Ultrahigh-Intensity Lasers: Nonlinear Optics in the Relativistic Regime

in the X-ray regime or of particles in the giga-electron-volt range. The energetic photons or particles may come in a shorter burst than the driving p...
0 downloads 0 Views 157KB Size
Waters Symposium: Lasers in Chemistry

Ultrahigh-Intensity Lasers: Nonlinear Optics in the Relativistic Regime for Future Applications in Time-Resolved Chemistry Gérard Mourou Center for Ultrafast Optical Science, University of Michigan, 2200 Bonisteel Boulevard, Ann Arbor, MI 48109-2099

It is the general trend that short pulses and X-ray or highenergetic particles are always associated with high intensities. The generation of high-energy radiation or particles from visible light relies on high-order nonlinear effects or, in the case of X-ray lasers, on the very short lifetime of the excited level, and requires high intensities. In turn, because pulses are ultimately limited by the period of light, shorter wavelengths will be the key to short pulses in the attosecond regime. Over the past ten years, a revolution has occurred in the generation of pulses in an unprecedented regime of intensities. The technique of chirped pulse amplification, with the recent ability to produce pulses composed of only a few optical cycles, has made possible the generation of pulses with extremely high intensity by a modest-sized laser. These pulses are characterized by large ponderomotive forces in the gigabar range and by the relativistic character of the electron motion. At 1018 W/cm2, for 1-µm light, the electron in the field of the laser oscillates with a quiver velocity close _ to the speed of light, leading to an increase in mass of √2. The Lorentz force applied to the electron is F = q[E + (v × B)/c], where v is the quiver electron velocity, q the charge of the electron, and E and B the laser electric and magnetic fields. In linear and nonlinear optics, the term (v × B)/c is always neglected because the ratio between the quiver velocity of the electron and the speed of light is very small and the magnetic term can be safely disregarded. At 1018 W/cm2, this approximation is not valid any more and new effects of relativistic characters become dominant. The very large increase in intensities is opening up new alternatives to the generation of high-energy radiation in the X-ray regime or of particles in the giga-electron-volt range. The energetic photons or particles may come in a

shorter burst than the driving pulse, down to the attosecond, and will make possible the investigation of chemical, biological, or physical phenomena in the time and spectral regime so far inaccessible. The Generation of Ultrahigh-Peak-Power Laser Pulses Until 1985, the amplification of ultrashort pulses was restricted to dye and excimer amplifying media, due to their ultrawide gain bandwidth and low saturation fluence. Optical nonlinearities, notably small-scale self-focusing, prevented the use of high-energy-storage media such as Nd:glass or alexandrite. In 1985, this state of affairs was drastically changed by the technique of chirped pulse amplification (CPA) (1), which made possible the amplification of ultrashort pulses to unprecedented power levels by using solidstate amplifying media with vastly superior energy-storage capability. The very large difference in storage capacity among amplifying media can be appreciated if we look at Figure 1, comparing, for the same amount of stored energy, the size of different amplifying systems such as dye, excimer, Ti:sapphire, Cr:LiSAF, Nd:glass, and Yb:glass. The detrimental nonlinearities were avoided by stretching the pulse in time before amplification and then compressing it, as shown in Figure 2. This technique, in parallel with the recent progress in ultrashort pulse generation (2), led to an explosion in laser peak power (see Fig. 3) by a factor of 103 to 104. This revolution in power was similar to that in the 1960s, when the peak power also increased by many orders of magnitude with the successive demonstration of Q-switching and mode-locking. With some refinements in the stretching and compression scheme, we will soon be capable of amplifying these short pulses up to their theoretical peak power limit (3). For superior energy-storage media, such as Yb:glass or alexandrite, this

Towards more compact CPA laser systems Ti:sapphire Jsat = 0.6 J/cm2 τ = 3 µs

hυ Jsat = σ

Nd:glass Jsat = 7 J/cm2 τ = 400 µs

(Drawings are scaled to the materials’ saturation fluences)

Rhodamine 6G Dye Jsat = 0.002 J/cm2 τ = 3.3 ns

Yb:silica Jsat = 32 J/cm2 τ = 840 µs

Alexandrite Jsat = 26 J/cm2 τ = 260 µs

Figure 1. Representation to scale of different amplifying media, for a given stored energy. Besides the vast differences in size, note the large differences in storage time τ. For instance, the large τ of Yb:silica makes it pumpable by inexpensive, free-running laser diodes.

oscillator

x 103-105

1011

: 3-105 –10

stretcher

amplifier

compressor

Figure 2. Chirped pulse amplification concept, showing the extensive pulse manipulation—stretching 104 times, amplification 1011, and recompression by 104.

JChemEd.chem.wisc.edu • Vol. 75 No. 5 May 1998 • Journal of Chemical Education

565

Waters Symposium: Lasers in Chemistry

limit, as indicated in Figure 3, is at the petawatt level for a beam cross section of 1 cm2 (tabletop). The diffractionlimited focused intensity corresponding to this limit is of the order of 1023 W/cm2. The CPA technique has been demonstrated with a variety of systems, from the small-size, doped-fiber amplifier system delivering microjoules (4 ) to building-size systems such as the ones existing at Limeil, France ( 5), at the RutherfordAppleton Laboratory (6 ), or at the Fusion Facility at Osaka (7 ), delivering pulses with energy of 100 to 1000 J. It is at Lawrence Livermore National Laboratory (LLNL) that the record power, exceeding a petawatt, was demonstrated (8). The technique has been demonstrated with a variety of materials—first with Nd:glass (1, 2), then alexandrite (9), Ti:sapphire (10–12), Cr:LiSAF (13), or a combination (14 ). Their reduced size makes them more favorable for cooling and for beam correction. CPA systems therefore have an average power and repetition rate 10 to 100 times greater than dye or excimer femtosecond systems and their energy can be concentrated over a diffraction-limited spot. This is not generally the case for large-scale terawatt lasers with a spot size several times the diffraction limit. Figure 4 illustrates the evolution of the peak intensity (W/cm2) as a function of year, along with the different laser techniques. The Chirped Pulse Amplification Technique: Cheating Mother Nature by 104 Times Within the five years that followed the inception of lasers in 1960, laser peak power increased rapidly from the kilowatt to the gigawatt by steps of three orders of magnitude (Fig. 3). These steps were due to the fact that the stored energy in the amplifying medium could be released in a shorter and shorter pulse duration: microseconds for the free-running (15), nanoseconds for the Q-switched (QS) (16 ), and picoseconds for the mode-locked laser (ML) (17 ).Once the intensities reached gigawatt power per square centimeter of beam, they could not be increased further owing to nonlinear effects such as intensity-dependent index of refraction, n = n0 + n2I (n0 is the index of refraction, n2 is the nonlinear

index of refraction, and I is the intensity), in laser components such as Pockels cells and amplifiers, windows, etc. This nonlinear effect leads mainly to wave-front deterioration (18), small-scale self-focusing (19) leading to beam filamentation and irreversible damage. It impels the laser amplifier to be run at intensities less than a GW/cm2. Conflicting with this condition is the requirement that, for efficient energy extraction, the input pulse energy per unit area (fluence), Fin (J/cm2), has to be as high as possible and of the order of the saturation fluence. This condition leads to an input intensity condition of the order of Fs /Tp, where Tp is the pulse duration and Fs the saturation fluence. Fs = hυ/σ, where h is the Planck constant, υ the transition frequency, and σ the emission cross section. The last condition says that for short Tp in the picosecond– femtosecond range, only inferior (small Fs , i.e., mJ/cm2) energystorage materials such as dyes and excimers are acceptable. It excludes excellent energy-storage media with large Fs in the J/cm2 range, which would require totally unacceptable input intensities in the 10-TW/cm2 range for picosecond-duration pulses—that is, 103 to 104 times larger than the level at the onset of nonlinear effects. We showed that we could circumvent these seemingly unreconcilable conditions—the simultaneous need for high input energy and low input intensity—by, prior to amplification, stretching the short pulse. In doing that, we could lower the intensity by a factor equal to the inverse of the stretching ratio without changing the input fluence necessary to efficient energy extraction. After being stretched, the pulse can be amplified safely in a high-energy-storage amplifying medium. Once the amplifier energy is efficiently extracted, the stretched pulse is recompressed—ideally, close to its initial value. We realized after the fact that this CPA technique, like most of the laser techniques—Q-switching, mode-locking, and the laser itself—was a transposition in the optical domain of the technique of chirped radar in the microwave regime. In radar, workers faced the same dilemma of reconciling the need for large pulse energy for long-distance ranging with the short pulses required for ranging accuracy. The main difference is that they do not recompress the chirped pulse after amplification but recompress the echo, instead, avoiding the 1030

theoretical limit PW

peak power

TW

chirped pulse amplification

GW mode locking MW Q switching

focused intensity (W/cm2)

1023

Nonlinear QED

E ⋅ e ⋅ λc = 2mo c 2

3 Laser intensity limit I = hν2 ⋅ ∆ν τ c

1020 Vosc ~ c Nonlinear relativistic optics Bound electron

e2 E=a 0

1015 CPA

KW free running

1010 1960

1970

1980

566

Q-switching

1990

Figure 3. Peak power vs year, showing the different laser techniques that led to increased power

mode-locking

1960

1970

1980

Figure 4. Laser focused intensity vs year.

Journal of Chemical Education • Vol. 75 No. 5 May 1998 • JChemEd.chem.wisc.edu

1990

Waters Symposium: Lasers in Chemistry

nonlinear propagation effect in the atmoTable 1. Theoretical Peak Power per cm2 of Beam in Amplifying Medium sphere. Chirped radar uses an impressive Tp Pth Storage Time τ Cross Section ∆ν Laser Type stretching/compression ratio of 200,000. (nm) (10᎑20 cm2) (fs) (TW/cm2) (µs) The complete analog of chirped radar in Nd:glass phosphate 4 22 80 60 300 optics was demonstrated by A. Braun et al. Nd:glass silicate 2.3 28 60 100 300 (20), when they showed that, by Ti:sapphire 30 120 ~8 120 3 recompressing the echo, we could increase Alexandrite 1 100 10 2000 200 optical ranging a thousand times while keeping submillimeter accuracy. Cr:LiSAF 3 50 15 300 70 Today the optical manipulations inYb:silica 0.5 200 8 3000 800 volved in CPA are almost as spectacular as the one in radar and show how robust and of material in the amplifier to a minimum. They succeeded general is this concept. Maybe the most sophisticated one has been demonstrated by Barty et al. (21) and Chambaret in amplifying 30 fs to the multiterawatt level (26). A different system, championed by Barty et al. (27 ), involves low grooveet al. (22). The initial pulse is as short as 10 fs with per-millimeter gratings, a cylindrical mirror, etc., and a careful subnanojoule energy. It is stretched almost 105 times and amplified by a factor of up to 1011 to the joule level, then positioning of all the components. This system has produced less-than-20-fs, 50-TW pulses. Finally, the groups of recompressed by 103 to 105 close to the initial pulse duration. Chambaret et al. (22) and S. Watanabe (28) demonstrated The main difference with chirped radar, chiefly concerned with an aberration-free stretcher of Offner type and have clearly energy and pulse duration, is that laser–matter interaction at shown (22) the importance of the surface quality of optics ultrahigh intensity demands that the pulse be of extraordinary in the generation of high-quality pulses. They produced pulses quality in both the spatial and temporal domains—that is, at the 30-TW level with 30-fs duration. clean over 8 to 10 orders of magnitude and with a wave front better than λ/10. Any prepulses will create a preformed plasma, so the laser will interact with the target in an unknown How High Can We Go? Toward the Theoretical state. The generation of this “perfect” pulse could only be Peak-Power and Intensity Limits accomplished with the invention of the conjugated stretcher– After the dramatic increase obtained recently in peak compressor–amplifier system (23). power, the logical question that we need to answer is how high can we go? This question can be answered quite simply Conjugate Stretching Compression: by saying that the maximum energy that can be extracted The Key to Efficient and Perfect CPA Systems from an amplifier is of the order of the saturation fluence, It is essential in CPA to use a large stretching–compression Fs = h ν/σ, and the shortest pulse duration Tp , is limited by ratio for efficient extraction. The larger the ratio, the better the gain bandwidth of the amplifying medium, ∆ν. The theothe energy storage materials we can use, leading to more retical peak power per cm2 of beam is therefore given simply compact systems for the same peak power. To reach this goal, by Pt h = h ν ∆ν/σ. It is worth noting that this intensity correwe must fulfill the condition that the sum of the phase funcsponds to the Rabi intensity of a π pulse, necessary to flip tions of the stretcher, amplifier, and compressor be equal to the population of the excited state, during the dephasing time zero over the entire frequency range of the amplified pulse. of the material. The maximum focusable intensity will then The first CPA system used fiber for stretching—using the be It h = h ν3 ∆ν/c 2σ. The highest peak power will therefore be positive group velocity dispersion of the fiber at 1.06 µm— obtained with the smallest transition cross section and the and gratings for compression, as demonstrated by B. Treacy largest-bandwidth amplifying materials (see Table 1). (24). In this embodiment, the stretcher and compressor were Pt h varies from 60 TW for Ti:sapphire to 3000 TW for matched over a very limited spectral range (10 Å) leading to Yb:glass. Using Yb:glass, a material that can be obtained in a stretching–compression ratio of only 100. The real breaklarge dimensions, a system with a beam size of 10 cm by 10 cm through was the discovery of the matched stretcher–comprescould produce peak power of 0.5 EW (an exawatt is 1018 W). sor demonstrated in 1987 (23). In this configuration the This power, focused over a diffraction-limited spot size of stretcher, originally proposed for optical communications by 1 µm2, could produce on-target intensities in the 1025/cm2 O. Martinez (25 ) as a compressor for pulses with a negative range. From Table 1, we can see that the storage time (fluochirp, is composed of a telescope of magnification of one rescent lifetime) varies over many orders of magnitude. Because between a pair of antiparallel gratings. Pessot et al. (23) of their very large storage time, Nd:glass and Yb:glass are good showed that this “compressor”, used as a stretcher, not only candidates for diode pumping. exhibited positive group velocity dispersion but also was matched over all orders with the standard Treacy compressor Average Power and Pulse Shaping of CPA Systems made of two parallel gratings, and consequently it could be used to amplify femtosecond pulses. This grating-based The largest average power is necessary for most applications stretcher–compressor today constitutes the basic architecture in order to increase the signal-to-noise ratio. CPA systems of all high-peak-power laser systems. have not only increased the peak power of lasers by three to For ultrashort pulses, the phase conjugation among four orders of magnitude, but have also contributed to imstretcher, amplifier, and compressor has to be accomplished proving the average power of femtosecond systems by at least exactly. The simplest approach, taken by Kapteyn two orders of magnitude. This is largely owing to the small et al., minimizes the stretching ratio by keeping the amount dimensions of the laser amplifier and, in the case of JChemEd.chem.wisc.edu • Vol. 75 No. 5 May 1998 • Journal of Chemical Education

567

Waters Symposium: Lasers in Chemistry

Ti:sapphire, to the excellent heat conduction. With previous femtosecond systems based on dyes or excimers, the average power was typically around 10 mW. With CPA Ti:sapphire systems, the average power went to 1 W (10). This power has since been increased to the 5-W level (29) by cooling the Ti:sapphire material to liquid nitrogen temperature. At this temperature the heat conduction becomes exceptionally high, equivalent to that for copper (30). In the future, with the use of materials with small quantum defects between the absorption and emission wavelengths (such as Yb:glass) pumped by laser diodes, and with rotation of the laser medium (31), average power of petawatt systems could reach the kilowatt level. Finally, one of the interesting features of CPA is the possibility of performing pulse shaping. In the pulse stretcher all the frequency components are spatially spread, so it is easy, by using a simple phase or/and amplitude mask in the Fourier plane, to spatially filter some frequency components (32, 33) and to obtain after pulse compression a pulse with a prescribed shape. Pulse shaping is important in the area of coherent control and, as we will show later, to drive a wake field efficiently. Laser–Matter Interaction at Ultrahigh Intensities Figure 4 shows the intensities over the years and the rapid progress that we have had and are experiencing at this moment, owing to the combination of ultrashort pulse generation and our ability to amplify pulses using chirped pulse amplification. Very much as with nonlinear optics in the early 1960s, new scientific thresholds are crossed. At intensities greater than 1015 W/cm2, we leave the regime of nonlinear optics of the bound electrons to penetrate new physical territories characterized by electric fields much larger than the Coulombic field. Under these conditions, the electron quiver velocity is close to the speed of light, producing large light pressures and, owing to the short duration, no hydrodynamic motion. These fundamentally new features open a new regime in laser–matter interaction. We will review them by giving some examples of the new physics and applications that become accessible: high harmonic generation, nonlinear relativistic effects, laser acceleration, thermonuclear ignition, and nonlinear QED. 1014-W/cm2 Nonlinear Optics of the Bound Electron— Multiphoton Effects and High Harmonic Generation The collisionless regime has been extremely active during the past 15 years. The main physical effects are known as multiphoton absorption (34–36), above-threshold ionization (37–44 ), and high harmonic generation (45 ). The last has real practical application with its possibility of providing the short, XUV, coherent pulses important for time-resolved spectroscopy in this spectral range. By shortening the pulse duration to minimize ionization and to increase the intensity, very high harmonics have been produced (46 – 48), reaching the water window. 1018 W/cm2—the Nonlinear Relativistic Regime

Interaction with Solids In this regime, light–matter interaction is dominated by the relativistic quiver motion of the electrons, which translates into large light pressure. In laser–matter interaction with long 568

pulses, the plasma produced during the interaction has a thermal pressure that is always greater than the light pressure. This is not true anymore for short pulses, where the light pressure associated with their intensity can easily be in the gigabar range, surpass the thermal pressure, and modify the critical surface. This effect was first observed by Liu and Umstadter (49). These authors were able to time-resolve the effect of the light pressure on the critical surface. This large light pressure was conveniently used by Kieffer et al. (50) to produce solid-density plasma of 4 × 1023 cm᎑3. Owing to their high temperature and high density, an ultrashort X-ray source in the kilo-electron-volt range has been produced this way. The high densities were obtained when excellent contrast pulses were used to eliminate prepulses that can cause a plasma before the main pulse turns on. A pulse duration of 300 fs has been demonstrated and used in a time-resolved EXAFS experiment (51). Recently, experiments at higher intensities have revealed even higher densities. The ultrahigh light pressures that can be produced by these pulses are at the crux of the so-called fast ignition concept ( 52), which decouples compression from ignition of the D-T target. First the D-T fuel is assembled (compression) by the long nanosecond pulse. At the point of maximum compression, the ultrashort pulse interacts with the target and an injected beam of high-energy electrons that deposit by collision their energy in the center of the target, igniting the thermonuclear reaction. A CPA laser delivering 1 kJ of energy in 1 ps (petawatt) has been built at LLNL to validate the fast ignitor concept.

Laser–Matter Interaction in Gases Focusing pulses with intensity of 1018 W/cm2 and above in a subcritical-density plasma drives the free electrons relativistically. Their quiver velocity is expressed by γ = (1 + a 2)1/ 2, where γ is the relativistic factor associated with the transverse motion of the electrons and a = γ υosc /c is the normalized vector potential, equal also to eE /moω c. Owing to the beam spatial intensity distribution, the electrons will experience a mass change according to their radial position. The mass change will translate into a modification of the index of refraction, n = [1 – (ωp /ω)2]1/2, where ωp = (4 πne e 2/γ mo)1/ 2 is the plasma frequency. We see that in the high-intensity part of the beam, the index of refraction will be higher, because of the higher electron mass, and the plasma frequency lower. This index change, in space and in time, will translate into a self-focusing ( 53) and self-phase-modulation ( 53) effect, analogous to the self-focusing, self-phase-modulation well known in “classical” nonlinear optics. This relativistic self-focusing is expected to occur at Pc = 17(ω/ωp )2 and has been observed by a number of groups ( 54 ), as have indications of relativistic self-phase modulation ( 55 ). The self-focusing will increase the energy in the channel to a new level of intensity, 1020 W/cm2. It is expected that at this intensity level, the very large transverse intensity gradient will push the electrons outside the channel, to create an evacuated intensity channel in a process called cavitation ( 56 ). Laser pressure, combined with ion inertia, can provide an electrostatic restoring force that can drive a high-amplitude electron plasma wave (EPA). By this process some of the laser energy is converted to a longitudinal electrostatic wake field traveling at nearly the speed of light, which can continuously accelerate the electrons ( 57–59) in the direction of the laser

Journal of Chemical Education • Vol. 75 No. 5 May 1998 • JChemEd.chem.wisc.edu

Waters Symposium: Lasers in Chemistry

beam to giga-electron-volts in a distance as short as a centimeter. That is 104 times larger than some of the highest field gradients produced by conventional means. Thus laser wake fields, combined with the significant laser size reduction obtained with CPA, could dramatically reduce the size of electron accelerators. It is worth noting that this potential reduction factor of 104 is the same as that obtained in lasers with CPA technology. Terawatt lasers used to be building-sized; today they fit on a tabletop. To help appreciate this factor of 104, let’s recall that it is the same factor as the one between a vacuum lamp and a CMOS in an integrated circuit. The generation of a large electrostatic wave amplitude was inferred from the presence of high-order satellites in the Raman forward scattering and by accelerated electrons from the self-modulated wake field (60, 61 ). The self-modulated regime (62) is the one where the plasma wavelength is shorter than the pulse duration, as opposed to the conventional wake field where the plasma period is longer than the pulse duration. Shortly after this, Umstadter et al. (63) demonstrated that the accelerated electrons were in fact coming out in a collimated beam with a transverse emittance and number of electrons (> 0.5 n C) comparable with the best photoinjector. They also showed that the accelerated electrons were exhibiting a very sharp threshold precisely at the critical power, Pc , for relativistic self-focusing. It might be surprising to see such a good emittance coming from such an unsophisticated system. This can be explained by the very large gradient, > 100 GeV/m, experienced by the electrons. Under this gradient, the electrons are quickly accelerated over a fraction of a millimeter to relativistic velocities before they have the time to be broadened by Coulombic repulsion. Let’s recall that the effect of Coulomb broadening in the relativistic regime goes as 1/ γ 2. So the technique of laser wake-field acceleration has the additional advantage of producing electron pulses with a time structure in the femtosecond time scale. A refinement of this technique is currently being tested, where the injection in the plasma wave, acting as an accelerating bucket, is done by an auxiliary pulse perpendicular to the main beam (see Fig. 5) (64 ). This technique, dubbed LILAC for LaserInjected Laser Accelerator, has generated an enormous interest in the scientific community because of its potential to revolutionize electron-beam technology. X-Ray Generation by Compton Scattering The Compton scattering of optical light from an energetic electron beam will upshift the frequency of laser light by a factor of 2 γ 2. This has been used as an electron beam diagnostic for many years with long laser pulses, but in current experiments by Kim et al. (65 ) and by Leemans et al. (66 ), an ultrashort (200-fs) pulse is being scattered from a 50-MeV beam at the Lawrence Berkeley Laboratory, in which case an X-ray (0.4-Å) pulse of similar duration is produced. Nonlinear Quantum Electrodynamics The possibility of creating electron–positron pairs from the laser field was proposed in the late 1960s (67 ). As seen in Figure 4, pair creation directly from the laser will require intensity of the order of 1030 W/cm2. With E as the laser field, it is the energy required over the Compton wavelength, λc = h /m 0 c, that it will take to create an electron–positron pair.

pump pulse

wake field

injection pulse

Figure 5. The LILAC (Laser Injected Laser Accelerator) is an all-optical injection technique. First, a pump beam creates acceleration buckets traveling close to c . An auxiliary probe beam, by ponderomotive force, dephases some of the electrons that will be trapped in the accelerating bucket.

The field is therefore equal to E = 2m0 c 2/e λc and corresponds to 1016 V/cm, which is about four to five orders of magnitude above the laser field of today’s laser. This enormous gap was bridged by using the electric field enhancement produced in the frame of superrelativistic electrons. Using the 50-GeV electron beam at the Stanford Linear Accelerator, corresponding to a γ of 105, with a currently available high-power laser, the field is enhanced to E ~ 1016 V/cm, exceeding the critical field. The multiphoton pair production, ω γ + n ω0 → e + + e ᎑, has been observed (68). The same group is studying nonlinear Compton scattering, e + n ω0 → e′+ ω γ . In this case, the highenergy γ-ray produced in the laser focus by an incident electron interacts before leaving the focal region. Up to the n = 4 process has been observed. Conclusion Our ability to amplify pulses to extreme levels has opened the door to fundamentally new opportunities in science and technology. In particular, laser–matter interaction in the ultrahigh-intensity regime makes possible the generation of energetic photons and particles in the X-ray and giga-electronvolt regime that will enable us to investigate ultrafast phenomena in the femtosecond and attosecond time scale in chemistry, physics, and biology with compact, relatively inexpensive, university-sized systems. Literature Cited 1. Strickland, D.; Mourou, G. Opt. Commun. 1985, 56, 219. Maine, P.; Strickland, D.; Bado, P.; Pessot, M.; Mourou, G. IEEE J. Quantum Electron. 1988, 24, 398. 2. Spence, D. E.; Kean, P. N.; Sibbett, W. Opt. Lett. 1991, 16, 42. 3. Perry, M.; Mourou, G. Science 1994, 264, 917. 4. Galvanauskas, A. Proc. SPIE 1995, 2377, 117. 5. Sauteret, C.; Husson, D.; Thiell, G.; Seznec, S.; Gary, S.; Migus, A.; Mourou, G. Opt. Lett. 1991, 16, 238. Rouyer, C.; Mazataud, E.; Allais, I.; Pierre, A.; Seznec, S.; Sauteret, C.; Mourou, G.; Migus, A. Opt. Lett. 1993, 18, 214. 6. Danson, C. N.; Barzanti, L. J.; Chang, Z.; Damerell, A. E.; Edwards, C. B.; Hancock, S.; Hutchinson, M. H. R.; Key, M. H.; Luan, S.; Mahadeo, R. R.; Mercer, I. P.; Norreys, P.; Pepler, D. A.; Rodkiss, D. A.; Ross, I. N.; Smith, M. A.; Taday, P.; Toner, W. T.; Wigmore, K. W. M.; Winstone, T. B.; Wyatt, R. W. W.; Zou, F. Opt. Commun. 1993, 103, 392. 7. Yamakawa, K.; Shiraga, H.; Kato, Y.; Barty, C. P. J. Opt. Lett. 1991, 16, 1593.

JChemEd.chem.wisc.edu • Vol. 75 No. 5 May 1998 • Journal of Chemical Education

569

Waters Symposium: Lasers in Chemistry 8. Perry, M.; Stuart, B. C.; Tietbohl, A.; Miller, J.; Britten, J. A.; Boyd, R.; Everett, M.; Herman, S.; Nguyen, H.; Powell, H. T.; Shore, B. W. In CLEO ’96; Technical Digest Series, Vol. 9; Optical Society of America: Washington, DC, 1996; pp 307–308. 9. Perry, M. D.; Patterson, F. G.; Weston, J. Opt. Lett. 1991, 16, 381. White, W. E.; Hunter, J. R.; Van Woerkom, L.; Ditmire, T.; Perry, M. D. Opt. Lett. 1992, 17, 1067. 10. Vaillancourt, G.; Norris, T. B.; Coe, J. S.; Mourou, G. A. Opt. Lett. 1990, 15, 317. Squier, J.; Mourou, G. Laser Focus World 1992, 28, 51. 11. Kmetec, J.; Macklin, J. J.; Young, J. F. Opt. Lett. 1991, 16, 1001. 12. Sullivan, A.; Hamster, H.; Kapteyn, H. C.; Gordon, S.; White, W.; Nathel, H.; Blair, R. J.; Falcone, R. W. Opt. Lett. 1991, 16, 1406. 13. Beaud, P.; Richardson, M.; Miesak, E.; Chai, B. T. Opt. Lett. 1993, 18, 1550. 14. White, W. E.; Hunter, J. R.; Van Woerkom, L.; Ditmire, T.; Perry, M. D. Opt. Lett. 1992, 17, 1067. 15. Maiman, T. H. Nature 1960, 187, 493. 16. McClung, F. J.; Hellwarth, R. W. Proc. IRE 1963, 51, 46. 17. Moker, H. W.; Collins, R. J. Appl. Phys. Lett. 1965, 7, 270. DeMaria, A. J.; Stetser, D. A.; Heynau, H. Appl. Phys. Lett. 1966, 8, 174. 18. Shen, Y. R. Principles of Nonlinear Optics; Wiley Interscience: New York, 1984. 19 Bliss, E. S.; Speck, D. R.; Holzrichter, J. F.; Erkkila, J. H.; Glass, A. J. Appl. Phys. Lett. 1974, 25, 448. 20. Braun, A.; Chien, C. Y.; Coe, S.; Mourou, G. Opt. Commun. 1994, 105, 63. 21. Barty, C. P.; Gordon, C. L., III; Lemoff, B. E. Opt. Lett. 1994, 19, 1442. 22. Chambaret, J. P.; Le-Blanc, C.; Cheriaux, G.; Curley, P.; Darpentigny, G.; Rousseau, P. Opt. Lett. 1996, 21, 1921. 23. Pessot, M. Opt. Commun. 1987, 62, 419. 24. Martinez, O. E. IEEE J. Quantum Electron. 1987, 23, 1385. 25. Treacy, E. B. IEEE J. Quantum Electron. 1969, QE-5, 454. 26. Zhou, J.; Huang, C. P.; Murnane, M. M.; Kapteyn, H. Opt. Lett. 1995, 20, 64. 27. Lemoff, B. E.; Barty, C. P. Opt. Lett. 1993, 18, 1651. 28. Itatani, J.; Nabekawa, Y.; Kondo, K.; Watanabe, S. Opt. Commun. 1995, 134, 134. 29. Backus, S.; Durfee, C., III; Mourou, G.; Kapteyn, H.; Murnane, M. Opt. Lett., in press. 30. CRC Handbook of Chemistry and Physics Weast, R. C.; Astle, M. J., Eds.; CRC: Boca Raton, FL, 1980; pp 4–79. 31. Biswal, S.; Nees, J.; Mourou, G. A. Advanced Solid Lasers; Technical Digest; Optical Society of America: Washington, DC, 1997; pp 102–104. 32. Weiner, A. M.; Leaird, D. E.; Patel, J. S.; Wullert, J. R. IEEE J. Quantum Electron. 1992, 28, 908. 33. Liu, X.; Wagner, R.; Maksimchuk, A.; Goodman, E.; Workman, J.; Umstadter, D.; Migus, A. Opt. Lett. 1995, 20, 1163. 34. L’Huillier, A.; Lompré, L. A.; Mainfray, G.; Manus, C. Phys. Rev. A 1983, 27, 2503. 35. Luk, T. S.; Pummer, H.; Boyer, K.; Shahidi, M.; Egger, H.; Rhodes, C. K. Phys. Rev. Lett. 1983, 51, 110. 36. Perry, M. D.; Landen, O. L.; Szoke, A.; Campbell, E. M. Phys. Rev. A 1988, 37, 747. 37. Agostini, P.; Fabre, F.; Mainfray, G.; Petite, G. Phys. Rev. Lett. 1979, 42, 1127. 38. Kruit, P.; Kimman, J.; Muller, H. G.; van der Wiel, M. J. Phys. Rev. A 1983, 28, 248. 39. Freeman, R. R.; Bucksbaum, P. H.; Milchberg, H.; Darack, S.; Schumacher, D.; Geusic, M. E. Phys. Rev. Lett. 1987, 59, 1092. Perry, M. D.; Szoke, A.; Kulander, K. C. Phys. Rev. Lett. 1989, 63, 1058. 40. Mohideen, U.; Sher, M. H.; Tom, H. W. K.; Aumiller, G. D.; Wood, O. R., II; Freeman, R. R.; Bokor, J.; Bucksbaum, P. H. Phys. Rev. Lett. 1993, 71, 509. Schafer, K. J.; Yang, B.; DiMauro, L. F.; Kulander, K. C. Phys. Rev. Lett. 1993, 70, 1599. 41. Kulander, K. C. Phys. Rev. A 1988, 38, 778.

570

42. Eberly, J. H.; Su, Q.; Javanainen,J. Phys. Rev. Lett. 1992, 62, 881. 43. Potvliege, R. M.; Shakeshaft, R. Phys. Rev. A 1989, 40, 3061. 44. Faisal, F. H. M. Theory of Multiphoton Processes; Plenum: New York, 1987, and references therein. 45. Li, X. F.; L’Huillier, A.; Ferray, M.; Lompré, L. A.; Mainfray, G.; Manus, C. Phys. Rev. A 1989, 39, 5751. L’Huillier, A.; Lompré, L. A.; Mainfray, G.; Manus, C. In Atoms in Intense Laser Fields; Gavrila, M., Ed.; Academic: Orlando, FL, in press. 46. Sarakura, N.; Hata, K.; Adachi, T.; Nodomi, R.; Watanabe, M.; Watanabe, S. Phys. Rev. A 1991, 43, 1669. Mayazaki, K.; Sakai, H. J. Phys. B 1992, 25, L83. 47. Macklin, J. J.; Kmetec, J. D.; Gordon, C. L. Phys. Rev. Lett. 1993, 70, 766. 48. Zhou, J.; Peatross, J.; Murnane, M.; Kapteyn, H.; Christov, I. Phys. Rev. Lett. 1996, 76, 752. 49. Liu, X.; Umstadter, D. Phys. Rev. Lett. 1992, 69, 1935. 50. Kieffer, J. C.; Chaker, M.; Matte, J. P.; Pépin, H.; Côté, C. Y.; Beaudoin, Y.; Johnston, T. W.; Chien, C. Y.; Coe, S.; Mourou, G.; Peyrusse, O. Phys. Fluids B 1993, 5, 2676. 51. Raksi, F.; Wilson, K.; Jiang, Z.; Ikhlef, A.; Côté, C. Y.; Kieffer, J. C. J. Chem Phys. 1996, 104, 6066. 52. Young, P. E.; Foord, M. E.; Hammer, J. H.; Kruer, W. L.; Tabak, M.; Wilks, S. C. Phys. Rev. Lett. 1995, 75, 1082. 53. Max, C.; Arons, J.; Langdon, A. B. Phys. Rev. Lett. 1974, 33, 209. 54. Borisov, A. B.; Borovskiy, A. V.; Korobkin, V. V.; Prokhorov, A. M.; Shiryaev, O. B.; Shi, X. M.; Luk, T. S.; McPherson, A.; Solem, J. C.; Boyer, K.; Rhodes, C. K. Phys. Rev. Lett. 1992, 68, 2309. Borisov, A. B.; Shi, X.; Karpov, V. B.; Korobkin, V. V.; Solem, J. C.; Shiryaev, O. B.; McPherson, A.; Boyer, K.; Rhodes, C. K. J. Opt. Soc. Am. B 1994, 11, 1941. Monot, P.; Auguste, T.; Gibbon, P.; Jakober, F.; Mainfray, G.; Dulieu, A.; Louis-Jacquet, M.; Malka, G.; Miquel, J. L. Phys. Rev. Lett. 1995, 74, 2953. Wagner, R.; Chen, S.-Y.; Maksimchuk, A.; Umstadter, D. Phys. Rev. Lett. 1997, 78, 3125. 55. Umstadter, D.; Chen, S. Y.; Maksimchuk, A.; Mourou, G.; Wagner, R. Science 1996, 273, 472. 56. Mora, P.; Antonsen, T. M., Jr. Phys. Rev. 1996, E53, R2068. 57. Tajima, T.; Dawson, J. M. Phys. Rev. Lett. 1979, 43, 267. 58. Marques, J. R.; Geindre, J. P.; Amiranoff, F.; Audebert, P.; Gauthier, J. C.; Antonetti, A.; Grillon, G. Phys. Rev. Lett. 1996, 76, 3566. 59. Siders, C. W.; LeBlanc, S. P.; Fisher, D.; Tajima, T.; Downer, M. C.; Babine, A.; Stepanov, A.; Sergeev, A. Phys. Rev. Lett. 1996, 76, 3570. 60. Coverdale, C. A.; Darrow, C. B.; Decker, C. D.; Mori, W. B.; Tzeng, K. C.; Marsh, K. A.; Clayton, C. Joshi, C. E. Phys. Rev. Lett. 1995, 74, 4659. 61. Modena, A.; Najmudln, Z.; Dangor, A. E.; Clayton, C. E.; Marsh, K. A.; Joshi, C.; Malka, V.; Darrow, C. B.; Danson, C.; Neely, D.; Walsh, F. N. Nature 1995, 377, 606. 62. Sprangle, P.; Esarey, E.; Krall, J.; Joyce, G. Phys. Rev. Lett. 1992, 69, 2200. Antonsen, T. M.; Mora, P. Ibid., 2204. 63. Umstadter, D.; Chen, S.-Y.; Maksimchuk, A.; Mourou, G.; Wagner, R. Science 1996, 273, 472. 64. Umstadter, D.; Kim, J. K.; Dodd, E. Phys. Rev. Lett. 1996, 76, 2073. 65. Kim, K.-J.; Chattopadhyay, S.; Shank, C. V. Nucl. Instrum. Methods Phys. Res. A 1994, 341, 351. 66. Leemans, W.; Behrsing, G.; Kim, K.-J.; Krupaick, J.; Matuk, C.; Selph, F.; Chattopadhyay, S.; Corneliussen, S. T. Proceedings of the 1993 Particle Accelerator Conference; IEEE, New York, 1993; Vol. 1, pp 83–85. 67. Nikishov, A. I.; Ritus, V. I. Sov. Phys. JETP 1964, 19, 529, 1191; 1965, 20, 757. 68. Bula, C.; McDonald, K. T.; Prebys, E. J.; Bamber, C.; Boege, S.; Kotseroglou, T.; Melissinos, A. C.; Meyerhofer, D. D.; Ragg, W.; Burke, D. L.; Field, R. C.;. Horton-Smith, G; Odian, A. C.; Spencer, J. E.; Walz, D.; Berridge, S. C.; Bugg, W. M.; Shmakov, K.; Weidemann, A. W. Phys. Rev. Lett. 1996, 76, 3116.

Journal of Chemical Education • Vol. 75 No. 5 May 1998 • JChemEd.chem.wisc.edu