14] D. J. Jones, S. A. Diddams, J. K. Ranka, A. Stentz, R. S. Windeler, J. L.
Hall, and S. T. Cundiff, Science 288, 635-639 (2000).
132 Chapter 5
Takayoshi Kobayashi
Department of Physics, Faculty of Science, University of Tokyo
Abstract: This chapter presents the basic principles for three parametric interactions that
enhance bandwidth to obtain short pulses while maintaining phase matching.
To extend the bandwidth, we introduced a noncollinear configuration between
the pump and signal. The idea is used in three different parametric processes:
optical parametric generation (OPG), optical parametric amplification (OPA),
and optical parametric oscillation (OPO). Using noncollinear phase matching,
we developed a noncollinear-optical-parametric amplifier (NOPA) that
delivers 4 fs visible-near-infrared pulses. We designed geometrical and
temporal configurations of the NOPA that broaden the gain bandwidth in
excess of 250 THz. The main requirements for bandwidth enhancement
include (1) phase matching, (2) group-velocity matching, (3) pulse-front
matching, and (4) optimization of the angular dispersion of the pump. To
achieve the extended-gain bandwidth, full phase adjustment is performed by
several compensators, including a prism pair, a grating-mirror system
equivalent to a grating pair, chirped mirrors, and a deformable mirror. By
adding these devices to the NOPA system, we obtained pulse widths of 3.9 fs
in the visible and NIR spectral range.
1. INTRODUCTION
2. ADVANCES OF NONCOLLINEAR-PHASE-
MATCHED OPTICAL PARAMETRIC
CONVERSION
signal and idler are arranged with larger noncollinear angles, as shown in
Figure 6-1.
ks signal
vs
α kp
β
vp pump
ki
idler
vi
38], and the NOPA continues to attract attention owing to its remarkable
advantages.
We will now describe the development of a noncollinear parametric
amplifier. First, we introduced a pulse-front-matching geometry [36] to
achieve a sub-5 fs light source using methods other than the conventional
continuum-compression scheme. In the nonlinear configuration, the group
velocity of a faster-traveling wave (the idler) is projected at an angle onto a
slower wave (the signal), thus improving the temporal overlap of the two and
achieving broadband phase matching. The importance of parametric
conversion in the visible and near infrared has been dramatically increased
by the discovery of the unique phase-matching conditions in a type-I BBO
crystal pumped by the second harmonic of the Ti:sapphire laser [36, 38]. As
a result, widely tunable sub-20 fs NOPAs have become more or less
routinely used [34, 39]. Tunable operation in a 10 fs regime in both the
visible and infrared has also been demonstrated [32, 36, 37, 40]. The
development of sophisticated pulse-compression schemes has made phase
correction over the entire parametric-amplification bandwidth possible,
resulting in the realization of sub-5 fs pulse generation in the visible–NIR
region [38, 41-43].
The great potential of sub-10 fs NOPAs was proven by several
spectacular applications of nonlinear spectroscopy such as time-domain
studies of ultrafast molecular dynamics in the condensed phase [42, 44].
Although other methods of ultrashort pulse generation may, in some cases,
be better than NOPAs for both pulse length and spectral width [45],
parametric amplifiers produce noticeably smoother spectra with high
stability. The smooth, stable spectra play a vital role in time- and frequency-
resolved spectroscopy with high time resolution and with relatively high
spectral resolution when the spectrometer is placed after the sample. The
demand for continuing improvement of the NOPA pulse quality is evident.
The second step in obtaining short optical pulses is to control the phase
of the field to be constant over the whole gain bandwidth. A constant field
phase leads not only to improved pulse compression, but also to easier
dispersion manipulation in phase-sensitive applications of nonlinear
spectroscopy. Several dispersion-control techniques have been employed to
obtain the shortest pulses from NOPAs. For instance, sub-5 fs pulses have
been achieved both by using a combination of a 45°-angled-prism
compressor and a set of ultrabroad-chirped mirrors (UBCM) [38] and by
using custom-designed UBCMs alone [43]. While it is feasible to fabricate a
set of “ideal” UBCMs, the fixed dispersion of such multilayer dielectric
structures makes it impossible to introduce fine wavelength-selective control
of group delay, which is required in the daily optimization of a pulse
138 Chapter 6
κ p = κ s + κi , (1)
ω p = ω s + ωi . (2)
Here κj and ωj are the wave vector and angular frequency with the suffixes j
= p, s, and i corresponding to the pump, signal, and idler, respectively. The
first is a phase-matching condition corresponding to momentum
conservation among the relevant three photons; the second is the
requirement of energy conservation. Hereafter, we assume a plane-wave
6. FEMTOSECOND NONCOLLINEAR PARAMETRIC 139
AMPLIFICATION AND CARRIER-ENVELOPE PHASE CONTROL
interaction among them. This assumption is well satisfied in the case when
the beam diameter (including the beam waist at the focal point) is much
larger than the wavelength. By neglecting the pump depletion and spectral
width, the well-known parametric gain G is [56, 57]
(
g = Γ 2 − (∆k 2 ) )
2 1/ 2
, Γ = 2d eff
ω sω iΦ
ε 0 n s ni n p c 3
, (3)
where α and β are the noncollinear angles between the wave vectors of the
pump and signal and those between pump and idler, respectively. Figure 6-1
illustrates the arrangement of the general noncollinear configuration.
In the conventional collinear geometry (α = β = 0), the expression is
simplified. By the Taylor expansion in powers of the angular frequencies
around the central frequencies ω s 0 − ω i 0 , ∆κ is given by
1 1 ∂ 2κ s δ 2κ i
(∆ω s )2 + ...
1
∆κ = − ∆ω s
− + (5)
ν s νi 2 δω 2 δω i2
which is defined as the full width at half maximum (FWHM) of the gain
spectrum G( ω s ). This represents the group-velocity mismatch (GVM)
between the signal and idler in the nonlinear optical crystal that determines
the bandwidth of an OPA. The GVM is an intrinsic effect for the ultrashort
pulse interaction and determines the parametric bandwidth around 200–500
cm-1 in standard OPAs [57]. Previously reported femtosecond OPAs with
collinear geometry suffered from this problem, and the shortest pulse
durations were limited to 40 fs in the visible [20] and to 30 fs in the NlR [16,
21]. Broadening of the parametric bandwidth is essential for obtaining a
tunable sub-10 fs light source. Sosnowski et al. [58] demonstrated a spectral
broadening of the signal to about 800 cm-1 in the visible region by multistage
amplification in which different spectral regions were amplified at each
stage with slightly different crystal angles. However, this method has
shortcomings, such as being rather complicated. In the late 1990s, novel NIR
OPAs were reported [59, 60]. For example, Fournier et al. used the
effectively cascading third-order process of second-harmonic generation of a
signal and obtained self-compression to 20 fs signal pulses [59]. Nisoli et al.
[60] demonstrated 14.5 fs signal pulse generation around 1.5 µm using an
ultrashort 18 fs pump source from a Ti:sapphire amplifier followed by a
hollow-fiber compressor. Both methods seem to be difficult and too sensitive
to the pump energy. Third-order effects play an essential role in the
characteristic pulse propagation in their schemes. Simpler and more robust
methods are strongly desired for shorter pulse generation.
δα 1 δ 2α
(∆ω s )2 + ...
cos α = cos α 0 − sin α 0 ∆ω s − cos α 0
2
(7a)
δω s 2 δω s 0
δβ 1 δ 2β
(∆ω s )2 + ...,
cos β = cos β 0 − sin β 0 ∆ω s − cos β 0
(7b)
δω s 2 δω 2
s 0
where the subscript 0 denotes the central value. From the phase-matching
condition of Equation (1) we obtain
6. FEMTOSECOND NONCOLLINEAR PARAMETRIC 141
AMPLIFICATION AND CARRIER-ENVELOPE PHASE CONTROL
By substituting Equations (7) and (8) into (4) we can expand the phase
mismatch as
∆κ = −
cos α
νs
0
−
cos β 0
νi
− κ s 0 sin α 0
∂α
+ κ i 0 sin β 0
∂β
∆ω s
+ K
∂ω s 0 ∂ω i 0
(9)
cos α 0 cos β 0 ∂α ∂β
− − κ s 0 sin α 0 + κ i 0 sin β 0
∂ω
=0 (10a)
νs νi ∂ω s 0 i 0
sin α 0 sin β 0 ∂α ∂β
+ + κ s 0 cos α 0 + κ i 0 cos β 0
∂ω
= 0. (10b)
νs νi ∂ω s 0 i 0
ν s =νi (11)
is fulfilled. This condition is only satisfied for the case of the degeneracy
(ω s = ωi = ω p 2) in a type-I interaction (e o + o or o e + e). In a type-
II interaction (e o + e or o o + e) the GVM is usually larger than that in
a type-I interaction because of the large birefringence in the crystals [22, 57]
142 Chapter 6
used in type II interactions. In a noncollinear geometry, the angular
dispersion can eliminate the GVM and a broadband phase matching can be
attained. The schematic of this configuration is shown in Figure 6-2.
ks(λs)
ki(λp
α β (λi) )
y
(λs) kp
θ
The angular dispersions satisfying this condition are calculated from (10) as
ν s = ν i cos(α + β ) , (13)
6. FEMTOSECOND NONCOLLINEAR PARAMETRIC 143
AMPLIFICATION AND CARRIER-ENVELOPE PHASE CONTROL
with
where the subscript 0 is dropped for simplicity. The physical meaning of this
condition can be stated as the projection of the idler group velocity on the
signal direction is equal to the signal group velocity. The phase mismatch in
Equation (9) can then be rewritten
cos(α + β ) 1
∆κ = − −
1
∆ω s +
cos β ν s ν i cos(α + β )
K. (15)
1 1
GVM s −i = − . (16)
νs ν i cos (α + β )
The GVM between the pump and signal, which limits the interaction length
in the crystal, is [27]
1 1
GVM p − s = − . (17)
νp ν s cos α
As clearly seen from Equations (16) and (17), both GVMs are dependent on
the noncollinear angle α. Each of the GVMs can be eliminated by selecting a
geometrical arrangement satisfying Equation (16) or (17). Group-velocity
matching between the pump and signal or between the signal and idler using
noncollinear geometry in the case of ν s > ν i has been reported; an order-of-
144 Chapter 6
magnitude longer interaction length than in the collinear geometry was also
obtained [27-29]. Our current interest is the signal-idler–group-velocity
matching, which is possible if ν s > ν i . The bandwidth is broadened by more
than tenfold and only limited by the group-velocity dispersion (GVD, as
indicated in Equation (5) [57]; ultrashort pulse generation is thus expected.
4. SIGNAL-WAVELENGTH-INSENSITIVE PHASE
MATCHING
k1e = k 20 + k 30 , (18)
ω1 = ω 2 + ω 3 . (19)
Here, ki and ωι (λi) are the wave vector and the frequency (wavelength),
respectively, of i-th beam, and i = 1, 2, and 3. Solving Equations (18) and
(19), we obtain the following expression for the phase-matching angle θ that
is defined as the internal angle made by the extraordinary (e) polarized pump
beam with the optic axis of the crystal:
θ = cos −1 (A Y ) e 2
−1
. (20)
(A e
A )
O 2
−1
Here,
Y=
(k ) + (k )
o 2
2
o 2
3 + 2k 2o k 3o cosψ ,
(21)
ψ =α + β , (22)
6. FEMTOSECOND NONCOLLINEAR PARAMETRIC 145
AMPLIFICATION AND CARRIER-ENVELOPE PHASE CONTROL
β = sin-1[(k2ok3o)sinα], (23)
and Ao = 2π(n1o/λ1), Ae = 2π(n1e/ λ1), k2o = k2o =2π(n2o/λ2), and k3o = k3o
=2π(n3o/λ3); n1, n2, and n3 are refractive indices of the three interacting
radiations with wavelengths λ1, λ2, and λ3. The superscripts o and e
correspond to the ordinary and extraordinary polarizations, respectively, and
α (β) is the noncollinear angle between the pump and the signal (idler)
beams. Here k2o and k3o are the wave numbers corresponding to the signal
and idler beams, respectively. Y is defined as the ‘average’ wave number of
the signal and idler beams parallel to the pump beam. Considering the phase
matching along the direction perpendicular to k1, Equation (23) is obtained.
In the case of monochromatic pump, from Equation (20) we observe that the
phase-matching angle θ will be independent (to first order) of the signal
wavelength or signal frequency (ω2) if ∂θ/∂ω2 = 0, for any value of ω2. In
Equation (20), all the parameters except Y are independent of ω2. Therefore,
the condition may be expressed as follows:
∂Y/∂ω2 = 0. (24)
If the incident angle of the tunable seed beam has no frequency or,
equivalently, wavelength dependency, the noncollinear angle α between the
pump and tunable seed pulses will remain unchanged even with the change
of the signal frequency, i.e., ∂α/∂ω2 =0. From Equations (24) and (25)
where, vgi = ∂ωi/∂ki, is the group velocity of the i-th beam (i = 2 and 3) with
frequencies of ωi. To obtain Equation (26), we have used ∂ω2 = −∂ω3, since
ω1 = constant because the pump is monochromatic. From Equation (23)
Equation (28) can also be derived [31] by equating the first-order derivative
of the phase mismatch with respect to ω2 to zero, assuming a monochromatic
pump and setting ∂α/∂ω2 = 0. Using this method, we found that for a NOPA,
the effective inverse group-velocity mismatch between the signal and the
idler pulses [31]
ϑ (λs ,α ) = cos −1
1
×
−2
( ) −2
n0 λ p − ne λ p ( )
1 1
−
(λ )
2
(30a)
2
2
nc2
λ 0 n0 (λs )cos α / λs + (n0 (λi ) / λi ) − (n0 (λs )sin β / λ )
2
p
β (λ s ,α ) = sin −1
n0 (λ0 ) λi sin α . (30b)
n0 (λi ) λs
Here the three wavelengths ( λ j ’s) satisfy the relation, λi = λ−p1 − λ−s 1 ( )
−1
, and
no and ne are the refractive indices of the ordinary and extraordinary rays,
respectively. The wavelength dependency of the refractive indices is
obtained by the Sellmeier equations.
3.0
2.5
Wavelength ( µm)
1.5 α=0°
α=6°
1.0
0.5
opacity region
0.0
22 26 30 34 38
θ (deg)
The curves of θ(λs, α) for various α with pump wavelength at 395 nm are
shown in Figures 6-3 and 6-4. There are two branches corresponding to the
signal and idler in the phase-matching curve for each non-zero α. These
figures show that the most characteristic feature of the phase-matching curve
is the broad spectral range. The region filled in gray in Figure 6-3 can satisfy
the phase-matching condition by a noncollinear interaction. For a given θ,
the signal and idler waves can be simultaneously emitted in a broad range of
wavelength as shown in Figure 6-5. For a smaller angle between the pump
beam and the crystal axis θ < 29.4° , the range of the spectrum is partially
limited, whereas for θ > 29.4° , optical modes extending from 450 nm to 3
µm are excited. This limit is due to the absorption of the idler in the crystal.
The broad bandwidth can be explained in terms of noncollinear
configuration with an extended group-velocity matching condition.
900
Signal Wavelength (nm)
800
GVM s-i =0
700 6°
600 5°
α =0°
4°
500 2°
3°
24 26 28 30 32 34 36 38
θ (deg)
from the Fourier transform of the gain spectrum shortest possible pulse
width is estimated to be 4.4 fs. Note again that the idler waves are also
generated with the same broad bandwidth in the NIR with the angular
dispersion to be phase-matched with the broad spectrum of the signal waves
(see Figure 6-6).
ks(λs)
ki(λp)
α = fixed β (λi)
y
θ kp
37 20
Idler Noncollenear Angle
Crystal Angleθ (degree)
36
35 15
G
34 G
T
I
G
33 10 F
32
β
31 5
30
29 0
450 500 550 600 650 700 750 800
Wavelength (nm)
Figure 6-7. Wavelength dependence of the parametric gain [Equation (3)] in the type-I
NCPM for the pump-signal noncollinear angle of α = 3.7°. Here the experimental parameters
Lc = 1 mm and Φ = 50 GW/cm2 are used.
and suited to be used as the seed of an OPA. The spectrum of the single-
filament continuum covers a broad spectral range from 450 nm to longer
than 1000 nm. There is a strong spike around 790 nm because of the
fundamental. Outside this region, the spectrum is smooth and flat, i.e., well-
suited for obtaining a smooth, short pulse without temporal structure because
of the Fourier-transformation relation.
BS 1kHz, 790nm
790nm SHG
HS 300µJ, 120fs
1µJ
395nm sapphire
100µJ plate white
light prism-pair
x compressor
NF y
pump z 550-690nm output
inverted 2-3µJ
D telescope signal αext PS
BBO
idler
Figure 6-8. Experimental setup for the noncollinear-phase-matched OPA (NOPA). SHG:
nonlinear crystal for second-harmonic generation; BS: beam splitter; HS: fundamental and
second-harmonic separator; NF: notch filter centered at 800 nm; D: variable optical delay
line; PS: periscope for rotating the polarization of the signal. The conelike parametric
fluorescence with the minimized dispersion (see text) is illustrated with the external cone
angle αext. Also illustrated are the crystal axes x, y, and z.
Figure 6-9. Spectra of the amplified signal. The center wavelength can be tuned only by
scanning the delay line of the pump by 50 µm.
Figure 6-10. Angular-dispersion property of the idler. Measured external noncollinear angles
βext with respect to the pump are shown (full circles at λs = 680 nm and open squares at λs =
600 nm). Also shown is the calculated phase-matching curve (solid curve).
2 2
Intensity (arb. units)
Intensity (arb. units)
1 0
500 600 700
Wavelength (nm)
0
-100 -50 0 50 100
Delay (fs)
Figure 6-11. Intensity autocorrelation trace of the amplified signal after pulse compression
(full circles). Autocorrelation measured with a second-harmonic generation crystal. The
sech2-fit (solid curve) pulse width is 14 fs (FWHM). The spectrum is shown in the inset.
154 Chapter 6
The measured pulse width and time-bandwidth product over the pump-
delay-tuning range are shown in Figure 6-12. The bandwidth is 700 cm-1 at a
central wavelength of 550 nm. The bandwidth reaches up to 2000 cm-1 when
the center wavelength is tuned to 690 nm. For both tunings, the pulse
duration is sub-20-fs. The time-bandwidth product varies from 0.6 to 1.1,
except at 550 nm where it is 0.4. In the long wavelength region, the spectrum
spreads to beyond 800 nm with a non-negligible spike that degrades the
time-bandwidth products to be larger than 1. The large time-bandwidth
product is partly due to the pulse-width measuring apparatus that may
overestimate the pulse width. The devices responsible could include the
autocorrelator composed of a 100 µm thick BBO with a group-velocity
mismatch of 40 fs at 600 nm and dispersive media, such as the lens and
beam splitter. Using the 3.3 fs/step delay stage, the time resolution of the
pulse stage used for the delay line also hinders accurate measurement of the
pulse width. The wings on both sides observed in the autocorrelation traces
also indicate higher-order dispersion. These problems have been eliminated
as discussed below.
U
-1
H
O
E
J
V
F
J
V K
F 9
K
Y
sub-20fs G
U
F
P N
C W
2
$
%GPVGT9CXGNGPIVJ
PO
Figure 6-12. Wavelength dependence of the pulse width (full circles, sech2-fit) and time-
bandwidth product (full squares). The dotted line indicates the product of transform-limited
pulse (0.315) in the case of a sech2-pulse envelope.
νs
tan α = tan γ . (31)
c
d tan γ
δttilt = , (32)
c
dε tan γ
=− . (33)
dλ λ
Figure 6-13. Schematic of the noncollinear interaction between the pump pulse and signal
pulse. The crystal is on the left side. The volumes occupied by the pump pulse are shown by
rectangular shapes elongated perpendicular to the propagation direction. Those for probe
pulse are shown by closed gray rectangles. The gain volume introduced by the pump pulse
causes the tilting of the wave front of the signal pulse by α in the crystal, resulting in the
special and angular dispersions of the exiting signal pulse.
2
total
slit position
0
500 550 600 650 700
Wavelength (nm)
Figure 6-14. Spatial chirp of the signal spectra after passing through the vertical slit. Slits are
located at three different positions (thin curves) after the nonlinear crystal of the NOPA. The
case of the fully open slit is shown in bold.
620
Wavelength (nm)
610
600
γ =6.3°
590 ( α=3.7°)
580
-5 -2.5 0 2.5 5
Relative Exit Angle (mrad)
dϕ 2 sin α apex dn
tan γ prism = −λ =− λ ', (34)
dλ cos φ1 cos φ 2 dλ '
'
158 Chapter 6
where n is the refractive index of the prism, α apex is the prism apex angle,
and φ1 ' is the internal incident angle. The tilt angle is then decreased and the
pulse width broadened during propagation in free space by the spectral
lateral walk-off over the beam cross section [69]. The following telescope
recollimates the spectral lateral walk-off and images the tilted fronts on the
focal plane with a longitudinal magnification factor, M = f1 / f 2 , which
gives the tilt angle γ ext at the crystal position
f1
tan γ ext = tan γ prism . (35)
f2
The internal tilt angle γ int is reduced by refraction with the relation
νg
tan γ int = tan γ ext . The pulse-front matching condition γ int = α is then
c
v g f1
tan α = tan γ prism . (36)
c f2
idler
130µJ, 150fs
100mm 88mm 270mm 50mm
@395nm
magnification
γext=6.4 γprism=2.3
Figure 6-16. Geometry of pulse-front matching optical system. L: lens; M: mirror; CYM:
cylindrical mirror; SM: spherical mirror; γint: internal tilt angle of pulse front of probe pulse in
the nonlinear crystal of NOPA; γext: external tilt angle of pulse front of probe pulse out of the
nonlinear crystal of NOPA; BBO: nonlinear crystal (β-barium borate).
6. FEMTOSECOND NONCOLLINEAR PARAMETRIC 159
AMPLIFICATION AND CARRIER-ENVELOPE PHASE CONTROL
dφ 1 d 2φ
φ (ω ) = φ (ω 0 ) + ( ) (ω − ω 0 )
ω − ω +
2
2
dω ω 0 2 dω ω
0
0
(37)
1d φ 1d φ
K
3 4
+ (ω − ω 0 ) + (ω − ω 0 ) +
3 4
3
3! dω ω 4! dω 4 ω
0 0
T (ω ) =
dφ d 2 φ
=
dω dω 2
(ω − ω 0 ) +
1 d 3φ
( ω − ω0
2 dω 3
)
2
ω0
(38)
1 d 4φ
+ (ω − ω 0 ) +
6 dω 4
3
L
is used for pulse-propagation characterization. The constant term induces
only a temporal shift during the pulse propagation without any shape change
L
and is thus neglected in Equation (38). The coefficients of the expansion
φ '' (ω ), φ ''' (ω 0 ), φ '''' (ω 0 ), are called GD dispersion (GDD), third-order
dispersion (TOD), fourth-order dispersion (FOD), and so on, respectively.
In the propagation through a medium with the path length l and wave
vector κ = (ω ) = n(ω )ω / c , the phase shift φ (ω ) = κ (ω )l yields
160 Chapter 6
T (ω ) =
l
, (39)
ν g (ω )
which gives the intuitive representation of the GD. However, the concept of
the GD can be represented in terms of wavelength-dependent optical paths
P(ω ) such as a grating pair [46] or prism pair [48].
Because the higher terms, such as a TOD and FOD, become significant in
a sub-10 fs regime with a broad spectral range, a more precise phase
correction is needed for the compression to the transform-limit.
Frequency (THz)
550 500 450 400
100
0
-100
-200 prism pair
-300 prism separation=1 M
material amount=6.7mm
200
100
-100
UBCM
8 reflections
150
air
100 4.5 M
50
50
40 beam splitter
30 0.5 mm
20
10
0
-10
100
50
0
-50
-100
-150
-200 Total
-250 -(measured GD)
-300
500 550 600 650 700 750 800
Wavelength (nm)
Figure 6-17. Group-delay property of the compressor. The group delays of the prism pair, the
UBCM pair with four round trips (thin, solid curve), the prism pair (dashed curve), air with a
4.5 m path length through the air, the beam splitter in the FRAC, and the whole compressor
system are shown from top to bottom. Also shown is the measured group delay of the signal
(full circles) in the bottom box with the sign reversed.
LBO HS 790nm
P2 395nm
TP
L1 <5fs, 5µJ
CM
He-Ne laser
VND WSM
BS pump
~1µJ 100µJ P1 CM
S L2 FRAC
PS
CM TBS
CF signal PD
CM BBO
idler BBO CCM
D BPF
2ω FD
120fs,400µJ PMT CCM
1kHz,790nm SM SM
Figure 6-18. Schematic of the visible sub-5-fs pulse generator. L: lens; M: mirror; CYM:
cylindrical mirror; SM: spherical mirror; BS: beam splitter; HS: harmonic separator; TP:
prism for pulse-front tilting; L1, L2: lenses for the telescope; SMs: spherical mirrors (γ = 100
mm); VND: variable neutral-density filter; WSM: spherical mirrors (γ = 120 mm); CF: cut-off
filter; D: optical delay line; PS: periscope; P1,P2: 45° fused silica prisms; and CCMs: corner-
cube mirrors.
beam interaction also avoids the spectral broadening of the signal caused by
beam divergence [66] that is accompanied by an undesirable spatial chirp.
Careful attention to both of these details is essential to obtain a signal
compressible to the transform-limit.
Frequency (THz)
600 500 400
2
Intensity (arb. units)
φ (ω ) = ∫
ω
[T pulse (ω ) + Tcompressor (ω )]dω . (40)
The scattered values of the measured group delay are caused partly by the
poor time resolution and by intensity fluctuation. The measured group delay
is smoothed by fitting to a cubic-polynomial function. Using a calculated
phase, the FRAC trace of the compressor output is then constructed and
depicted by open circles in Figure 6-20.
164 Chapter 6
8
DIFT
7 (4.7fs)
Normalized Intensity
6 sech2-fit
5 (3.5fs)
4
3
2
1
0
-15 -10 -5 0 5 10 15
Delay Time (fs)
8
Normalized Intensity
1
0
-40 -30 -20 -10 0 10 20 30 40
Delay Time (fs)
The second approach is to fit the measured FRAC trace with parameters
describing the phase. The complicated wavelength dependence of the group
delay of the UBCMs is included. The residual phase can be reasonably
assumed to have a smooth property and is determined by fitting the trace to a
cubic-polynomial function of the frequency [5, 6]. Because an
autocorrelation is used, the direction of the time axis is not determined by
this process alone. By comparison with the calculated phase mentioned
above, the direction can be determined. The fitted FRAC trace is shown in
Figure 6-20 with full circles.
Both phase profiles show similar behavior. The deviation is within π/2
radian over the whole spectral range from 510 to 790 nm. The oscillations
are due to the group delay of the UBCMs. The large deviations in the regions
of < 500 nm and > 800 nm are due to the limited bandwidth of the UBCMs.
However, UBCMs are less effective for pulse-width broadening because the
spectral intensity in these regions is weak. The difference between both
phases is mainly caused by the uncertainty of the measured group delay
6. FEMTOSECOND NONCOLLINEAR PARAMETRIC 165
AMPLIFICATION AND CARRIER-ENVELOPE PHASE CONTROL
2
Intensity (a. u)
1 4.7 fs
(TL 4.4 fs)
0
-20 -10 0 10 20
Time (fs)
Figure 6-21. Intensity profile of the compressed pulse. The transform-limited pulse width
determined by the spectrum is 4.4 fs (thin solid curve), whereas the calculated (thick solid
curve) and Fourier-transform-fit to FRAC with phase parametric (dashed curve) pulse widths
are 4.7 fs and 4.7 ± 0.1 fs, respectively.
9. SECOND-GENERATION NONCOLLINEAR
PARAMETRIC AMPLIFIER
To spectrometer
Figure 6-22. Schematic of experimental setup of sub-4 fs NOPA. λ/2: 800 nm wave-plate; SP:
2 mm sapphire plate; P1, P2: 45°quartz prisms; P3: 69° quartz prism (the distance from P3 to
the NOPA crystal is 80 cm); CM1,2: ultrabroadband chirped mirrors; GR: 300 lines/mm ruled
diffraction grating (Jobin Yvon); SM: spherical mirror, R = -400 mm; BS1,2: chromium-
coated d = 0.5 mm quartz beam splitters; SHG crystal: 0.4 mm θ = 29° BBO (EKSMA);
NOPA crystal: 1 mm θ = 31.5°BBO (Casix); and SHG FROG crystal: θ = 29° BBO wedge
plate d = 5÷20 µm (EKSMA). Spherical mirrors around the NOPA crystal are R = -200 mm.
Thick arrows on the left indicate the data flow from the pulse diagnostic setup (SHG FROG)
and the feedback to the flexible mirror.
amplified visible pulses with a bandwidth of nearly 200 THz [36, 37, 41].
The uniqueness of this pumping arrangement comes from the fact that the
pump inclination to the direction of the seed (~ 3.7°) matches almost
perfectly the angle of birefringent walk-off between the o and e waves inside
the crystal. As a result, BBO as long as 1–2 mm can be employed even in 5
fs NOPAs. The implications of noncollinear phase matching are well
understood and discussed in numerous papers [26, 29, 31, 32, 34, 36-40, 66,
72].
We paid further attention to the subtleties of parametric amplification in a
noncollinear configuration [36, 38], by considering the effect of pulse-front
tilting of the signal wave on the ability to compress the signal into a sub-5 fs
pulse. To prevent the tilting of the signal pulse in space, an effect that also
results in angular dispersion of the amplified pulse, several researchers
proposed to use a pump beam with a tilted wave front. This configuration,
named pulse-front-matching, was implemented by sending the pump beam
through a prism and adjusting the pulse tilt with a telescope consisting of
two convex lenses [38, 41, 73].
We took care to find a balanced Mach-Zehnder interferometric
autocorrelator [74] to characterize the NOPA pulses [38, 41, 42]. This
particular autocorrelator does not have spatial resolution and, therefore, is
insensitive to wave-front tilt. This does not necessarily mean the real pulse
width does not suffer from broadening of the pulse width due to pulse-front
tilting, however. The required spatial sensitivity can also be achieved by
using an asymmetric number of reflections in the interferometer.
Together with the effect of pulse-front matching, the angular dispersion
of the pump beam in the sub-5 fs NOPA is an important factor in terms of
enhancing the phase-matching bandwidth. Even with relatively thick second-
harmonic crystals (1–2 mm LBO or BBO) and comparatively long (120–150
fs) pulses from standard regenerative amplifiers employed to pump the
NOPAs [37, 38, 41, 42], the resulting second-harmonic radiation has a
bandwidth of several nanometers. Wide-bandwidth parametric amplification
can subsequently be achieved by pointing the seed beam in a specific
direction for each spectral component of the pump, as schematically shown
in Figures 6-23 and 6-24. The adjustment of the pump beam dispersion,
required for phase-matching optimization, is obtained by selecting the apex
angle of a prism in the second-harmonic pathway [Figure 6-23(b)] and the
distance from the prism to the focusing optic. The required second-harmonic
(SH) dispersion
168 Chapter 6
l
δα (λ ) = atan tan(γ (λ0 ) − γ (λ ) ) , , (41)
f
sin γ 0
γ (λ ) = asinn(λ ) sin θ apex − asin . (42)
n(λ )
In Equation (42), γ 0 is the angle between the SH beam and the normal to the
input face of the prism (incidence angle onto the prism), θ apex stands for the
apex angle of the prism, and n(λ ) is the refractive index of glass.
(a)
(b)
Frequency [THz]
650 600 550 500 450 400
λp1=389 nm
λp2=392 nm
λ p3=395 nm
32
Phase-matching angle [deg]
31 1
Intensity
6 nm
Direction of
seed light
30
0
375 380 385 390 395 400 405 410
SH wavelength [nm]
500 550 600 650 700 750 800
Wavelength [nm]
Figure 6-24. Phase matching in a 1 mm Type I BBO crystal, θ = 31.5°. Dark-shaded contours
indicate FWHM of the angular phase matching for individual monochromatic pump
wavelengths after optimization of their incidence angles. Dashed line denotes the direction of
the seed beam. Inset depicts experimentally measured second-harmonic spectrum used to
pump the NOPA (shaded contour) and calculated conversion efficiency of a 0.4 mm Type I
BBO second-harmonic-generation crystal.
ACKNOWLEDGEMENTS
REFERENCES
[1] A. Baltuška, Z. Wei, M. S. Pshenichnikov, D. A. Wiersma, and R. Szipocs,
Appl. Phys. B 65, 175-188 (1997).
[2] A. Baltuška, Z. Y. Wei, M. S. Pshenichnikov, and D. A. Wiersma, Opt.
Lett. 22, 102-104 (1997).
[3] A. Baltuška, M. S. Pshenichnikov, and D. A. Wiersma, Opt. Lett. 23, 1474-
1476 (1998).
[4] M. Nisoli, S. DeSilvestri, and O. Svelto, App. Phys. Lett. 68, 2793-2795
(1996).
[5] M. Nisoli, S. DeSilvestri, O. Svelto, R. Szipocs, K. Ferencz, C. Spielmann,
S. Sartania, and F. Krausz, Opt. Lett. 22, 522-524 (1997).
[6] M. Nisoli, S. Stagira, S. DeSilvestri, O. Svelto, S. Sartania, Z. Cheng, M.
Lenzner, C. Spielmann, and F. Krausz, Appl. Phys. B 65, 189-196 (1997).
172 Chapter 6
[7] R. Szipocs, K. Ferencz, C. Spielmann, and F. Krausz, Opt. Lett. 19, 201-
203 (1994); R. Szipocs and A. Köházi-Kis, Appl. Phys. B 65, 115-135
(1997); R. Szipocs and F. Krausz, Dispersive dielectric mirror, USA Patent
5734503 (1998).
[8] T. Brabec and F. Krausz, Rev. Mod. Phys. 72, 545-591 (2000).
[9] G. P. Agrawal, Nonlinear Optics (Academic Press, San Diego, 2001).
[10] J. K. Ranka, R. S. Windeler, and A. J. Stentz, Opt. Lett. 25, 25-27 (2000); J.
K. Ranka, R. S. Windeler, and A. J. Stentz, Opt. Lett. 25, 796-798 (2000).
[11] I. D. Jung, F. X. Kärtner, N. Matuschek, D. H. Sutter, F. Morier-Genoud,
G. Zhang, U. Keller, V. Scheuer, M. Tilsch, and T. Tschudi, Opt. Lett. 22,
1009-1011 (1997); Y. Chen, F. X. Kärtner, U. Morgner, S. H. Cho, H. A.
Haus, E. P. Ippen, and J. G. Fujimoto, J. Opt. Soc. Am. B 16, 1999-2004
(1999); U. Morgner, F. X. Kärtner, S. H. Cho, Y. Chen, H. A. Haus, J. G.
Fujimoto, E. P. Ippen, V. Scheuer, G. Angelow, and T. Tschudi, Opt. Lett.
24, 411-413 (1999); G. Steinmeyer, D. H. Sutter, L. Gallmann, N.
Matuschek, and U. Keller, Science 286, 1507-1512 (1999).
[12] R. K. Shelton, L. S. Ma, H. C. Kapteyn, M. M. Murnane, J. L. Hall, and J.
Ye, Science 293, 1286-1289 (2001); R. K. Shelton, S. M. Foreman, L. S.
Ma, J. L. Hall, H. C. Kapteyn, M. M. Murnane, M. Notcutt, and J. Ye, Opt.
Lett. 27, 312-314 (2002).
[13] Z. Y. Wei, Y. Kobayashi, Z. G. Zhang, and K. Torizuka, Opt. Lett. 26,
1806-1808 (2001).
[14] P. E. Powers, R. J. Ellingson, W. S. Pelouch, and C. L. Tang, J. Opt. Soc.
Am. B 10, 2162-2167 (1993).
[15] V. Petrov, F. Seifert, O. Kittelmann, J. Ringling, and F. Noack, J. Appl.
Phys. 76, 7704-7712 (1994); M. K. Reed and M. K. Steiner-Shepard, IEEE
J. Quantum Electron. 32, 1273-1277 (1996).
[16] K. R. Wilson and V. V. Yakovlev, J. Opt. Soc. Am. B 14, 444-448 (1997).
[17] V. G. Dmitriev, G. G. Gurzadyan, and D. N. Nikogosyan, Handbook of
Nonlinear Optical Crystals (Springer-Verlag, Berlin, 1997).
[18] F. Seifert, V. Petrov, and F. Noack, Opt. Lett. 19, 837-839 (1994); M.
Nisoli, S. De Silvestri, V. Magni, O. Svelto, R. Danielius, A. Piskarskas, G.
Valiulis, and A. Varanavicius, Opt. Lett. 19, 1973-1975 (1994).
[19] M. K. Reed, M. K. Steiner-Shepard, and D. K. Negus, Opt. Lett. 19, 1855-
1857 (1994).
[20] M. K. Reed, M. S. Armas, M. K. Steiner-Shepard, and D. K. Negus, Opt.
Lett. 20, 605-607 (1995).
[21] V. V. Yakovlev, B. Kohler, and K. R. Wilson, Opt. Lett. 19, 2000-2002
(1994).
[22] S. R. Greenfield and M. R. Wasielewski, Opt. Lett. 20, 1394-1396 (1995).
[23] S. Takeuchi and T. Kobayashi, J. Appl. Phys. 75, 2757-2760 (1994).
[24] E. S. Wachman, W. S. Pelouch, and C. L. Tang, J. Appl. Phys. 70, 1893-
1895 (1991).
[25] Q. Fu, G. Mak, and H. M. van Driel, Opt. Lett. 17, 1006-1008 (1992); A.
Shirakawa, H. W. Mao, and T. Kobayashi, Opt. Commun. 123, 121-128
6. FEMTOSECOND NONCOLLINEAR PARAMETRIC 173
AMPLIFICATION AND CARRIER-ENVELOPE PHASE CONTROL
Abstract: The new alliance between the time and frequency domains in laser
spectroscopy has made it possible to develop ultrafast counting schemes that
can keep track of single optical oscillations. With this achievement, counting
optical oscillations of more than 1015 cycles in one second has become a
simple task. High-resolution spectroscopy for basic research and metrology
greatly benefit from this technology as it has allowed the highest possible
precision. This development has also led to the construction of all-optical
atomic clocks that are expected to eventually outperform current state-of-the-
art cesium clocks. The possibility to measure almost any frequency ratio with
very high precision can be used to search for possible variations of natural
constants.
Key words: hydrogen spectroscopy, optical atomic clocks, optical frequency metrology,
optical frequency synthesis
1. FREQUENCY COMBS
nonlinear devices usually produce weak signals, at least when they are
driven with a continuous wave (cw). Electronic phase-locked loops can be
used to stabilize any kind of oscillator, even lasers, provided their intrinsic
stability is sufficient so that there is no need for very rapid frequency
corrections. Repeating the multiplication and phase-lock procedure many
times makes it possible to convert a reference radio frequency, say from an
atomic clock, to much higher frequencies. Because of the large number of
steps necessary to build a long harmonic frequency chain, it was not before
1995 when visible laser light was first referenced phase coherently to a
cesium atomic clock using this method [2].
The disadvantage of these harmonic frequency chains was not only that
they could easily fill several large laser laboratories at one time, but that they
could be used to measure only a single optical frequency. Even though
mode-locked lasers for optical frequency measurements were used in
rudimentary form in the late 1970s [3], this method did not become practical
until the advent of femtosecond (fs) mode-locked lasers. Such a laser
necessarily emits a very broad spectrum, comparable in width to the optical
carrier frequency. Currently the field’s work horse is the Ti:sapphire Kerr-
lens mode-locked laser, but fiber-based lasers are expected to take over for
frequency metrology applications.
In the frequency domain, a train of short pulses from a femtosecond
mode-locked laser is the result of a phase-coherent superposition of many cw
longitudinal cavity modes. These modes at ωn1 form a series of frequency
spikes called a frequency comb. The individual modes can be selected by
phase locking other cw lasers to them. As has been shown, the modes are
remarkably uniform, i.e., the separation between adjacent modes is constant
across the frequency comb [4, 5]. This strictly regular arrangement is the
most important feature used for optical frequency measurement and may be
expressed as:
ω n = nω r + ω CE . (1)
Here the mode number n of some 105 may be enumerated such that the
frequency offset ωCE lies in between 0 and ωr = 2π/T. The mode spacing is
thereby identified with pulse repetition rate, i.e., the inverse pulse repetition
1
The notation used in this chapter corresponds to that of other chapters through the following
relationships:
ωn = 2πνn
ωr = 2πfr
ωCE = 2πf0
∆φ = ∆φce
178 Chapter 7
time T. With the help of Equation (1), two radio frequencies ωr and ωCE are
linked to the optical frequencies ωn of the laser. For this reason, mode-
locked lasers are capable of replacing the harmonic frequency chains of the
past.
To derive these frequency comb properties [6, 7], as detailed by Equation
(1), it is useful to consider the electric field E(t) of the emitted pulse train. If
the pulses were exact time-shifted copies, E(t) = E(t - T), a simple Fourier
transformation would yield a strictly periodic spectrum with a mode
separation of ωr and a zero comb offset ωCE = 0. However, this is not what
occurs in a real laser. Because of intracavity dispersion, the group and phase
velocities do not match for the pulse that is stored in the cavity. This causes
the carrier wave to continuously shift with respect to the pulse envelope (see
Figure 7-1). The pulses that emerge at the output-coupling mirror after each
round trip show a discrete pulse-to-pulse carrier-envelope phase shift of ∆φ.
E (t)
∆7 2∆7
Fourier
transformation
E (1)
1CE
1r 1c 1
Figure 7-1. Three consecutive pulses of the pulse train emitted by a mode-locked laser and
the corresponding spectrum. The pulse-to-pulse phase shift ∆φ results in an offset frequency
ωCE = ∆φ/T because the optical carrier wave at ωc moves with the phase velocity while the
envelope moves with the group velocity.
strongly suppressed. For this reason, the properties of the pulse train are
most conveniently derived by separating the optical carrier wave that
propagates with its phase velocity at ωc from the pulse-envelope function:
E (t ) = A(t )e −iω ct . The pulse repetition time T = 2π / ω r is determined from
the group velocity by demanding that A(t) = A(t - T). Since it is strictly
periodic it may be written in terms of a Fourier series
The cesium D1 frequency can be used to derive a value for the fine
structure constant α by taking advantage of the extremely well-known
Rydberg constant, R∞ = α 2 cme / 2h . Steven Chu and collaborators at
Stanford University have measured the photon recoil shift of the D1 line
f rec = f D21 h / 2mCs c 2 with an atom interferometer aiming for an accuracy
near the parts-per-billion (ppb) level. Together with the very precise values
for the proton–electron mass ratio mp /me [9] and the cesium–proton mass
ratio mCs /mp [10], the fine structure constant is derived according to
2 R∞ h 2 f c mCs m p
α2 = = 2 R∞ rec . (3)
c me f D21 m p me
The atomic mass ratios are measured with high precision in terms of the
ratios of their cyclotron frequencies in Penning traps. Please note that all
necessary ingredients to Equation (3), including the Rydberg constant that is
determined by spectroscopy on atomic hydrogen, are based on intrinsically
accurate frequency measurements. Because α scales all electromagnetic
interactions, it can be determined by a variety of independent physical
methods. Unfortunately, different values measured with comparable
accuracy disagree with each other by up to 3.5 standard deviations [11], and
the derivation of the currently most accurate value of α from the electron g -
7. OPTICAL FREQUENCY MEASUREMENT 181
3. OPTICAL SYNTHESIZERS
I(1)
n1r+1CE 2n1r+1CE 1
x2
2(n1r+1CE)
Figure 7-3. The offset frequency ωCE that displaces the modes of an octave-spanning
frequency comb from being exact harmonics of the repetition rate ωr is measured by
frequency-doubling some modes at the “red” side of the comb and beating them with modes
at the “blue” side.
In that case, a group of modes from the low-frequency end of the comb
around nLω r + ω CE is frequency doubled in a nonlinear crystal and beat
notes with another group of modes around nH at the high-frequency end of
the comb are created. As sketched in Figure 7-3, this technique creates
signals at the difference frequency of two optical waves on a photodetector:
2(n Lω r + ω CE ) − ( n H ω r + ω CE ) = ( 2nL − nH )ω r + ω CE . These beat
frequencies are created for all possible combinations of nL and nH. However,
the detected beat notes cannot exceed the bandwidth of the photodetector.
One can further restrict the number of signals to precisely one by using an rf-
182 Chapter 7
low-pass filter. If this filter has a cut-off frequency of ωr/2, only the beat
note at ωCE, belonging to all combinations with 2n L − nH = 0 , remains. The
octave-spanning comb ensures that these combinations exist, i.e., that the
modes nL and nH with nH = 2nL are simultaneously active modes. Knowing
both ωCE and ωr means that frequencies of all modes according to Equation
(1) are known and may be used for optical frequency measurements. This is
accomplished by creating another beat note ωb between an unknown laser at
ωl and a nearby mode of the comb. The frequency of that laser can then be
determined by ω l = nω r + ω CE + ω b . The mode number n may be
determined by a coarse measurement of ωl with a wave meter or by
repeating the measurement with slight variation in ωr.
This technique is called self-referencing [5, 14] and requires a frequency
comb spanning a full optical octave in its simplest form. Alternatively, any
combination of two different harmonics from the same comb will serve that
purpose. Self-referencing has been performed with the beating of the 2nd
and 3rd harmonic of comb components [15] and by comparing modes with
frequency ratio of 7:8 [16]. The frequency comb, even without the
knowledge of ωCE, can be thought of as a ruler in frequency space that
allows the precise measurement of large optical frequency differences. The
initial utilization of frequency combs was for that purpose [8].
To stabilize the frequency comb to a precise rf reference, it is not
sufficient to measure the two comb parameters ωr and ωCE; they also have to
be controlled separately. The repetition frequency is most conveniently
controlled via the cavity length through a piezoelectric transducer. For a
typical laser, the intrinsic stability of the repetition rate is high enough that
the attainable bandwidth of the transducer is sufficient to keep it in phase
with a good rf reference. The other comb parameter ωCE is controlled by
changing the intensity of the stored pulse. It was shown that for a soliton-like
laser such as the Ti:sapphire Kerr-lens mode-locked laser, the group and
phase velocity depend in a different way on the peak intensity of the stored
pulse [17]. Therefore one can adjust the pulse-to-pulse slippage of the
carrier-envelope phase, and thereby ωCE, by controlling the average laser
power. For a Ti:sapphire Kerr-lens mode-locked laser, this can be done by
controlling the power of the pump laser. In this way the comb offset, just
like the repetition frequency, can be adjusted to stay in phase with the rf
reference. The two controls are not completely independent, but they affect
the round-trip group delay and the round-trip phase delay differently.
Keeping the servo bandwidths for ωr as low as possible and for ωCE as high
as possible usually decouples the two servo systems. Having two two-phase
servo systems operational that lock ωr and for ωCE to a radio frequency
reference such as an atomic clock ensures that, by virtue of Equation (1) and
7. OPTICAL FREQUENCY MEASUREMENT 183
Φ NL (t ) = − n2 I (t )ω c l / c , with I (t ) = A(t ) .
2
(4)
Because Φ NL (t ) has the same periodicity as A(t), the comb structure of the
spectrum, as previously derived, is not affected.
In an optical fiber, this process can be quite efficient as compared to bulk
material, even though the nonlinear coefficient in fused silica is
comparatively small. This efficiency arises because of the fiber core carrying
a high-intensity pulse over an extended length. The details of fiber actions on
the pulse train are, in fact, more complicated because other nonlinear effects
such as Raman scattering may contribute. Whatever the fiber does to the
pulses, however, a periodic input should produce a periodic output of some
kind if the fiber action is the same for all the pulses. Therefore the comb
structure, as derived from the periodic envelope function, is maintained.
Effective self-phase modulation in a regular single-mode fiber takes
place only over a limited fiber length. Dispersion causes the pulse to broaden
temporally. This broadening continuously lowers the peak intensity as the
pulse travels along the fiber until the nonlinear interaction comes to a halt.
184 Chapter 7
5. APPLICATION TO HYDROGEN
laboratory was the use of optical-frequency interval dividers [23, 24]. The
concept was to phase lock the second harmonic of a laser 2f to the sum
frequency generated by two additional lasers at f1 + f2. In doing so, the laser
frequency f is set at the precise midpoint between f1 and f2: f = (f1 + f2)/2. In
this way, an arbitrary large frequency gap f2 - f1 can be divided in two. These
dividers can be cascaded such that n stages would divide an optical
frequency gap by 2n. With a 5-stage interval-divider chain, we have divided
the aforementioned gap to become accessible to radio frequency counting
techniques. At the time, the accuracy of the experiment was limited by the
reproducibility of our He-Ne laser standard to about 840 Hz [6].
Already then it was clear that, given a way to measure frequency gaps
that are comparable in size to the optical frequency itself, absolute optical
frequency measurements could be performed using the ν-to-2ν method
described above [23]. The first electro-optic frequency comb was brought to
our lab by Motonobu Kourogi from the Tokyo Institute of Technology in
1997. This device was capable of measuring frequency gaps of several THz
using a single laser. It replaced the much-more-complicated interval divider
stages but was itself soon replaced by a femtosecond frequency comb. With
the frequency comb, it was possible for the first time to measure frequency
gaps comparable to the optical carrier frequency and thereby convert the 1S–
2S transition frequency phase coherently to the radio frequency domain [16].
Given a precise radio frequency reference, the accuracy was limited by the
hydrogen spectrometer.
Figure 7-4. Sketch of the first (still somewhat complicated) self-referenced frequency comb.
While the second harmonic of the dye laser at 486 nm is used to excite the 1S–2S transition,
its subharmonic 3.5f-∆f and the 4th harmonic of the He-Ne laser are compared to the
frequency comb. The ratio of these frequencies is 7:8 instead of the 1:2 in a much simpler ν-
to-2ν self-referencing scheme. Because the 3.39 µm He-Ne laser could not be tuned precisely
to the 28th subharmonic of the dye laser, a second, much smaller frequency gap ∆f was present
in this setup and was measured simultaneously with the same frequency comb. The oval
symbol represents an optical interval divider that fixes the center frequency 4f – ∆f relative to
its inputs 4f and 7f - 2∆f. All optical connections represent phase-locked lasers except where
the rf counters are located. These are used to measure beat notes with the nearest comb
modes. Once the two frequency gaps 0.5f and ∆f have been precisely measured with a Cs
clock-referenced frequency comb, the dye laser at 14f + 8∆f is readily calculated. In addition,
we can calculate the frequency of any other mode in the frequency comb and fix the carrier-
envelope phase by locking the comb with ωCE = 0.
7. OPTICAL FREQUENCY MEASUREMENT 187
The cw dye laser near 486 nm that is used to excite the hydrogen atoms
can be locked to a high-finesse reference cavity to narrow its line width and
fix its frequency. A new cavity made from Ultra-Low-Expansion (ULE)
glass was used for the 2003 measurement. Its drift was measured to be less
than 1 Hz s-1 for the entire time of the measurement. Due to better thermal
and acoustic isolation and improvements in the laser-locking electronics, the
laser line width was narrower than it had been in 1999. An upper limit for
the laser line width was deduced from an investigation of the beat signal
between two laser fields locked separately to independent Zerodur and ULE
cavities to about 120 Hz at a laser wavelength of 486 nm.
dye laser
486 nm
BBO
stabilization servo frequency doubler
AOM
253 nm
vacuum chamber
reference resonator
nozzle
frequency chopper
measurement Lα-detector rf dis-
charge
fork H
chopper
H2
Figure 7-5. Hydrogen spectrometer. The 486 nm light from the dye laser is frequency doubled
in a Barium β-Borate (BBO) crystal. The resulting radiation is used to drive the 1S–2S two-
photon transition that is free of the first-order Doppler effect if the two photons are counter-
propagating. The latter is ensured by a linear enhancement cavity in the vacuum chamber that
is held at a pressure of about 10-8–10-7 mbar. The acousto-optic modulator (AOM) allows the
dye laser to be tuned away from the resonance of the reference resonator to scan the hydrogen
line.
14 delay 1000 µs
0
8 -10 0 10 5.2
delay 200 µs
6
5.0
4 delay 1000 µs delay 400 µs
2 4.8
0
-40 -30 -20 -10 0 10 20 0 1 2 3 4 5 6 7 8
detuning @ 121nm [kHz] power @ 243 [a.u.]
Figure 7-6. Left: simultaneous fit of a 1S–2S transition spectrum recorded at different
detection delays. The nozzle temperature was 7 K. Right: AC Stark shift extrapolation to zero
power for all the lines recorded during one day of data taking.
600
fcavity - 616 516 065 769 800 Hz [Hz]
400
10-13
200
∼τ−1/2
0
-200
10-14
-400
-600
0 1000 2000 3000 4000 5000 1 10 100 1000
Figure 7-7. Left: Absolute frequency of the standing light field inside the reference resonator
(AOM shifted dye laser, see Figure 7-5) as a function of time. The solid line is a parabolic fit
to the data. The linear contribution to the drift is less than 0.1 Hz per s in this example and
below 1 Hz per s at all times. Right: Normalized Allan variance vs averaging time computed
from a time series of 1 s counter readings with considerable dead time. The straight line
indicates the τ -1/2 dependence, which is the signature of the Cs fountain clock. The raw data
analysis (solid squares) shows that the stability for averaging times longer than 20 s is limited
190 Chapter 7
by the drift of the ULE reference cavity. Open squares represent data corrected for the
parabolic cavity drift.
2
The result of 2 466 061 102 474 870 Hz was inadvertently described in [25] as “the
weighted mean value” but was calculated without consideration of the daily statistical
uncertainties.
7. OPTICAL FREQUENCY MEASUREMENT 191
Table 7-1. Results of the (1S, F = 1, mF = ±1 2S, F' = 1, m'F = ±1) transition frequency→
measurement (νH,1999, νH,2003) and uncertainty budgets (σH,1999, σH,2003) for the 1999 and 2003
measurements, respectively.
5200 44 months
f(1S2S) - 2 466 061 102 470 000 Hz [Hz]
5100
5000
4870(36) Hz
4851(25) Hz
average
4900
average
4800
4700
4600
June 1999 - July 1999 February 2003
4500
→
Figure 7-8. Experimental results and averages for the 1999 and 2003 measurements of the
(1S, F = 1, mF = ±1 2S, F' = 1, m'F = ± 1) transition frequency in atomic hydrogen. Each
data point represents one day of averaging. The two results are unweighted mean values of the
daily averages.
192 Chapter 7
where Ry is the Rydberg frequency expressed in hertz and Frel(α) takes into
account relativistic and many-body effects. The constant A is a pure
numerical factor and does not depend on any natural constant. For hydrogen,
one would use the Schrödinger energies with A = 1/n2 and Frel(α) = 1 to
describe the first-order scaling in α, which is sufficient here. The
corresponding expression for hyperfine transitions reads:
µn
ν hf = A′ R y α 2 ′ (α ) ,
Frel (7)
µB
α∂α ln FH (α ) ≈ 0 (9)
194 Chapter 7
α&
x = ∂ t ln(α ) = = ( −0.3 ± 2.0) × 10 −15 yr −1 , and (14)
α
Figure 7-9 shows the current limit for the drift rates from laboratory
measurements. The uncertainties of these limits reach within a factor of ten
to the uncertainties of the relative drift rates derived from astronomical
observations [30, 32]. In the near future, one can expect a significant
7. OPTICAL FREQUENCY MEASUREMENT 195
4
Hg+
2 H
-2 Yb+
y=
-4
-6 -4 -2 0 2 4 6
x= ln(α) [10-15 yr-1]
Figure 7-9. Combining the drift rates of optical transitions in hydrogen, Hg+ and Yb+. The
experimentally allowed region for the drift rates of α and µCs/µB is shown in the form of an
ellipse.
It has also been argued that a possible drift of the coupling constants
must be correlated [37], at least if one believes in the existence of a Grand
Unified Theory. In such a theory, the fundamental interactions are described
within the same framework. The drift rates must therefore be synchronized
in some sense. According to Calmet et al. [37], the fractional time variation
of hadron masses and along the same line the nuclear magnetic moments,
should change about 38 times faster than the fractional time variation of the
fine structure constant α. The analysis presented here is independent of such
3
If there was such a model that describes the drift of natural constants in terms of other
constants, it would simply mean that our current constants are not fundamental.
196 Chapter 7
REFERENCES
[1] L. O. Hocker, A. Javan, D. R. Rao, L.Frenkel, and T. Sullivan, Appl. Phys.
Lett. 10, 147-149 (1967); K. M. Evenson, J. S. Wells, F. R. Petersen, B. L.
Danielson, and G. W. Day, Appl. Phys. Lett. 22, 192 (1973).
[2] H. Schnatz, B. Lipphardt, J. Helmcke, F. Riehle, and G. Zinner, Phys. Rev.
Lett. 76, 18-21 (1996).
[3] J. N. Eckstein, A. I. Ferguson, and T. W. Hänsch, Phys. Rev. Lett. 40, 847-
850 (1978).
[4] T. Udem, J. Reichert, R. Holzwarth, and T. W. Hänsch, Opt. Lett. 24, 881-
883 (1999); S. A. Diddams, L. Hollberg, L. S. Ma, and L. Robertsson, Opt.
Lett. 27, 58-60 (2002); L. S. Ma, Z. Y. Bi, A. Bartels, L. Robertsson, M.
Zucco, R. S. Windeler, G. Wilpers, C. Oates, L. Hollberg, and S. A.
Diddams, Science 303, 1843-1845 (2004).
[5] R. Holzwarth, T. Udem, T. W. Hänsch, J. C. Knight, W. J. Wadsworth, and
P. S. J. Russell, Phys. Rev. Lett. 85, 2264-2267 (2000).
[6] T. Udem, A. Huber, B. Gross, J. Reichert, M. Prevedelli, M. Weitz, and T.
W. Hänsch, Phys. Rev. Lett. 79, 2646-2649 (1997).
[7] J. Reichert, R. Holzwarth, T. Udem, and T. W. Hänsch, Opt. Commun.
172, 59-68 (1999).
[8] T. Udem, J. Reichert, R. Holzwarth, and T. W. Hänsch, Phys. Rev. Lett. 82,
3568-3571 (1999).
[9] P. J. Mohr, Rev. Modern Phys. 72, 351-495 (2000).
[10] M. P. Bradley, J. V. Porto, S. Rainville, J. K. Thompson, and D. E.
Pritchard, Phys. Rev. Lett. 83, 4510-4513 (1999).
[11] A. Wicht, J. M. Hensley, E. Sarajlic, and S. Chu, Physica Scripta T102, 82-
88 (2002).
[12] T. Kinoshita, Reports on Progress in Physics 59, 1459-1492 (1996); T.
Kinoshita, IEEE Trans. Instrum. Meas. 50, 568-571 (2001).
[13] J. M. Hensley, Ph.D Thesis, Stanford University (2001).
[14] S. A. Diddams, D. J. Jones, J. Ye, S. T. Cundiff, J. L. Hall, J. K. Ranka, R.
S. Windeler, R. Holzwarth, T. Udem, and T. W. Hänsch, Phys. Rev. Lett.
84, 5102-5105 (2000).
[15] T. M. Ramond, S. A. Diddams, L. Hollberg, and A. Bartels, Opt. Lett. 27,
1842-1844 (2002).
[16] J. Reichert, M. Niering, R. Holzwarth, M. Weitz, T. Udem, and T. W.
Hänsch, Phys. Rev. Lett. 84, 3232-3235 (2000).
[17] H. A. Haus and E. P. Ippen, Opt. Lett. 26, 1654-1656 (2001).
[18] P. Russell, Science 299, 358-362 (2003).
7. OPTICAL FREQUENCY MEASUREMENT 197
Key words: frequency comb, frequency chains, optical atomic clocks, optical frequency
metrology, optical frequency references
1. INTRODUCTION
Ca Frequency Standard
PLL @ 60 MHz
455986.240494 GHz
2x
Diode Laser
227993.090247 GHz
12C18O
2x 2 -Laser P*(20)
Counter
28534.284 GHz
2x
CO2-Laser P(14) 3x DRO
28464.684 GHz 23.171 GHz
8x
PLL @ 10 MHz Gunn 126.5 GHz
phase control
CO2-Laser R(32)
Gunn
29477.165 GHz 3x 55.4 GHz
4x
PLL @ 100 MHz Klystron 73.35 GHz
Figure 8-1. Harmonic frequency chain to the Ca frequency standard. (PLL: phase-locked
loop; ⊗: harmonic mixers.)
202 Chapter 8
-12
10
-13
10
σy( 2, τ )
-14
10
1 10 100 1000 s
integration time
Figure 8-2. Allan standard deviation of Ca (■) and CH4 (□) frequency standards. For
comparison, the typical stability of an H maser is represented by the line without
measurement points.
ν1 + ν2. Complete frequency chains have been planned using this approach
[6]. However, difficulties were encountered in obtaining the lasers in certain
ranges of the visible spectrum.
To measure the frequency of the 1S–2S transition in atomic hydrogen,
Hänsch’s group in Garching built a frequency chain starting from a CH4-
stabilized reference laser at 3.39 µm and reaching the UV at 243 nm. The
technique combined conventional harmonic mixing with optical interval
bisection [13, 14]. The CH4-reference frequency itself had to be calibrated
using the harmonic-mixing frequency chain at PTB and a transportable CH4-
frequency standard.
Laser 1 ν1
Laser 2 ν2
SUM
ν1+ν2 - 2ν3
Detector
Servo
SHG
Laser 3 ν3 = (ν1+ν2) / 2
The second method (b) was conceived early [8]. Using an electro-optic
modulator, driven at a microwave frequency (typically 5–10 GHz) in an
optical resonator, a wide spectrum of modulation sidebands is produced. If
the round-trip time of the light in the optical resonator is synchronized with
the microwave period, the phase modulation of the light adds up over a
number of round trips. This means that the phase excursion of the phase
modulation is increased by a factor equal to the optical resonator finesse,
thereby increasing the number and strength of the modulation sidebands. In
this way, Kourogi [16] was able to generate frequency combs of 30 THz
widths. Such a frequency comb generator was used by Huber et al. [17] to
bridge a gap of ~2 THz in the 800 nm range. The width of the comb was
further increased by passing the optical radiation through a nonlinear fiber
that effects self-phase modulation.
During the last decade, a number of measurements of the properties of
hydrogen, such as the 1S–2S transition frequency [18], the isotope shift in
deuterium [19], the 1S Lamb shift [20], and the Rydberg constant [13, 21]
were performed via optical frequency measurements. Incorporating novel
techniques like Kurogi’s optical comb generator [16], the frequency chain in
Garching could be modified to measure the frequency of the In+ clock
transition at 236.5 nm [22]. More details of the various frequency
measurements performed with different concepts at different laboratories can
be found in recent review articles [23]. Although frequency-interval division
had proved its usefulness, no complete chain was ever built using this
principle.
spectral span of this comb reflects the duration of an individual pulse, while
the spacing between the lines is determined by the pulse repetition
frequency.
We start with a time domain description of the pulses emitted by a mode-
locked laser, as shown in the upper part of Figure 8-4. Every time the pulse
circulating inside the cavity hits the output coupler, the laser emits a pulse.
This process results in a train of pulses separated by time τ = lc / vg, where lc
is the round-trip length of the cavity and vg is the group velocity. Because of
unavoidable dispersion in the cavity, the group and phase velocities vp are, in
general, not equal. This inequality results in a shift of the carrier phase with
respect to the peak of the envelope for each round trip.
∆Φ
∆φce
Time domain:
time
τ == 1/f
∆t 1/frep
r
Frequency domain:
νν(m)
(m) ν(2n)
ν(2n)
νfceo
0
fr
frep
ν(m)
ν(m) = ν=ceof0++mmfrep
fr
SH frequency
ν=0
νfceo
0 == 2 ν(n)- -ν(2n)
2ν(n) ν(2n) f0
νceo νν(n)
(n) 2ν(n)
2ν(n)
Figure 8-4. Representation of the output field of a mode-locked laser in the time domain or
frequency domain, including the definition and measurement of the carrier-envelope-offset
frequency f0.
m ⋅ 2π ∆φce
ωm ≡ + . (1)
τ τ
∆φ
ν ( m) = mf r + f r = mf r + f 0 . (2)
2π
Therefore, the frequency of any line of the spectral comb emitted by such
a laser, as shown in the lower part of Figure 8-4, is given by an integer m,
the pulse repetition frequency fr of the laser, and the carrier-envelope offset-
frequency, f0, which accounts for an offset of the entire comb with respect to
the frequency zero. The frequency f0 represents the time derivative of the
optical (carrier) phase ∆φce as measured with respect to the pulse envelope.
The determination of the frequency of the mth comb mode ν(m) therefore
requires precision measurements of fr, f0, and m. The determination of m can
be performed by a wavelength measurement, for example. The measurement
of the pulse repetition rate fr is straightforward to carry out. However, since
fr enters the optical frequency measurement process with the large factor m,
the noise properties of the measurement are demanding.
The measurement of f0 is also a demanding task. Several more or less
complex schemes, depending on the comb span available, have been
proposed for the measurement of f0 [32]. The simplest concept requires an
octave span of the frequency comb that is not directly available from the
laser used. The available span has to be expanded, e.g., by external spectral
broadening. Microstructure air-silica fibers, which allow the tailoring of the
group-velocity dispersion (GVD) properties [26], are highly suited for SPM
of moderate peak power pulses. They provide both lateral and temporal
confinement of the pulses over long interaction lengths. If the octave span is
achieved, then the f0 measurement can be accomplished by second-harmonic
generation (SHG) of the spectral lines in the low-frequency wing of the
comb, as shown in Figure 8-4. This method has been developed at JILA [33]
and MPQ [34].
Whereas the frequencies of the comb lines are shifted by f0 with respect
to the origin, their second harmonics are shifted by 2f0 (Figure 8-4). Thus,
the beat notes between these harmonics and the corresponding high-
frequency-wing lines of the comb shows the desired component at f0 (n=2m)
When all three quantities fr, f0, and m are known, any unknown external
optical frequency νext of a laser emitting within the spectral range of the
comb can, in turn, be determined by measuring the beat-note frequency ∆ν
with a suitable comb line. The absolute frequency νext is then expressed by
νext = ν(m) + ∆ν. Furthermore, it is possible to directly reference the comb
8. OPTICAL FREQUENCY MEASUREMENT USING FREQUENCY 209
MULTIPLICATION AND FREQUENCY COMBS
spacing (fr) and position (f0) to the microwave cesium time standard, thereby
determining the absolute optical frequencies of all of the comb lines.
This recent advance in optical frequency measurement technology has
facilitated the realization of the ultimate goal of a practical optical frequency
synthesizer: it forms a phase-coherent grid linking the entire optical
spectrum to a microwave standard, or vice versa.
PD
SESAM
PZT2 Filter
PZT1 Grating
LBO
MS Fiber
OC
Self-Phase-
10 GHz PM
fr -Servo
f REP Modulation
Etalon
& Meas.
PD
Counter PLL
PLL
f/U
10 GHz ν CEO
f0 -Measurement
× 96 ×3 Counter
DRO
Figure 8-5. Femtosecond laser as frequency comb generator and generated comb spectrum.
LBO: Lithium β-Borate crystal and SESAM: Semiconductor Saturable Absorber Mirror.
The optical standards that are used for absolute optical frequency
measurement at PTB are located in two different buildings, and the comb
generator in a third building. The standards’ radiation is brought to the comb
generator by ~250 m long single-mode fibers and the rf standard signal by
air-dielectric cables. Measurements have shown that the changes in the cable
and fiber lengths due to temperature changes, which produce temporal phase
changes (corresponding to frequency shifts), remain below ∆ν/ν = 10-15. At
the present level of precision for absolute frequency measurements, these
changes can be neglected.
300
νCa - 455 986 240 494 000 Hz
Hz
200
100
1995 1997 Sep '00 Jun '01 Oct '01 Oct '03
40
Hz
30
ν - 688 358 979 309 312 Hz
20
10
0
-10
-20
-30
-40
12 / 00 01 / 01 11 / 03
Figure 8-7. Frequency measurements of laser-cooled Yb+ optical frequency standards over
the past three years. The first three data points represent published data from Reference [41].
standard [42] and the 171Yb+ frequency standard. The dominant contribution
to the total uncertainty of the frequency value is the statistical uncertainty of
the frequency measurement of 5 Hz because of the limited time of
observation. The most significant systematic contribution to the uncertainty
of 3 Hz is due to a quadrupole shift caused by unknown electric field
gradients in the trap. This quadrupole Stark effect has recently been
measured [43]. The reproducibility of frequency comparisons of two traps
over an extended time, which included reloading traps, seems to indicate that
this quadrupole Stark shift would, with the traps used presently, be below 1
Hz.
The weighted mean of the frequency of electric-quadrupole clock
transition of the 171Yb+ ion is
f3 stabilized
b2,3 = mf2 - f3 m
*m 1
f2 transfer osc.
f2 - f3/m
b1,2 = nf1 - f2
*n
Figure 8-8. Using an intermediate oscillator in the transfer mode. The signal fc is independent
from f2.
m1 xxffr
m ffLO
1 rep LO
Local Oscilator
Local Oscillator
Harmonic
Harmonic PLL
PLL m33
xxm
A ν = (m f – f
1 r LO ) m3
νA= (m 1frep- fLO)m3
ννCC==νAνA++ννBB
νfCEO
0
νB = (νBCEO
ν = (f + ∆ν ) / m2
f0 +0 ∆ν) /m2
÷m
÷ m22
ν
∆∆ν
νν == m22ννCC ++ m2m3ffLO
LO
= m1 ·m2 ·m3 fr + f0 + ∆ν
with the larger instability. Thus, frequency ratios for oscillators with better
stability can be determined with smaller uncertainty if a technique is
available to realize the frequency ratio without introducing additional noise.
Such a technique is the transfer oscillator concept, described above, which
has been generalized for signals with rational frequency ratios [29, 46].
∆1
ννCC==νν –ν
A A- νB B
νν
A A== ff00 ++∆∆11
νCEO ==νν – ν x m1/m
1 1 ν2 ×2 m1 /m
- 2 3
ν2ν2
f0 νν11
∆2 ∆
∆
∆1 ∆2
2
νB = (ννCEO
B0 = ( f02 + ∆2 )1 x m
f + ∆ ) × m /m2 /m 1
1 2
Figure 8-10. Modified transfer concept for the measurement of optical frequency ratios. νC
represents the beat note between arbitrary optical frequencies ν1 and ν2.
If, as with the optical comb generator, two frequencies are related by the
comb through a “rational” harmonic ν2/ν1 = m2/m1, then the two beats have to
be treated accordingly. One of the beat frequencies has to be multiplied by
m2/m1, and the result has to be subtracted from the other beat frequency.
Figure 8-10 shows the scheme for the measurement of the ratio of the
frequencies of two optical standards whose frequencies are given by
Three signals are detected: f0 and two beat signals of two external
frequency standards with the modes of the femtosecond frequency comb, ∆1
and ∆2. The sum frequency of ∆1 and f0 is generated by a mixer. The sum
frequency of ∆2 and f0 is additionally processed with a direct-digital
synthesizer (DDS). Such a device generates an output signal from an input
signal with a frequency ratio given by a long digital tuning word. It
numerically approximates the ratio of the two integers, m1 and m2, by j/2n,
where j is an integer and n the bit length of the tuning word. The resulting
possible error is negligible for commercially available n = 48 bit devices.
Generating the difference frequency of the signals νA and νB with the
help of a mixer, results in the ratio
218 Chapter 8
ν 2 m2 1 m ν
+ ∆2 + f 0 − 2 (∆1 + f 0 ) = 2 + C .
m
=
ν 1 m1 ν 1 m1 m1 ν1
-13
10
-14
10
-15
10
σy( 2, τ )
-16
10
-17
10
-18
10
0,01 0,1 1 10 100 s 1000
integration time
Figure 8-11. Allan standard deviation of the Yb+ frequency standard (full dots) with respect
to an H-maser. The frequency instability of a typical hydrogen maser reference (solid line) is
given for comparison. The open squares represent the stability of an I2-stabilized–frequency-
doubled Nd:YAG laser as derived from the frequency ratio with respect to Yb+ standard. The
triangles show the Allan standard deviation for the frequency ratio between the second
harmonic and the fundamental of the I2-stabilized Nd:YAG laser.
The triangles in Figure 8-11 show the Allan standard deviation for this
measurement. The instability is more than two orders of magnitude smaller
than for the νYb+/ν0 measurement, for which high-quality oscillators were
already used. The optical frequency ratio can be measured at averaging times
of 100 s to a level of ~10-18 corresponding to mHz fluctuations. The residual
fluctuations at the large averaging times do not indicate a limitation of the
220 Chapter 8
frequency measurement but are the result of the slow optical phase
variations caused by the changes in the optical path lengths between the
optical elements on the table where the system was set up, in temperature, in
the motion of the air, and so forth. These variations could evidently be
reduced further so that 10-18 does not represent a limit in measurement
precision.
In any case, the lower curve shows unambiguously that the transfer
method works well and that a frequency comb used in the transfer mode is
adequate for forming a part of future atomic clocks based on optical
transitions; it would not limit the transfer of optical stability to the rf range,
even at the projected precision and stability of 10-18.
Figure 8-12 shows a Fourier transform of the measurements represented
by the lowest curve of Figure 8-11. The 9 mHz spectral line representing the
frequency measurement uncertainty corresponds to the “virtual measurement
frequency” of 43 THz. Converted to the measured optical frequency of 282
THz it would show a spectral line of 60 mHz width (still remarkably narrow)
at 282 THz as characterizing the measurement, or the coherence of the
measurement process.
Figure 8-12. Power spectrum of the beat note between sub-harmonics of the outputs of an I2
and an Yb+-frequency standard taken at a virtual frequency of 43 THz as calculated from the
recorded phase excursions by fast Fourier transform.
7. SUMMARY
Frequency chains have been developed and operated for many years in
several laboratories. However, during the last few years, great progress has
been achieved in the development of optical frequency standards and optical
8. OPTICAL FREQUENCY MEASUREMENT USING FREQUENCY 221
MULTIPLICATION AND FREQUENCY COMBS
ACKNOWLEDGEMENTS
REFERENCES
[1] C. O. Weiss and A. Godone, IEEE J. Quantum Electron. 20, 97-99 (1984).
[2] D. J. E. Knight, K. I. Pharaoh, and M. Zucco, in 5th Symposium of
Frequency Standards and Metrology, Woods Hole, MA, 1995), p. 465-467.
222 Chapter 8
[46] H. R. Telle, B. Lipphardt, and J. Stenger, Appl. Phys. B 74, 1-6 (2002).
[47] C. Tamm, D. Engelke, and V. Bühner, Phys. Rev. A 61, 053405 (2000).
[48] P. Cordiale, G. Galzerano, and H. Schnatz, Metrologia 37, 177-182 (2000).
[49] R. K. Shelton, S. M. Foreman, L. S. Ma, J. L. Hall, H. C. Kapteyn, M. M.
Murnane, M. Notcutt, and J. Ye, Opt. Lett. 27, 312-314 (2002); R. K.
Shelton, L. S. Ma, H. C. Kapteyn, M. M. Murnane, J. L. Hall, and J. Ye,
Science 293, 1286-1289 (2001).
Chapter 9
FEMTOSECOND LASERS FOR OPTICAL
CLOCKS AND LOW NOISE FREQUENCY
SYNTHESIS
Key words: optical atomic clocks, optical frequency metrology, mode-locked lasers,
frequency comb, optical frequency synthesis
1. INTRODUCTION
Nonlinear Frequency
Mode-locked Conversion and
Solid State Lasers Spectral Broadening
Figure 9-1. The development of optical clocks and optical frequency metrology has arisen out
of many different sub-fields of atomic, molecular, and optical physics. While optical clocks
have their roots in the precision optical-frequency spectroscopy of cooled and trapped ions
and atoms, the emergence of robust mode-locked solid-state femtosecond lasers and nonlinear
spectral broadening/conversion techniques has provided new opportunities and invigorated
the field as a whole.
A tom s νo
Detector
& Laser ∆ν ν
Control
Atom ic Resonance
Frequency
Laser
Counter
O ptical Cavity
Figure 9-2. Schematic diagram of an optical clock. A laser is first stabilized to a Fabry-Perot
optical cavity that provides a means to narrow the laser linewidth leading to good short-term
stability. The center of a narrow resonance in an appropriate atomic sample then provides a
stable reference to which the frequency of the laser can be steered. Once its frequency is
locked to the center of the atomic resonance, a predetermined number of optical cycles are
counted to mark a second of time.
9. FEMTOSECOND LASERS FOR OPTICAL CLOCKS AND LOW 229
NOISE FREQUENCY SYNTHESIS
to provide both laser distance ranging (length metrology) and time keeping
simultaneously for satellites [27].
For the near future, it is already clear that optical clocks and optical
frequency metrology will provide interesting and new scientific avenues to
study our universe—pushing the limits on tests of the most fundamental
physical laws to new levels. Of particular interest is the continued
application of optical frequency standards in spectroscopy and the improved
determination of the fine structure constant α and the Rydberg constant R∞
[14, 28]. As measurement accuracy improves, metrologists may find
themselves in the unique position of being able to observe physical
“constants” evolve in time [29]. Indeed, laboratory tests on the possible
divergence of clocks based on different atomic transitions already provide
some of the most stringent constraints of the variation of α [30, 31].
Experiments of fundamental importance for which precision
clocks/oscillators are of value include searches for variations in the isotropy
of space, a preferred reference frame, and Lorentz and Charge-Parity-Time
(CPT) symmetry violation [32]. Following the recent trapping of cold
antihydrogen at CERN [33], in the coming years it may be possible to
compare optical clocks based on the 1S–2S transitions in both hydrogen and
anti-hydrogen [34]. Such measurements would provide precisions tests of
the fundamental symmetry between matter and antimatter [35].
∆ν rms 1 T
σ y (τ ) ≈ ≈ , (1)
ν0 τ
π ⋅ Q τN
-13
10
Allan Deviation — Instability — σy (τ)
Cs Fountain Clock
Iodine
-14 (predict) Iodine
10
-15
10
Optical Calcium
-16
10 Cavity
Ca, Sr Hg+ (predict)
-17
(predict)
10 1 day 1 month
-2 0 2 4 6
10 10 10 10 10
Averaging Time (s)
observed. All of these factors are critically important for the achievement of
the highest accuracy.
The technical challenges of making an optical frequency standard based
on a single ion are formidable but single-ion standards have now been
achieved in a handful of laboratories around the world. At NIST, an optical-
frequency standard based on a single trapped 199Hg+ ion is being developed.
The performance of this standard is immediately competitive with the
performance of the best microwave standards and has the potential to surpass
those standards in terms of stability, frequency reproducibility, and accuracy.
For this standard, a single 199Hg+ ion is trapped in a small rf Paul trap (≈1
mm internal dimensions) and laser cooled to a few milliKelvins using 194
nm radiation. A highly stabilized dye laser at 563 nm with a linewidth of
less than 1 Hz is frequency doubled to 282 nm (1,064 THz) to probe the
clock transition [41]. Measured linewidths as narrow as 6.7 Hz on the 282
nm transition have been reported [42]. For an averaging time τ in seconds,
the projected instability of an optical frequency standard using a single Hg+
ion is 1 x 10-15τ−1/2, and fractional frequency uncertainties approaching
1x10-18 seem feasible.
One apparent difficulty of modern optical standards is the requirement
that suitable transitions for laser cooling, fluorescence detection, and the
clock itself be present in the same atom or ion. While some atoms have
good clock transitions, their cooling and detection transitions might be less
than ideal. Wineland et al. have proposed an efficient solution to this
problem in the case of ion-based clocks that involves simultaneously
trapping a “clock” ion and a “logic” ion [43]. Using techniques developed
for quantum computation applications [44], the logic ion would provide the
functions of sympathetic cooling and detection, leaving more flexibility in
choosing the best clock ion with a narrow, unperturbed and accessible
transition. This approach is currently be implemented at NIST using 27Al+
as the clock ion and 9Be+ as the logic ion.
Some neutral atoms also have narrow optical transitions that are
relatively insensitive to external perturbations and are thus attractive as
optical frequency standards. Neutral atoms have some advantages and
disadvantages relative to ions. Using the well-established techniques of laser
cooling and trapping, they are fairly easy to confine and cool to low
temperatures. However, in contrast to ions, the trapping methods for neutrals
perturb the atomic energy levels, which is unacceptable for use in a
frequency standard. To avoid the broadening and shifts associated with the
trap, neutral atoms are released from the trap before the clock transition is
9. FEMTOSECOND LASERS FOR OPTICAL CLOCKS AND LOW 235
NOISE FREQUENCY SYNTHESIS
probed. The atoms fall from the trap under the influence of gravity and
expand with low thermal velocities (typically a few cm/s). The resulting
atomic motion brings with it limitations in accuracy (and even stability) that
are associated with velocity-dependent frequency shifts. Two of the more
troublesome effects are the limited observation time and the incomplete
cancellation of the first-order Doppler shift associated with wave-front
curvature and wave-vector mismatch. Reduced observation times limit the
line Q, the stability, and the accuracy. However, neutral atoms do have at
least one significant advantage: large numbers of atoms can be used,
producing a large signal-to-noise ratio (SNR) in a short time and the potential
for exceptional short-term stability.
The atomic Ca optical frequency standard [45, 46] is one of the most
promising and extensively studied cases. It has a 400 Hz wide clock
transition at 657 nm (1S0 – 3P1) that is reasonably immune from external
perturbations. It is readily laser cooled and trapped, and it is experimentally
convenient because the relevant transitions are accessible with tunable diode
lasers. Cooling and trapping of about 107 Ca atoms to mK temperatures can
be accomplished in a magneto-optic trap (MOT) with frequency doubled
diode laser tuned to the 423 nm 1S0 – 1P1 transition. With the cooling
radiation turned off, an injection-locked and stabilized diode laser at 657 nm
(456 THz) then probes the clock transition with the separated excitation
method of optical Ramsey-Bordé spectroscopy. Optical fringes with high
signal-to-noise ratio are observed using shelving detection on the cooling
transition. With this technique, the present Ca standard can provide short-
term fractional frequency instability of about 4 x 10-15 in 1 s of averaging.
Second stage cooling on the narrow 1S0 – 3P1 to temperatures ~10 µK has
been achieved with the aid of 552 nm light that quenches (i.e., depopulates)
the long-lived 3P1 state [47]. The 10-µK temperatures reduce velocity-
related systematic shifts, and it appears that uncertainties at or below 10-15
will be attainable with such a system.
Ultimate neutral atom-based systems with high accuracy will demand a
stringent separation between the external degrees of freedom (controlled by
the trapping potential) and internal level structure (clock transitions), similar
to that obtained with single trapped ions. A crucial step towards high
reproducibility and accuracy is to confine neutral atoms in the Lamb-Dicke
regime, while at the same time limiting the effect of the confining potential
to only the external degrees of freedom [48]. The fermionic isotope of 87Sr
has a nonzero nuclear magnetic moment (I = 9/2) that gives rise to magnetic
substructure in both the ground and excited states [49]. Moreover, 87Sr
possesses a doubly forbidden J = 0 to J = 0 clock transition, 1S0 (F = 9/2) –
3
P0 (F' = 9/2), which has a ~1mHz linewidth (corresponding to an intrinsic
236 Chapter 9
wavelength is decreased from 532 nm to near the dissociation limit [56]. The
high SNR results indicate that I2 transitions in the wavelength range 501–
532 nm hold great promise for future development of optical frequency
standards, especially with the advent of all-solid-state Yb:YAG lasers. One
exciting candidate is the 514.67 nm standard that has a projected instability
<10-14 at 1 s [57].
Table 9-1. Some of the promising optical frequency references for emerging and future
optical clocks. Presently achieved values of instability and uncertainty are listed, while the
projected values are given in parentheses. The research institute is given in the last column.
Frequency 1-s Fractional Fractional
Institute
(THz) Instability Uncertainty
Ions
Hg+ 1064 3µ10-15 1µ10-14 NIST
(1µ10-15) (<1µ10-17)
In+ 1267 2µ10-13 MPQ
Sr+ 445 5µ10-15 NRC, NPL
Al+ 1124 NIST
Yb+ 688 9µ10-15 PTB
(<1µ10-17)
Yb+ 642 2µ10-12 NPL
Neutral
Atoms
H 2466 2µ10-14 MPQ
Ca 456 4µ10 -15
2µ10-14 PTB, NIST
(<2µ10-16) (<1µ10-15)
Sr U. Tokyo,
JILA, SYRTE,
LENS
Yb KRISS, NIST
Molecules
OsO4 29 3µ10-13 Paris Nord
CH4 88 Lebedev, NPL,
PTB,
Novosibirsk
I2 563 4µ10-14 9µ10-12 JILA, PTB,
NMIJ, BIPM
238 Chapter 9
νn = nfr + f0 . (2)
The first experiments [13, 15, 16] verifying the valuable contribution of
femtosecond lasers to optical frequency metrology all employed
femtosecond lasers that had been developed over the previous decade for
experiments in ultrafast science [66]. While such lasers worked very well
for the initial proof-of-principle experiments, and continue to be used in
many cases, they lacked other desirable qualities for frequency metrology.
The ultimate femtosecond laser for optical frequency metrology and optical
clocks is likely to be a continuously evolving device, but nonetheless, in this
section we attempt to lay out some of the desirable characteristics of these
devices as they relate to their use in frequency metrology and optical clocks.
1
10 (c)
-1
(a) 10
-3
P rism P air 10
Power per mode (microWatt)
Pum p Ti:S 10
-5
1
10 (d )
-1
10
-3
10
-5
(b) 10
Pump 1
600 800 10 00 120 0
Ti:S 10 (e )
-1
10
-3
10
-5
10
600 800 10 00 120 0
W avelength (nm )
Figure 9-4. Common Ti:sapphire femtosecond lasers used for frequency metrology and their
output spectra. (a) Conventional linear cavity design employing a combination of chirped
mirrors and prisms produces the broadband spectrum shown in (c). The ring laser shown in
(b) uses only chirped mirrors and can produce the spectrum show in (d). The spectrum shown
in (e) is produced using either a laser (a) or (b) in combination with nonlinear microstructure
optical fiber.
Repetition rate: There are a few factors that drive the consideration of
the best repetition rate. The most obvious feature is that for a constant
9. FEMTOSECOND LASERS FOR OPTICAL CLOCKS AND LOW 241
NOISE FREQUENCY SYNTHESIS
average output power from the laser, the mode comb of a high repetition rate
laser will have a proportionately higher power per mode. This is a clear
advantage for frequency domain applications where we rely on a good SNR
in the heterodyne beats between a cw laser and an individual tooth of the
comb itself. However, one cannot take this point to the extreme. For
example, if we again assume a constant output power, as the repetition rate
increases, the power per pulse decreases. This is an important consideration
if one is to rely on nonlinear processes to broaden the optical spectrum to an
octave. As a rule of thumb, with the typical microstructure fibers used with
Ti:sapphire lasers around 800 nm, one needs roughly a few hundred
picojoules of energy in a ~50 fs pulse to obtain an octave of spectral
broadening. This can be achieved with a 1 GHz repetition rate laser, but
given the same average power, it is not likely to be possible if the laser could
operate with a 10 GHz repetition rate.
Photodetectors & electronics: Since many uses of the femtosecond
laser comb ultimately involve a connection to the rf/microwave domain, the
photodetectors that provide this connection as well as the subsequent
electronics that are used must also be a consideration. While photodetectors
and microwave electronics with bandwidths up to ~50 GHz are
commercially available, devices that operate much above 10 GHz tend to be
more costly, more difficult to use, and less sensitive. This last factor can
actually result in a loss of SNR that one had hoped to achieve precisely by
choosing a higher repetition rate in the first place. This can be particularly
true with the photodetector that is used to detect the optical clock output at fr,
where extremely high SNR is required to generate stable electrical signals
[68]. Another important electronics-related consideration surrounding the
choice of repetition rate is the frequency one might ultimately choose for
connecting the fr output of the femtosecond laser to existing rf/microwave
sources and standards. The best choice of low-noise electronic synthesizers
that would divide or multiply fr up or down to common frequencies of 5, 100
or 9,192.631770 MHz will certainly be a consideration for the most
demanding applications.
Size, simplicity & robustness: While size, simplicity, and robustness
are not necessarily synonymous, it is often true that a smaller device of
simple construction can be more robust. For example, small size lends itself
to improved temperature stability, which directly impacts the operation of
any femtosecond laser and the required dynamic range for the various servo
controls. This is a point in favor of higher repetition rate systems, or at least
systems with smaller footprints. When it comes down to the actual design of
the laser, this factor would also tend to favor the use of chirped mirrors or
fully integrated optics (such as a fiber laser) over the use of prisms for
242 Chapter 9
optical
µ-wave out
reference Optical Synthesizer
1 3
I(f)
f0
2 fr 4
µ-wave
0 νn = nfr + f0
f optical out
reference
of the output is given by the repetition rate fr or one of its harmonics [46, 60,
61]. The synthesizer can also function to effectively synthesize new optical
frequencies that are displaced from the input optical reference by a freely
Ø
chosen offset plus-or-minus harmonics of fr (ports 1 4) [80].
(a) f
f0 = fb2 – 2fb1
fb1 fb2
(b) f
νs 2νs
Phase Ref.
PZT
Detector fr or n fr
(c) f
Phase Ref.
PZT
Detector
fb = ν N - ν s
(d) f
νs
Figure 9-6. Techniques for measuring and controlling f0 (a, b) and fr (c,d). See text for
details.
control have so far been used. The first employs modulation of intracavity
laser power [79, 86], and the second (which is applicable in linear cavities
employing prism pairs) involves the tipping of one of the cavity mirrors [18,
87]. While the merits and physical mechanisms of these differing techniques
remain a topic of discussion, both have been successfully implemented to
reduce the residual fluctuations of f0 to the millihertz level relative to a stable
reference. When f0 is phase-locked to D.C. or a harmonic of fr, precise
stabilization has the consequence of maintaining the relative carrier-
envelope phase at a fixed value for times approaching one hour [88]. This
has enabled the subsequent amplification and use of such lasers for
experiments in phase-sensitive extreme nonlinear optics experiments (see,
for example, Reference [64] and Chapter 10 of this volume).
Control of fr: As already mentioned, fr can be detected with a sufficiently
fast photodetector. The rf spectrum of the resulting photocurrent consists of
a series of harmonics of fr, one of which can be filtered and phase-locked to
a low-noise rf source (Figure 9-6(c)). In this case, it is common to use a
piezo-mounted cavity mirror as an actuator for changing fr. A more careful
consideration of Equation (2) highlights some of the difficulties and
disadvantages of controlling fr in the rf/microwave domain as just described.
Because the optical frequency νn scales as the multiplicative mode number n,
the spectral density of phase noise on the rf reference source to which fr is
phase-locked is multiplied by a factor of n2. For a mode in the optical
domain with n º500,000 (fr = 1 GHz), this implies that the phase noise of the
rf reference is increased by 115 dB when it is effectively multiplied up to the
optical domain using the femtosecond laser comb. A high-quality quartz-
based rf reference at 1 GHz might have a typical noise floor of –110 dBc/Hz,
but when it serves as the reference for fr, there will be nothing remaining of
the phase-coherent carrier in the optical mode νn. This is similar to the well-
known problem of “carrier collapse,” which places extreme demands on the
rf reference (as well as all intermediate electronic components) for fr.
There are, however, a few ways to minimize this phase-noise
multiplication problem. First of all, one can rely on the relatively good
short-term stability of the femtosecond laser itself. On time scales less than
~1ms, the noise of a typical Ti:sapphire laser has been measured to fall
below that of most high quality microwave sources [20, 68]. This means
that if one controls fr relative to a microwave reference, a control bandwidth
of §1 kHz is all that is required to remove the low-frequency thermal and
acoustic noise of the mode-locked laser. Use of higher bandwidth will
simply transfer the noise of the microwave reference to fr, which will
subsequently be multiplied up to the optical comb elements. Although the
observed optical linewidth may still be on the order of 0.1–1 MHz, one
generally finds that the fractional stability of the optical comb elements can
9. FEMTOSECOND LASERS FOR OPTICAL CLOCKS AND LOW 247
NOISE FREQUENCY SYNTHESIS
-80
High Quality Quartz
-100
L(f) [dBc/Hz]
-140 Ca Optical
Sapphire Resonator
-160
-180
Hg+ Optical
Cavity
-200
0 1 2 3 4 5
10 10 10 10 10 10
Frequency [Hz]
Figure 9-7. Phase noise of various oscillators and synthesizers at 1 GHz. The quartz curve
corresponds to a high-quality 5 MHz oscillator and assumes noise-free multiplication to 1
GHz. The commercial synthesizer is residual synthesizer noise that would be added to the
reference oscillator (typically quartz). The sapphire resonator curve represents a state-of-the-
art commercial product with its frequency divided down from 10 GHz to 1 GHz. The dotted
Ca curve is a projection of what should be achievable, and the Hg+ optical cavity is estimated
from experimental data [41].
test was repeated with four systems, two of which employed octave-
spanning spectra generated in nonlinear microstructure fiber and two which
generated broad spectra directly from the laser [100]. With improved short-
term instability and longer averaging times, a fractional frequency
uncertainty limit of near 1 × 10-19 was achieved. Using the “transfer
oscillator” technique already discussed above [94], Stenger et al. tested a
femtosecond laser comb by measuring the ratio of the frequency of a cw
laser to its second harmonic with an uncertainty of 7 × 10-19.
In the few years since the introduction of femtosecond lasers into field of
frequency metrology, the palette of available sources has already begun to
broaden beyond that of Ti:sapphire lasers (see Figure 9-7). The reliability of
the mature Ti:sapphire laser has made it the natural place to begin this
exciting field, and it is likely that femtosecond laser-based synthesizers
employing Ti:sapphire will continue to be used in many applications.
However, the present size and cost of the pump laser for Ti:sapphire-based
systems has motivated the search for alternative systems. In the future, it
would not be surprising to find that femtosecond laser-based synthesizers
and optical clocks will be widely used in science and technology and will be
commercially available in compact packages similar to today’s microwave
synthesizers and clocks. The availability of robust, low-priced femtosecond-
laser synthesizers will be particularly important for applications in air- or
space-borne platforms or widespread applications (such as communications
systems) for which cost and rugged packaging are of greatest importance.
To date, some of the promising femtosecond lasers that have produced
octave-spanning spectra include diode-pumped Cr:LiSAF [82], a fiber-laser-
pumped Cr:Forsterite [101], and an Er-doped fiber laser [83, 85, 102]. Each
of these has some advantages and disadvantages relative to Ti:sapphire. For
example, both Cr:LiSAF and Cr:Forsterite have more convenient and
compact pumping schemes either directly with diode lasers or with a Yb-
doped fiber laser. One of the trade-offs here is that these laser hosts are not
as broadband and efficient as Ti:sapphire and tend to have worse thermal
properties. An Er-fiber laser-based comb generator has a number of
advantages over Ti:Sapphire. It can be much more compact, robust, lighter,
and power-efficient than a bulk optic solid-state laser system, and would
require less alignment. Additionally, it can be easily integrated into a
telecommunication system in the important 1300–1600 nm regime.
However, at this point, the Er-fiber-based systems only operate at repetition
252 Chapter 9
rates of 50–100 MHz and can have excess noise that is not fully understood
[85].
g h
2004
f i
2002 e
c d
2000
Year
b
1998
1996 a
1994
Figure 9-7. Spectral extent of some sources used as combs in frequency metrology. (a)
Electro-optic modulator-based comb developed by Kourogi et al. [103]; (b) Ti:sapphire
femtosecond laser comb [13, 14]; (c) Octave-spanning femtosecond laser comb generated
using microstructure fiber [15, 16]; (d) Broadband spectrum generated directly from a
Ti:sapphire laser [69]; (e) Cr:LiSAF femtosecond laser plus microstructure fiber [82]; (f)
Octave-spanning comb generated with femtosecond Cr:forsterite and nonlinear optical fiber
[104]; (g) Octave-spanning comb generated with Er-fiber laser and nonlinear optical fiber
[83]; (h) Offset-frequency-free comb near 3400 nm generated via difference-frequency
generation [93]; (i) Tunable frequency comb generated via difference-frequency generation
between synchronized Ti:sapphire lasers [105].
With the advent of optical atomic clocks and the associated superior
short-term frequency stability, transfer of signals linked to such
clock/frequency standards over an appreciable distance with a minimal loss
of stability has become an active subject for research [107, 108]. The
stability of the best microwave and optical frequency standards, for times
less than ~ one week, can now exceed the capabilities of the traditional
transmission systems (i.e., GPS, Two-Way Time transfer [109]) used to
distribute these time/frequency reference signals to remote sites. The
instabilities in transmission channels can contribute a significant fraction of
the overall uncertainty in the comparison of high-performance standards that
are not co-located. While improvement in the transfer process over large-
scale signal paths remains challenging, researchers have started
experimenting with optical fibers as effective means for local networks of
dissemination or comparison of time and frequency standards, both in the
microwave and optical domains. The attractiveness of this approach lies in
the fact that an environmentally isolated fiber can be considerably more
stable than open-air paths, especially at short time scales. Furthermore, the
same advantages enjoyed by communication systems in optical fiber (e.g.,
low loss, scalability, etc.) can be realized in a time/frequency distribution
system. Active stabilization of fiber-optic channels for the distribution of
reference frequencies can also be employed to improve the stability of the
transmitted standard. Besides the obvious benefit of more precise
time/frequency transfer, an actively stabilized optical fiber network can play
a critical role in the implementation of long-baseline coherent interferometry
or ultralow timing jitter in particle-accelerator-based novel light sources.
An rf signal can be transferred in an optical fiber network by amplitude
modulating a suitable optical carrier (for example at 1.55 µm) used for
transmission and then recovering the modulation frequency at the remote
end as the transferred reference signal. However, the instability of this rf
modulation-based frequency transfer protocol seems to rise to a few parts in
1013 at 1-s over a 6.9 km long fiber linking JILA to NIST. Jet Propulsion
Laboratory colleagues have demonstrated the rf transfer instability of parts
in 1014 at 1 s for a 16 km-long fiber under active noise cancellation control
[110]. In comparison, direct transfer of a cw laser-based optical frequency
standard through the same 6.9 km fiber suffered an instability of a few parts
in 1014 at 1 s. This instability is further reduced to 3 × 10-15 at 1 s after active
optical-phase-noise cancellation is implemented [107]. A femtosecond laser
located at the remote end can be phase locked to this incoming cw laser
254 Chapter 9
ACKNOWLEDGEMENTS
The work presented in this chapter would not have been possible without
the valuable contributions of many people. In particular, we would like to
thank Albrecht Bartels, Jim Bergquist, Martin Boyd, Sebastien Bize, Lisheng
Chen, Steven Cundiff, Anne Curtis, Seth Foreman, Rich Fox, John Hall,
Kevin Holman, Darren Hudson, Tetsuya Ido, Erich Ippen, Euguene Ivanov,
David Jones, Jason Jones, Franz Kärtner, Thomas Loftus, Andrew Ludlow,
Long-Sheng Ma, Adela Marian, Kevin Moll, Oliver Mücke, Nathan
Newbury, Chris Oates, Jin-Long Peng, Tanya Ramond, Matthew Stowe,
Thomas Udem, Kurt Vogel, Guido Wilpers, Robert Windeler, and Dave
Wineland.
256 Chapter 9
REFERENCES
[1] H. Dehmelt, Rev. Mod. Phys. 62, 525-530 (1990).
[2] H. G. Dehmelt, IEEE Trans. Instrum. Meas. 31, 83-87 (1982).
[3] J. L. Hall, Science 202, 147-156 (1978); J. L. Hall, M. Zhu, and P. Buch, J.
Opt. Soc. Am. B 6, 2194-2205 (1989).
[4] T. W. Hänsch, I. S. Shahin, and A. L. Schawlow, Nature-Physical Science
235, 63 (1972).
[5] S. N. Bagayev, Y. D. Kolomnikov, V. N. Lisitsyn, and V. P. Chebotayev,
IEEE J. Quant. Electr. 4, 868 (1968).
[6] V. P. Chebotayev, V. G. Goldort, V. M. Klementyev, M. V. Nikitin, B. A.
Timchenko, and V. F. Zakharyash, Appl. Phys. B 29, 63-65 (1982).
[7] D. J. Wineland, Science 226, 395-400 (1984); D. J. Wineland, J. C.
Bergquist, W. M. Itano, F. Diedrich, and C. S. Weimer, in The Hydrogen
Atom, edited by G. F. Bassani, M. Inguscio and T. W. Hänsch (Springer-
Verlag, Berlin, 1989), p. 123-133.
[8] A. Clairon, A. Vanlerberghe, C. Salomon, M. Ouhayoun, and C. J. Borde,
Opt. Commun. 35, 368-372 (1980).
[9] C. O. Weiss, G. Kramer, B. Lipphardt, and E. Garcia, IEEE J. Quantum
Electron. 24, 1970-1972 (1988).
[10] H. Schnatz, B. Lipphardt, J. Helmcke, F. Riehle, and G. Zinner, Phys. Rev.
Lett. 76, 18-21 (1996).
[11] J. E. Bernard, A. A. Madej, L. Marmet, B. G. Whitford, K. J. Siemsen, and
S. Cundy, Phys. Rev. Lett. 82, 3228-3231 (1999).
[12] H. R. Telle, D. Meschede, and T. W. Hänsch, Opt. Lett. 15, 532-534
(1990).
[13] T. Udem, J. Reichert, R. Holzwarth, and T. W. Hänsch, Opt. Lett. 24, 881-
883 (1999).
[14] T. Udem, J. Reichert, R. Holzwarth, and T. W. Hänsch, Phys. Rev. Lett. 82,
3568-3571 (1999).
[15] S. A. Diddams, D. J. Jones, J. Ye, S. T. Cundiff, J. L. Hall, J. K. Ranka, R.
S. Windeler, R. Holzwarth, T. Udem, and T. W. Hänsch, Phys. Rev. Lett.
84, 5102-5105 (2000).
[16] D. J. Jones, S. A. Diddams, J. K. Ranka, A. Stentz, R. S. Windeler, J. L.
Hall, and S. T. Cundiff, Science 288, 635-639 (2000).
[17] T. Udem, R. Holzwarth, and T. W. Hänsch, Nature 416, 233-237 (2002).
[18] S. T. Cundiff, J. Ye, and J. L. Hall, Rev. Sci. Instrum. 72, 3746-3771
(2001).
[19] J. L. Hall, J. Ye, S. A. Diddams, L.-S. Ma, S. T. Cundiff, and D. J. Jones,
IEEE J. Quantum Electron. 37, 1482-1492 (2001).
[20] L. Hollberg, C. W. Oates, E. A. Curtis, E. N. Ivanov, S. A. Diddams, T.
Udem, H. G. Robinson, J. C. Bergquist, R. J. Rafac, W. M. Itano, R. E.
Drullinger, and D. J. Wineland, IEEE J. Quantum Electron. 37, 1502-1513
(2001).
[21] S. T. Cundiff and J. Ye, Rev. Mod. Phys. 75, 325-342 (2003).
[22] J. Ye, H. Schnatz, and L. W. Hollberg, IEEE J. Sel. Top. Quantum
Electron. 9, 1041-1058 (2003).
9. FEMTOSECOND LASERS FOR OPTICAL CLOCKS AND LOW 257
NOISE FREQUENCY SYNTHESIS
[35] R. Bluhm, V. A. Kostelecky, and N. Russell, Phys. Rev. Lett. 82, 2254-
2257 (1999).
[36] A. Javan, Annals of the New York Academy of Sciences 168, 715 (1970).
[37] P. Gill, Proc. 6th Symp. on Freq. Standards and Metrology, World
Scientific, Singapore (2002).
[38] D. W. Allan, Proc. IEEE 54, 221 (1966).
[39] W. M. Itano, J. C. Bergquist, J. J. Bollinger, J. M. Gilligan, D. J. Heinzen,
F. L. Moore, M. G. Raizen, and D. J. Wineland, Phys. Rev. A 47, 3554-
3570 (1993); C. Santarelli, P. Laurent, P. Lemonde, A. Clairon, A. G.
Mann, S. Chang, A. N. Luiten, and C. Salomon, Phys. Rev. Lett. 82, 4619-
4622 (1999).
[40] J. J. Bollinger, J. D. Prestage, W. M. Itano, and D. J. Wineland, Phys. Rev.
Lett. 54, 1000-1003 (1985); A. A. Madej and J. E. Bernard, Frequency
Measurement and Control: Advanced Techniques and Future Trends 79,
153-194 (2000).
[41] B. C. Young, F. C. Cruz, W. M. Itano, and J. C. Bergquist, Phys. Rev. Lett.
82, 3799-3802 (1999).
[42] R. J. Rafac, B. C. Young, J. A. Beall, W. M. Itano, D. J. Wineland, and J.
C. Bergquist, Phys. Rev. Lett. 85, 2462-2465 (2000).
[43] D. J. Wineland, M. Barrett, J. Britton, J. Chiaverini, B. DeMarco, W. M.
Itano, B. Jelenkovic, C. Langer, D. Leibfried, V. Meyer, T. Rosenband, and
T. Schatz, Philos. Trans. R. Soc. Lond. Ser. A 361, 1349-1361 (2003).
[44] M. D. Barrett, B. DeMarco, T. Schaetz, V. Meyer, D. Leibfried, J. Britton,
J. Chiaverini, W. M. Itano, B. Jelenkovic, J. D. Jost, C. Langer, T.
Rosenband, and D. J. Wineland, Phys. Rev. A 68 (2003).
[45] C. W. Oates, E. A. Curtis, and L. Hollberg, Opt. Lett. 25, 1603-1605
(2000); T. Udem, S. A. Diddams, K. R. Vogel, C. W. Oates, E. A. Curtis,
W. D. Lee, W. M. Itano, R. E. Drullinger, J. C. Bergquist, and L. Hollberg,
Phys. Rev. Lett. 86, 4996-4999 (2001).
[46] G. Wilpers, T. Binnewies, C. Degenhardt, U. Sterr, J. Helmcke, and F.
Riehle, Phys. Rev. Lett. 89 (2002).
[47] E. A. Curtis, C. W. Oates, and L. Hollberg, J. Opt. Soc. Am. B 20, 977-984
(2003).
[48] M. Takamoto and H. Katori, Phys. Rev. Lett. 91 (2003).
[49] X. Y. Xu, T. H. Loftus, J. W. Dunn, C. H. Greene, J. L. Hall, A. Gallagher,
and J. Ye, Phys. Rev. Lett. 90 (2003).
[50] H. Katori, M. Takamoto, V. G. Pal'chikov, and V. D. Ovsiannikov, Phys.
Rev. Lett. 91 (2003).
[51] T. Mukaiyama, H. Katori, T. Ido, Y. Li, and M. Kuwata-Gonokami, Phys.
Rev. Lett. 90 (2003).
[52] J. L. Hall, C. J. Borde, and K. Uehara, Phys. Rev. Lett. 37, 1339-1342
(1976); V. P. Chebotayev, Radio Science 14, 573-584 (1979); A. Clairon,
B. Dahmani, A. Filimon, and J. Rutman, IEEE Trans. Instrum. Meas. 34,
265-268 (1985).
[53] M. A. Gubin, D. A. Tyurikov, A. S. Shelkovnikov, E. V. Kovalchuk, G.
Kramer, and B. Lipphardt, IEEE J. Quantum Electron. 31, 2177-2182
(1995); J. von Zanthier, J. Abel, T. Becker, M. Fries, E. Peik, H. Walther,
R. Holzwarth, J. Reichert, T. Udem, T. W. Hänsch, A. Y. Nevsky, M. N.
Skvortsov, and S. N. Bagayev, Opt. Commun. 166, 57-63 (1999); B. Frech,
L. F. Constantin, A. Amy-Klein, O. Phavorin, C. Daussy, C. Chardonnet,
9. FEMTOSECOND LASERS FOR OPTICAL CLOCKS AND LOW 259
NOISE FREQUENCY SYNTHESIS
Key words: carrier-envelope phase, ultrashort pulses, Gouy phase shift, high harmonic
generation, above threshold ionization, phase-stabilized amplifiers
264 Chapter 10
1. INTRODUCTION
0 0 0
Figure 10-1. Numerical simulation of laser-driven soft x-ray emission from noble gas atoms.
(a) Electric field of a 20 fs laser pulse. (b) Electric field of a 5 fs laser pulse optimized for
generation of a single subfemtosecond x-ray pulse. (c) Electric field of a 5 fs pulse producing
the two highest energy x-ray pulses. (d) – (f) Time-domain structures of x-ray radiation
emitted in a 10 eV bandwidth at half the cutoff energy. (g) – (i) Time-domain structures of x-
ray radiation emitted in a 10 eV bandwidth at the cutoff energy. The lower x-ray peak yield in
(d) and (g) in comparison to (e), (f), (h), and (i) is the consequence of higher ionization by
more numerous field peaks of the 20 fs pulse. In total, gas concentration loss due to ionization
was 6.7% and 1.7% for the 20 and 5 fs pulses, respectively. Horizontal bars in (a)–(c) show
the peak intensity threshold required to yield cutoff-frequency x-ray radiation.
E L (t ) = AL (t ) cos[ω L (t )t + φ ] . (1)
ω n = nω r + ω CE , (2)
where n is the mode number, ω r is the repetition rate of the laser, and ω CE is
a frequency shift from an exact integer multiple of ω r . Equation (2)
describes the comb of frequencies across the mode-locked bandwidth that is
of interest in frequency-domain metrology [17, 42]. The physical origin of
the frequency mismatch ω CE lies in the difference between the group delay
(the cavity round-trip time of the envelope) and the phase delay (the round
trip of the carrier wave). Consequently, the carrier-envelope phase (i.e., the
position of the carrier-wave oscillation with respect to the envelope that is
advancing at a different velocity) is shifted with each successive laser pulse
by
ω CE
∆φ = 2π . (3)
ωr
value of φ. Reduction of the repetition rate (e.g., by selecting every mth pulse
with a pulse picker in front of a chirped-pulse-amplifier) leads to an m-fold
increase of the mode density but does not scale the frequency mode
mismatch, which remains mod ω r m {ω CE } . In fact, the generation of new
frequency modes is the result of applying a modulation that is nonlinear in
the time domain (the fast switching of the pulse picker). This nonlinearity
does not destroy the properties of the frequency comb as all the modes
remain rigorously locked to each other. The same reasoning applies to the
case of a single-isolated laser pulse ( ωr = 0 ) that corresponds to a
continuously filled spectrum, I L (ω ) . In this treatment, φ can be interpreted
as a frequency-independent offset of the spectral phase φ L (ω ) . This
relationship is immediately apparent from the following Fourier-transform
link between the time- and frequency-domain electric fields
I L (ω ) exp[i(φ L (ω ) − ωt + φ )] dω + c. c. .
1
E L (t ) = (4)
2π ∫
Here the intensity spectrum I L (ω ) and the phase φ L (ω ) fully determine the
quantities AL (t ) and ω L (t ) [2] but do not specify the carrier-wave offset.
The spectral manifestation of φ, given by Equation (4), is particularly
valuable for describing various single-shot measurements of the carrier-
envelope-phase throughout the chapter.
Polarizer
(b)
VND Spectrometer
SHG
Epulse<1 µJ
WLG (sapphire)
Figure 10-2. Setup for phase characterization of amplified pulses by nonlinear spectral
interferometry. VND: variable neutral density filter; SHG: second harmonic generator; WLG:
white light generator.
Initial amplifier phase tracking schemes [7, 20, 40] took advantage of
abundant pulse energy and used different types of single-shot nonlinear
spectral interferometry. The objective was to use broadband detection of the
ν-to-2ν beat as opposed to the narrowband scheme acceptable in the case of
272 Chapter 10
E SHG (ω ) ∝ exp(i 2ϕ )∫ χ (ω : ω ′, ω − ω ′) I WL (ω ′) I WL (ω − ω ′)
2
where φWL (ω ) is the spectral phase of the white light and χ is the second-
2
S (ω ) = (1 − a) I WL (ω ) + aI SHG (ω )
, (7)
+ 2 a(1 − a) I WL (ω ) I SHG (ω ) cos(φ SHG (ω ) − φWL (ω ) + ωτ 0 + ϕ )
where the coefficient a stands for the polarizer transmission for the
polarization of the SHG light. For a sufficiently large pulse separation τ 0 ,
corresponding to many spectral fringes in the interferogram S (ω ) , the
argument of the cosine in Equation (7) can be recovered using the standard
algorithm of spectral interferometry [47]. The phase of the newly obtained
complex spectral function then directly yields the differential phase of the
two fields, mod 2π {φ SHG (ω ) − φWL (ω ) + ϕ }. In principle, such an interference
measurement can be used not only to characterize the pulse-to-pulse changes
of φ, but also to find the actual value of φ. In practice, however, this is not
feasible because with current methods, φ SHG (ω ) and φWL (ω ) can only be
characterized with the precision of an arbitrary constant [2]. Only by fully
accounting for the evolution of the spectral phase in the SHG crystal, which
includes both linear and nonlinear pulse propagation, would it be possible to
retrieve the value of φ in this scheme. Therefore, the ν-to-2ν interferometer
can only measure the full carrier-envelope phase after calibration by an
independent external experiment, as shown in Section 7. Otherwise, this
device reflects the phase jump in the jth laser pulse with respect to another
(e.g., 0th) pulse arriving earlier from the same pulse train and chosen as
reference, i.e., the interferometer measures δφj = φj − φ0. The difference
between the method outlined above and Fourier-transform-spectral-
interferometry based on the use of parametric waves, as reported in
Reference [22], is that, in the latter case, one measures the interference
between the frequency-doubled idler and the signal pulses. The advantage is
that the white-light continuum injected into the parametric device does not
10. GENERATION AND MEASUREMENT OF INTENSE PHASE- 273
CONTROLLED FEW-CYCLE LASER PULSES
have to span an octave and the interference is observed close to the spectral
peak of the signal rather than in the continuum wing.
In this work, however, we opted for a simpler setup, as shown in [Figure
10-2], which is similar to the scheme published in Reference [20]. For
further simplification, we replaced white-light generation in a hollow
waveguide with bulk continuum generation in a 2 mm of sapphire. Because
of the dispersion of the sapphire and of the 1 mm BBO (β-BaB2O4)
frequency-doubling crystal, the white-light and SHG are sufficiently
separated in time to employ spectral interferometry. The spectral
interferogram is measured by a fiber-coupled spectrometer behind a
polarizer cube that selects the common polarization of the beams. With a 2
MHz acquisition card, single-shot ν-to-2ν interferograms can be detected for
every 3rd laser shot at a 1 kHz repetition rate. In the ν-to-2ν unit shown in
Figure 10-2, the timing jitter, τ = τ 0 + δτ , has an immediate impact on the
fringe pattern of the measured interferogram, as will be shown in Section
3.3. However, for the collinear arrangement, depicted in Figure 10-2, the
problem of timing jitter is negligible as the interfering beams follow the
same path.
and reference pulses. The shape of the intensity spectrum I L (ω ) [solid curve
in Figure 10-3(a)] and differential phase φ (ω ) [dotted curve in Figure 10-
3(c)] are selected arbitrarily. The values of δτ and δφ are chosen so that the
interferograms in Figure 10-3(b) closely resemble each other. Nevertheless,
there is a fundamental difference because the timing variation changes the
fringe period. This change can be clearly observed in an interferogram
consisting of a large number of fringes. Therefore, a charged coupled device
(CCD) camera-based spectrograph (which covers the relevant spectral width
and has a sufficiently high resolution) makes it possible to separate the
contributions of δφ and δτ as long as Fourier-transform–spectral-
interferometry noise criteria [49] are met. We implemented the time jitter
correction numerically in our Fourier-transform–spectral-interferometry
code by extracting a linear phase obtained by back-transformation into the
frequency domain.
(a) FT φ(ω)+ωτ0
Intensity
Intensity
-τ0 0 τ0
Time
(b)
Intensity
φ(ω)+ωτ0+δϕ
φ(ω)+ω(τ0+δτ)
(c) φ(ω)+ωδτ
φ(ω)+δϕ
Phase
δϕ
φ(ω)
δτ
0
0 ω0
Frequency
Figure 10-3. Discrimination between phase and timing jitter in Fourier-transform spectral
interferometry. (a) Intensity spectrum (solid curve) and spectral interferogram (dotted curve).
(b) Spectral interferograms modified by a timing shift (solid curve) and a frequency-
independent phase shift (dashed curve). (The fringe period changes in the case of timing
shift.) (c) Differential spectral phases after time-delay slope subtraction. Dotted and solid and
dashed curves correspond to interferograms in (a) and (b), respectively. Inset shows the power
276 Chapter 10
spectrum of the Fourier transform to the time domain. Dashed contour corresponds to a
possible filter required in the phase extraction to isolate the temporal peak at τ0.
4.1 Phase lock between the input pulse and the white-
light continuum
[∇ 2
]
+ k 2 (ω ) E (r , ω ) =
ω 2 nl
ε 0c 2
P (r, ω ) , (9)
P nl (r, ω ) = ∫∫ χ (ω : ω1 ,ω 2 , ω1 + ω 2 − ω )
3
where χ
3
represents the effective third-order susceptibility and the
conjugation symbol marks emission [52]. The propagation-dependent
change of dispersion is included in φ (r , ω ) . Here we assume that all mixing
frequencies are positive [53]. The appearance of the conjugation operator in
the convolution integral in Equation (10) accounts for the preservation of the
initial phase offset φ in the output field. After solving Equation (9) over the
interaction length of the white-light-generation medium r0 → r1 , the
resultant output pulse will generally carry a different value of the carrier-
envelope phase [6, 54] than the input pulse (i.e., φWL ≠ φ), which is also the
case for any type of dispersive linear or nonlinear pulse propagation.
Nevertheless, the important criterion for an accurate carrier-phase-phase drift
determination is the ability of the white-light continuum to repeat (without
distortion) the input phase offset. This ability is the consequence of the rule
for phase summation in Equation (10). After plugging Equation (10) into
Equation (9), φ cancels out, i.e., it has no effect on φ (r1 , ω ) of the output
field regardless of the microscopic origin of the nonlinearity. The presence
of a resonant or nonresonant fifth-order parametric wave mixing
( EE * EE * E ) [55] does not lead to a φ-dependent propagation either.
However, the solution for pulse propagation does become φ-dependent if
second-harmonic generation and/or third-harmonic generation cannot be
neglected. To account for these contributions, P nl (r, ω ) should also include
the terms describing the three-wave mixing of second-harmonic generation
and the four-wave mixing of third-harmonic generation, with their phase
summation rules given by the field products EE and EEE , respectively.
We would like to underscore two points. First, the phase shift added by
both the nonresonant and/or resonant white-light generation does not depend
on the carrier-envelope phase of the input pulse. Therefore, the techniques of
spectral broadening can be readily applied to schemes for measuring the
278 Chapter 10
Spectrometer Adjustable
delay
CCD 1 CCD 2
BS
VND
BS BS WLG
Onset of Breakdown of
Nonlinear phase shift
conditions of
f-to-2f measurement
-5π
Figure 10-5. Intensity-phase coupling of the white-light generation in bulk sapphire. Solid
dots depict phase changes extracted from single-shot interferograms. Straight line shows a
linear fit marking the slope of phase intensity-dependent variation in the vicinity of the pulse
energy used for ν-to-2ν characterization.
5. CONCEPTS OF PHASE-CONTROLLED
AMPLIFICATION
WLG seed
(b)
Epulse~5 nJ 3 nJ
Oscillator PP Ampl-
ifier
Trigerring
2 nJ signal
NLI Comparator
(c) Epulse~60 nJ
cw 3 nJ
Oscillator ampl- PP Ampl-
ifier ifier
2 nJ
Cavity
GD <0.1 µJ
control NLI
Linear
PLL Σ FTSI
(d) Epulse~60 nJ
kHz rep. rate 3 nJ
Ampl-
Oscillator PP ifier
2 nJ
Cavity MHz rep. rate <0.1 µJ
GD NLI
control Linear
FTSI
PLL Σ
(e)
Epulse=5 nJ 3 nJ
Oscillator PP Ampl-
ifier
~1 µJ
Cavity
GD 2 nJ NLI
control NLI
Nonlinear
FTSI
PLL Σ
Figure 10-6. Multiple concepts of phase-stable pulse amplification. (a) All-optical (passive)
phase stabilization with optical parametric amplification (OPA). (b) Phase-selective seed by a
free-running oscillator. (c)–(e) Schemes employing a phase-stabilized seed oscillator with a
secondary phase-correction loop behind the amplifier based on linear (c), (d) and nonlinear (e)
phase-drift detection. Concepts (c) and (d) require high-energy seed that can be attained with
a cw-amplified or a long-cavity oscillator (c) and a cavity-dumped oscillator (d). PP: pulse
picker; PLL: phase-locked loop; NLI: nonlinear ν-to-2ν interferometer; and FTSI: Fourier-
transform spectral interferometry.
284 Chapter 10
The schemes drawn in Figure 10-6(c) and (d) are based on the use of a
carrier-envelope-controlled oscillator [12, 13]. They explore the possibility
of sending a reference (pilot) beam for the measurement of the phase offset
in the amplifier. Information on the slow phase drift behind the amplifier is
obtained with the linear spectral-interferometry described in Section 3.3. It
can then be plugged into the phase-stabilization loop of the oscillator. The
core of the phase-stabilizing electronics is the phase-lock loop (PLL in
Figure 10-6) that forces the beat signal ω CE to oscillate in phase with an
external local oscillator via a feedback to the intracavity dispersion. The
inverse of the phase offset measured by spectral interferometry at the
amplifier output can then be combined with the phase of the local oscillator
signal ( 1 4 ω r ), thus precompensating this offset already in the oscillator.
The use of the reference beam for (a potentially single-shot) spectral
interferometry imposes the demand for high-seed pulse energy. Pulses of
tens of nanojoules can be obtained by intermediate cw amplification of the
seed pulse train before ν-to-2ν characterization, by employing long-cavity
oscillators [13, 65] or using a cavity-dumped oscillator [66]. This type of
laser has a dual output. It emits a pulse train through the output coupler at
the full repetition rate, which can be conveniently used to drive the ν-to-2ν
interferometer. At the same time, the pulses ejected by the intracavity pulse
picker are directly synchronized with the amplifier and satisfy the needs for
a kHz seed and a spectral interferometry reference.
The remaining concept for phase control of an amplifier [Figure 10-6(e)]
offers the most compatibility with existing CPA systems and is well suited
for retrofitting them. This scheme requires implementation of a standard
phase-lock loop of the seed oscillator, which consists of a ν-to-
2ν interferometer for the oscillator, electronics, and an additional secondary
feedback loop. The feedback loop consists of a high-energy ν-to-2ν
interferometer and computer for the spectral interferometry. Neither the
oscillator nor the amplifier requires significant modification. We have
followed this blueprint to upgrade a standard multipass Ti:sapphire CPA
system [46] for carrier-envelope–phase-stabilized operation.
with both feedbacks active. The device exhibits a rms jitter of ~75 mrad.
This figure, however, should not be confused with the actual precision with
which the carrier-envelope phase can be maintained behind the amplifier.
Since the phase characterization in the nonlinear interferometer is influenced
by the device noise (cf. discussion in Section 3), the feedback loop would
not be able to discriminate this measurement noise from the real phase jitter
in the amplifier. Therefore, an external (out-of-loop) experiment is required
to confirm the carrier-envelope phase stability. It will be presented in
Section 7.
Delay [fs] Wavelength [nm]
(a) -20 -10 0 10 20 1000 800 600
Measured 8 1
Electric field strength [arb. units]
Retrieved
Intensity
Intensity
1
0 0
1.5 2.0
Photon energy [eV]
AL(t) AL(t)cos(ωL(t))
-1
-15 -10 -5 0 5 10
Time [fs]
+
π
_ 0 0
500 520 540 500 520 540
2 Wavelength [nm] Wavelength [nm]
-π_2 (oscillator
Simple phase-lock
loop only)
Combined phase-lock
(oscillator+amplifier loop)
0 5 10 15
Time [s]
Figure 10-7. Performance overview of the phase-stabilized amplified system. (a) Pulse
properties behind the hollow-fiber–chirped-mirror compressor: retrieved pulse envelope
(dashed curves) and a possible electric field (solid curve) of the compressed pulse. Insets
show measured intensity spectrum (right) and measured and retrieved second-order
interferometric autocorrelation (left). (b) Carrier-envelope phase dynamics with oscillator-
only phase stabilization (dotted curve) and combined oscillator-amplifier phase stabilization
(solid curve). Insets show ν-to-2ν interference spectrograms recorded with the nonlinear
interferometer at the amplifier output [Figure 10-6(e)].
288 Chapter 10
process had been developed in several excellent models [23, 24, 69] that
have been tested and confirmed by a large number of numerical studies and
experiments [29-31, 70].
The relevant aspects of the process can be summarized as follows. The
highest-energy XUV or soft x-ray photons are emitted around the zero
transition(s) of the laser electric field near the peak of the envelope. This is
implicit in semiclassical models [6, 23, 24, 69] and was recently verified in
an attosecond experiment [27]. The energy of the emitted photon is
essentially determined by the intensity of the laser half-cycle preceding the
emission. For a few-cycle pulse with φ ≈ 0 or π [the “cosine” wave, shown in
Figure 10-1(b)], there is only one most intense half cycle that implies the
emission of an isolated x-ray burst at the highest photon energies
[Figure 1(h)]. The presence of such an isolated feature in the time domain
corresponds to a smooth spectrum. In stark contrast, an identical pulse with a
“sine” carrier [φ ≈ ± π/2, Figure 10-1(c)] exhibits two most intense half
cycles, giving rise to a pair of soft x-ray bursts separated by T0/2 in time
[Figure 10-1(i)]. In the frequency domain, this temporal structure implies a
modulated (quasi-harmonic) spectrum up to the highest photon energies.
To check the validity of this simple intuitive picture, we developed a
computer code simulating propagation of an intense few-cycle laser pulse
through an ionizing medium. The program solves Maxwell’s equations in
three space dimensions and calculates the emergence of high-order
harmonics using a refined version [6] of the quantum theory of Lewenstein
et al. [24, 71]. Figure 10-8 summarizes the results of the simulations relevant
to the experiments described below. In the simulation, the laser pulse
duration is τL = 5 fs, the pulse energy is 0.2 mJ, the beam diameter is 2wL =
122 µm, the medium is neon, the pressure is 100 mbar, and the length is
2 mm. Figure 10-8(a)–(d) displays the computed spectral distribution of few-
cycle-driven coherent soft x-ray emission from neon in the cutoff range for
four different values of the carrier-envelope phase of the driving laser pulse.
Figure 10-8(e) and (f) shows the temporal intensity profile of cutoff x-rays
transmitted through a Gaussian filter [solid contour in Figure 10-8(a)] for φ =
0 and φ = π/2 (solid curves) along with the instantaneous intensity of the
driving laser [i.e., E L (t ) 2 ], depicted by dashed curves. Detailed numerical
simulations show that the smoothest and most-modulated spectral-cutoff
features actually correspond to the values of ~20° and ~110°, respectively,
rather than to the intuitive values of 0° and 90°. However, this discrepancy is
insignificant in our experiments, because it lies within the uncertainty with
which we can identify different carrier-envelope phase values.
The results of our simulations, together with previous numerical studies
[6, 30, 31, 72], support our intuitive analysis and show a strong spectral
290 Chapter 10
dependence between the temporal structure of the x-ray bursts and the
spectral shape of the cutoff region. The width of the smooth spectral
continuum, ∆(hω)cont, that corresponds to an isolated x-ray burst depends on
the ratio of instantaneous intensity E L (t ) 2 of the strongest half cycle and its
neighboring peak. The corresponding difference of instantaneous intensity is
marked by ∆ in Figure 10-8(e). Because of the linear dependence between
the highest-emitted x-ray photon energy and the peak laser intensity, the
continuum bandwidth can be expressed as:
hω cutoff − I p
∆ (hω ) cont = ∆,
(E L (t )
2
)
max
(11)
(a) ϕ=0
(b) ϕ=π/6
(c) ϕ=π/3
(d) ϕ=π/2
(e) 2
(EL)max ϕ =0 EL(t) [10 W/cm ]
1
∆
2
15
0
(f) ϕ=π/2
2
0
-2 -1 0 1 2 3
Time [fs]
Figure 10-8. Numerical simulations of few-cycle-driven coherent soft x-ray emission from
ionizing atoms. (a)–(d) Cutoff-range spectra are shown for different carrier-envelope-phase
settings of the driving laser pulse. (e)–(f) The solid curves depict the temporal intensity profile
of the cutoff harmonic radiation filtered through a Gaussian bandpass filter with a full width
at half maximum of 7 eV [solid contour in (a), whereas the dashed curves plot EL(t)2]. These
results support the intuitive analysis that the field carrying φ = 0 is predicted to produce a
single soft x-ray burst [filtered in the cutoff (e)]. Deviation of φ from zero gradually
suppresses the magnitude of the main burst and gives rise to a satellite spike. The latter
becomes most prominent for φ → π/2 (f). In the frequency domain, the isolated pulse
emerging for φ = 0 implies a continuous spectrum shown in (a) that becomes increasingly
modulated with the appearance of the second burst for φ → π/2 (b)–(d).
292 Chapter 10
Table 10-1. Expected spectral width of the cutoff continuum for a φ = 0 wave as a function of
pulse duration.
τ L T0 † ∆ (E L ( t )2 )max ‡
∆(hω )cont @125 eV *
2 0.16 20 eV
3 0.07 8.8 eV
4 0.04 5.0 eV
5 0.03 3.8 eV
† Number of oscillation cycles within the laser pulse duration
‡ Instantaneous intensity ratio
* width of continuum.
(a)
ϕ=ϕ0-π/2
(b)
ϕ=ϕ0+0
Soft-X-ray spectral intensity [arb. units]
(c) ϕ=ϕ0+π/3
(d) ϕ=ϕ0+π/2
(e) unlocked
Figure10- 9. Measured spectral intensity of soft x-ray emission from ionizing atoms driven by
5 fs pulses. (a)–(d) Data obtained with phase-stabilized pulses for different carrier-envelope-
phase settings are shown. (e) Spectrum measured without carrier-envelope-phase stabilization.
X-ray CCD exposure was 0.5 s in all cases.
ϕ=ϕ0+π/4
ϕ=ϕ0+π/2
ϕ=ϕ0+3/2π
ϕ=ϕ0+0.9π
Unlocked
90 100 110 120 130 140 82 84 86 88 90 110 112 114 116 118
Photon energy [eV] Photon energy [eV] Photon energy [eV]
Figure 10-10. Measured spectral intensity of soft x-ray emission from ionizing atoms driven
by 10 fs pulses. (a) Full-range spectra. (b) and (c) Close-ups on spectral regions with
distinctly different spectral dependencies on the carrier-envelope phase.
Both the experiments with 5 and 10 fs pulses underscore the benefits that
carrier-envelope-phase stabilization brings to the study of coherent laser-
driven x-ray sources.
296 Chapter 10
8. CARRIER-ENVELOPE-PHASE MEASUREMENT
WITHOUT AMBIGUITY
Figure 10-11. Electric field (short-dashed curve) and envelope (long-dashed curve) of a 5 fs
laser pulse for CE phase 0 (left panel) and –π/2 (right panel). The drift momentum of high-
energy photoelectrons as a function of the ionization time t0 is shown in black (emission to the
right) and gray (emission to the left). It was calculated with a classical model. Note that
highest momentum does not necessarily coincide with highest electron yield: In the right
panel the ionization probability at times labeled with 1 is considerably lower than those
labeled with 2 because of lower field strength at t0.
Figure 10-13. (a) Photoelectron spectra for different CE phases controlled by fine movement
of one of the wedges. ∆x indicates the glass added hereby. The black curves correspond to
emission to the right (positive direction), the gray ones to the opposite direction. The insets
show the corresponding real time variation of the electric field, as deduced from the phase
assignment shown in Figure 10-14. Only without phase stabilization were identical spectra
measured to both left and right as expected. (b) Left-right ratio of the total electron yield
(solid black curve) and high-energy electrons (dashed gray curve) as a function of glass
thickness ∆x added or subtracted by moving one of the wedges. ∆x = 0 corresponds to optimal
dispersion compensation, i.e., the shortest pulses. Maximal left/right ratio for the total yield
does not coincide with that for high-energy electrons. Note the different scales for low- and
high-energy electrons. The upper horizontal scale indicates the CE envelope phase of the
pulse, as deduced from the comparison with theory shown in Figure 10-14. (c) Fine-increment
phase scan indicating long-term CE phase stabilization drift of the laser setup. The
measurement was performed over ≥10 min by slowly scanning the wedges forth (triangles
pointing to the right) and back (triangles pointing to the left) using a fine increment of the
glass thickness. The mismatch at the start and the end indicates a phase drift of ~50 mrad/min.
300 Chapter 10
experiment does not affect the laser beam. Thus a stereo above-threshold-
ionization phase meter can be placed anywhere along a laser beam line.
z
φ = − arctan ,
(12)
z R (λ )
Equation (12). Indeed, the details of the phase change in the focus depend on
the spatial profile of the laser beam and on the focusing geometry. In all
cases, an overall π phase shift is expected between two symmetric positions
far away from a spherical focus. Because of the continuous phase change of
the pulse propagating through the focal region, the corresponding electric
field of light takes on all possible phase variations. For example, a focused
“+cos” E-field will be a sinelike wave in the focal region and will reemerge
as a “-cos” field beyond the focus. This variation, of course, can affect the
outcome of a phase-sensitive light–matter interaction that is not confined to a
length much shorter than the Rayleigh range. Thus, precise control of the
spatial variation of the carrier-envelope phase within the whole focal region
is crucial to any phase-dependent experiment.
We present the experimental determination of the evolution of the
carrier-envelope phase in the focus of few-cycle laser pulses. Together with
the determination of the carrier-envelope phase presented in Section 8, this
result constitutes a full and unambiguous characterization of the electric field
of laser pulses in space and time within the paraxial approximation (i.e.,
neglecting the small longitudinal component of the electric field). We
studied the spatial variation of the carrier-envelope phase using the same
technique that was presented in Section 8, i.e., by simultaneously detecting
above-threshold-ionization photoelectron spectra in opposite directions
(Figure 10-12). The pulse energy of carrier-envelope-controlled 5 fs pulses
was attenuated to 20 µJ, and the beam focused with a f/30 geometry into a
low-density xenon gas jet. Under these focusing conditions, the electric field
is expected to undergo a π phase shift over a few millimeters. To reveal the
the Gouy phase, one has to selectively detect the electrons generated at a
well-defined position of the focus. A pair of moveable slits perpendicular to
the beam (z direction) and to the polarization axes allows the entire focal
region to be scanned. The slit width is 250 µ m, well below the Rayleigh
range (≈1 mm). To achieve optimum spatial resolution in the z direction, the
slits are placed at a distance of only 1 mm from the beam (Figure 10-12).
With this setup, the angular distribution of the emitted photoelectrons does
not affect the phase resolution, which is estimated to be ≈0.1 rad.
Figure 10-15 shows the asymmetry of the electron count rate (left/right
ratio) as a function of the glass thickness introduced by moving the fused-
silica wedges in front of the above-threshold-ionization spectrometer. We
made the measurement by moving the pair of slits to a distance of ≈2 mm
before (dashed line) and after (solid line) the focus. The phase shift of π
between the two curves is a direct measurement of the Gouy phase shift in
the focus.
304 Chapter 10
Figure 10-15. Variation of glass thickness in the beam path to change the carrier-envelope
phase of the pulses. The left/right ratio of the electron yield exhibits clear oscillations with a
periodicity consistent with glass dispersion at the wavelength of the laser. The measurement
was performed before (dashed line) and after (solid line) the focus. The π phase shift is due to
the Gouy carrier-envelope phase shift resulting from the passage through the focus.
measurements were made by approaching the focus from the outer part and
by moving the slits alternatively before and after the focus. With this
procedure, measurements of symmetric positions around the focus are
consecutive, thus reducing detrimental effects from possible long-term phase
drifts.
Figure 10-16. Left/right asymmetry maps (logarithmic scale) for different longitudinal
positions as a function of the electron energy and glass thickness introduced. Lighter shades
indicate dominant left emission; darker shades— dominant right emission. The maps (a) –(f )
correspond, respectively, to the positions: z=-1.75 mm, z=-1.0 mm, z=-0.25 mm, z=+0.25 mm,
z=+1.0 mm, z=+1.75 mm (positive values represent positions after the focus). The phase
difference is determined by evaluating the shift of the characteristic structures (indicated by
the dashed lines) of the asymmetry pattern. The extension of the electron yield to higher
energies in the central part of the focus is due to the higher intensity.
Figure 10-16(a) and (f) corresponds to the outer part of the focal range.
The strong asymmetry in the high-energy part of the spectra [above-
threshold-ionization plateau (dashed area)] changes sign while passing
through the focus, confirming the π phase shift already discussed (see Figure
10-15). More interestingly, Figure 10-16(b)–(e) corresponds to positions in
the central part of the focus. The asymmetry in the plateau is partly smeared
out, but another clear asymmetric area appears in the low-energy part of the
306 Chapter 10
Figure 10-17. Retrieved carrier-envelope Gouy phase shift as a function of the propagation
distance in the focus. The solid line is the Gouy phase of a cw Gaussian beam, shown for
comparison. In the outer part of the focus, the electron count rate rapidly decreases, making
detection of additional experimental points difficult.
The pulses undergo the π phase shift within a few Rayleigh distances.
Because of the rapid decrease of electron yield at lower intensities, the
measurements were stopped at a distance of ≈2 mm before and after the
focus. This prevented observation of the expected area of constant carrier-
10. GENERATION AND MEASUREMENT OF INTENSE PHASE- 307
CONTROLLED FEW-CYCLE LASER PULSES
envelope phase in the outer part of the focus. However, the region of interest
for all experiments is entirely covered, and the estimated error for the
experimental data is relatively low (≤0.1 rad). The phase changes smoothly
with a constant slope and does not exhibit any wiggles or irregularities,
which is particularly important for experiments.
ACKNOWLEDGEMENTS
REFERENCES
[1] A. M. Weiner, Rev. Sci. Instrum. 71, 1929-1960 (2000).
[2] R. Trebino, K. W. DeLong, D. N. Fittinghoff, J. N. Sweetser, M. A.
Krumbugel, B. A. Richman, and D. J. Kane, Rev. Sci. Instrum. 68, 3277-
3295 (1997).
[3] C. Iaconis and I. A. Walmsley, Opt. Lett. 23, 792-794 (1998).
[4] E. Constant, Ph.D Thesis, University of Sherbrooke (1997).
[5] G. Steinmeyer, D. H. Sutter, L. Gallmann, N. Matuschek, and U. Keller,
Science 286, 1507-1512 (1999); B. Schenkel, J. Biegert, U. Keller, C.
Vozzi, M. Nisoli, G. Sansone, S. Stagira, S. De Silvestri, and O. Svelto,
Opt. Lett. 28, 1987-1989 (2003).
[6] T. Brabec and F. Krausz, Rev. Mod. Phys. 72, 545-591 (2000).
[7] A. Baltuška, T. Fuji, and T. Kobayashi, Opt. Lett. 27, 306-308 (2002).
[8] L. Xu, C. Spielmann, A. Poppe, T. Brabec, F. Krausz, and T. W. Hänsch,
Opt. Lett. 21, 2008-2010 (1996).
[9] T. Udem, Ph.D Thesis, Ludwigs-Maximilians Universität (1997).
[10] J. Reichert, R. Holzwarth, T. Udem, and T. W. Hänsch, Opt. Commun.
172, 59-68 (1999).
[11] H. R. Telle, G. Steinmeyer, A. E. Dunlop, J. Stenger, D. H. Sutter, and U.
Keller, Appl. Phys. B 69, 327 (1999).
[12] D. J. Jones, S. A. Diddams, J. K. Ranka, A. Stentz, R. S. Windeler, J. L.
Hall, and S. T. Cundiff, Science 288, 635-639 (2000).
[13] A. Apolonski, A. Poppe, G. Tempea, C. Spielmann, T. Udem, R.
Holzwarth, T. W. Hänsch, and F. Krausz, Phys. Rev. Lett. 85, 740-743
(2000).
[14] T. Udem, J. Reichert, R. Holzwarth, and T. W. Hänsch, Opt. Lett. 24, 881-
883 (1999); T. Udem, R. Holzwarth, and T. W. Hansch, Nature 416, 233-
237 (2002); M. Zimmermann, C. Gohle, R. Holzwarth, T. Udem, and T. W.
Hansch, Opt. Lett. 29, 310-312 (2004).
[15] R. Holzwarth, T. Udem, T. W. Hänsch, J. C. Knight, W. J. Wadsworth, and
P. S. J. Russell, Phys. Rev. Lett. 85, 2264-2267 (2000).
[16] S. A. Diddams, D. J. Jones, J. Ye, S. T. Cundiff, J. L. Hall, J. K. Ranka, R.
S. Windeler, R. Holzwarth, T. Udem, and T. W. Hänsch, Phys. Rev. Lett.
84, 5102-5105 (2000).
10. GENERATION AND MEASUREMENT OF INTENSE PHASE- 309
CONTROLLED FEW-CYCLE LASER PULSES
[17] S. T. Cundiff, J. Ye, and J. L. Hall, Rev. Sci. Instrum. 72, 3746-3771
(2001).
[18] S. T. Cundiff, J. Phys. D. 35, R43 (2002).
[19] MenloSystem-GmbH (http://www.menlosystem.com)
[20] M. Kakehata, H. Takada, Y. Kobayashi, K. Torizuka, Y. Fujihira, T.
Homma, and H. Takahashi, Opt. Lett. 26, 1436-1438 (2001).
[21] A. Baltuška, T. Fuji, and T. Kobayashi, Phys. Rev. Lett. 88, art. no.-133901
(2002).
[22] A. Baltuška, T. Fuji, and T. Kobayashi, Opt. Lett. 27, 1241-1243 (2002).
[23] P. B. Corkum, Phys. Rev. Lett. 71, 1994-1997 (1993).
[24] M. Lewenstein, P. Balcou, M. Y. Ivanov, A. Lhuillier, and P. B. Corkum,
Phys. Rev. A 49, 2117-2132 (1994).
[25] M. Drescher, M. Hentschel, R. Kienberger, G. Tempea, C. Spielmann, G.
A. Reider, P. B. Corkum, and F. Krausz, Science 291, 1923-1927 (2001); P.
M. Paul, E. S. Toma, P. Breger, G. Mullot, F. Auge, P. Balcou, H. G.
Muller, and P. Agostini, Science 292, 1689-1692 (2001).
[26] M. Hentschel, R. Kienberger, C. Spielmann, G. A. Reider, N. Milosevic, T.
Brabec, P. Corkum, U. Heinzmann, M. Drescher, and F. Krausz, Nature
414, 509-513 (2001).
[27] R. Kienberger, M. Hentschel, M. Uiberacker, C. Spielmann, M. Kitzler, A.
Scrinzi, M. Wieland, T. Westerwalbesloh, U. Kleineberg, U. Heinzmann,
M. Drescher, and F. Krausz, Science 297, 1144-1148 (2002).
[28] M. Drescher, M. Hentschel, R. Kienberger, M. Uiberacker, V. Yakovlev,
A. Scrinizi, T. Westerwalbesloh, U. Kleineberg, U. Heinzmann, and F.
Krausz, Nature 419, 803-807 (2002).
[29] I. P. Christov, M. M. Murnane, and H. C. Kapteyn, Phys. Rev. Lett. 78,
1251-1254 (1997).
[30] A. de Bohan, P. Antoine, D. B. Milosevic, and B. Piraux, Phys. Rev. Lett.
81, 1837-1840 (1998).
[31] G. Tempea, M. Geissler, and T. Brabec, J. Opt. Soc. Am. B 16, 669-673
(1999).
[32] C. Spielmann, N. H. Burnett, S. Sartania, R. Koppitsch, M. Schnurer, C.
Kan, M. Lenzner, P. Wobrauschek, and F. Krausz, Science 278, 661-664
(1997).
[33] C. G. Durfee, A. R. Rundquist, S. Backus, C. Herne, M. M. Murnane, and
H. C. Kapteyn, Phys. Rev. Lett. 83, 2187-2190 (1999).
[34] G. G. Paulus, F. Grasbon, H. Walther, P. Villoresi, M. Nisoli, S. Stagira, E.
Priori, and S. De Silvestri, Nature 414, 182-184 (2001).
[35] C. Lemell, X. M. Tong, F. Krausz, and J. Burgdorfer, Phys. Rev. Lett. 90, -
(2003); V. S. Yakovlev, P. Dombi, G. Tempea, C. Lemell, J. Burgdorfer, T.
Udem, and A. Apolonski, Appl. Phys. B 76, 329-332 (2003).
[36] A. Poppe, R. Holzwarth, A. Apolonski, G. Tempea, C. Spielmann, T. W.
Hansch, and F. Krausz, Appl. Phys. B 72, 977-977 (2001).
[37] Y. Kobayashi and K. Torizuka, Opt. Lett. 25, 856-858 (2000).
310 Chapter 10
[38] P. Dietrich, F. Krausz, and P. B. Corkum, Opt. Lett. 25, 16-18 (2000).
[39] I. P. Christov, Opt. Lett. 24, 1425-1427 (1999).
[40] A. Baltuška, T. Udem, M. Uiberacker, M. Hentschel, E. Goulielmakis, C.
Gohle, R. Holzwarth, V. S. Yakoviev, A. Scrinzi, T. W. Hansch, and F.
Krausz, Nature 421, 611-615 (2003).
[41] J. Eckstein, Ph.D Thesis, Stanford University (1978).
[42] S. T. Cundiff and J. Ye, Rev. Mod. Phys. 75, 325-342 (2003).
[43] U. Morgner, R. Ell, G. Metzler, T. R. Schibli, F. X. Kärtner, J. G. Fujimoto,
H. A. Haus, and E. P. Ippen, Phys. Rev. Lett. 86, 5462-5465 (2001).
[44] J. C. Knight, T. A. Birks, P. S. Russell, and D. M. Atkin, Opt. Lett. 21,
1547-1549 (1997); J. K. Ranka, R. S. Windeler, and A. J. Stentz, Opt. Lett.
25, 25-27 (2000); J. K. Ranka, R. S. Windeler, and A. J. Stentz, Opt. Lett.
25, 796-798 (2000).
[45] R. Ell, U. Morgner, F. X. Kärtner, J. G. Fujimoto, E. P. Ippen, V. Scheuer,
G. Angelow, T. Tschudi, M. J. Lederer, A. Boiko, and B. Luther-Davies,
Opt. Lett. 26, 373-375 (2001); A. Bartels and H. Kurz, Opt. Lett. 27, 1839-
1841 (2002); T. M. Fortier, D. J. Jones, and S. T. Cundiff, Opt. Lett. 28,
2198-2200 (2003).
[46] S. Sartania, Z. Cheng, M. Lenzner, G. Tempea, C. Spielmann, F. Krausz,
and K. Ferencz, Opt. Lett. 22, 1562-1564 (1997).
[47] M. Takeda, H. Ina, and S. Kobayashi, J Opt Soc Am 72, 156-160 (1982).
[48] C. Dorrer and F. Salin, J. Opt. Soc. Am. B 15, 2331-2337 (1998).
[49] C. Dorrer, J. Opt. Soc. Am. B 16, 1160-1168 (1999); C. Dorrer, N. Belabas,
J. P. Likforman, and L. Joffre, Appl. Phys. B 70, S99-S107 (2000).
[50] F. W. Helbing, G. Steinmeyer, U. Keller, R. S. Windeler, J. Stenger, and H.
R. Telle, Opt. Lett. 27, 194-196 (2002).
[51] G. P. Agrawal, Nonlinear Fiber Optics (Academic Press, San Diego, 1995);
P. V. Mamyshev and S. V. Chernikov, Opt. Lett. 15, 1076-1078 (1990); J.
Botineau and R. H. Stolen, J Opt Soc Am 72, 1592-1596 (1982).
[52] N. Bloembergen, Nonlinear Optics (World Scientific, Singapore, 1996).
[53] S. M. Gallagher, A. W. Albrecht, T. D. Hybl, B. L. Landin, B. Rajaram,
and D. M. Jonas, J. Opt. Soc. Am. B 15, 2338-2345 (1998).
[54] T. Brabec and F. Krausz, Phys. Rev. Lett. 78, 3282-3285 (1997).
[55] S. Mukamel, Principles of Nonlinear Optical Spectroscopy (Oxford
University Press, New York, 1995).
[56] H. A. Haus and E. P. Ippen, Opt. Lett. 26, 1654-1656 (2001).
[57] X. Gu, L. Xu, M. Kimmel, E. Zeek, P. O'Shea, A. P. Shreenath, R. Trebino,
and R. S. Windeler, Opt. Lett. 27, 1174-1176 (2002); P. Baum, S.
Lochbrunner, J. Piel, and E. Riedle, Opt. Lett. 28, 185-187 (2003).
[58] M. Bellini and T. W. Hänsch, Opt. Lett. 25, 1049-1151 (2000).
[59] T. M. Fortier, J. Ye, S. T. Cundiff, and R. S. Windeler, Opt. Lett. 27, 445-
447 (2002); T. M. Fortier, D. J. Jones, J. Ye, S. T. Cundiff, and R. S.
Windeler, Opt. Lett. 27, 1436-1438 (2002).
[60] D. Strickland and G. Mourou, Opt. Commun. 56, 219-221 (1985); M. D.
Perry and G. Mourou, Science 264, 917-924 (1994).
10. GENERATION AND MEASUREMENT OF INTENSE PHASE- 311
CONTROLLED FEW-CYCLE LASER PULSES
Abstract: We discuss the physical processes involved in the generation and optimization
of extreme ultraviolet and soft x-ray light though the process of high-order
harmonic generation. We show that by manipulating the sub-optical-cycle
attosecond dynamics of this process using optimized waveguide structures and
pulse shapes, we can control the energy of the emitted photons, the phase
matching of the conversion process, and the spatial and temporal coherence of
the light. High-order harmonic generation is a useful source of short
wavelength light with ultrashort time duration. Thus, optimization and
manipulation of high-order harmonic generation demonstrates control of
electron dynamics on attosecond time scales.
the ionized electron recollides and then recombines with its parent ion,
releasing a high energy photon. The energy and phase of the emitted high-
harmonic light depend on the detailed electron trajectory in the laser field—
dynamics that occur on a fraction of an optical cycle, or attosecond,
timescale. Thus, the properties of HHG are sensitive to the field evolution of
the driving laser [4]; in contrast, traditional nonlinear processes, such as
second-harmonic generation, are insensitive to many aspects of the driving
field such as the carrier-envelope offset (CEO). Detailed studies of the
dependence of high-harmonic generation on the time-history of the driving
laser field represent the first results from the field of attosecond science [5,
6], an area of research that has received considerable attention in the early
2000s [7]. Consequences of the attosecond dynamics of HHG include the
ability to manipulate electron dynamics with attosecond precision [8-10], the
phase matching of the conversion process (attosecond engineering) [11, 12],
and the ability to generate pulses of light with subfemtosecond duration. The
strong dependence on the laser field also means that HHG is sensitive to the
CEO of the driving laser pulse [13-15]. By combining pulse shaping
techniques with recently developed methods to control the CEO of an
ultrafast pulse by locking the mode frequencies to a cw reference [16], it
becomes possible to control the complete electric field of the pulse in the
time domain with attosecond precision. This allows us to access the fastest
time scales that are possible with modern laser technology.
phase of 18 degrees after the peak of the laser cycle. This cutoff photon
energy is then predicted to be:
U p = e 2 E 2 / 4mω 2 ∝ Iλ 2 , (2)
ϕ ≈ qωt f − τ sU p , (3)
Figure 11-1. Illustration of intra-atomic phase matching. When the optimized pulse shape is
used, a particular harmonic constructively adds over each half cycle (from [9]).
matching. In its simplest manifestation, phase matching will occur when the
driving laser and the harmonic signal travel with the same phase velocity, so
that signal generated throughout the conversion region adds constructively.
In HHG, since the ionization level is changing throughout the pulse and the
increasing density of free electrons has a large effect on the phase velocity of
the driving laser, phase matching can only be obtained in limited time
windows in the pulse. Enhancement of the flux from HHG is possible by
guiding the driving laser with a hollow-core, gas-filled waveguide [28, 29].
The laser light is guided by glancing incidence reflection, allowing
propagation over an extended interaction length with a well-defined intensity
and phase profile. The fundamental and harmonic light propagate through
the waveguide with phase velocities determined by the dispersion of the
neutral gas, the plasma, and the waveguide at the two different frequencies.
The difference in the wave vectors of the fundamental and harmonic light
results in a phase mismatch
q112 λ 2π (1 − η )
∆k = + Pη N atm re (qλ − λ q ) − ∆δ , (4)
4π a 2
λ
−1
N atm re λ 2
η cr = 1 + . (5)
2π∆δ
No Phase-
Matching
∆k = 0
Figure 11-2. Plot of pressure for phase matching as a function of normalized ionization
fraction (η/ηcr). Beyond critical ionization, phase matching is no longer possible.
For low intensities, and therefore lower harmonic orders, where the
ionization fraction is small, phase matching is primarily achieved by
balancing the waveguide dispersion with the neutral gas dispersion. At the
higher intensities required to generate higher energy harmonics, however,
the plasma term becomes significant and higher neutral gas pressures are
needed to compensate. Beyond the critical ionization, phase matching is not
possible [13, 31]. The optimum pressure also depends on the diameter of the
waveguide; for the case of “plane-wave” propagation without a waveguide
in a uniform-density gas, phase matching will occur when the fractional
ionization equals the critical ionization [30]. However, this requires either
very high pulse energy or high gas pressure and results in spatially varying
phase matching that can create a complex spatial mode of the harmonic
emission. In contrast, the use of the hollow waveguide allows for better
phase matching, resulting in build-up of an EUV beam with a spatial mode
322 Chapter 11
of extremely high coherence [26, 32]. In 2004, stable beams of EUV light
with full spatial coherence had only been generated using the hollow
waveguide geometry.
Another consequence of the time-varying–phase-matching conditions is
sensitivity to the CEO of the pulse, both for few-cycle pulses and also in the
case of light pulses with durations of tens of femtoseconds [13, 15, 33]. At
the peak of the pulse, the ionization fraction is increasing significantly in a
series of steps over each half cycle, so that the phase-matching conditions
vary rapidly with time. The ionization as a function of time for a particular
driving laser pulse can be calculated using the Ammosov-Delone-Krainov
(ADK) tunneling ionization rates [34]. Figure 11-3 shows the fractional
ionization of argon gas for two different values of the CEO. The amount of
ionization created at each half-cycle is dependent on the CEO; however, the
final level of ionization is the same for both pulses.
1.0
0.10
0.8
Normalized Intensity
Ionization Fraction
0.08
0.6
0.06
0.4
0.04
0.02
0.2
0.00 0.0
-20 -10 0 10 20
Time (fs)
Figure 11-3. Calculation of the fractional ionization of argon using ADK ionization rates for a
20 fs pulse with peak intensity of 2.2 x 1014 W/cm2 for a cosine and sine pulse.
For a given harmonic order, tuning the pressure can allow phase
matching at different half cycles of the pulse. Durfee et al. observed
evidence of the CEO effect on phase-matching [13]. Figure 11-4(a) shows
the calculation of the pressure dependence of the flux of the 29th harmonic
for one value of the CEO for a 20 fs pulse. The different peaks correspond to
optimal phase matching at different ionization “steps.” The exact position of
these peaks depends on the intensity, the pressure, and the CEO of the pulse.
For a fixed gas pressure, the positions of the peaks move around as the
phase-matching conditions change with the CEO. Figure 11-4(b) shows the
experimentally measured 29th order harmonic signal. In the pressure region
corresponding to phase matching at the peak of the pulse, where the amount
of ionization is dependent on the CEO, the signal shows a strongly
11. QUANTUM CONTROL OF HIGH-ORDER HARMONIC 323
GENERATION
Figure 11-4. (a) Calculation of the flux of the 29th harmonic for a 20 fs laser pulse and peak
intensity of 2.2 x 1014 W/cm2 as a function of argon pressure using a 3 cm waveguide. (b)
Experimentally measured flux for the same conditions as (a) (from [29]).
dipole moment. Figure 11-5(b) shows the same calculation using a hollow-
core waveguide with a 0.5 mm period sinusoidal corrugation that changes
the laser intensity by 5%. The modulated waveguide causes a dramatic
enhancement in the final signal (note the different scales of the two plots).
We can understand the basic physics for enhancement of the HHG signal
by using classical nonlinear optics theory for quasi-phase-matching. In a
simplified model of harmonic generation, the field of harmonic order q, after
propagating a distance L in a nonlinear medium, is related to the phase
mismatch, ∆k , by
L
E q ∝ ∫ Eωn ( z ) d ( z )e −i∆kz dz , (6)
0
∑D e
∞
d (z) = m
iK m z
, (7)
m = −∞
Figure 11-5. (a) Calculation of the signal of the 95th harmonic for a straight waveguide as a
function of propagation distance. (b) Calculation for a 0.5 mm period modulated waveguide
(from [11]).
Figure 11-6. Experimental harmonic spectra from 111 Torr He with a driving pulse of 25 fs
and peak intensity ~5 x 1014 W/cm2 for different modulated-waveguide periodicities (from
[12]).
qne e 2 λ
∆k plasma ≈ , (8)
4πmeε o c 2
where λ is the laser wavelength and ne is the electron density. For example,
for fully ionized argon at a pressure of 1 Torr, ne = 3.5 x 1016 cm-3, giving ∆k
~7550 m-1 for the 95th harmonic order (150 eV). Therefore, the coherence
length, Lc, given by Lc = π/∆k, is ~0.4 mm. Thus, very substantial levels of
ionization can be compensated for by using QPM with modulation periods in
the range of 1–0.25 mm that can be readily manufactured with glass-blowing
techniques.
In initial experiments, the ionization level in helium was still relatively
low (~1%) and only ~4% in argon. However, more recent experiments have
demonstrated QPM in fully ionized gas. The results of more recent
experiments performed at higher laser intensities of 1.6 x 1015 W/cm2, in 9
328 Chapter 11
Modulated
fiber (0.25mm)
Carbon K-edge
Straight
fiber
Figure 11-7. Comparison of experimental harmonic spectra for neon for straight and
modulated fibers [40].
5. CONCLUSION
ACKNOWLEDGEMENTS
REFERENCES
[1] S. Backus, C. G. Durfee, M. M. Murnane, and H. C. Kapteyn, Rev. Sci.
Instrum. 69, 1207-1223 (1998).
[2] T. Brabec and F. Krausz, Rev. Mod. Phys. 72, 545-591 (2000).
[3] A. McPherson, G. Gibson, H. Jara, U. Johann, T. S. Luk, I. A. McIntyre, K.
Boyer, and C. K. Rhodes, J. Opt. Soc. Am. B 4, 595-601 (1987); M. Ferray,
330 Chapter 11
Key words: mode-locked laser, ring laser, sensors, optical parametric oscillator, optical
cavity stabilization
1. MODE LOCKING
(a)
Delay of
fs pulses 1 s??
Interferences?
(b)
Beat
fs pulses note
A
time
D
B
-0.8
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10
seconds
(b)
0.5
0.4
Amplitude
0.3
FWHM ~ 3Hz
0.2 Coherence time 0.3 s
0.1
0.0
200 250 300 350 400 450 500
Frequency (Hz)
Figure 12-2. (a) Beat note recorded between two femtosecond pulse trains of the same
repetition rate. (b) Fourier transform of the same recording, showing a bandwidth of only 3
Hz, or 0.8 Hz broader than the sampling time limit.
Round-trip i
ϕ0
Envelope
crossing
∆ϕ
Round-trip i+1
Moving interference
pattern (standing wave
if the two carrier frequencies are equal)
Figure 12-3. Topological representation of a mode-locked ring laser. The pulse envelopes
overlap at the same location of the cavity at each round trip.
∆ν ∆P
= , (1)
ν P
where P is the perimeter of the ring laser and ν the optical frequency. A beat
note ∆ν of 3 Hz corresponds to a perimeter difference ∆P of P x 3 Hz/ν ≈
10-14 m. A mirror position fluctuation of 10-14 m during a round-trip time of
10 ns corresponds to a motion of 0.01 µm in a mechanical resonance period
of 10 ms. To further reduce the bandwidth of the beat note, it is necessary to
suppress the fluctuations that cause a difference in time of arrival that, in
turn, will cause a difference in perimeter between the two circulating pulses
during any round trip. Stabilization of the cavity perimeter may achieve the
goal of reducing the beat note bandwidth. The type of stabilization required
will be discussed in Section 7, which follows a discussion of the applications
of the ring laser.
The detection of the beat note between the two outputs of a mode-locked
ring laser leads to numerous sensing applications of exceptional sensitivity.
The “sensor” consists of a mode-locked ring laser operating in a
bidirectional mode and a detector located at an equal optical path distance
from the intracavity pulse crossing point, as sketched in Figure 12-4. To
avoid gain competition and have equal energy for the two counter-
circulating intracavity pulses, the amplifying medium should be located 1/4
cavity perimeter away from the pulse crossing point.
12. APPLICATIONS OF ULTRAFAST LASERS 339
Synchronized
S excitation
Detector
Amplitude
coupling
Gain medium
∆ν ∆P
=
ν P
∆ν = ∆ϕ /τRT
Figure 12-4. The general configuration of the sensor is that of a mode-locked ring laser and a
detector located at equal optical path from the pulse crossing point, via the clockwise or
counterclockwise direction of circulation. The gain medium is located at 1/4 cavity perimeter
from the pulse crossing point. S is the sample of which the characteristics have to be
determined. An eventual excitation of the sample is applied exactly at the cavity round-trip
time.
Without the need for adding any intracavity element, the device of Figure
12-4 can be used for rotation sensing. The rotational beat note response ∆νR
is related to the rotation of the laser (angular velocity ΩR) around its axis by:
4A
∆ν R = ΩR , (2)
Pλ
where A is the ring area. A large area to perimeter (A/P) ratio is desirable to
have an optimum sensitivity to rotation. For applications where rotation
sensing is not desired, a figure-of-8 laser with a net-zero area is the preferred
geometry. Since the femtosecond ring laser acts as laser gyroscope without a
dead band, the sensitivity to rotation is only limited by the bandwidth of the
beat note, itself directly related to the stability of the resonator. Without
active stabilization of the cavity, the 3 Hz bandwidth of the beat note of the
Ti:sapphire ring laser shown in Figure 12-2 corresponds to a sensitivity to
rotation of 0.2 degrees/hour, which compares favorably with the best
navigational optical gyroscopes. Stabilization of the cavity perimeter would
improve this figure by at least three orders of magnitude.
Rotation is not the only effect that causes a differential change in carrier
frequency for the counter-circulating pulses. The beat note is extremely
sensitive to air flows through the Fresnel Drag effect. The first evidence of
nonreciprocal response in a ring mode-locked dye laser was a measurement
of air flow through the Fresnel drag [6]. Very small air currents (air
velocities of the order of 1 mm/s) can also account for the observed beat note
bandwidth in Figure 12-2.
Magnetic field detection requires the insertion of a material with a high
Verdet constant in the ring cavity, as with sample S sandwiched between two
quarter-wave plates in Figure 12-4. The clockwise and counter-clockwise
circulating pulses become, respectively, right and left circular polarized in
the sample. In presence of a magnetic field, there is a difference in optical
path ∆P between the two senses of circulation, resulting in a beat note
proportional to the magnetic field. This effect can be particularly large in the
presence of resonance. Optimal magnetic field sensitivity requires the use of
a material for which the index change due to a magnetic field is maximal.
Atomic vapors, such as used in potassium magnetometers [7], have a large
resonant Faraday rotation, hence a large magnetic field-induced change in
index for circularly polarized light. These vapor magnetometers have been
used successfully in aerial mapping of the earth’s magnetic field and to
locate minerals, in particular magnetic iron deposits and nickel deposits [8].
The beat note response of the ring laser is significantly more sensitive to
12. APPLICATIONS OF ULTRAFAST LASERS 341
∆φ 2πd
∆ν = ≈ (n(I + ) − n(I −)) , (3)
τ RT λ
where I+ and I- are the beam intensities at the sample for the clockwise and
counterclockwise circulating pulses, respectively. Equation 3 is valid only in
the approximation that the sample thickness d is sufficiently small so that the
intensities of each beam can be assumed to be constant in the sample. The
difference n(I + ) − n(I −) can be varied by translating the sample along the
beam axis between two lenses or spherical mirrors [11]. An example of
measurement of the nonlinear index of lithium niobate with an optical
parametric oscillator ring laser can be found in Reference [12].
342 Chapter 12
Either in the ring laser or in its linear limit, an electro-optic device can be
inserted in the cavity to split the path of two circulating pulses, sending one
through the sample to be analyzed and the other through a reference path [5],
which can be stabilized to a reference cavity. Unlike the previous
measurement, a stable cavity is not a sufficient attribute to perform accurate
measurements. As will be seen in Section 5 on stabilization, the reference
cavity may drift; hence there would be an ambiguity in the measurement: is
the change in beat note reflecting an elongation of the sample or of the
reference cavity? Thus this type of measurement requires accuracy (long-
term stability) in addition to short-term stability through the reference cavity.
The solution is to use properties of atomic transition to accurately fix a
reference repetition rate and carrier frequency [14].
12. APPLICATIONS OF ULTRAFAST LASERS 343
4. INTRACAVITY-PUMPED OPTICAL
PARAMETRIC OSCILLATOR AS A MODE-
LOCKED RING LASER
Ti-Sapphire
OPO crystal
D
L2
MQW
Figure 12-5. Illustration of the intracavity optical parametric oscillator (OPO) pumped by the
Ti:sapphire laser. Four LaKL21 prisms are incorporated in the pump cavity to compensate for
the group-velocity dispersion (GVD) from the Ti:sapphire crystal, the Periodically Poled
Lithium Niobate (PPLN) crystal, and other intracavity elements such as lenses and mirrors.
This four-prisms configuration was necessitated by the desire to have large GVD
compensation (needed because of the large positive GVD of LiNbO3) and a reasonably short
cavity length (1/2 of the perimeter of the OPO cavity). The main control of the GVD
compensation is the prism spacing L2. Two quantum wells (MQW) of AlGaAs on top of a
mirror structure are used in the cavity as a saturable absorber to mode lock the laser.
Most measurements with ring laser sensors have been performed with
either dye lasers [6] or Ti:sapphire lasers [4, 11, 13] mode locked with
flowing saturable absorbers. Such systems are appropriate for demonstration
purposes, but the future is for all-solid-state compact systems that can be
readily stabilized. Some initial results on the most promising system, an
intracavity-pumped Optical Parametric Oscillator (OPO) ring, are presented
here. A synchronously pumped OPO offers the possibility of decoupling
relative phase and repetition rates of the oscillating signals. The repetition
rate of the OPO signal is equal to that of the pump laser and can therefore be
controlled through stabilization of the mode spacing of the pump laser. The
carrier frequency of the OPO is directly determined by the OPO cavity
length. The position of the crossing point of the two circulating pulses in the
OPO is simply determined by the timing of two successive pump pulses sent
344 Chapter 12
in opposite directions. It is essential that the OPO crystal be part of the pump
cavity. Only such an intracavity-pumped configuration can ensure that the
OPO gain volumes for either direction are directly superimposed (because
they share a common mode of the pump cavity). The operation of such a ring
OPO has been demonstrated [15]. The cavity configuration of such an OPO-
pumped intracavity by a linear Ti:sapphire laser is sketched in Figure 12-5.
The Ti:sapphire laser radiation consists of 200 fs pulses centered at 785
nm, with a repetition rate of 95MHz. The OPO crystal is a 3 mm Brewster-
cut PPLN crystal (HC Photonics, Taiwan) with a period of 19.4 mm (quasi-
phase matching for the signal near 1.36 mm) that is temperature stabilized at
408 K to prevent photorefractive damage. The cavity mirrors and crystal are
located away from the two crossing points of the ring.
0 .3 5
(a.u.)
0 .3 0
Amplitude
(a.u.)
0 .2 5
Amplitude
0 .2 0
- 0 .0 4 - 0 .0 2 0 .0 0 0 .0 2 0 .0 4
T Time
i m e (s)
(s )
0 .0 8
68.64
Amplitude (a.u.)
0 .0 6
0 .0 4
Amplitude (a.u.)
0 .0 2
0 .0 0
0 5 0 1 0 0 1 5 0 2 0 0
B e aBeat
t f r frequency
e q u e n c (Hz)
y (H z )
Figure 12-6. Beat frequency (a) taken by a sensitive InGaAs detector (D in Figure12-5) and
(b) its Fourier transform.
PZT1
Ti:S laser
PZT2
To main
AOM experiment
BS
BS
Reference Cavity PM
L2 L1
Grating
In from
L.O.
PD2 PD1
Servo1, Servo2,
νo frR
fm= 10.7MHz
To PM
0 .6
0.6
-
Voltage
Voltage
0 .0
0.0
-0.6
-0 .6
+
Frequency
Frequency
Figure 12-7. Schematic of experimental setup. The laser output, protected from feedback by a
Faraday isolator, is sent through an acousto-optic modulator that controls average frequency
before being sent through a phase modulator and mode matched to the reference cavity.
AOM: acousto-optic modulator; PM: phase modulator; B: beam splitter; PD1 and PD2:
photodiodes 1 and 2; and PZT1–3: piezo-electric transducers.
The error signals obtained in this way are the composite signals from all
longitudinal modes detected within the spectral regions seen at PD1 and
PD2. One detector can provide the error signal to lock the position of the
comb about some average frequency by adjusting AOM and PZT1 (i.e.,
controlling the cavity length). The difference between the error signals from
PD1 and PD2 determines fluctuations in the laser repetition rate that can be
stabilized by PZT2, which selects the average group velocity [18]. The
sensitivity of this discriminator is equal to the slope of the error signals
multiplied by the number of modes between them. This technique provides
an extremely sensitive discriminator for the repetition rate by effectively
locking a high harmonic of it to the cavity. The reference cavity sees the
348 Chapter 12
120
Frequency- 385.2845017 THz (kHz)
100
80 drift rate = 1.48 kHz/second
60 standard deviation = 707 Hz
40
20
0
-20
-40
-60
-80
-100
06 06 06 06 06 06 06 06 06
:38 :38 :39 :39 :39 :40 :40 :40 :40
:38 :55 :12 :29 :46 :03 :20 :37 :54
Time Stamp
Figure 12-8. Measurement of optical frequencies of the femtosecond comb. One hundred
consecutive measurements are made showing the slow, but easily measurable, drift of the
reference cavity. Fractional Allan deviation from these counts is σ (1s) ~3 x 10-12.
accurate 100 MHz frequency reference source provides the repetition rate
value. The 1-s Allan deviation for these counts is ~3 x 10-12. This number is
limited by the stability of the low-noise frequency reference (100 MHz).
A reference cavity and a laser have mutual dependency: each one can be
a source and reference for the other. As mentioned before, the nonzero offset
frequency is due to the phase change upon reflection from the two cavity
mirrors. In the frequency domain, this phase shift is usually frequency
dependent, i.e., the reference cavity mode spacing changes with frequency.
In the time domain, the pulse’s frequency component goes through different
depths in the mirror coating; thus each frequency component “sees” a
different length of the reference cavity, known as an "external cavity."
The stabilized femtosecond laser can act as a frequency comb to
characterize the dispersion of the external cavity mirrors. This ruler can
measure the dispersion of gases and low-density materials inside the external
cavity over a broad bandwidth. Furthermore, the high peak-intensity build-
up in the cavity provides a tool to study nonlinear phase shifts. An
experimental setup is presented in the following section.
mirrors. Higher quality mirrors with lower GVD increase the bandwidth of
the transmitted laser and improve locking stability of femtosecond lasers to
high-finesse reference cavities.
Mode-locked laser
fm= 10.7
MHz
CW laser
KLM laser
grating m-1 m m+1
CW laser
Frequency
BS
PBS
Error signals
Figure 12-9. Experimental setup to measure longitudinal modes of reference cavity. PBS:
polarizing beamsplitter and BS: non-polarizing beamsplitter.
200
Beat Frequency (kHz)
5
Cavity FSR - fr (Hz)
4
100
0
2
1
-100
0
360 365 370 375 380 385
Frequency (THz)
Figure 12-10. Measured deviation (∆) of the reference cavity longitudinal modes under
vacuum from the equidistant positions of the frequency comb (left ordinate; squares) and the
calculated mode spacing of the cavity vs frequency (right ordinate; dashed curve).
12. APPLICATIONS OF ULTRAFAST LASERS 351
∂Φ n
τn = , (4)
∂Ω
where Φ n is the total round-trip phase shift inside the cavity. The change in
∂σ ∂ 1 1 ∂τ 1 ∂ 2Φ n
= =− 2 n =− 2 , (5)
∂Ω ∂Ω τ n τ n ∂Ω τ n ∂Ω 2
and therefore the round-trip GVD can be related to the measured FSR by
1 ∂σ
Φ n′ = − . (6)
σ 2 ∂Ω
Figure 12-11. Measured deviation of the reference cavity longitudinal modes at atmospheric
pressure from the equidistant positions of the frequency comb (left ordinate; data points) and
the calculated mode spacing of the cavity vs frequency (right ordinate; dashed curve).
+
0000000
SERVO-LOOP FOR
REPETITION RATE
-
10.6 MHz
EO
BS
M2
M1
DBS
M3
C
M4
Figure 12-12. Sketch of a possible ring OPO geometry and the stabilization loop.
The beams of the pump laser and the OPO signal are combined by a
dichroic beam splitter (DBS). Both beams are given the same phase
modulation at 10.5 MHz by an electro-optic crystal. The reflection of the
reference cavity (1/2 the cavity length of the pump) is dispersed by a grating.
The difference between the two error signals taken at the edges of the pump
spectrum are used to control the cavity length (mirror M1) of the pump laser
and hence the repetition rate of the system. The wavelength of the OPO
signal is unaffected, since it is solely determined by the OPO cavity
perimeter. An error signal is thus derived from the dispersion of the grating
at the signal wavelength and mixed with the modulation. The correction
signal is sent simultaneously to the piezo drivers of mirrors M3 and M4,
located symmetrically with respect to the pulse crossing point C. It is
essential to symmetrically apply the correction for the signal cavity
simultaneously to the two circulating pulses to prevent a broadening of the
beat-note bandwidth by the motion of the cavity length-correcting mirrors.
As mentioned earlier, for applications that require long-term stability or
accuracy (monitoring or slow drifts in the picometer range), the OPO
frequency and the pulse repetition rate have to be locked to some atomic
transition. Because of the general tunability of an OPO, there is a large
choice of atomic systems available. One approach [14] is to use a Λ level
structure that has a two-photon resonance at the carrier frequency of one of
the OPO pulses and a ground-level splitting resonant with the pulse
repetition rate. The main challenge is to find an error signal (fluorescence,
dispersion, absorption) that has a well-defined peak at the proper optical
354 Chapter 12
REFERENCES
[1] B. C. Young, F. C. Cruz, W. M. Itano, and J. C. Bergquist, Phys. Rev. Lett.
82, 3799-3802 (1999).
[2] R. K. Shelton, L. S. Ma, H. C. Kapteyn, M. M. Murnane, J. L. Hall, and J.
Ye, Science 293, 1286-1289 (2001).
[3] R. J. Jones and J. C. Diels, Phys. Rev. Lett. 86, 3288-3291 (2001).
[4] S. Diddams, B. Atherton, and J. C. Diels, Appl. Phys. B 63, 473-480
(1996).
[5] J. C. Diels, M. Bohn, J. Jones, and T. T. Dang, (United States Patent, Filed
April 27 2000., 2000), Vol. Patent Pending, (UNM Docket No. UNM-490).
[6] M. L. Dennis, J. C. M. Diels, and M. Lai, Opt. Lett. 16, 529-531 (1991).
[7] P. J. Hood and H. Ward, Airborne Geophysical Methods (Academic Press,
1969).
[8] D. J. Teskey, P. J. Hood, L. W. Morley, R. A. Gibb, P. Sawatzky, M.
Bower, and E. E. Ready, Can. J. Earth Sci. 30, 243-260 (1993).
[9] P. J. Hood, Northern Miner, 1-14 (1991).
[10] S. A. Diddams, J. C. Diels, and B. Atherton, Phys. Rev. A 58, 2252-2264
(1998).
[11] M. Bohn, Ph.D Thesis, University of New Mexico (1998).
[12] X. M. Meng, J. C. Diels, D. Kuehlke, R. Batchko, and R. Byer, Opt. Lett.
26, 265-267 (2001).
[13] M. J. Bohn, J. C. Diels, and R. K. Jain, Opt. Lett. 22, 642-644 (1997).
[14] L. Arissian, J. Jones, and J. C. Diels, J. Mod. Opt. 49, 2517-2533 (2002).
[15] X. M. Meng, Ph.D Thesis, University of New Mexico (2003).
[16] A. I. Ferguson and R. A. Taylor, Opt. Commun. 41, 271-276 (1982); E.
Kruger, Rev. Sci. Instrum. 66, 4806-4812 (1995).
[17] T. Udem, J. Reichert, R. Holzwarth, and T. W. Hänsch, Phys. Rev. Lett. 82,
3568-3571 (1999).
[18] J. Reichert, R. Holzwarth, T. Udem, and T. W. Hänsch, Opt. Commun.
172, 59-68 (1999).
[19] R. J. Jones and J. Ye, Opt. Lett. 27, 1848-1850 (2002).
[20] R. J. Jones, J. C. Diels, J. Jasapara, and W. Rudolph, Opt. Commun. 175,
409-418 (2000).
[21] R. W. P. Drever, J. L. Hall, F. V. Kowalski, J. Hough, G. M. Ford, A. J.
Munley, and H. Ward, Appl. Phys. B 31, 97-105 (1983).
[22] D. J. Jones, S. A. Diddams, J. K. Ranka, A. Stentz, R. S. Windeler, J. L.
Hall, and S. T. Cundiff, Science 288, 635-639 (2000).
[23] Handbook of Chemistry and Physics (The Chemical Rubber Co.,
Cleveland, OH, 1985/1986).
Appendix
AUTHOR ADDRESSES
Arissian, Ladan: Department of Physics and Astronomy, 800 Yale Blvd. NE,
University of New Mexico, Albuquerque, NM 87131
Hall, John L.: JILA, National Institute of Standards and Technology and
University of Colorado, University of Colorado, Boulder, CO 80309-
0440
Windeler, Robert S.: OFS Laboratories, 600 Mountain Avenue, Murray Hill,
NJ 07974