You are on page 1of 14

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/253017647

Future trends and applications of ultrafast laser technology - art. no. 61000H

Article  in  Proceedings of SPIE - The International Society for Optical Engineering · February 2006
DOI: 10.1117/12.661032

CITATIONS READS
2 651

7 authors, including:

Jason Eichenholz Ian Read


University of Central Florida Newport Corporation
52 PUBLICATIONS   670 CITATIONS    13 PUBLICATIONS   264 CITATIONS   

SEE PROFILE SEE PROFILE

Jim Kafka
Newport Corporation
43 PUBLICATIONS   857 CITATIONS   

SEE PROFILE

All content following this page was uploaded by Ian Read on 08 June 2016.

The user has requested enhancement of the downloaded file.


Invited Paper

Future Trends and Applications of Ultrafast Laser Technology

Jason M. Eichenholz*a,b, Mingwei Lib, Ian Readb, Stefan Marzenellb, Philippe Férub, Rich Boggyb
and Jim Kafkab
a
Newport Corporation, 1791 Deere Ave., Irvine, CA 92606;
b
Spectra-Physics – A Division of Newport Corporation, 1335 Terra Bella Avenue,
Mountain View, CA 94043
*Jason.eichenholz@newport.com; Phone (407) 380-7300; http://www.newport.com

ABSTRACT

In this talk we will present an overview of recent development of ultrafast lasers sources and their applications. This talk
will highlight some recent state of the art ultrafast pulse results from Ti:Sapphire and Ytterbium based laser systems.
There are significant advantages in being able to directly diode pump Ytterbium materials resulting in more compact
bulk solid state and fiber based laser systems. Several newly emerging technologies such as Optical Parametric Chirped
Pulse Amplification, and Supercontinuum Generation have generated great excitement in recent years. The evolution of
more compact and user friendly ultrafast laser systems has enabled completely new fields that take advantage of the
extremely high peak powers and very short time duration of ultrafast laser pulses. Recent results in the fields of
multiphoton microscopy, micromachining, 3-D fabrication, and spectroscopy will be discussed.

Keywords: Lasers, ultrafast lasers, fiber lasers, ytterbium lasers, supercontinuum, photonic crystal fibers, multiphoton
microscopy

1. ULTRAFAST LASER DEVELOPMENT

The recent advances in the generation and amplification of ultrafast laser pulses (<1 ps) have opened up many new
possibilities in laser–matter interaction, spectroscopy, bio-photonics, and nano-photonics. Ultrafast lasers have
undergone a revolution over the past 30 years as lasers pulses have become shorter and the laser systems have become
more reliable and much more user friendly1-3. Figure 1 below shows how the pulse durations have evolved over time
and how multiple gain media have been used. As ultrafast technology enters the mature stage and significant practical
applications are starting to emerge, the demand for practical turn-key femtosecond lasers has become one of the
dominant forces shaping the current technological development of ultrashort-pulse sources.

Ultrafast laser technology involves the use of femtosecond (10-15 s) laser pulse traveling at the speed of light. These
ultrafast lasers can also be used to produce laser pulses with enormous peak powers and power densities. This leads to
applications such as transient absorption spectroscopy, multiphoton imaging, laser micromachininig machining and
ablation, and generation of electromagnetic radiation at unusual wavelengths (such as supercontinuum, millimeter waves
and X-rays). The difficulty in applying femtosecond laser pulses is that the broad frequency spectrum can lead to
temporal broadening of the pulse on propagation through the experimental set-up.

A state of the art tabletop ultrafast laser system can generate more than 1 Joule of energy is less than 20 femtoseconds.
This huge amount of laser energy in such a short period of time that when focused down can generate peak fluences at
the focus that can exceed 1020W/cm2. In contrast, the total solar flux at the Earth is only 1017 W. An exciting era has
begun, which will allow the possibility of using these intense ultrashort laser pulses as sources of short X-ray and
electron pulses. These will reveal the positions of atoms as a function of time as reactants proceed to products through
the transition states.

In the past, these systems were quite large, occupying a full of several large optical tables. A typical higher end
amplified system produces 1mJ per pulse at a 1-kHz repetition rate or 100 mJ at 10–20Hz with a pulse duration between
20 and 100 fs. Moreover, a Ti:Sapphire laser oscillator and amplifier combination can be purchased in a single box less

Solid State Lasers XV: Technology and Devices, edited by Hanna J. Hoffman, Ramesh K. Shori,
Proc. of SPIE Vol. 6100, 61000H, (2006) · 0277-786X/06/$15 · doi: 10.1117/12.661032

Proc. of SPIE Vol. 6100 61000H-1


than 1m square, operating at 110 V with no external water-cooling requirements. Total hands-off operation is a reality
and, in fact, the complete laser system can be computer controlled.

1.1. Ultrafast oscillators

In the 1970’s, the earliest commercial ultrafast lasers required a significant amount of optical table space and were based
on messy dyes. These dye lasers were pumped by large Argon Ion lasers that required 400V 3 phase power and ~5
gallons per minute of cooling water. The dye gain media was messy and carcinogenic. The average powers from
commercial dye lasers at that time were in the 10’s of mW with pulse durations in the 10-20 picoseconds

The introduction of Ti3+-doped sapphire, Ti:Al2O3 had a dramatic impact on the development of ultrafast laser systems.
Laser action in Ti:Sapphire was first report by Moulton4 in 1982, cw laser operation5 (1986) and mode-locked laser
operation6 (1991) was demonstrated soon after. In 1991 the first commercial mode-locked Ti:sapphire laser was
available: the Spectra-Physics Tsunami7. At the time, this laser could be configured for femtosecond or picosecond
operation, but it required a manual adjustment of the wavelength and dispersion. When the Spectra-Physics Millennia
laser, a frequency doubled, diode-pumped Vanadate laser, was introduced in 1996, it opened up a new realm of ultrafast
lasers research by significantly reducing the size and utility requirements of an ultrafast laser system. The figure below
is modified from the Report on HED with Ultra-Short-Pulse Lasers II.

z0 Nd:glass
Nd
ode
=
D'\5P
C SOLID-STATE
LU
C,
REVOLUTION
-J Cw
=
0-
L0 BI
LU CE [is ( C )AF
CR1 Dye
I
0
CO
\t:hEYAG
w/Compressio Cr: to F ste rite
Ofs b

Ti:sapphire
I I

ifs 1970 1975 1980 1985 1990 1995 2000 2005


YEAR

Figure 1: Trends of ultrafast laser pulse duration as a function of time

Another revolution was the introduction of the first computer-controlled, tunable, one-box mode-locked Ti:sapphire
laser, the Spectra-Physics Mai Tai®, in 1999. The first model had a tuning range of 750nm to 850nm and a maximum
average output power of 1 W. Because operation of the laser was completely hands free and computer controlled. the
Mai Tai was revolutionary because it allowed non-laser experts to begin to use these systems for experiments in biology
and chemistry. Over the last 5 years, commercially available one-box titanium doped sapphire laser systems have made
huge improvements in performance and reliability, and have become excellent fully automated widely-tunable
femtosecond sources for a growing number of scientific and industrial applications.

The changes over time at an ultrafast laser laboratory at Cornell University can be shown in Figure 2. In each picture,
one can see how the laser system has benefited by advances of diode pumped solid state lasers and more reliable hands

Proc. of SPIE Vol. 6100 61000H-2


free operation. Since then, several new models that increase the average output power to more than 2.9W and the tuning
range from under 700nm to over 1020nm have become available.
Ti: Sapphire Lasers
1991: Argon pump 1997: Diode pumped 1998: Diode pumped 1999: Diode Pump and
and Ti:8 Nd-vanadate SW Nd-vanadate lOW Ti:S resonator in a
pump pump single case.

Figure 2: Changes in the use of Ti:Sapphire lasers oscillators in the 1990’s

1.2. Ultrafast amplifiers

Usually the ultrafast laser oscillator operates at a pulse repetition rate of about 80 MHz, and the pulse energy is in the
range of nanojoules. Amplification to the microjoule and millijoule level is needed for many applications, including
nanosurgery and micromachining. However, the pulses cannot be amplified directly in an additional amplification step,
as the high peak power of the short pulses inside the amplifier will damage the optical components. Self-focusing of
high-intensity pulses makes the design of systems even more complicated. To overcome this limitation, the technique of
chirped-pulse amplification (CPA) was developed by Strickland and Mourou8 in 1985. In CPA, the pulsewidth of the
oscillator is stretched by a factor of 1,000–10,000 times in a pulse stretcher to reduce the peak power. The stretched
pulse is then amplified in the amplifier. The gain medium in the amplifier must also support the bandwidth of the short
pulses. Finally, the amplified long pulse is recompressed back to its short pulse state in the compressor. Normally the
pulse stretcher and compressor are made of a pair of optical gratings. The dispersive properties of the gratings are
exploited to change the pulsewidth of an ultrafast pulse having a broad-band spectrum. Starting from a short pulse,
where all frequency components that make up the pulse travel together, different frequency components are dispersed
into different directions, hence have different optical path lengths. The frequency component that has a shorter optical
path length will come out earlier than the frequency component that travels a longer path, hence the pulse is stretched
and vice-versa compressed.

2. NEW MATERIALS AND TECHNIQUES

2.1. Bulk ytterbium materials

As discussed above, for the past decade or more, most ultrafast laser sources have been based on Ti:Sapphire, with its
emission centered at 800 nm. Because the gain material’s absorption falls within the visible wavelength range, they must
be pumped by the green output of intracavity frequency-doubled solid-state lasers. This complexity is even more
noticeable in high-energy systems, where the output pulses from a mode-locked Ti:Sapphire oscillator are used to seed a
Ti:sapphire regenerative amplifier. Both the oscillator and the amplifier need separate green pump lasers. In recent years,
tremendous advances in system design and engineering have enabled all these lasers to be packaged in a single laser
head, nonetheless, the complexity of these systems can still be daunting and limiting for those who are not laser experts.

Proc. of SPIE Vol. 6100 61000H-3


Fortunately, this situation has been improved, thanks to ytterbium doped materials such as ytterbium tungstate
(Yb:KGW) and the high-power femtosecond lasers based on this new material.

Ytterbium is a lanthanide element similar to erbium, and like erbium it has a broad emission spectrum when doped into
various glasses and crystals. Of the various ytterbium containing crystals, ytterbium tungstate (Yb:KGW)9 is a
particularly attractive laser material because it combines a high cross-section for stimulated emission with a broad
emission spectrum, from ~ 1020 to ~ 1060 nm. Such a broad emission spectrum is essential for mode-locking with pulse
durations in the femtosecond regime. In addition, the absorption spectrum of ytterbium tungstate (Fig. 3) shows that it
can be efficiently pumped with laser diodes at 940 or 980 nm, wavelengths where high power laser diodes are readily
available with proven reliability.

Figure 3: Normalized spectral absorption cross section of Yb:KGW for the three crystal axis

Although several low power lasers based on Yb:KGW have previously been developed, high power amplifiers are only
now becoming available. One main challenge has been the thermal management of this new material. Intense pumping
of most laser materials can produce significant thermal gradients and as a result, thermal stress, which potentially can
lead to fracture and failure. We have been able to eliminate this thermal damage issue with a multi-pronged approach,
allowing the full potential of Yb:KGW to be accessed. The various aspects of their approach include: selection of
defect-free Yb:KGW crystals with the optimum doping concentration, fabrication of high quality rods from these
crystals, thorough investigation of the thermal properties of Yb:KGW, selection of the optimum rod orientation, design
of a proprietary crystal mounting and cooling scheme, and design of a proprietary optical coating that eliminates surface
damage.

The new Spectra-Physics Eclipse® regenerative amplifier10 consists of a laser head, and a compact rack-mounted power
supply and control unit (Fig. 4). The system also includes a closed-loop chiller and does not require external cooling
water. There are two self-contained active building blocks inside the laser head – an oscillator and a regenerative
amplifier – as well as a module for stretching and compressing the pulses, before and after amplification utilizing
diffraction gratings. The mode-locked oscillator is a factory sealed unit with a single rod of Yb:KGW. It is end-pumped
(at both ends) by the output from two fiber-coupled single-emitter laser diodes that are located in the control unit and
operated at de-rated power levels. Despite lifetimes in excess of 10,000 hours, the laser diodes are designed to be field-
replaceable without requiring any optical alignment. Utilizing standard Chirped Pulse Amplification (CPA) the output
pulses from the sealed oscillator are stretched in a single grating pulse stretcher, before entering the sealed amplifier
module. The gain in the amplifier module cavity is supplied by a single rod of Yb:KGW, pumped at both ends by the
output of two 40 W diode bars, which are each de-rated to less than 30 W. After the number of round trips around the
amplifier cavity required to achieve gain saturation, a high-speed switcher deflects the amplified pulse out of the cavity.
This amplified pulse is then recompressed in time by a single grating pulse compressor. The end result is an amplified
output pulse train in a TEM00 beam with an average power of up to 3 watts at repetition rates of up to 5 kHz. The pulse
width is less than 500 femtoseconds, and the contrast ratio exceeds 1,000:1. This output can be frequency doubled by an
optional integrated second harmonic generator, which produces green output at 524 nm with an average power of up to
1.5 watts.

Proc. of SPIE Vol. 6100 61000H-4


r
V

r
r
r

Figure 4: The Eclipse Yb:KGW femtosecond amplifier includes a) a sealed oscillator module, b) a stretcher/compressor, c) an
amplifier module, d) an optional second-harmonic generator, e) a safety shutter and f) a high-speed switcher electronics rail. The
pump diodes are in the control module and fiber-coupled to the oscillator and amplifier modules.

2.2. Optical parametric chirped pulse amplification (OPCPA)

Optical parametric chirped pulse amplification (OPCPA)11,12 is a rapidly developing technology that is ideally suited
for achieving high peak and average powers. The OPCPA principle has been proposed and demonstrated for the first
time by Dubietis et al.11 in 1992. OPCPA is a nonlinear optical process in which the photon energy and momentum of a
pump beam are converted into two photons, resulting in the generation of two beams denoted as “signal” and “idler.”
The amplified signal beam retains the spectral phase and spatial wavefront characteristics of the seed pulse, which means
that the high-energy pulse can be recompressed and focused with high fidelity. Extremely high gains can be achieved in
monolithic, single-pass configurations involving minimal optical components. Since this nonlinear optical process
deposits virtually no heat in the gain medium, thermal management is decoupled from the ultrafast amplification.
OPCPA also offers tunability since broadband gain can be achieved across a wide range of wavelengths. Extensive
development of the OPCPA technique has been continued by Ross et al.12

3. FIBER LASERS

Optical fibers have been viewed as a very attractive medium for the generation and manipulation of ultrafast pulses for a
long time. Ultrafast fiber lasers offer several advantages over bulk solid-state lasers, including greater stability and
freedom from misalignment, compact design, and relative ease of construction with inexpensive components. Fiber
lasers provide enormous gains (small-signal gains as high as 105) and low threshold operation. The high single pass gain
is a key factor allowing for efficient diode pumped operation with the fiber set up as either oscillator or amplifier.

Fibers owe their excellent heat dissipation properties to their high surface-to-volume ratios which facilitates heat
removal, allowing for room temperature operation without complex chillers. One consequence of the outstanding
thermo-optic properties of the actively doped fiber is that the beam quality is determined by the effective refractive index
profile of the active core, independent of the launched pump power. This is a key difference between fiber and bulk
lasers since thermal lensing and beam distortions due to heat accumulation in the gain medium represent the principal
limitations to power scaling in bulk laser devices.

Proc. of SPIE Vol. 6100 61000H-5


Currently, most high power fiber sources based on step index fibers generally incorporate a double clad structure, in
which the rare-earth doped core is surrounded by much larger and higher NA undoped inner cladding of slightly lower
refractive index. The inner cladding is, in turn, encased in an outer cladding of yet-lower index of refraction. Typically
both core and the inner cladding are composed of glass, with the glass composition of the clad selected so that its index
is slightly lower than that of the core. A dual clad Photonic Crystal Fiber typically has a higher NA in the inner cladding
due by substituting the outer low index material with air.

The performance advantages of ultrafast fiber laser are due to the excellent heat dissipation characteristics of the long
and thin gain medium, along with the high efficiencies (often more than 80%). Double-clad fiber design with CW
powers in excess of 1 kW in a single spatial-mode beam has been reported for diode pumped systems.13,14

Despite the rapid growth of high power CW laser systems, fiber lasers have lagged behind solid-state lasers with regard
to pulse energy, duration, and quality. Recent progress in ultrafast fiber lasers has primarily focused on increasing pulse
energy and shorter pulse durations.

3.1. Fiber oscillators


There has been much activity in the development of ultrafast fiber lasers over the past few years utilizing passive and
active–passive mode locking techniques as well as linear and nonlinear fiber amplifiers.15-18 Fiber lasers can be mode-
locked actively or passively. In active mode-locking, a modulator produces amplitude or phase modulation. In passive
mode locking, an intensity fluctuation acts in conjunction with the fiber nonlinearity to modulate the cavity loss without
external control. The upper-state lifetimes of rare-earth-doped fibers are typically long (~ ms) implying that the gain
does not react significantly within the cavity round-trip time. A fast saturable absorber is therefore required to clean up
both the leading and trailing edges of the pulse. Rare-earth-doped fiber lasers are also susceptible to Q-switching due to
their long upper-state lifetimes.16

A Yb fiber laser operating in the self-similar mode recently produced greater than 10-nJ pulses at ~100-fs duration,
directly from a femtosecond fiber oscillator17. In applications involving time-resolution, high-quality, short pulses are
important. A Yb fiber laser operating in the stretched-pulse mode has produced 36-fs pulses; however, the pulse quality
was low, with significant energy extending out to 150 fs18. The primary limitation to achieving shorter pulses was
attributed to uncompensated third-order dispersion (TOD) inside the laser cavity. Wise et al.19 recently demonstrated the
use of a prism-grating sequence to reduce third-order dispersion inside a Yb fiber oscillator (Figure 5). Pulses as short as
33-fs, the shortest from a fiber laser, were generated with extremely clean profiles.

600 g/mm
Ml

/SF
HWP QWP
.L1 M2
600 g/mm\i0Th\

975 nm
Figure 5 Experimental setup for generation of 10-cycle pulses from a Yb fiber laser using cubic phase compensation

Proc. of SPIE Vol. 6100 61000H-6


3.2. Fiber amplifiers

The ultimate performance of ultrafast fiber systems will always be limited by pulse distortions that are due to the onset of
nonlinear effects, such as excessive self-phase modulation or Raman scattering, that arise as a result of the tight spatial
modal confinement of the pulses within the fiber gain medium. Nonlinear evolution, however, need not always result in
pulse degradation. The combined interaction of normal dispersion, gain, and nonlinearity (self-phase modulation) can
create linearly chirped parabolic pulses. It has been shown that high-power pulses in a fiber amplifier with normal
dispersion evolve toward a self-similarly propagating pulse with a parabolic intensity profile20. As the parabolic pulses
propagate, they expand both temporally and spectrally while maintaining a linear chirp across the central region of the
pulse. This permits recompression of the output pulses to short durations despite significant self-phase modulation in the
fiber. Parabolic pulses from Yb fiber amplifiers were previously demonstrated with average powers of up to 17 W (pulse
energy 230 nJ) utilizing a bulk glass seed laser21 and up to 13 W (pulse energy 260 nJ) by use of a fiber-based seed
laser.22 In 2004 Malinowski et al. reported both a higher average power of more than 25 W and higher pulse energy of
410 nJ from an all Yb fiber laser system23. A conventional bulk-grating compressor was used to remove the linear chirp,
resulting in 100-fs pulses.

The use of the chirped-pulse amplification technique, in which pulses are stretched to durations of the order of 1 ns
before amplification to reduce peak power and thus avoid nonlinear effects, are also effective in producing high average
powers24,25. The opposite approach is sufficient pulse stretching in the time domain and enlargement of the mode-field
diameter of the fiber to reduce the peak power and therefore nonlinear effects such as stimulated Raman scattering and
self-phase modulation. This approach leads to rare earth doped fiber based chirped-pulse-amplification systems
employing low-NA large-mode-area (LMA) fibers. Recently a Yb doped photonic crystal fiber based chirped-pulse
amplification system produced 131 W of average power 220 fs pulses at 1040 nm center wavelength in a diffraction-
limited beam25. The system displayed in Figure 6 had a pulse repetition rate of 73 MHz, corresponding to a pulse energy
of 1.8 µJ and a peak power as high as 8.2 MW.

Higher femtosecond pulse energies of up to 100 µJ from fiber lasers were recently demonstrated by Galvanauskus et al.
26
. However many of these higher energy systems utilize >1000:1 pulse stretching ratios were produced using stretchers
and compressors based on bulk optical gratings with matched dispersion compensation. The size and alignment
tolerances associated with bulk gratings, make them impractical for use in industrial laser systems. Thus, in order to
realize the principal integration advantage of amplified ultrafast fiber lasers, fiber-based alternatives are required to
replace bulk gratings.

150 fs, 73 MHz


1040 nm, 8 nm
100 mW Transmission grating compressor
diode laser diode laser
Yb:KGW
oscillator
5W 175 W

OI
Λ=833 nm Yb-doped Yb-doped Λ=800 nm, 96.5% Eff.
preamplifier power amplifier => 75% in 4pass
output

grating stretcher
120 ps
50 mW
Figure 6: Photonic crystal fiber based CPA system producing 131 W of ultrafast laser power

While both fiber based stretchers and compressors can be implemented, the current state of technology limits the energy
handling of fiber based compressors. Thus bulk grating compressors are used for the generation of the highest pulse
energies. The combination of fiber stretcher and bulk grating compressor results in a significant third-order dispersion

Proc. of SPIE Vol. 6100 61000H-7


mismatch for large stretching ratios. Shah et al27, demonstrated for the first time the generation of cubicons at multi-watt
level average powers and pulse energies of 100 µJ in an Yb fiber based CPA system. In that paper they demonstrated that
this third-order dispersion mismatch can be compensated by nonlinear self-phase modulation in the amplifier fiber
through the use of cubicons. Cubicons reflects that the pulse evolution in the nonlinear power amplifier is related to that
of similaritons, however whereas similaritons (or parabolic pulses20) induce a quadratic nonlinear phase delay, cubicons
also induce a significant component of cubical nonlinear phase delay. Unlike similaritons, cubicons are strongly chirped
pulses with asymmetric temporal and spectral distribution for which SPM leads to a negative dispersion slope during
amplification, thus cubicons allow for compensation of large values of third order dispersion mismatch between pulse
stretcher and compressor.

4. ULTRAFAST APPLICATIONS

4.1. Supercontinuum generation in photonic crystal fibers


Photonic crystal fibers (PCFs) were first demonstrated in 199628 and have generated much attention since then. PCFs,
also called micro structured fibers, are a newly developed type of specialty fiber optic with a highly structured cross-
section of glass and air-space running along the length of the fibers. Although PCFs have only recently become
commercially available, there are already many different types, supporting a fast growing range of applications29,30.
Formation of broad continuous spectra through propagation of short femto- or picosecond-range high power pulses
through nonlinear media, also known as super continuum generation, (SCG) was first observed in 1970 and has since
then been studied extensively in many different materials. The term supercontinuum does not cover a specific
phenomenon but rather a plethora of nonlinear effect leading to considerable spectral broadening of optical pulses and
thereby potentially octave-spanning output. The nonlinearities depend on the dispersion in the material and nonlinear
effects such as self-phase modulation (SPM), Raman scattering, phase matching and soliton formation. Results on SCG
in PCFs have previously been presented with pumping in the anomalous dispersion regime or at the zero-dispersion
wavelength in both the visible and the infrared wavelength range.
Although SCG can be observed in glass or water given enough pumping power, PCFs are ideal media for SCG as the
dispersion can be designed to facilitate continuum generation in a specific region. Recently, numerous investigations on
supercontinuum (SC) generation in highly non-linear photonic crystal fibers have been published with pump pulses
durations from nanoseconds to femtosecond31-35.
Highly non-linear fibers are constructed by producing a very small core diameter (down to 1 µm), in combination with a
high index ratio between core and cladding. This is often accomplished using a construction consisting of a honeycomb
array of large air spaces surrounding a small central core. This extremely small and well-confined mode volume
confines a high power density along the fiber length producing the plethora of non-linear effects. The large air-filled
cladding in this PCF also produces large waveguide dispersion. This waveguide dispersion can be tailored to either
enhance or cancel out the inherent dispersion of the fiber material, providing for more flexibility in terms of overall
dispersion profile than is achievable in step-index fibers. This, in turn, can be used to enhance or reduce various non-
linear processes. Most experiments today utilize femtosecond pumping as this results in spectacularly broad spectra. The
flatness and width of the generated continua depends highly on the pump parameters such as pump pulse duration,
wavelength, polarization and the fiber characteristics, in particular the fiber zero dispersion point and slope. Much of the
recent attention in SC is a direct result of the unique properties of the SC in PCFs, such as low energies and
comparatively large input-pulse durations, its coherence, high spatial brightness as well as effectively single-mode
operation. Due to these properties, SCG in PCFs is very interesting for applications such as femtosecond carrier envelope
phase stabilization36, frequency metrology37, optical coherence tomography38, as a novel light source for spectroscopy.

4.2. Multiphoton microscopy


Multiphoton microscopy (MPM)39 has revolutionized life sciences by enabling long-term imaging of living preparations
in highly scattering tissue while minimizing photodamage. MPM is exceptionally well suited for fluorescence imaging
several hundred microns deep inside intact biological tissue. Multiphoton imaging systems typically use laser raster scanning
to assemble an image. The underlying principle of MPM is that at very high-photon densities, an excitable molecule may
simultaneously absorb two or more lower energy photons (near IR) with the same effect as the absorption of a single photon

Proc. of SPIE Vol. 6100 61000H-8


with an energy equal to the sum of the individual lower photon energies. In the case of two-photon imaging, the excitation
wavelength is set to about twice that of the one-photon absorption peak of the fluorophore being observed, which would not
normally produce any appreciable fluorophore excitation. However, if a high peak power, ultrafast laser is used, it is possible to
achieve instantaneous photon densities that give rise to a significant yield of two-photon events at the focal volume of an
objective lens, with a mean power level that will not produce optical trapping or damage to a specimen.

Confocal microscopes typically require an optical element such as a pinhole to eliminate out-of-focus noise and interference.
MPM systems have simplified optical systems, as pinhole apertures do not have to be used because the fluorophore
excitation is confined to the focal volume where the photon density is highest. This is because the photon density is
significantly lower away from the focal region. This is illustrated in Figure 7 where a fluorescent dye is excited using CW 488
nm light on the left and 900 nm femtosecond light on the right. The figure clearly shows how the multiphoton excitation is
confined to the focal volume where the photon density is highest. Optical sectioning is achieved because there is no appreciable
fluorophore excitation above or below the plane of focus. MPM is particularly effective in obtaining images of optical
sections from deep within a specimen as the longer wavelength near-infrared photons experience less scattering of the
excitation photons compared to the UV and blue wavelength light (such as from an argon ion laser) that is typically used
for fluorophore excitation in biological microscopy.

1-photon 2-photon
488 nm 900 nm

Figure 7 Fluorescent dye excitation using 488 nm cw and 900 nm femtosecond laser irradiation

4.3. Micro machining


Femtosecond laser micromachining is a rapidly advancing area of ultrashort laser applications. It utilizes the ultrashort
laser pulse properties to achieve an unprecedented degree of control in sculpting the desired microstructures internal to
the materials without collateral damage to the surroundings. Using femtosecond rather than picosecond or nanosecond
light pulses, laser energy is deposited into small volumes by multiphoton nonlinear optical absorption followed by
avalanche ionization. The reason for this is that typical heat diffusion time is on the order of nanosecond to microsecond
time scales whereas the electron-phonon coupling time of most materials is in picosecond to nanosecond time scale.
Therefore when laser energy is deposited at a time scale much shorter than both the heat transport and the electron-
phonon coupling, the light-matter interaction process is essentially frozen in time. The affected zone altered from solid to
vapor phase and to plasma formation almost instantaneously. Unlike conventional laser machining, femtosecond laser
machining reduces collateral damages to the surroundings. Because the machining process is not dependent on the linear

Proc. of SPIE Vol. 6100 61000H-9


absorption at the laser wavelength, virtually any dielectric, metals, and mechanically hard materials can be machined by
the same laser beam. However all the benefits of ultrafast micromachining are not limited to metals and dielectrics, if
fact they are also very useful for biological materials.
For example Shaffer40 et al. recently developed techniques to precisely image and creates blood clots in the brain in real
time using multiphoton microscopy and femtosecond micromaching. This research could help to understand the
mechanisms behind small strokes implicated in many forms of dementia, including Alzheimer’s disease. The
researchers used a multi photon microscope to image vascular architecture, to select a vessel for injury and to measure
blood-flow dynamics as shown in Figure 8 below. Once an area was characterized, an amplified femtosecond laser was
used to trigger the formation of individual blood clots via photodisruption in tiny arteries of the brains of anesthetized
rats near the focus of the beam. The photodisruption causes clotting in the area of the focus and some leakage of blood
from the vessel.

lOOts
y-11J

optical breakdown shock wave, vessel damage


in vessel lumen cavitation bubble triggers clotting
Figure 8 Laser induced vessel lumen damage and clotting

After the vascular structures were irradiated, the researchers were then able to monitor the resulting changes in blood
flow using multi-photon microscopy image stacks before and after induction of the clot in a deep microvessel as shown
in Figure 9. The clotting procedure is accompanied by leakage from the vessel, so that the target vessel can be seen as the
extra bright vessel in the post-clot images.

baseline post-clot

top-view
side-view 400 jim down
Figure 9 - Multiphoton microscopy images of clotting of vascular structures

4.4. Spectroscopy
Recently a tremendous expansion of the research activity in the field of femtosecond laser spectroscopy has occurred.
Many research groups around the world operating in different fields from physics and chemistry to biology, medicine
and material science, utilize laser setups capable of providing subpicosecond pulses. Femtochemistry is one of the
revolutionary ideas in the field of chemistry and A. H. Zewail has been recently awarded with the Nobel Prize for his
work41. Femtosecond or ultrafast phenomena have to do with the chemical reactions (breakage of the chemical bonds,
molecule’s orientation, energy transfer, etc.) as well as the behavior of the electrons and nuclei in the molecules. For the

Proc. of SPIE Vol. 6100 61000H-10


study of such phenomena several techniques based on the femtosecond laser pulses have been developed. Ultrafast laser
pulses can now be used to observe and even control the outcome of reactions in real time.

Femtosecond laser systems have opened the door to studying ultrafast dynamical processes, which range from nuclear
motion in molecules to relaxation mechanisms of charge carriers in solids. A characteristic of the approach is that the
system investigated is no longer in thermodynamic equilibrium. It is rather in an excited state whose decay into other
degrees of freedom is being probed. The extremely short pulse durations allows one to create, detect and study very
short-lived transient chemical reaction intermediates and transition states. Most electronic devices cannot measure
transients much faster than about 1 ns. Although there are specialized electronic devices such as streak cameras that can
resolve picosecond or even hundreds of femtosecond transients in real time, these systems have limitations and
significant disadvantages. The two of the most widely used methods are the femtosecond fluorescence upconversion42
and the pump-probe femtosecond transient absorption 43.

Significantly more information is available when the probe pulse is a white light continuum such as that discussed above
in Section 4.1 and a spectrograph is used to measure the spectral distribution of the beam transmitted by the sample.
Under these conditions one obtains an absorption spectrum of transient species formed, or ground state species removed
by the initiating pulse. This spectrum will be registered at a time that is given by the difference in arrival times at the
sample of the pump and probe pulses and it will be an average over the temporal width of the probe pulse. An
alternative method of obtaining spectral information is to employ a probe beam of a selected, but changeable
wavelength. Such a beam can be isolated from a continuum by one of a set of interference filters; or it can be generated
by an optical parametric amplifier which can be scanned to provide spectral coverage.

4.5. 3-D lithographic microfabrication

3D lithographic microfabrication (3DLM) technology is a promising new approach involving the use of two photon
absorbing dyes to activate chemical or physical processes with high spatial resolution in three dimensions. Like multi
photon microscopy, under tight-focusing conditions, absorption is confined to a 3D pixel sizes (voxel) as small as 0.04
µm3 and limited only by the laser wavelength. Using two photon absorbing materials, it is also possible to localize
chemical reactions, including polymerization, within this excitation volume. Consequently, by computer-controlled
scanning of a laser-beam focus within a photo-chemically active precursor material, virtually any three-dimensional
structure can be fabricated with submicron resolution. Specifically, by employing photo-polymerizable systems that
become insoluble via photocrosslinking upon two-photon excitation, 3DLM technology has been demonstrated with sub-
200 nm resolution. This achievement was made possible by recent breakthroughs in the design of two-photon absorbing
dyes. Perry et al.44-47 reported the development of molecules with exceptionally large two-photon absorptivities, yielding
chromophores that advance the sensitivity 20–50 times that of state-of-the-art photopolymers. Such highly-sensitive two-
photon absorbing molecules provide a means of efficiently activating chemical or physical processes with laser energies
well below the damage thresholds of most patterning materials. This technology has subsequently been used to create
high resolution three dimensional structures, and in the fabrication of channel structures for microfluidic applications as
seen in Figures 10a and 10b respectively.

L:.

xl.Ok 0003 25kV SOum


xZQQ 0007
V 200izu1
(a) (b)
Figure 10 – 3-D nanofabricated devices made with two photon polyermization (a) chain, (b) microfluidic device

Proc. of SPIE Vol. 6100 61000H-11


5. SUMMARY

There has been an explosion of activity in ultrafast lasers over the past few decades. Today’s ultrafast laser sources are
more efficient and user friendly than ever. Completely hands free, computer controlled turn key systems are now
commercially available from multiple suppliers. Due to this new simplicity and reliability, ultrafast lasers systems are
coming out of the pure physics research laboratory and into the biology and chemistry labs and even the industrial
factory floor. These new systems have enabled completely new fields such as multiphoton microscopy, spectroscopy,
micromachining, and 3-D microfabrication that utilize the extremely high peak powers and very short time duration of
ultrafast laser pulses. The recent development of Ytterbium doped bulk crystals and fiber lasers are expected to
accelerate the development of more compact, efficient, and powerful ultrafast laser systems, ad therefore new
applications

REFERENCES

1. G. Mourou “The ultrahigh-peak-power laser: present and future” Appl. Phys. B 65, 205 (1997).
2. S. Backus, C.G. Durfee III, M.M. Murnane, H.C. Kapteyn “High power ultrafast Lasers” Rev. Sci. Instrum. 69,
1207 (1998).
3. U. Keller, “Recent developments in compact ultrafast lasers” Nature 424, 831 (2003).
4. P. Moulton, “Ti-doped sapphire: a tunable solid-state laser”, Opt. News 8, 9 (1982).
5. P. Albers, E. Stark, and G. Huber, “Continuous-wave laser operation and quantum efficiency of titanium-doped
sapphire”, J. Opt. Soc. Am. B 3(1), 134 (1986).
6. D. E. Spence, P. N. Kean, and W. Sibbett, “60-fsec pulse generation from a self-mode-locked Ti:sapphire laser”,
Opt. Lett. 16(1), 42 (1991).
7. J. D. Kafka, M. L. Watts, and J.-W. J. Pieterse, “Picosecond and Femtosecond Pulse Generation in a Regeneratively
Mode-Locked Ti:Sapphire Laser”, IEEE J. of Quantum Elect. 28 2151 (1992).
8. D. Strickland and G. Mourou, “Compression of amplified chirped optical pulses,” Opt. Comm., 56, 219 (1985).
9. F. Brunner, G. J. Sphler, J. Aus der Au, L. Krainer, F.Morier-Genoud, R. Paschotta, N.Lichtenstein, S.Weiss,
C.Harder, A.Lagadsky, A.Abdolvand, N.V.Kuleshov, U.Keller, “Diode-pumped femtosecond YbKGd(WO4)2 laser
with 1.1W average power”, Opt. Lett. 25, 1119 (2000).
10. A. Kruger, P. Feru, “Ytterbium Tungstate Revolutionizes the Field of High-Power Ultrafast Lasers” Photonics
Spectra, April 2004
11. A. Dubietis, G. Jonusauskas, and A. Piskarskas, “Powerful femtosecond pulse generation by chirped and stretched
pulse parametric amplification in BBO crystal,” Opt. Comm., 88, 437 (1992).
12. I.N. Ross, P. Matousek, M. Towrie, A.J. Langley, J.L. Collier, “The prospects for ultrashort pulse duration and
ultrahigh intensity using optical parametric chirped pulse amplifiers”, Opt. Comm. 144, 125 (1997).
13. Y. Jeong, J. K. Sahu, D. N. Payne, and J. Nilsson, “Ytterbium-doped large-core fiber laser with 1.36 kW continuous-
wave output power” Opt. Express 12, 6088 (2004).
14. A. Tünnermann, T. Schreiber, F. Röser, A. Liem, S. Höfer, H. Zellmer, S. Nolte, J. Limpert, “The renaissance and
bright future of fibre lasers,” J. Phys. B: At. Mol. Opt. Phys. 38, 681 (2005).
15. M.E. Fermann, A. Galvanauskas, G. Sucha, and D. Harter, “Fiber-lasers for ultrafast optics” Appl. Phys. B 65, 259
(1997).
16. L.E. Nelson, D.J. Jones, K. Tamura, H.A. Haus, and E.P. Ippen, “Ultrashort-pulse fiber ring lasers” Appl. Phys. B
65, 277 (1997).
17. J.R. Buckley, F. Ö. Ilday, T. Sosnowski and F.W. Wise, “Femtosecond fiber lasers with pulse energies above 10 nJ,”
Opt. Lett, 30, 1888 (2005).
18. F.Ö. Ilday, J. Buckley, L. Kuznetsova and F.W. Wise, “Generation of 36-fs pulses from a ytterbium fiber laser,”
Opt. Express 11, 3553 (2003).
19. J. R. Buckley, S.W. Clark, F. W. Wise, “Generation of 10-cycle pulses from a Yb fiber laser using cubic phase
compensation” In OSA Proceeding of Advanced Solid State Photonics (ASSP) 2006, paper ME4
20. M. E. Fermann, V. I. Kruglov, B. C. Thomsen, J. M. Dudley, and J. D. Harvey, Phys. Rev. Lett. 84, 6010 (2000).
21. J. Limpert, T. Schreiber, T. Clausnitzer, K. Zollner, H. J. Fuchs, E. B. Kley, H. Zellmer, and A. Tünnermann, Opt.
Express, 10, 628 (2002).
22. A. Galvanauskas and M. E. Fermann, in Conference on Lasers and Electro-Optics (CLEO), Vol. 39 of OSA Trends
in Optics and Photonics Series (Optical Society of America, Washington, D.C., 2000), post-deadline paper PD3.

Proc. of SPIE Vol. 6100 61000H-12


23. A. Malinowski, A. Piper, J. H. V. Price, K. Furusawa, Y. Jeong, J. Nilsson, and D. J. Richardson, “Ultrashort-pulse
Yb3+-fiber-based laser and amplifier system producing >25-W average power”, Opt. Lett. 29, 2073 (2004).
24. J. Limpert, T. Clausnitzer, A. Liem, T. Schreiber, H. J. Fuchs, H. Zellmer, E. B. Kley, and A. Tunnermann, Opt.
Lett. 28, 1984 (2003).
25. F. Röser, J. Rothhard, B. Ortac, A. Liem, O. Schmidt, T. Schreiber, J. Limpert, and A. Tünnermann, “131 W 220 fs
fiber laser system”, Opt. Lett. 30, 2754 (2005).
26. A. Galvanauskas, G.C Cho, A. Hariharan, M.E Fermann, and D. Harter, “Generation of high-energy femtosecond
pulses in multimode-core Yb-fiber chirped-pulse amplification systems,” Opt. Lett. 26, 935 (2001).
27. L. Shah, Z. Liu, I. Hartl, G. Imeshev, G. C. Cho and M. E. Fermann, “High energy femtosecond Yb cubicon fiber
amplifier”, Optics Express, 13, 4717 (2005).
28. J. Knight, T. Birks, P. Russell, and D. Atkin, “All-silica single-mode optical fiber with photonic crystal cladding”,
Opt. Lett. 21, 1547 (1996)
29. P. Russell, “Photonic Crystal Fibers”, Science, 299 358 (2003).
30. Photonics Crystal Fibres, A. Bjarklev, J. Broeng, and A.S. Bjarklev, Kluwer Academic Publishers (2003).
31. J.K. Ranka, R.S. Windeler, and A.J. Stentz, "Optical properties of high-delta air-silica microstructure optical fibers,"
Opt. Lett. 25, 796 (2000).
32. I. Hartl, X.D. Li, C. Chudoba, R.K. Ghanta, T.H. Ko, J.G. Fujimoto, J.K. Ranka, and R.S. Windeler, “Ultrahigh-
resolution optical coherence tomography using continuum generation in an air silica microstructure optical fiber”
Opt. Lett. 26, 608 (2001).
33. R. Holzwarth,M. Zimmermann, Th. Udem,T.W. Hansch, P. Russbuldt, K. Gabel, R. Poprawe, J.C. Knight, W.J.
Wadsworth, and P.St.J. Russell, “White-light frequency comb generation with a diode-pumped CrLiSAF laser “
Opt. Lett. 26, 1376 (2001).
34. St. Coen, A.H. Lun Chau, R. Leonhardt, J. D. Harvey, J. C. Knight, W. J. Wadsworth, and P. St. J. Russell, “White-
light supercontinuum generation with 60-ps pump pulses in a photonic crystal fiber”Opt. Lett. 26, 1356 (2001).
35. P. A. Champert, S. V. Popov, J. R. Taylor, “Generation of multiwatt, broadband continua in holey fibers” Opt. Lett.
27, 122 (2002).
36. R. Holzwarth, Th. Udem, T.W. Hänsch, J.C. Knight, W.J. Wadsworth, and P.St.J. Russell, “Optical Frequency
Synthesizer for Precision Spectroscopy” Phys. Rev. Lett. 85, 2264 (2000)
37. D.A. Jones, S.A. Diddams, J.K. Ranka, A. Stentz, R.S. Windeler, J.L. Hall, and S.T. Cundiff, “Carrier-envelope
phase control of femtosecond mode-locked lasers and direct optical frequency synthesis”, Science 288, 635 (2000).
38. N. Nishizawa, Y. Chen, P. Hsiung, E. P. Ippen, and J. G. Fujimoto, “Real-time, ultrahigh-resolution, optical
coherence tomography with an all-fiber, femtosecond fiber laser continuum at 1.5 mm”, Opt. Lett. 29, 2846 (2004)
39. W. Denk, J. H. Strickler, and W. W. Webb, "Two-photon laser scanning fluorescence microscopy," Science 248, 73
(1990).
40. N. Nishimura, C. B. Schaffer, B. Friedman, P. S. Tsai, P.D. Lyden, and D. Kleinfeld, “Targeted insult to subsurface
cortical blood vessels using ultrashort laser pulses: three models of stroke”, TBP Nature Methods February 2005
41. http://nobelprize.org/chemistry/laureates/1999/
42. J. T. M. Kennis, B. Gobets, I. H. M. van Stokkum, J.P. Dekker, R. van Grondelle, and G. R. Fleming, “Light
harvesting by chlorophylls and carotenoids in the photosystem I core complex of Synechococcus elongatus: a
fluorescence upconversion study” J. Phys. Chem. B 105, 4485 (2001).
43. K. Wynne, R. M. Hochstrasser “The Theory of Ultrafast Vibrational Spectroscopy” Chem. Phys. 193, 211 (1995).
44. M. Albota, D. Beljonne, J.-L. Brédas, J. E. Ehrlich, J.-Y Fu, A. A. Heikal, S. E. Hess, T. Kogej, M. D. Levin, S. R.
Marder, D. McCord-Maughon, J. W. Perry, H. Röckel, M. Rumi, G. Subramaniam, W. W. Webb, C. Xu, and X.-L.
Wu, Science 281, 1653 (1998).
45. B. H. Cumpston, S. P. Ananthavel, S. Barlow, D. L. Dyer, J. E. Ehrlich, L. L. Erskine, A. A. Heikal, S. M. Kuebler,
I-Y. S. Lee, D. McCord-Maughon, J. Qin, H. R. Röckel, M. Rumi, X.-L.Wu, S. R. Marder, and J. W. Perry, “Two-
photon polymerization initiators for three-dimensional optical data storage and microfabrication”, Nature 398, 51
(1999).
46. S. M. Kuebler, K. L. Braun, W. Zhou, J. K. Cammack, T. Yu, C. K. Ober, S. R. Marder, and J. W. Perry, “Design
and application of high-sensitivity two-photon initiators for three-dimensional microfabrication” J. Photochem. and
Photobio. A,: Chemistry, 158, 163 (2003).
47. W. Zhou, S. M. Kuebler, K. L. Braun, T. Yu, J. K. Cammack, C. K. Ober, J. W. Perry, and S. R. Marder, “An
efficient two-photon-generated photoacid applied to positive-tone 3D microfabrication”, Science 296, 1106 (2002).

Proc. of SPIE Vol. 6100 61000H-13

View publication stats

You might also like