:l n- ..z i . 19's,)». A .. O 8 r8 . ‘ .z ,. . . V .. ‘ . "bun...” Jaw,qu....$4::. . v 2L i 3.)!“ . :hvf‘k’? ,. Ev. . fdw . .n in...“ (45.5. .lta... . gt... pg 1...] fi‘ .3 . 3.3 . .wbtg),v¥84nuat h ‘1. VJ} .6 dual: .... b Lb .' U I'll I a «a I 3.11 A . .0! .1... v : new. , 1?... Al Fri]! .. . 31. x 3.39:}! I .iIV-m . V! ‘ V. 4£|LI t! Al'l, ‘ s a? .i . ‘ . [$an . i... . . ixiuutfiuuvaq 3.33:... ha. h. $0.3: u (kg? ‘ :1 ha. flunk} 3.5.1.11 .ILI. .. . I ‘1! . o-n “ p’Dr’ t .3 . s L‘ curt. o ‘ s 9.1.20..- n :Evfl rung...) . . log!!! .1 s1. 141th .. ,9 .fl 2, {onwmflufino .3..vln.huv.:. .5! law 5. _ . «PJHJSLHJ l: . Illnmmh‘uuui ufifiaflufli . "f l 11“.)ng .1 u y. I! . xlllfluoxt £11322! 3-} .I . .l-VJ Niftiubznlo 'l’l I! . I XIII» v.3 In ,lu.11ntfiul.6lhr.re 9...! hot,” . . of. £5.11)... DIE-5‘}: f,“ In 14. vallIsllllR It"..- v1..v\! III}... 4.1.. . a“?! ‘1’” . < v . .. .5: 010:!»an V’Ilw' o. I“! h 13:33.!!! :13» 5!! 1.. 13.533: :.I...:¢h!..v~!vucl $4.912!» .Ioi 3.ch .L. 1... «4|! .. $531.” .3 .c .l f“ Pk“: {lSflhl 1. .0! L n 11:21.. . . .57 i 1&1 . . is. 1. etvut‘h‘ . slur!!! I. ifimulg e ‘ z , .. 3.. aria: ‘ . Ufnfvnn i0!)3i.!3...tt 25:15.1: .. .11., nut ’vl‘ql 34: ‘I ’III ‘...Sio$!!.} .. 9-- Hr; 0:111: $353.»-.. I “to 5-3.1:.--311.‘ {0.3.3. . hdl..¥3..l..ri.§2€l£. .Ji ..f.¥.!)€l I'm! ’Ibovo-uoip, F '.O'§l...vtv|.’v .4... .111 . Vt».l:..v .YA...I.' 1.:4 . 3.91“ .‘t.’I~I-£|v!§r it. Ill..nAO ix mph“- '1! r In. if . 0tpllllv.‘\.|)l.. UltvD'lL . ‘ . ‘ ‘1‘): O}; ..‘Xllll..nll¢‘l.vs1[‘0. ‘ . . ‘ . . T ,Alnvri .III\ 1’. . 4 ' mm M‘ two, n..! a c . . t y .. w . . . ’ 5... .. . I . . ‘ \ rl...tJ-idunulfd.hu I (Khan... . Wow . .LH. iiuwumuflJou. #1." fiuk..uq...u.m.mwmuhul. .th..n....1u ‘ 4;. in , «J. . I | r I I! ll' . Illllll THF‘HS 1003 This is to certify that the dissertation entitled Observation and Control of Intramoiecuiar Dynamics of Gas-Phase Molecules by Ultrafast Laser Puises presented by . Igor Pastirk has been accepted towards fulfillment of the requirements for Docterai degree in (:hemj SIC)! {gm}: Major professor Date 07/25/2002 MSU i: an Affirmative Action/Equal Opportunity Institution 0-12771 LIBRARY Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 cJCIRC/DateDuopes-sz Observation and Control of Intramolecular Dynamics of Gas-Phase Molecules by Ultrafast Laser Pulses by Igor Pastirk A DISSERTATION Submitted to Michigan State University in partial fitlfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 2002 ABSTRACT Observation and Control of Intramolecular Dynamics of Gas-Phase Molecules by Ultrafast Laser Pulses by Igor Pastirk The work presented here describes experimental femtosecond three pulse four- wave mixing techniques as a tool in observation and laser control of molecular dynamics. The ability to control population and coherence transfer in molecular systems is of great importance in the quest for controlling the outcome of laser-initiated chemical reactions. The approach is based on the use of different femtosecond pulse sequences and applied field’s phase characteristics for providing a fundamental understanding of underlying non-linear optical processes. It relies on a large body of experimental measurements on a model system consisting of isolated (gas-phase) iodine molecules that have resonant X 'ng (ground) to B 3110+u (excited) electronic transition with applied laser fields. Control over population and coherence transfer is demonstrated by selecting specific pulse sequences. The technique is also used to sort and measure the processes that contribute to coherence relaxation of gas-phase 12. The observed relaxation times for vibronic coherence are determined using photon echo and reverse transient grating measurements at different temperatures to isolate inhomogeneous and homogeneous components. Different pulse sequences are used to select ground or excited state vibrational coherences. Measurements of ground and excited state wave packet spreading times due to anharrnonicity, a process that does not involve energy dissipation or phase relaxation, are also presented. Femtosecond ground state dynamics of gas phase N204 and N02 are studied by non-resonant four-save mixing (FWM) technique, applying a transient grating pulse sequence. Both fast vibrational dynamics (< 1 ps) and long time (up to 100 ps after initial excitation) rotational revivals are recorded for both samples. Observed fast (133 fs) beats in N204 are assigned to stretching mode of N-N bond. The average rotational constant for NO; in the ground state was determined from long time measurements. Saneli \\ ht Lo >0 ext idc Acknowledgments During my graduate studies at Michigan State University I met a number of people that have helped me, in one way or another. The level of cooperation and willingness to help that is present at the Department of Chemistry would make the list here very long. I would like to thank them all. I would also like to thank the past and present members of the research group. Specifically, Emily Brown, Matt Comstock, Irene Grimberg, Peter Gross, Vadim Lozovoy, Una Marvet, Kathy Walowicz, and Qingguo Zhang — thank you for sharing your knowledge in the lab and your advice out of the lab. I would particularly like to thank my advisor Marcos Dantus, for introducing the exciting field of ultrafast spectroscopy and laser control to me, being always open to new ideas and supporting me both professionally and personally during my graduate studies. 9) SI.‘ TABLE OF CONTENTS LIST OF TABLES ........................................................................................................... viii LIST OF FIGURES ........................................................................................................... ix 1. INTRODUCTION ...................................................................................................... 1 1.1. Observing chemical reactions in real-time ......................................................... 3 1.2. Controlling the dynamics of molecules .............................................................. 6 1.3. Outline of the thesis .......................................................................................... 10 2. EXPERIMENTAL (ultrashort lasers and techniques) .............................................. 13 2.1. Introduction ....................................................................................................... 13 2.2. F emtosecond pulses, characterization and experimental techniques ................ 14 2.2.1. Femtosecond pulse characterization ......................................................... 18 2.2.2. One-pulse excitation ................................................................................. 25 2.2.3. Three—pulse FWM .................................................................................... 27 2.3. Colliding Pulse Mode-locked Laser System (CPM) ......................................... 28 2.4. Titanium sapphire laser system ......................................................................... 33 3. QUANTUM CONTROL OF THE YIELD OF A CHEMICAL REACTION BY SINGLE FEMTOSECOND PULSES .............................................................................. 38 3.1. Introduction ....................................................................................................... 38 3.2. Experimental ..................................................................................................... 40 3.3. Results and discussion ...................................................................................... 42 4. PULSE SEQUENCES AND CHIRP AS A TOOL IN FEMTOSECOND QUANTUM CONTROL OF MOLECULAR EXCITATION AND INTRAMOLECULAR DYNAMICS ............................................................................... 48 4.1. Introduction ....................................................................................................... 48 4.2. Four wave mixing ............................................................................................. 53 4.2.1. The role of pulse sequences in controlling ultrafast intramolecular dynamics with FWM ................................................................................................. 53 4.2.2. Formal description of nonlinear optical processes involved in sequences’ control scheme .......................................................................................................... 55 vi 1(1 5. 4.3. 4.2.3. Experimental ............................................................................................. 61 4.2.4. Overlapped Pulses ..................................................................................... 65 4.2.5. Pulse Sequence I: Virtual Photon Echo and Stimulated Photon Echo ...... 68 4.2.6. Pulse Sequence II : Virtual Photon Echo and Stimulated Photon Echo 72 4.2.7. Cascaded Free-Induction Decay FWM ..................................................... 75 4.2.8. Conclusions ............................................................................................... 77 Control of quantum phase with coherent light .................................................. 84 4.3.1 . Introduction ............................................................................................... 85 4.3.2. Theory ....................................................................................................... 87 4.3.3. Experimental ............................................................................................. 98 4.3.4. Spectral information in femtosecond three-pulse F WM ........................... 99 4.3.5. Two-dimensional three-pulse F WM measurements ............................... 101 4.3.6. Three-pulse F WM with chirped laser pulses .......................................... 107 4.3.7. Conclusions ............................................................................................. 1 13 FEMTOSECOND PHOTON ECHO AND VIRTUAL ECHO MEASUREMENTS OF THE VIBRONIC AND VIBRATIONAL COHERENCE RELAXATION TIMES OF IODINE VAPOR ............................................................................................................ 116 5. 1 . Introduction ..................................................................................................... 1 16 5.2. Experimental ................................................................................................... 1 1 9 5.3. Results and discussion .................................................................................... 119 5.4. Conclusions ..................................................................................................... 126 6. FEMTOSECOND GROUND-STATE DYNAMICS OF GAS-PHASE DINITROGEN TETROXIDE AND NITROGEN DIOXIDE ........................................ 127 6.1. Introduction ..................................................................................................... 127 6.2. Experimental ................................................................................................... 129 6.3. Results and discussion .................................................................................... 130 6.4. Conclusion ...................................................................................................... 138 7. SUMMARY AND CONCLUSIONS ..................................................................... 140 8. APPENDIX I .......................................................................................................... 143 9. REFERENCES ....................................................................................................... 146 vii LIST OF TABLES Table 1. Summary of the observed phenomena for the different pulse sequences possible with the three laser pulses when two pulses overlap in time. For the experimental arrangement used here, there are only four distinguishable sequences (shown in bold face). Because beams Ea and Ec are equivalent, two additional (though not unique) sequences are possible. Sequences for which experimental data are presented have been labeled with a star. The phenomena include transient grating (TG), reverse-transient grating (RTG), reverse photon echo (RPE), and photon echo (PE) .................................. 63 Table 2. Summary of the observed phenomena and the different pulse sequences possible with three time-separated laser pulses. For the experimental arrangement used here, there are only six distinguishable sequences (shown in bold face). Because beams E3 and Ec are equivalent, additional (though not unique) sequences are possible. Sequences for which experimental data are presented have been labeled with a star. The phenomena include virtual photon echo (VE), stimulated photon echo (PE), and cascaded free-induction decay FWM (C-FID-FWM) .............................................................................................. 64 viii LIST OF FIGURES Figure 2.1 Schematic of the experimental setup for an intensity autocorrelation measurement. M - mirrors, BS — 50/50 beamsplitter, HCC - hollow comer cube mounted on a computer controlled variable delay stage, SHG - doubling crystal, L — focusing and collimating lens, UV — filter. ............................................................................................ 19 Figure 2.2. Quasistatic interaction length (thickness) of selected crystals as a function of the duration of transform limited pulses for Type I SHG at 800 nm. ............................... 20 Figure 2.3. Typical autocorrelation trace (triangles) obtained for amplified titanium- sapphire fs laser pulses with a Gaussian function fit. Duration of the pulses is 50 fs. ..... 21 Figure 2.4. Typical P-FROG trace obtained for transform-limited (a) and positively chirped (b) pulses from an amplified CPM laser with central wavelength of 620 nm. 23 Figure 2.5. a) Experimental SHG FROG trace of near TL pulses from an amplified titanium sapphire laser, obtained by CCD and spectrometer. Darker shades correspond to the higher intensity. b) Temporal profile (thin line) and phase dependence (thick line) retrieved from data in a) by FROG software. c) Spectral (thin line) and phase (thick line) profiles of the pulses retrieved by FROG software. Pulses were 55 fs with time- bandwidth product of 0.4646. Phase variation of about 0.1 radians across most of the pulse duration in b) is considered to be a minor deviation from TL. The spectral phase characteristic in c) signifies minor higher order chirp. ..................................................... 24 Figure 2.6. Experimental setup for one-beam experiment. P1 and P2 — double-pass compressor prisms. P2 is mounted on an actuator. Mirrors (M) direct the beam to an SHG crystal where the frequency of the pulses is doubled, and pulses are focused on the sample cell. rPD — reference photodiode used for discrimination. ................................... 26 Figure 2.7. Beam arrangement for the three-pulse four-wave mixing experiment. BS - beamsplitters; HCC — hollow comer cubes mounted on variable delay stages; L —focusing and collimating lenses. Detail of the spatial overlap is also shown (top) with the direction of the signal, detected in the phase-matching direction ks. ............................................... 28 Figure 2.8. Experimental setup for CPM dye laser with general detection scheme. M and CM - mirrors and curved mirrors, respectively; P - prisms; OC — output coupler. Nd:YAG laser and detection are triggered and synchronized to the pulses detected by photodiode PD. ................................................................................................................. 31 Figure 2.9. Titanium-sapphire laser oscillator (seed laser). Light from a Nd:YVO4 laser is focused by lens (L) through one of the curved mirrors (CM) on the titanium-sapphire crystal. CM are transparent for 532 nm light, but 100 % reflective in the 760-850 nm region. Emission from the crystal is collected by CM and directed through prisms P1 and P2 to the end mirror (EM). Length of the cavity (distance from EM to output coupler 0C) is ~ 2m. Pulses shorter than 15 fs when compressed and centered around 800 nm with FWHM of more than 50 nm were obtained. For the proper operation of the ix amplifier, presented in the next figure, much less bandwidth is needed (25-30 nm) and the oscillator is adjusted accordingly. For fast determination of the bandwidth, one reflection of P2 is directed from the diffraction grating to an inexpensive CCD and TV monitor. The image can be calibrated and used for rough estimation of pulse bandwidth. ............ 35 Figure 2.10. Schematic for regenerative amplification of pulses from titanium-sapphire laser. Beam from the seed laser is directed through the Faraday isolator (F1) to diffraction grating G(1) in the stretcher. F I acts as filter preventing the back-reflected pulses from the amplifier from reentering the seed laser cavity. The beam diffracted from G] reflects from silver mirror SM to long mirror LM and back. Still expanded, the beam travels to periscope P and back for another round trip (G1, SM, LM and back). Spatially recompressed, but chirped to ~100 ps, the beam is sent to an amplifier (see text) ........... 36 Figure 3.1 (a) Experimental measurement of the yield of the molecular pathway producing 12 from the multiphoton dissociation of CH2I2 with 624 nm laser pulses as a function of chirp. The insert shows the relevant energetics for the reaction. 0)) Fluorescence yield as a function of chirp following the multiphoton excitation of 12 vapor .................................................................................................. 44 Figure 3.2 Experimental measurement of the yield of the molecular pathway producing 12 from the multiphoton dissociation of CH2I2 with 312 nm laser pulses as a function of experimentally available chirp range. The inserts show the relevant energetics for the reaction, as well as a plot of the maximum 12 yield enhancement recorded at 2400 fs2 chirp, as a function of laser peak intensity (measured at zero chirp). ......................... 46 Figure 4.1. (a) Beam arrangement in the forward box configuration for the three-pulse FWM experiment. The three laser fields are applied in a given temporal sequence and overlapped spatially in the sample. The signal is detected in the direction of the wave vector ks that satisfies the phase-matching condition. The dashed line indicates a plane of symmetry that makes fields Ea and Ec equivalent. (b) Temporal sequence of the three laser pulses for a virtual echo measurement. 2'12 and 2'23 are the time delays between first and second or second and third pulses, respectively. (c) Temporal sequence of the three laser pulses for a virtual photon echo measurement. Notice that the rephasing occurs at a time t3 + ((2 — (I) for the stimulated photon echo signal and t3 — ((2 — II) for the virtual photon echo signal. ........................................................................................................... 56 Figure 4.2. Diagrammatic representation of the processes that contribute to the virtual echo signal when the time interval between the first two pulses is fixed. (a) The double- sided Feynman diagrams show explicitly the time evolution of the density matrix when the fields act on the bra (right) or ket (left) side of the diagram and the time increases from bottom to top. The diagram on the left, R4, is responsible for the observation of the ground-state dynamics. The one on the right, R1, is responsible for the excited state dynamics. (b) The corresponding ladder diagrams illustrate explicitly the energy levels involved in each electronic interaction. A vertical dashed or solid line represents interaction on the bra or the ket respectively. In these diagrams, time evolves from left to right. .................................................................................................................................. 60 Figure 4.3. Experimental F WM data obtained for three different sequences having two pulses overlapped in time. The signal corresponds to: TG, RTG and PE depending on the temporal arrangement of the electric fields. The pulse sequence and the DSFD of the processes contributing to each phenomenon are included in each case. The dominant oscillation period in all cases, 307 fs, corresponds to excited state vibrations. The TG and the RTG transients in (a) and (b) show some additional contribution of ground state vibrations. The RTG transient presents a prominent slow modulation (2 ps) corresponding to rotational dephasing and a large background signal. Notice that the PE data show only excited state vibrations with very little rotational dephasing. ................. 66 Figure 4.4. Experimental transients for sequences with fixed time delay between the first two pulses, PS I, with Tab: 25,5460 fs or Tab: 258:614 fs. The VE signal is obtained when the initial electric field acts on the ket side of the diagram and a PE signal is obtained if the initial electric field acts on the bra side of the diagram. (a) When rab=460 fs, the oscillation period of the VE signal corresponds to the vibrational motion of the excited state molecules. (b) When rab=614 fs, the VE signal oscillates with the ground state vibrational frequency and the rotational dynamics (slow 2 ps modulation) are clearly observed. Notice that the predominant feature in both PE signals (c and (I) shows 307 fs oscillations corresponding to the excited state .................................................................. 69 Figure 4.5. Top: Experimental data for VE (thick line) and PE (thin line) measurements with Tab: 21,5460 fs. Notice that the signals, corresponding to excited state dynamics, are exactly out of phase. The signal intensity has not been re-scaled, but the VEI signal has been shifted vertically. Note that the photon echo signal peaks at 154 fs. The 180° phase difference between these transients can be understood by looking at the dynamics of the wave packet in the B state of 12 resulting from the interaction with pulse Eb. Bottom: Simulation of the wave packet motion on the B state of 12 under the conditions of the VB and PE signals for PS I when Tab: 153:460 fs. In both cases signal formation depends on de-excitation of ‘Pbm (dotted curve). At time delay z=0 fs, which corresponds to the time delay between last two pulses, ‘I’bm is in the Franck-Condon region for VB 1, maximizing the transition probability when the third pulse is applied there. However, for case PE 1, LPb“) is at the outer turning point of the excited state potential when 1:0 fs, minimizing the transition probability when the third pulse is applied in the Franck—Condon region.. 71 Figure 4.6. VE transients obtained for a sequence having the time delay between the last two pulses fixed to rba=460 fs or 21,5614 fs. The vibrational excited-state motion dominates both transients. A very small contribution from the ground state vibrational motion is apparent when Iba=614 fs. A larger rotational dephasing is observed (2 ps modulation) in the 253:614 fs signal and the signal intensity is higher in this case. The DSFD of the processes that contribute to the signal are included with the pulse sequence. ........................................................................................................................................... 73 Figure 4.7. PE transients obtained for PS 11 when rac=460 fs or Tac=6l4 fs. These transients, mostly dominated by the excited state motion, do not show the selectivity of the molecular dynamics with the selection of the fixed time delay. In addition to the pulse sequence, the DSFD of the processes that contribute to the signal are included. ............. 74 xi \\ 01 Figure 4.8 Insert: Transient grating and double-sided Feynman diagrams for the case when pulse b is the last to arrive at the sample. The grating formed by pulses a and c (indicated by the diagonal lines) cannot diffract pulse b in the signal direction. Violation of the RWA is identified in the double-sided Feynman diagrams by the crossed arrows. C-FID-FWM transients for rab=460 fs (a) and rab=6l4 fs (b). When rab=460 fs, the oscillation period of the C-FID-F WM signal corresponds to the vibrational motion of the excited state molecules, whereas when rab=614 fs the signal oscillates with the ground state vibrational frequency and shows a larger rotational modulation (2 ps). Notice that the data are very similar to those observed for VB 1 (Figure 4.4a). This implies that the dominant pathway for signal formation can be illustrated using similar DSFD for the two ensembles. ......................................................................................................................... 76 Figure 4.9. Cascaded free-induction decay F WM transients for PS I for rac=460 fs (a) and rac=614 fs (b). Notice that the signal intensity for these sequences is quite low, especially after the first picosecond. The transients do not show selectivity between the ground and excited state molecular dynamics with the fixed time delay. The pulse sequence and the coupled DSFD are included for each case. ........................................... 78 Figure 4.10. (a) Representation of the density matrix for a multilevel system composed of two electronic states, each one with a manifold of vibrational levels. The indices g, g' range over the number of vibrational levels of the ground electronic state (in the present simulation g, g' = 1, 2). The indices e, e' range over the number of vibrational levels of the excited electronic state (in the present simulation e, e' = 3, 4). The diagonal matrix elements, pgg and pee, represent the population of the vibrational levels (dark shade). The off-diagonal matrix elements of the diagonal blocks, pgg' and pee, represent the vibrational coherence in the respective electronic state (light shade). The elements of the off-diagonal blocks, pge and peg, represent the vibronic coherence between the electronic states. (b) Double-sided Feynman diagrams for the processes described in the text afier the first two electric field interactions only. ..................................................................... 88 Figure 4.11. Simulations of spectrally dispersed FWM. At the top of the figure are diagrams of the averaged density matrix elements <|p,j(2)|2> calculated with Equations (4.7-4.10). The corresponding spectrograms calculated with Equation (4.11) are shown with wavelength plotted along the horizontal axis and time delay between the second and third pulses (r) plotted along the vertical axis. The darker regions in spectrograms correspond to higher signal intensity. For these calculations, the initial population distribution parameter (w) in the ground state was assumed to be zero and the pulses were unchirped. (a) The delay between first two pulses (rah) is 400 fs, corresponding to 5/2 vibrational periods of the ground state, 18. The density matrix diagram for this pulse sequence (top left) shows population and coherence terms only in the upper diagonal block, corresponding to ground-state matrix elements alone. The spectrogram (left) clearly shows 160 fs periods, reflecting the vibration of the ground state (rug). (b) For Tab = 460 fs, 3/2 vibrational periods of the excited state 22,, only the matrix elements of the excited state (lower diagonal block) are present in the diagram. In the corresponding spectrogram (right), the excited-state vibrations of 307 fs are evident ............................. 95 xii 11".1 ur‘: 0.. fs b} CO! Sir 9X ev Figure 4.12. Simulations of spectrally dispersed FWM for Tab = 460 fs, w = O, and two values of laser pulse chirp. Again, the density matrix diagrams are time independent throughout the region shown in the spectrograms. (a) For a small value of chirp (¢" = 1000 fsz), the changes in the dynamics are obvious (compare to Fig 2b). These differences are expected, as seen in the density matrix diagram by the appearance of vibrational coherences in both the ground and excited states (upper and lower diagonal blocks). (b) It is possible to calculate the appropriate chirp where ground state elements dominate the contribution to the signal even for Tab = 460 fs where unchirped pulses result in seeing only excited-state dynamics. With ¢" = 4000 fs2 (Figure 3b), 160 fs oscillation periods appear in the simulation, showing primarily ground-state dynamics. 97 Figure 4.13. Experimental spectra (dots) and fit (lines) of the three-pulse FWM signals and the transform-limited laser. (a) F WM signal observed when all three beams coincide in time ( Tab = 0 fs; 1: 0 fs). This spectrum is very broad but shows some spectroscopic transitions. The spectrum is fit with six Gaussian functions. (b) Spectrum of the unchirped femtosecond laser pulses used for these experiments. This profile is fit with a Gaussian function with width 8 nm (FWHM). (c) FWM signal observed when rag, = 614 fs and I: 614 fs. This trace shows well-resolved spectroscopic transitions which were fit by six Gaussian functions. The spacing between the peaks in spectra (a) and (c) is 4 nm, corresponding to the vibrational spacing of the B state of iodine ................................... 100 Figure 4.14. Spectrally dispersed experimental data for three-pulse FWM with Tab = 0 fs and transform-limited pulses. Transients taken at six different wavelengths for 3.3 ps are represented as (a) a topographical plot and (b) a contour plot. As before, the darker areas in the contour plot correspond to the higher signal intensity whereas the topographical plot uses lighter shades to show higher intensity. Spectral profiles were approximated by Gaussian profiles (see Figure 5a and c) to generate the experimental spectrograms. Both ground- and excited-state dynamics are observed with these experimental conditions. 104 Figure 4.15. Spectrally dispersed experimental and theoretical spectrograms of three- pulse F WM for Tab = 614 fs and ¢" = 0 fsz. (a) The spectrogram shows the experimental transients taken at five different wavelengths where the spectral profiles were approximated by Gaussian profiles. For all wavelengths, the dominant feature is the 160 fs vibrational motion reflecting the ground-state dynamics. (b) The initial population distribution was taken to be w = 0.33 for the calculation of the theoretical spectrogram of the experiment shown in (a). The 160 fs vibrations from the ground state dynamics are evident in both the theoretical and experimental spectrograms. ..................................... 105 Figure 4.16. Spectrally dispersed experimental and theoretical spectrograms of three- pulse FWM for Tab = 460 fs and (15" = 0 fsz. (a) The spectrogram generated from the transients taken at five different wavelengths shows only 307 fs oscillations from the excited-state dynamics. There is no evidence of ground state dynamics. (b) The simulation of the experiment again used w = 0.33 for the calculation. The theoretical and experimental spectrograms are in excellent agreement, with the 307 fs oscillations clearly evident in both ................................................................................................................. 107 xiii Figure 4.17. Three-dimensional representation of the experimental spectra taken using differently chirped pulses for the pulse sequence Tab = 460 fs and I: 600 fs. Chirp values ranged from —2800 fs2 to 2800 fsz. Note the change in intensity towards shorter wavelengths as the chirp values increase from negative to positive values. .................. 109 Figure 4.18. Spectrally dispersed experimental and theoretical spectrograms of three- pulse FWM for Tab = 460 fs and (11" = 3300 fsz. (a) The experimental spectrogram was generated from transients obtained with five different wavelengths as above. The signal observed for wavelengths 624 nm and shorter originates from mixed ground- and excited- state dynamics. It is clear when compared with Figure 8a that the addition of chirp alters the observed signal by introducing ground-state dynamics for shorter wavelengths. (b) The theoretical simulation, where w = 0.33, shows good agreement with the experiment both in the dynamics features and in the intensities of the beats. ................................... 110 Figure 4.19. Diagrams of averaged density matrix elements <|p,-,-(2)|2> and corresponding theoretical spectrograms for time delay rag, = 614 fs and w = 0.33 with ¢" = 1000 fs2 and ¢" = 4000 fsz. (a) For small chirps, the dynamics at shorter wavelengths reflect the ground state while higher wavelengths represent signal predominantly from the excited state. (b) When the pulses are chirped to the critical value, (15" = (Zn/(0,22) = 4000 fsz, the signal for all wavelengths reflects excited state dynamics only. Notice that the density matrix diagram in this case shows population but no coherence in the ground state. 112 Figure 5.1. (a) Pulse sequence, corresponding DSFD, and PE measurements of the homogeneous vibronic relaxation between the X and B electronic states of 12. The time delay between pulses b and a(c) was scanned. The DSFD indicates that the signal depends on the off-diagonal terms pge. The logarithm of the amplitude normalized signal FWM emission is plotted as a function of time delay I. The solid lines are exponential fits to the experimental data obtained at different temperatures. (b) Pulse sequence, corresponding DSFD, and RTG measurements of the inhomogeneous vibronic relaxation between the X and B electronic states of 12. The time delay between pulses c and a(b) was scanned. The DSFD shows that signal depends on the off-diagonal terms peg. The logarithm of the amplitude-normalized signal F WM emission is plotted as a function of time delay 1. The solid lines are fits to a Gaussian function of the data obtained at different temperatures. The dashed line corresponds to the exponential decay measured for the PE data at 108 °C (in Figure 5.1a). ...................................................................... 121 Figure 5.2. (a) RTG measurements of the vibronic relaxation between the X and B electronic states of 12. The FWM emission is plotted as a function of time delay I. The solid line is an exponential fit to the experimental data (dots). (b) The pulse sequence and corresponding DSFD for TG measurements of vibrational coherence in the X and B electronic states of 12. The DSF D shows that the signal depends on the elements in the diagonal blocks pgg, pggg pee and peer. The FWM emission is shown as a function of time delay 2'. The solid line is an exponential fit to the experimental data (dots). The dashed line indicates the long-lived component of this signal, which decays with T1. The data in (a) and (b) were obtained in a special cell at 2000 C (see text), and include inhomogeneous and homogenous broadening. ............................................................... 123 xiv th. fur 31 F1: (b Fit Fr. in: 1111 ind Figure 5.3. (a) The sequence and corresponding DSFD for the VE pulse sequences. The DSFD show that the signal depends on the peev and pg'g elements. (b) VE measurements of vibrational coherence in the B (112 =460 fs) and X (112 =614 fs) electronic states of 12. The intensity of FWM emission is shown as a function of time delay I. The solid line is an exponential fit to the experimental data (dots). The data were obtained in a special cell at 200 0C (see text). ......................................................................................................... 124 Figure 5.4. Measurements of vibrational wave-packet spreading: (a) excited state (only the first 12 ps), (b) ground state with the VE pulse sequences as in Figure 3. The decay in modulation amplitude caused by wave packet spreading was fit to an exponential function and is included here (dashed lines). The data in (a) show a wave packet revival at 10 ps. The corresponding power FFT is shown. ........................................................ 125 Figure 6.1. Sketch of the N204 molecule. The N-O distance (a) is 1.19 A , the N-N bond (b) is 1.776 A , and the angle at is 1346", from reference 234. ...................................... 128 Figure 6.2. Experimental time-resolved FWM signal obtained during the first 2 ps: a) Frequency-dispersed signal obtained as a function of time delay for N02, b) Frequency - integrated N02 data as a function of time delay. c) Frequency-dispersed signal as a function of time delay for N204, d) Frequency-integrated N204 data as a function of time delay. For all scans, the signal intensity is plotted on a logarithmic scale. .................... 132 Figure 6.3. Experimental time-resolved and frequency-integrated FWM signal (a) N02 and (b) N204 taken at 363 K and 294 K, respectively. Note the subtle differences in the shape of the rotational recurrences obtained for both cases. The signal intensity is plotted on a logarithmic scale. The vertical lines indicate the position of the rotational recurrences obtained from a simulation using Bave=(B+C)/2. ........................................ 134 Figure 6.4. FFT of the data in Figure 6.3b. The observed peak at 234 cm”1 was assigned to the N-N stretching mode. ............................................................................................ 136 Figure 6.5. Experimental data obtained at 363 K (a) and 294 K (b). The main difference between the two transients, after the first two picoseconds, occurs at 76.265 ps and is indicated by an arrow. Both intensity axes are plotted on a logarithmic scale. .............. 138 XV 1. INTRODUCTION Some of the main questions that preoccupied scientists and eventually led to the development of chemistry as one of the fundamental natural sciences were: How do substances react with each other? Why do they react as they do? Is there a way to control and direct the reaction pathways towards a desired outcome? Answering these questions, chemists (and physicists) generated new ones and developed numerous branches of chemistry. This thesis belongs to the branch of physical i. Reproduced in part with permission from I. Pastirk, E. .I. Brown, Q. Zhang and M. Dantus, J. Chem. Phys. 1998, 108, 4375. Copyright 1998 American Institute of Physics. ii. Reproduced in part with permission from I. Pastirk, V. V. Lozovoy, B. I. Grimberg, E. J. Brown, and M. Dantus, J. Phys. Chem. A, 1999, 103, 10226. Copyright 1999 American Chemical Society. iii. 1. Pastirk, E. J. Brown, I. Grimberg, V. Lozovoy, and M. Dantus, Faraday Discussion 1999, 1 13, 401. Reproduced in part by permission of The Royal Society of Chemistry. iv. Reproduced in part with permission from V.V. Lozovoy, B. I. Grimberg, E. J. Brown, I. Pastirk, and M. Dantus, J. Raman Spectroscopy, 2000, 31, 41. Copyright 2000 John Wiley & Sons v. Reproduced in part with permission from V. V. Lozovoy, I. Pastirk, E. J. Brown, B. I. Grimberg, and M. Dantus, Int. Rev. Phy. Chem. 2000, 19, 531. Copyright 2000 Taylor & Francis vi. Reproduced in part with permission from I. Pastirk, Vadim V. Lozovoy, and Marcos Dantus, Chem. Phys. Lett, 2001, 333, 76. Copyright 2001 Elsevier Science vii. Reproduced in part with permission from V. V. Lozovoy, Igor Pastirk, M. Comstock and M. Dantus, Chem. Phys. Lett, 2001, 266, 205. Copyright 2001 Elsevier Science viii. Reproduced in part with permission from I. Pastirk, M. Comstock, and M. Dantus, Chem. Phys. Lett., 2000, 349, 71. Copyright 2001 Elsevier Science ()1 \K‘. 10' Ill.‘ at: C01” CqL dcr C0}: Th. chemistry known as molecular dynamics. It evolved from studies of rates of chemical reactions and chemical kinetics rooted in van’t Hoffs and Arrhenius’s equations."2 In essence molecular dynamics is concerned about motions of molecules before, during, and after chemical transformations and the rates at which those changes occur. When I started my graduate studies at Michigan State University in the spring of 1997, and shortly after joined Professor Dantus’ research group, the field of laser control of molecular dynamics encompassed a quite wide array of laser techniques with an even wider spectrum of goals. The first attempts at using lasers to drive molecular behavior towards particular aims appeared in late 1960’s, but the investigators encountered numerous problems, such as intramolecular vibrational redistribution (IVR), and the attempts were abandoned. The field of laser control was reestablished in the late 1980’s by several new theoretical approaches, and its experimental achievements were sparked by rapid advancement of ultrashort laser techniques in the early 1990’s. As in any relatively new scientific field laser control was vaguely defined as an investigation of light—matter interaction with the aim of altering the dynamics of evolution of a molecule. The coherence of laser light and quantum interference make coherent and quantum control of chemical reactions possible. Over the last five years the field of laser control of molecular dynamics has grown and diversified rapidly, primarily as a result of the availability of commercial ultrafast equipment, but more importantly as a result of a number of successful experimental demonstrations of its possibilities. The motivation for research in this exciting field is compelling: the idea of using light to alter the outcome of natural molecular processes. The achievements in the field have been reviewed recently3'7 and applications are Ct ft. dc Ill; qu C31 SUL‘ SCQ Eng. numerous, from the pioneering work dealing with selective bond breaking in selected systems,8 through molecular alignment910 and selective excitationll'18 to recent demonstrations of optimal control schemes”23 and possibilities in molecular computing.24 There are two ultimate goals in laser control. The first is to understand the fundamental processes and light-matter interactions to the level that would allow design of the fields required to drive a chemical reaction towards one out of many possible pathways. The second, which comes as a natural extension of the former one, is to ‘produce a mole of product’, and make the method applicable in the chemical industry. Although still experimentally unachieved, this goal now seems realizable, unlike a decade ago. Before practical applications in chemistry, it is likely that applications in a number of other fields such as micromachining or optoelectronics (control of a single quantum dot or construction of a fast optical switch) will develop. The work presented in this thesis deals with exploration and development of the experimental techniques that would allow control of the outcome of molecular evolution (an excitation or dissociation pathway) by controlling the properties of electric field that cause the observed process. 1.1. Observing chemical reactions in real-time It has been established in the 1930’s that the fundamental chemical processes, such as bond breakage and bond formation, occur on the femtosecond (fs = 10'l5 s) time scale.25 Semi-empirical calculations showed that the molecules traveling on the potential energy surface (PES) from their reactant to their product state pass through the transition state of an activated complex. Formulation of the transition state theory26»27 showed that Arrhenius’s pre-exponential factor A (corresponding to frequency of collisions) contains the probability that the reaction will occur. In other words, it describes the probability that the activated complex will cross from the reactant valley to the product valley. The ‘frequencies’ of these passages through the activated complex are on the order of 1013 Hz, which corresponds to the 10'13 s in the time domain. This time line represents the approximate time resolution needed (both in calculations and experiments) in order to follow the evolution of molecules on the reaction coordinate and also the approximate period of molecular vibrations. This observation confirmed the need for the development of time-resolved spectroscopy. The dynamic processes of molecules involve displacement of electrons and nuclei that are three orders of magnitude apart in energy. Applying the Bom-Oppenheimer approximation allows investigation of molecular motion by observing the dynamics of the nuclei. The tools for measurement of such fast processes became available with the discovery of lasers followed by the rapid development of pulsed laser sources. Pulsed lasers allowed progress of time-resolved spectroscopy from the nanosecond in 1960’s to femtosecond techniques in 1980’s. The shape of a particular electronic PES as well as energy spacing between the energy levels can be determined from frequency-resolved spectroscopy. High-resolution information about the investigated molecular systems in the energy (frequency) domain can be obtained using non-coherent sources (flashlamps, white-light bulbs) or coherent sources (continuous-wave, nanosecond pulsed lasers). The time resolution in time-domain spectroscopy depends on the inverse of the coherent bandwidth of the excitation source used. The shorter the excitation pulse, the faster a ll.’ IL" 51. \K. 5? frt brt I‘lU the col no} Sta Um Exr molecular processes can be observed and also the broader the spectral bandwidth. Therefore in order to resolve molecular vibrations or follow the evolution of an activated complex, ultrashort fields are required. The groundbreaking experiment in femtosecond time-resolved spectroscopy was the study of the transition state in the photodissociation reaction of ICN performed at Caltech.28 It demonstrated that the observation of transition states is possible and directly clocked the time duration of a bond breaking reaction. It was soon followed by a series of experiments on several molecular systems that established a new field in molecular dynamics — femtosecond transition-state spectroscopy (FTS).29'32 FTS was successfully applied to numerous systems ranging from diatomics to proteins and to processes in both gas and condensed phases. The basic principles33s34 and achievements of femtosecond spectroscopy in clocking chemical bond breaking or formation, in solving the structure of molecules, or in solvation dynamics are numerous and have been reviewed frequently.”40 Pump-probe and other multipulse methods allow one to study processes as fast as the pulses themselves. The shortest laser pulses obtained so far are 4.8 fs.4"42 The coherence and large spectral bandwidth of fs pulses lead to the formation of nonstationary nuclear wave packets involving translational, vibrational and/or rotational states. The high peak power associated with femtosecond pulses permits investigation of multiphoton and nonlinear processes. The properties of femtosecond pulses that allow time-resolved studies of elementary molecular processes will be described in the experimental chapter of this thesis. {'1 “c mi of b} m Sc co» C0i elk Cht COI 1.2. Controlling the dynamics of molecules The road to laser control of chemical reactions has been one with many important tums.33~43 The invention of the laser gave scientists the hope of controlling chemical reactions that went beyond photochemistry. The monochromatic quality of continuous- wave (CW) lasers fostered the idea that chemical bonds could be selectively broken by multiple photon excitation (MPE) at a specific frequency determined by the vibrational frequency of the particular bond.44'48 It was soon discovered that the energy being deposited in one vibrational mode was quickly dissipated to other intramolecular chemical bonds by a process termed intramolecular vibrational redistribution (IVR).49'Sl This realization meant that the MPE technique inevitably resulted in breakage of the weakest bond. Therefore, the main use of MPE would be to deposit energy in selected molecules in a mixture, based on spectral absorption. Based on a thorough understanding of this method, selective bond-breakage would require excitation of a local mode in a molecular system with extremely slow IVR. This excitation would need to be followed by prompt bond fission achieved with UV excitation or a reactive collision. Vibrationally mediated chemistry was first proposed and demonstrated by Letokhov and coworkers.52 Selective bond breaking of the OH or OD bonds in HOD was proposed by Irnre and coworkers53 and carried out experimentally by Crim and coworkers8 and Valentini and coworkers.54 With the realization that IVR prevented the accumulation of energy in a given chemical bond for periods longer than a picosecond, two different approaches to control chemical reactions were developed independently. One approach was based on the continuous and coherent excitation of two competing reactive pathways. This method at CT 811 C 0 on the of Cht the lec' ma. COIT dubbed coherent control (CC), is based on the constructive or destructive interference among the different photochemical pathways.55~56 This approach has yielded some very remarkable successes in controlling atomic ionization,57 controlling the Outcome of chemical reactions,58 controlling the flow of current in semiconductors,59 and controlling the directionality of photoelectron emission in the gas phase.60 Coherence relaxation is an important parameter that limits the efficiency of this technique. In general, coherence relaxation is slower for gas-phase samples, especially samples cooled by molecular beam expansion. On the other hand, the inhomogeneous and homogeneous broadening that accompanies condensed phase measurements makes this the most challenging environment for CC. To achieve coherent control in the condensed phase, one must ensure that the sample-laser interactions occur on a time scale that is faster than IVR and coherence relaxation. The other approach to control chemical reactions circumventing IVR is founded on the use of ultrafast laser pulses. Conceptually, the laser-sample interactions take place in a time that is short compared to IVR. This concept was embodied in the pump-dump theoryf’l’62 On the experimental front, Zewail, who had been active in the determination of IVR rates,63 realized that shorter pulses would be critical for achieving control of chemical reactions.50 The development of lasers with femtosecond pulse duration made the pursuit of this work possible.‘54'67 Zewail’s group quickly incorporated these techniques and dedicated their work to the study the ultrafast dynamics of chemical reactions in the gas phase.36‘38~68 Ultrafast lasers have been used to make important contributions to the study of chemical reactions in gases, liquids, solids and surfaces.35 The experimental observation of vibrational wave packet dynamics caused by impulsive excitation of multiple vibrational levels using femtosecond pulses indicated that pump-dump control could be used to achieve mode-selective chemistry using ultrafast lasers.69 This is true provided that the timing between the laser pulses is carefully controlled. Most of the femtosecond pump-probe experiments can be considered control experiments, even though the coherence between the laser pulses does not play a role in the observed signals. The timing between the laser pulses controls the species that is excited with the probe laser. Limitations of these types of experiments relate primarily to wave packet spreading and the lack of a well-defined phase between pump and probe pulses. Even for isolated molecules, wave packet spreading results from the anharmonicity of the potential energy surface. This limitation can be addressed with phase altered (chirped) laser pulses, whereby the laser chirp can be used to cause a focusing of the vibrational wave packet at a particular point in time and space.“~70»7I Single chirped laser pulses, in fact, have been used to control multiphoton excitation and chemical reactions.'8’72'74 The main limitation of the use of ultrafast, precisely timed, chirped pulses has been our limited knowledge of the molecular Hamiltonian. A new approach to control of chemical reactions was the proposition that an electric field exists such that a specific target can be optimally achieved.75 This approach to control depends on the search for the optimal laser pulse. The search, initially carried out by computer algorithms43’76 has reached experimental realization in recent years."~12-~19~23~77'80 Optimal control has the distinct advantage that the molecular Hamiltonian does not need to be known. The search is carried out by a number of statistical methods that operate on a trial and error basis. The solution is found by a number of iterations as the optimal field is found by convergence to a maximum dt 01 co lOt arr CO! thL TEL] m:- adi “a 16C} has attainment of the target. This iterative process, without a priori knowledge, implies that the technique can be applied to very complex systems. Interestingly enough, some of the theoretical work in the field of optimal control has concluded that optimal solutions in many cases converge on combinations of a small number of pulses with a specific time delay and phase relationship.“82 This observation highlights the importance of every one of the methods that have been used to control chemical reactions: (a) multiphoton interactions, (b) short pulses, (c) pulse phase coherence, (d) timing between laser pulses, and (e) chirp. The above combination of characteristics can be achieved by a setup combining collinear, phase-locked femtosecond laser pulses. Scherer et al. showed that phase locking provided control over the laser-molecule interaction in a collinear pump-probe arrangement.83 When they changed the phase of the laser pulses by it, their data showed control over the excited state wave packets of iodine. The experimental realization of these measurements required active phase control over the two laser pulses in order to ensure a specific phase between the pulses in the interaction with the sample. This requirement makes the technique difficult. Bergmann and coworkers have developed a method for laser control of rovibrational population transfer based on stimulated Raman adiabatic passage (STIRAP). This technique uses two collinear pulses with different wavelengths that are partially overlapped in time (such that the dump pulse precedes the pump pulse) to achieve near 100% efficiency.84 The combination of degenerate femtosecond laser pulses can be achieved using four—wave mixing methods. These techniques, developed over the last three decades, involve nonlinear optical processes, based on the coherent combination of laser pulses without requiring active phase 0‘. ex be Dc Ch; dis 0111 stabilization.85'87 Coherent Anti Stokes Raman Scattering (CARS) measurements can be considered examples of control over excitation pathways, as described in the recent work of Schmitt et al.“,89 In some cases the emission of signal can be controlled in four-wave mixing (FWM) experiments. This is achieved through the interference of different nonlinear FWM signals.9O 1.3. Outline of the thesis The experimental tools and basic design of the femtosecond lasers used in experiments presented in this thesis are described in Chapter 2. Ultrafast experiments can be performed with a number of different beam arrangements and experimental setups. Details on the ones applied in this work are also presented along with pulse- characterization techniques. Use of phase-altered (chirped) pulses for controlling the outcome of the dissociation of CH212 is demonstrated in Chapter 3. It is shown that even simple changes in ultrashort pulse phase characteristics, such as chirp, can be used to enhance the yield of one over several other possible channels in chemical reaction. The Dantus research group at Michigan State University has recognized nonlinear optical techniques, which use phase matching detection as a means to ensure coherence among a number of laser beams, to be an ideal platform for the coherent combination of laser pulses. Work on FWM, presented in this and in part in Dr. Emily Brown’s thesis,” resulted in number of publications92'103 showing that pulse sequences can take advantage of the vibrational dynamics of the sample molecules for achieving quantum control of electronic and Sill FY- COI qua col~ cor ENC 0rd mm vibrational processes with three degenerate laser pulses. These experimental observations, as well as a theoretical understanding developed,'04 show how the pulse sequences can be used to select among different excitation pathways. It is demonstrated in Chapter 4 and used throughout this thesis that pulse sequences can be used to control the observation of ground- or excited-state vibrational dynamics in three-pulse FWM experiments. One of the reasons for the success of this technique arises from the realization that laser-molecule interactions can take place with two components of the electric field exp[-i(kx-ax)] or exp[i(kx-at)]. Which one interacts first with the molecule changes the nonlinear optical processes that can be observed and controlled. This dual interaction is typically explained in terms of bra and ket interactions. It is described here how this selection is done in the laboratory and what effects the different pulse sequences have on total interaction of the electric field and the sample. Further, Chapter 4 describes the experimental aspects of the spectrally dispersed FWM technique, where coherently combined chirped femtosecond pulses are used to control the intramolecular dynamics of the system.96'98 As with most laser control methods, FWM based techniques are limited by the rate of coherence loss. Realizing that quantum control can only be achieved while the system maintains its coherence, coherence relaxation measurements were carried out (Chapter 5), which use some of the control pulse sequences we have identified to provide information from ground- or excited-state coherenceloo The measurement of the rate of coherence loss is presented in order to illustrate differences caused by inhomogeneous broadening. The methods in Chapters 3-5 represent basic experimental explorations of a model system (the I2 molecule). An extension to somewhat larger molecules is presented 11 in Chapter 6 where short- and long-time dynamics of N204 and N02 are probed. '01 12 am pul alo 2. EXPERIMENTAL (ultrashort lasers and techniques) 2.1. Introduction The first lasers capable of producing sub-picosecond pulses were developed in the early 1980’s.64'67,105 Those original ultrafast laser systems were dye-based lasers and were generally home-built and relatively large. Over the last two decades, the duration of the pulses shortened and systems shrank to the size of desktop computers. The experiments presented in this thesis were performed on two different femtosecond laser systems in Professor Dantus’s laboratory at Michigan State University. The first system, the colliding-pulse mode-locked laser (CPM), is one of the few still operable femtosecond dye lasers of the original design that sparked the development of ultrafast science. It is capable of producing 50-60 fs pulses at 30 Hz. These pulses are typically about 0.3 ml. The second system, a regeneratively amplified titanium-sapphire laser, represents the latest generation of ultrafast technology. It produces sub 50 fs pulses at 1 kHz, with energy close to 1 mJ/pulse. A substantial part of my work on this thesis involved construction, maintenance and characterization of the laser systems and various experimental setups. Ultrashort pulse characterization and both laser systems are described in some detail in this Chapter, along with applied experimental setups and the corresponding detection schemes. l3 2.2. Femtosecond pulses, characterization and experimental techniques The amplification of light in the laser medium is possible if the gain exceeds the loss inside the resonator.‘06'108 This condition depends on the laser medium properties, and it is generally fulfilled only for a narrow range of wavelengths. The number of modes with wavelengths A that can be amplified is determined by the length of the light path L inside the resonator: L=m7t/2, where m is an integer. Wavelengths that satisfy this relation are called longitudinal modes. Using the relation k=c/v, the difference in frequency between two neighboring modes can be expressed as Av=c/2L, where c is the speed of light and v is the frequency. This would mean that an average femtosecond laser centered at 800 nm (pulse duration ~100 fs, with bandwidth of 10 nm and L~2 m) should operate with more than 105 modes. If the laser operates simultaneously in several modes with random phases - the resulting output would be a continuous-wave (CW) beam. In order to obtain a stable train of ultrashort pulses the phase between the different modes operating inside the resonator must be held constant. This process is referred to as mode- locking. Following the development of ultrafast laser sources a large number of experimental ultrafast techniques have been developed. Some of the techniques applied in the exploration of ultrafast phenomena are more suitable for condensed phase studies while others are preferred methods in gas phase and molecular beam experiments. The common denominator in virtually all time-resolved ultrafast experiments is the century old pump-probe principle that was first applied to study a chemical reaction with femtosecond pulses in 1987.28 Its beauty lies in the simplicity of the principle; one pulse initiates a change in a system that is investigated (pump), the system allowed to evolve in time and then another pulse is applied to investigate the state of the system (probe). The work presented in this thesis does not involve the use of the pump—probe technique directly. It involves three pulse methods generically known as four-wave mixing (FWM), which have some similarities and are much more powerful. Starting from the Heisenberg’s relation for the uncertainty between position and momentum for a wave packet, Avath/Z, and the velocity of that wave packet v (group velocity), the uncertainty in the energy can be calculated as AE=A0)2/2m) arm/m) :ApgwAp. The uncertainty in position can be viewed as approximate width of the wave packet in space, which implies that it will be also spread in time. In other words it will take a time AtgAx/v for a wave packet to pass through a certain region. Rearranging these relations yields AEtAIZh/2. Given that E =fl'a), we obtain a frequency- time uncertainty relation for electromagnetic waves, AwwAtzl/Z. This is a very important relation for the Work in this thesis, because it can be used to describe ultrafast pulses. It suggests that for a given spectral width the pulses have a lower limit (minimum) in duration. Such pulses, where the time-bandwidth product reaches its theoretical minimum, are referred to as Fourier transform-limited, or TL pulses. The exact numerical value of this product can vary because the light pulses can assume several temporal profiles.107 Experimentally it is easier to determine the full width at half-maximum (FWHM) of the spectral and temporal profiles. Ultrashort pulses are usually represented by a Gaussian profile 15 fr pr 1hr pu? M, frc f(t)=exp[—t221n2/rz], (2.1) where 2' is the pulse duration. For TL Gaussian pulses, the product AwAt=21n2/7r, where Avis the frequency FWHM (in Hz) and At temporal FWHM (in s). This value is a consequence of Gaussian shape of the pulse and equals 0.4413. The AVvAt product takes generally higher values if different pulse shapes are assumed.‘07 The electric field of a femtosecond pulse can be represented in time as'092110 E(t) = E0f(t)exp[—ia)0t — tag], (2.2) where E; is the amplitude, fit) is function describing the time dependence of the pulse envelope, coo is the pulse central frequency and a is the phase modulation of the pulse in frequency space (temporal linear chirp). Chirp is a natural phenomenon related to propagation of light through the media. For a linear phase function across the spectrum, the pulses are TL. If the function is quadratic, then the pulses are referred to as linearly chirped. The spectral content of the pulses stays unchanged, but the time duration lengthens. The pulse envelope fit) is proportional to the square root of the pulse intensity [(1) and can have a number of different shapes. As mentioned earlier, the shorter the laser pulse the larger its spectral content will be, therefore the pulse duration and its spectral width are related quantities. They are Fourier transforms of each other giving in the frequency domain 16 (60— (1)0)2 —ia'(w —a)0)2 2F 2 6(a)) = .90 exp[- 1. l: 50 exp[— (0) - wo )2 27 (2.3) where y is the complex frequency width defined as y2 =1/rz + ia with Re[1/72]=I/F2 and Im[1/)’7]=a’. The linear chirp a’ is often labeled as ¢” (the notation with ¢”is used throughout the thesis). The bandwidth of the pulses is given by definition as F2 =1/r2+a 212. The temporal and frequency chirp are connected through ¢”I‘2=ar‘?. The phase variation of a TL pulse has a linear time dependence (usually referred to as ‘chirp-free’). It has been demonstrated in numerous experiments that knowledge of both frequency and time profiles of the ultrashort pulses and one’s ability to alter them is necessary in order to obtain correct information about dynamical response of the molecule. One of the most important characteristics of an ultrashort pulse is the phase relation among the different frequencies that are present in the pulse. The phase characteristic of the pulse is caused by the fact that light of different frequencies travels through dispersive media with different speeds (red propagates faster than blue light). The frequency dependent phase Wm) of the pulse can be expanded in a Taylor series around the central frequency (00. The time delay of the frequencies relative to one another is given by the derivative of the total phase with respect to €015,107; 1 d3¢((00) 61mm): demo) +ldzczxwo) 6 dw3 dw do) 2 dwz (w-wo)2+... (2,4) (co—coo)+ The first term represents a constant time-delay, the second term is linear spectral l7 chirp (a’ or ¢ ") and the third term is quadratic chirp. Under experimental conditions, the frequency chirp is more often used (in units of fsz) given that the pulses have a fixed power spectrum with bandwidth F. The pulse duration is therefore proportional to the extent of linear frequency chirp 12 =l/1"2+(¢”1)2,109 2.2.1. Femtosecond pulse characterization Characterization of ultrashort laser pulses requires the determination of the temporal and spectral profile. There are several methods for ultrashort pulse characterization: cross-correlation, interferometric autocorrelation, and frequency resolved optical gating in all of its variations. These techniques share a common principle in that pulses are measured against themselves. This is a consequence of the very short time duration of pulses and the fact that no electronic device can have the response time required. These pulse characterization methods are basically pump-probe ultrafast experiments themselves, relying in general on the instantaneous response of non-linear materials. Since the methods presented in this thesis emphasize the time-energy profile (‘shape’) of ultrashort fields, it is obvious that each experiment requires precise knowledge of the character of the applied femtosecond pulse. All of the methods described below were used. The simplest and fastest optical method for characterization of the time profile of the pulse is known as intensity auto-correlation, also referred to as second-order autocorrelation. The pulsed laser beam is split into two arms of equal intensity by a 1 mm 50/50 beamsplitter in a modified Michelson interferometer. Two beams are arranged to propagate parallel to each other and focused on a thin non-linear doubling (second 18 harmonic generation-SHG) crystal (Figure 2.1). Depending on the wavelength, pulse- duration characterized, and availability, the following crystals were used: lithium triborate (LBO), B-barium borate (BBQ) and potassium dihydrogen phosphate (KDP). There are two major classes of phase matching and SHG crystals. Type I is characterized by the fundamental beams being polarized in either an ordinary (O) or extraordinary (E) direction. The SHG formed is then either E or 0, respectively. Type 11 phase matching requires two fundamental photons of different polarization (0+E), and can give either E or 0 output. PMT HCC or CCD Figure 2.1 Schematic of the experimental setup for an intensity autocorrelation measurement. M - mirrors, BS — 50/50 beamsplitter, HCC - hollow comer cube mounted on a computer controlled variable delay stage, SHG - doubling crystal, L — focusing and collimating lens, UV — filter. All crystals used in described experiments were of Type I. LBO has a wider acceptance bandwidth but also a greater non-linear coefficient than KDP.lll A larger non-linear coefficient implies a larger temporal walk-off, making the crystal less suitable for characterization of very short pulses. The choice of crystal for a particular application is therefore necessarily a trade-off. The ideal thickness of a SHG crystal used in 19 £11 or at autocorrelation measurements ranges from 10 to 300 um, depending on the material and duration of the pulses measured (see Figure 2.2). For ~ 50 fs pulses, 200-300 um crystals are successfully used, while for pulses of 30 fs and shorter, 50-100 um crystals are necessary due to the huge spectral bandwidth of such a short pulse. For characterization of pulses from titanium-sapphire seed laser (13 fs), 50 um LBO is used due to its large acceptance bandwidth, while for measuring the amplified pulses (~50 fs), 300 pm BBC is used. As the interaction length in the crystal increases, the group velocity mismatch can greatly affect efficiency of SHG and the recorded signal. 2.6 _' e 2.4 :' —I—BBO ~22 - —A—LBO £20 I —e—KDP "€018 j 531.6 f .51-4 r g 1.2 _- e 9-»; 1.0 f .9 0.8 _' g O_6 : O/: §O.4 " e/./ 00 /...i....1..,.1....1. O 50 100 150 200 duration of TL pulse [fs] Figure 2.2. Quasistatic interaction length (thickness) of selected crystals as a function of the duration of transform limited pulses for Type I SHG at 800 nm. It is very important for the fidelity of the measurements that the two beams are of equal intensity and they experience an identical optical path. In order to avoid unwanted 20 non—linear processes in the SHG crystal (self phase modulation, self-focusing or continuum generation) the beams should be attenuated. For the same reason (avoiding strong fields), the two beams should cross at the crystal slightly before their focal point. The time delay of one pulse with respect to the other is changed by altering its optical path using a hollow comer cube mounted on a computer controlled actuator (See Figure 2.1). The autocorrelation signal appears between the two incident beams after summation in the SHG crystal. It is detected after spectrally filtering the fundamental frequency with a UV-filter, and spatially filtering the two doubled beams by a fast photo-diode or a PMT. A typical autocorrelation trace is shown in Figure 2.3. The pulse duration is determined by fitting the curve and determining the FWHM of the autocorrelation function, Ar. If the function has a Gaussian profile the pulse duration is A1=AT/\/2. For a sech2 profile, At=At/1.54. 1.0 0.8 0.6 0.4 Intensity [arb. units] 0.2 -200 A o ’ 200 time delay [fs] Figure 2.3. Typical autocorrelation trace (triangles) obtained for amplified titanium- sapphire fs laser pulses with a Gaussian function fit. Duration of the pulses is 50 fs. 21 This method can be used for fast determination of the extent of linear chirp ¢’ ’. A series of experiments where the compressor prism (see following sections) is translated along its bisector can be used. The shortest pulse duration is assumed to be transform- limited (7,") and linear chirp (in fsz) is determined from the measured pulse duration (row) by:112 2 2 _ 2 ¢H=\/Tin(rri;f2 Tin) (25) where F4ln2. Another common method used for the characterization of ultrashort pulses is frequency resolved optical gating (FROG), which relies on the instantaneous response of a non-linear optical medium.“3" ‘7 There are numerous variations of this method. The one presented here and used routinely in our lab for pulse characterization is based on the polarization grating effect P-FROG. Two beams of equal path length and different intensities are focused by a cylindrical lens on a non-linear material, usually a thin quartz plate. The strong beam (pump) and weak beam (probe) are polarized at 45° relative to each other. When the pulses coincide in space and time, the intensity of the probe produces an optical Kerr effect (an anisotropic change in the refractive index) in the quartz plate, which rotates the polarization of the probe pulse and allows it to transmit through a polarizer. The beams are focused in a vertical line, thus producing a time delay in the vertical axis. Temporal information (the pulse profile) is then carried in the vertical dimension. The signal is dispersed in a grating to obtain the spectrum along the horizontal axis. The spectrally resolved signal is detected by a charge-coupled device 22 _ as.» I’r‘ [er pla ol‘t iter; retri tauls chirr latt 5mm 31m( (CCD) camera. The software that captures the signal from the CCD changes the axes to produce the image on the monitor, with spectral information encoded vertically and temporal information encoded horizontally. A FROG trace is thus obtained in the (co, 1) plane. Typical responses of FROG for transform limited and linearly chirped pulses obtained on a commercially available FROG apparatus (Clark-MXR) are shown in Figure 2.4. The two-dimensional array of data is then analyzed by commercially available iterative algorithm on the PC,1 ‘7 and the pulse duration and its phase characteristics are retrieved. The method provides fast identification of the sign of the linear chirp. If the pulses are positively chirped (see Figure 2.4b), the slope is positive while for the negative chirp the image has a negative slope. The only downsides of P—FROG are the relatively low dynamic range (only pulses of ~ 10 u] and greater can be characterized) and somewhat time-demanding alignment. especially if its intended for analyzing the pulses at more than one central wavelength. frequency time Figure 2.4. Typical P-FROG trace obtained for transform-limited (a) and positively chirped (b) pulses from an amplified CPM laser with central wavelength of 620 nm. 23 m] length In wave FigUr lllanf the h rem e The SHG-FROG is a more sensitive method to characterize laser pulses. It can be used to measure very weak (< 1 n1) pulses. The setup is identical to a second-harmonic autocorrelator, with the difference that the signal is spectrally resolved. The two- dimensional (or, 1) signal is recorded. To obtain SHG-FROG traces with beams of different wavelengths, the conversion requires a small change in the incidence angle of the SHG crystal (phase-matching angle). A typical trace of SHG FROG with phase information retrieved by sofiware is presented in Figure 2.5. The crystal used for this measurement was 300 pm thick BBQ, and the 1 kHz amplified titanium sapphire pulses centered at 804 nm were attenuated to 1 14.1. 1.0 1.0 , . . , 410 a) b) c) g 405- '_.‘ .—, ._. E ,3. E) 30.5- no.5? - E 8 g 400 l- ? J 395 . fi . 0,0 - I 0.0 . 4 1 -100 0 100 -100 0 100 780 800 820 time 115] time [fs] wavelength [nm] Figure 2.5. a) Experimental SHG FROG trace of near TL pulses from an amplified titanium sapphire laser, obtained by CCD and spectrometer. Darker shades correspond to the higher intensity. b) Temporal profile (thin line) and phase dependence (thick line) retrieved from data in a) by FROG software. c) Spectral (thin line) and phase (thick line) 24 u _-- P” ham pul~ cha init: con> Prep The to in profiles of the pulses retrieved by FROG software. Pulses were 55 fs with time- bandwidth product of 0.4646. Phase variation of about 0.1 radians across most of the pulse duration in b) is considered to be a minor deviation from TL. The spectral phase characteristic in c) signifies minor higher order chirp. 2.2.2. One-pulse excitation The experiments presented in Chapter 3 use a simple well-characterized pulse to initiate control over the yield of a known photochemical reaction. The experiment consists of fs pulse preparation, excitation of the sample and fluorescence detection. Preparation of the pulses provides a fs pulse of known phase and spectral characteristics. The pulses can be positively or negatively chirped by a prism-pair arrangement in order to introduce positive or negative frequency chirp. Femtosecond pulses at the fundamental frequency or their second harmonic (See Figure 2.6) were focused on a gaseous sample. The extent of the chirp introduced in pulses was controlled by a prism-pair arrangement from the pulse compression part of the laser setup. One of the prisms was mounted on a computer-controlled actuator. Advancing or retarding the prisms introduces more or less glass in the optical path, and therefore introduces linear chirp. This setup provides precise control of the amount of the dispersive medium that the fs pulse encounters and allows one to chirp the pulses linearly. The intensity of the laser pulses was monitored by a reference photodiode (rPD) in order to compensate for system instabilities and perform discrimination. Pulses were characterized by FROG as described in the previous section. The fluorescence produced 25 by fin: tub r651 Figu C0m | \ Samr by the sample was collected perpendicular to the beam propagation. The fluorescence is further spectrally filtered through an imaging spectrometer, detected by a photo multiplier tube (PMT) and sent to a boxcar averager. The analog signal is digitalized by an A/D converter. The signal was recorded on a personal computer as a function of variable parameter (chirp), using LabView programs. This kind of setup does not provide time resolved data (on the ultrafast scale). l\%__ signal rPD ‘ ‘ detection shg \ BS 36am ‘/ : cell : i / ~310 nm " ' ' ' ' M from laser ~620 nm Figure 2.6. Experimental setup for one-beam experiment. P1 and P2 - double-pass compressor prisms. P2 is mounted on an actuator. Mirrors (M) direct the beam to an SHG crystal where the frequency of the pulses is doubled, and pulses are focused on the sample cell. rPD — reference photodiode used for discrimination. 26 2.2.3. Three—pulse FWM Several experiments presented in this thesis are based on the three-pulse F WM method. The beam arrangement is shown in Figure 2.7. Its principles are described in detail in chapter 4. The main experimental requirements and considerations are discussed here. The beam from an fs laser is split into three arms by two consecutive 1 mm thick beamsplitters. The optical pathway of two of the beams is variable in order to allow time ordering of the pulses. By changing the length of the path along which the pulse travels across the optical table, one controls the delay of the pulses with respect to each other. One of the hollow comer cubes (HCC) is mounted on a computer-controlled actuator (stepper motor) capable of advancing 0.2 pm per step. This translates into 4/3 fs per step. The second variable delay with HCC was mounted on a micrometer and its delay with respect to other pulses was manually controlled. The three beams were arranged in the forward box geometry (Figure 2.7 top) and focused by a 2 inch diameter 50 cm focal length lens onto a sample cell. Two aluminum templates were machined and positioned perpendicular to the axis of beam propagation, in order to ensure that the beams are parallel before focusing and after the collimation lenses. Alignment is relatively simple; three beams have to be parallel to each other and have to pass through three holes of the template that is perpendicular to the beam propagation. If the beams are aligned properly they will spatially overlap at the focal point. For temporal overlap it is necessary to find the ‘time zero’, the length of each of the three arms at which the pulses coincide in time. This is performed by placing the SHG crystal or a thin (~100 um) quartz plate at the focal point, and by retarding or advancing variable delays with hollow corner cubes until the diffraction pattern is observed at the screen. The FWM signal appears in the fourth comer 27 of the square at the screen; it is collimated, spatially filtered (to avoid any scattering signal), and detected either by the PMT or the CCD through a spectrometer. from laser HCC HCC A» /‘\$ ks=ka'kb+kc , , ,' signal ks spatial filter ,’ k ' sample cell ' / . BS BS\ ’ ' L L Figure 2.7. Beam arrangement for the three-pulse four-wave mixing experiment. BS — beamsplitters; HCC — hollow comer cubes mounted on variable delay stages; L —focusing and collimating lenses. Detail of the spatial overlap is also shown (top) with the direction of the signal, detected in the phase-matching direction ks. 2.3. Colliding Pulse Mode-locked Laser System (CPM) The system used for the experiments described in Chapters 3-5 is a home-built 28 amplified colliding-pulse mode-locked dye laser (CPM). It is described in detail elsewhere.64'67-~9'-‘01l 13,119 In short, the laser oscillator is a ring cavity consisting of 7 mirrors (M and CM), 4 prisms (P), and 2 dye jets. Rhodarnine tetrafluoroborate (R6G), a gain dye, dissolved in ethylene glycol, flows through one of the dye jets. The 532 nm line of a continuous wave Spectra-Physics Millenia laser (typically operating at 1.8 W) is used to pump the R6G dye optically to an excited vibrational level in the first excited state. The excited dye molecules emit broadband fluorescence as they relax to the ground state. The ring-cavity arrangement (Figure 2.8) allows light to propagate in both counter-clockwise and clockwise directions. The output coupler at the right edge of the CPM cavity is partially transmissive (2%) and some light is allowed to exit the cavity. The counter-clockwise propagating beam is used as a reference while the clockwise propagating beam is sent to the four-stage amplifier. This continuous-wave (CW) laser can operate in the 580-620 nm range. In this laser cavity a large number of different frequencies can exist, all oscillating with unrelated phases (multimode operation); thus the laser is continuous wave. In order to produce pulses, the phase relationship between all the different frequencies must be set and maintained constant in time. If the initial phases of all of the modes are equal, one can obtain pulses and the laser is referred to as c mode-locked’ .64-66, 105.120-122 Mode-locking is achieved using a saturable absorber dye. In the CPM it is 3,3’- diethyloxadicarbocyanine iodide 2,3 (DODCI) dissolved in ethylene glycol circulating through the other dye jet (absorber jet). The saturable absorber dye’s absorbance is intensity-dependent, resulting in no light passing through it when the light intensity is 29 low. When a higher intensity light, caused by a fluctuations in the laser cavity, encounters 30 computer ’4 ‘ —' detection Tbox- system car sync. ' ....... ___‘_l I experimental sample cell ’ l set-up Nd:YAG laser ........ j, _ _ _ _ _ _ _ |Nd:Yvo4|4x \ 532 “m I 532 nm T PD P N / M P P .. P OC . ________ ‘ r“! 1"] L_] r pt” U LJ pressor ’ ’>CM 620 nm . M M amplifier absorber jet gain jet ! \ o t Figure 2.8. Experimental setup for CPM dye laser with general detection scheme. M and CM - mirrors and curved mirrors, respectively; P — prisms; OC — output coupler. Nd:YAG laser and detection are triggered and synchronized to the pulses detected by photodiode PD. the saturable absorber dye-jet it saturates the absorbance transition. This allows passage of some light through the saturable absorber. This light passes through the gain dye where it is amplified and then easily passes through the saturable absorber on the next trip around the cavity. This regimen through the gain and saturable absorber dyes repeats until the gain dye is saturated. The leading edge of the formed pulse is absorbed in subsequent round-trips by the saturable absorber (because of its lower intensity) while the trailing edge is not amplified by the gain dye (because of saturation). This allows shortening of the time duration of the pulse resulting in femtosecond pulses. The final 31 time-energy profile of the pulse is obtained when the pulse reaches a steady-state condition and does not change as it propagates through the cavity. A requirement for stable and efficient operation of the CPM laser is that the two counter-propagating laser beams must collide in time and space at two points in the cavity. The optimal situation occurs when they collide at the saturable absorber dye. In this case there is less loss of intensity at the saturable absorber. Collision at the gain dye would be the worst situation, as two counter-propagating pulses would be competing with each other for amplification in the gain jet. The maximum recovery time for the gain dye is achieved if the pulses collide first at the saturable absorber jet, and later as far from the gain jet as possible. This occurs if two jets are placed at a distance apart equal to 'A of the length of the cavity. Once mode-locked, the laser operates at the central wavelength of about 622 nm with a bandwidth of up to 10 nm (245-300 cm"). The repetition rate is 100 MHz and the energy per pulse is on the order of 200 pJ. The refractive indices of different media vary differently with respect to the wavelength (frequency) of the light. This results in a naturally occurring phenomenon known as group velocity dispersion (GVD). It arises when different wavelengths travel at different velocities in a particular medium. It is often referred to as chirp. This phenomenon will be used as a tool in the control of molecular dynamics as described in Chapters 3 and 4. The prisms (P) in the laser oscillator are used to compensate for temporal broadening of the pulses that is introduced by media (ethylene glycol and mirror coatings) in the ring cavity, to produce the shortest pulses that thegiven bandwidth will support (transform-limited pulses). 32 The energy of these femtosecond pulses is generally not large enough for gas phase experiments, therefore the pulses need to be amplified. A four-stage dye-laser amplifier was built in our laboratory by Marcos Dantus and Una Marvet. Una’s thesis contains a more detailed description of the amplification process and construction requirements. In short, the amplifier (Figure 2.8) consists of four Bethune cells (BC) filled with circulating gain dyes. Cells are transversely pumped by the frequency-doubled (532 nm) output of a 30 Hz flash-lamp pumped Nd:YAG laser (typically 10 W or 333 mJ energy per pulse). The Nd:YAG laser is synchronized with the CPM by the SMl synchronization electronics, which ensures efficient amplification of the femtosecond pulses. The gain media is kiton red dye dissolved in water in the first cell, and rhodamine 640 dissolved in water in the other three cells. Femtosecond pulses from the CPM are amplified by stimulated emission from the excitation processes of these dyes. During the amplification process, the pulses travel through water and glass in the amplifier cells and are broadened by GVD. This is corrected after the amplification by introducing negative GVD, which is achieved by a double-pass prism pair (see Figure 2.5 bottom).123 The pulses finally available for experiment were transform limited, 47-60 fs in duration, and had energy of 0.4 ml per pulse. 2.4. Titanium sapphire laser system Titanium doped sapphire (A1203) is a great laser material for ultrashort pulse generation because of its broad gain cross-section. When pumped by a CW frequency doubled laser (NszVO4) or an Ar+ laser, the TizSapphire emission band (fluorescence), centered around 800 nm with a 230 nm bandwidth, is capable of sustaining pulses as 33 short as 5 fs. That corresponds to only 2-3 optical cycles of the electric field. The first sub-100 fs Ti:Sapphire lasers were built a decade ago, exploiting the material’s high damage threshold.‘24'127 The mechanism necessary for modelocking Ti:Sapphire lasers with a large number of longitudinal modes is called Kerr lens modelocking (KLM) or self-modelocking. The Kerr lens effect relies on a relative index of refraction that is intensity dependent. If pulses of relatively high intensity are present in the laser resonator, they will cause self-focusing in the lasing medium (Ti:Sapphire rod). This effect is present in all longitudinally pumped lasers to some extent and by careful construction of the laser oscillator it can be made a preferred mode of operation. In order to initiate and maintain modelocked operation and Kerr lensing, the pulses within the Ti:Sapphire crystal must be kept short. During the round trip though the laser cavity the pulse traverses the crystal and disperses (gains in duration). This group velocity dispersion (GVD) is compensated by a prism-pair located in one arm of the laser cavity. A typical Ti:Sapphire laser resonator is shown in Figure 2.9. Pulses generated in resonators similar to the one described here have been reported with pulses as short as 4.8 fs when compressed.“42 The one built in our laboratory and used in the experiments described in Chapter 6 produces 13 fs pulses when compressed. The repetition rate of the laser is 87 MHz, and the pulses are centered at 800 nm with an energy of 6 nJ/pulse. The oscillator was constructed from a commercially available kit (KM-labs) and is relatively easy to setup and maintain. The titanium-sapphire rod in the oscillator is pumped by 3.5-4.5 W of frequency-doubled CW Nd:YVO4 laser light (532 nm). Its fluorescence is collimated by curved mirrors and directed along two arms of the oscillator. The laser readily operates in CW mode, and by 34 cw NszVO4 laser \{VI ’ CCD Figure 2.9. Titanium-sapphire laser oscillator (seed laser). Light from a NszVO4 laser is focused by lens (L) through one of the curved mirrors (CM) on the titanium-sapphire crystal. CM are transparent for 532 nm light, but 100 % reflective in the 760-850 nm region. Emission from the crystal is collected by CM and directed through prisms P1 and P2 to the end mirror (EM). Length of the cavity (distance from EM to output coupler 0C) is ~ 2m. Pulses shorter than 15 fs when compressed and centered around 800 nm with FWHM of more than 50 nm were obtained. For the proper operation of the amplifier, presented in the next figure, much less bandwidth is needed (25-30 mm) and the oscillator is adjusted accordingly. For fast determination of the bandwidth, one reflection of P2 is directed from the diffraction grating to an inexpensive CCD and TV monitor. The image can be calibrated and used for rough estimation of pulse bandwidth. the mode adjustment it starts self mode-locking. The mode-locking can be achieved by introducing a small physical perturbation to the oscillator, such as jiggling one of the prisms. If properly aligned, the central wavelength is tunable in the 780-820 nm range even in the mode-locked regime. Peak to peak stability is about 1%. Although potentially very high at the experimental sample (~109 W/cmz), these energy densities are not enough to perform non-linear experiments on gas-phase samples, and therefore need to be amplified. One of the best ways to date to amplify ultrashort pulses generated in solid- state media is a technique called chirped pulse amplification.128 In the CPA technique, pulses from a high repetition rate (~90 MHz) femtosecond oscillator (seed pulses) are stretched in time by introducing the frequency chirp to the pulse (see Figure 2.10). This is performed in the stretcher that consists of a prism pair (folded one-prism arrangement is 35 usually used to save space and cost of gratings). This increases the duration of the pulses by a factor of 104 (100 ps), and allows the extraction of high energy from the gain medium (another Ti:Sapphire crystal) in the amplifier, while keeping the peak power below the damage threshold. Afier temporal broadening the pulses are injected into the 1 kHz frequency doubled NszLF (523.5 nm) pumped regenerative amplifier. Two intra- cavity pockels cells (PC) and half-waveplate (M2) regulate the number of the pulse’s passes through the amplifier resonator (Figure 2.10). from T \J from ns 1 kHz pump Hui PCl laser Figure 2.10. Schematic for regenerative amplification of pulses from titanium-sapphire laser. Beam from the seed laser is directed through the Faraday isolator (F1) to diffraction grating G(l) in the stretcher. F1 acts as filter preventing the back-reflected pulses from the amplifier from reentering the seed laser cavity. The beam diffracted from G1 reflects from silver mirror SM to long mirror LM and back. Still expanded, the beam travels to periscope P and back for another round trip (G1, SM, LM and back). Spatially recompressed, but chirped to ~100 ps, the beam is sent to an amplifier (see text). After the pulse has gained energy in the resonator, it is ejected by PC2 and directed through a pulse compressor that consists of another grating pair that compensate the chirp introduced in the stretcher. This configuration of the amplifier generally allows pulse energies up to 1 mJ at a 1 kHz repetition rate and duration of <50 fs, centered at 800 nm. In order to obtain stable TL pulses the alignment of the stretcher-amplifier must be perfect. Higher order chirps are easily introduced. The unit used in Dantus laboratory is commercially available Spitfire (SP Lasers) regenerative amplifier. 37 3. QUANTUM CONTROL OF THE YIELD OF A CHEMICAL REACTION BY SINGLE FEMTOSECOND PULSES 3.1. Introduction The goal for active laser control is to devise electromagnetic fields that drive the outcome of a chemical reaction in the desired direction.‘l3~‘29'l32 There are basically two approaches to this problem. The frequency-resolved scheme (also known as coherent control), proposed by Brumer and Shapiro, utilizes quantum interference between different pathways to a final state to exert control over the outcome.'33’134 One of the most striking demonstrations of this scheme is found in the work of Gordon and coworkers where they control autoionization versus predissociation in H1 and DI molecules.‘35 The time-resolved scheme (also known as pump-dump), proposed by Tannor and Rice, exploits the time-dependent motion of wave packets created by ultrafast (usually femtosecond) laser pulses to manipulate the outcome of the reaction.62~I36 Experimental demonstrations of this control scheme are found in many pump-probe time- resolved experiments, for example, the excitation of 12 to produce either the D (’53:) or the F (’Zu”) states,‘37 or the production of Na’" or Na2+ as a function of time delay between pump and probe pulses.‘38 The application of chirped pulses to shape nuclear wave packets and enhance vibrational coherence was proposed by Ruhman and Kosloff.‘39 Rabitz first formalized the search for an optimal electromagnetic field in terms of spectral and temporal composition to control the outcome of a chemical 38 reaction,140 while Wilson, Mukarnel and coworkers generalized this analysis to obtain a formalism that is more amenable for the study of thermal ensembles of molecules.1 L73 Boers et al. demonstrated the use of chirped pulses to enhance the population transfer in the three-state ladder, Ss—>5p—>5d, of the rubidium atom. 1‘“ Experimental and theoretical studies on the effect of chirped pulses on the multiphoton excitation of molecules showed that the traditional saturation limits could be exceeded, thereby facilitating population inversion. The groups of Shank, Wilson, Gerber, and Leone have recently shown experimental evidence that tailored and phase-altered femtosecond pulses can be used to modify the initial wave packet formed by the excitation laser'3~'6’l42'I46 and, in some cases, tailored pulses can be used to enhance single and multiphoton excitation.'7,'3 The results presented here provide evidence of the effect of chirp on the yield of a photodissociation pathway for polyatomic molecules.72 The product state (i. e. electronic, vibrational, rotational and translational) distribution resulting from the photodissociation of polyatomic molecules depends upon the potential energy surfaces participating in the fragmentation process, along with their couplings and the characteristics of the incident electromagnetic field. The field is characterized by its frequency, duration, intensity and chirp. Chirp (see section 2.2) is naturally caused by the propagation of laser pulses through matter, which leads to group velocity variations as a function of frequency within the pulse, thus causing a frequency sweep.l ‘2 In most cases the group velocity variation causes a positive chirp in which the leading edge of the pulse is red-shifted and the trailing edge is blue-shifted with respect to the central frequency of the pulse. Negative chirp corresponds to the opposite effect. Increases in absolute chirp lead to a temporal broadening of the pulse and are usually 39 considered detrimental for the study of ultrafast phenomena, where the best time resolution is required. Recently, chirp has been recognized as an important parameter that can be used to control the dynamics of a system excited by ultrafast laser pulses.13’16' l8,l 10,142-144,l47 The shape and time evolution of a quantum mechanical wave packet Mt) produced through absorption of an ultrafast laser pulse are determined, in part, by the phase factors in the following expression, —‘E,, M W) =Xane ’ ’ c0. (1) n where E, and (0,, denote the eigenvalue and eigenfunctions of each level n. The quantity a" is given by the Franck-Condon overlap between the initial and each final state n. The initial phase factor for each eigenstate, the quantity in the exponential, is equal for all states when excitation takes place with transform limited pulses, i. e. no chirp. However, chirped pulse excitation opens the possibility for introducing different initial phases. Therefore, the shape and dynamics of the wave packet can be controlled by chirped pulses with the goal of affecting the outcome of a chemical process. 3.2. Experimental The following experiments were carried out with a home built femtosecond laser system consisting of a colliding pulse mode—locked dye laser (CPM) described in detail in Chapter 2. The CPM output was amplified in a pulsed dye amplifier. The amplified pulses were compressed by a two—prism double—pass pulse compressor. The laser system 40 produced pulses with 47 fs temporal width and with the central wavelength at 624 nm. Typical pulse energy was 0.4 mJ (30 Hz repetition rate), but was attenuated for the experiments to 70 al. In cases where 312 nm femtosecond pulses were necessary, the 624 nm beam was frequency—doubled by a KDP crystal (0.1 mm) producing ~ 7 pl pulses. The pulses were then focused into a quartz cell containing the gaseous sample. Fluorescence signals originating from the cell were collected along the direction perpendicular to the laser beam propagation through a 0.27 meter spectrometer. The laser intensity of the 312 nm beam was continuously monitored by a reference photodiode. One of the compression prisms was mounted on a computer-controlled actuator (See Figure 2.5). Translation of the prism along its bisector varied the amount of glass through which the beam propagated and permitted the introduction of linear chirp. Because the amount of chirp was computer controlled, scans of signal intensity as a firnction of chirp could be achieved in a convenient and fast manner. The extent of positive or negative chirp was limited by thickness of the prisms. Characterization of the femtosecond pulses was carried out by frequency resolved optical gating (FROG). FROG measurements in our laboratory provided direct measurements of the frequency components of each pulse as a function of time, from which the vale of linear chirp was determined. A continuous chirp scale as a function of prism position was compiled using the expression for group velocity dispersion as a function of prism position derived by Fork et al.121 and simplified by Salin and Brun.‘23 Quartz cells containing iodine, 12, or methylene iodide, CH212 (Aldrich 99%), were prepared on a vacuum line, and were degassed to less than 10‘6 Torr. Iodine 4l scavenging agents (sodium thiosulphate and copper) were introduced in the CH212 cell to ensure that the signal derived only from nascent iodine. Experiments were carried out at room temperature (21° C) with vapor pressures of 0.25 and 1.2 Torr for 12 and CH212 respectively. 3.3. Results and discussion Spectral profiles for the 624 nm and 312 nm pulses measured at zero, negative and positive chirps were obtained in order to confirm experimentally that the introduced chirp did not affect the spectrum of the pulses. The intensity variations of the 624 nm and 312 nm pulses as a function of chirp were also measured. The intensity of the fundamental remained essentially constant, whereas a small variation in the UV intensity was observed. This variation can be understood in terms of the dependence of second harmonic generation on the peak intensity of the incoming pulses. As absolute chirp increases, the pulse width increases, resulting in a reduction of peak intensity. The concerted elimination of 12 molecules following irradiation of CH212 is known to occur for excitation energies larger than ~ 9.4 eV.‘48 The yield of this molecular process has been measured to be on the order of 1%. Most other pathways produce atomic iodine in its ground (I) and spin orbit excited states (1.). The nascent I2 molecules are formed primarily in the D' state and have been detected following single and multiphoton excitation.‘49‘l5| In our laboratory, we have explored the femtosecond dynamics of the molecular detachment process and have found it to take place in less than 50 fs.‘52 We have also found that other halogenated alkanes with formulas CX2Y2 42 for (X = H, D, F and Y = Cl, Br and I) undergo similar molecular detachment processes.'53~'54 Detection of the molecular product is selectively carried out by dispersing the laser induced fluorescence and collecting only a spectral window that contains the majority of the D'—>A' emission. Multiphoton excitation of CH212 with 624 nm pulses produces the well known 12 D'—>A’ emission at 342 nm. Figure 3.1a presents variations in the relative yield of the molecular detachment pathway as a function of linear chirp. The data have been normalized to unity at zero chirp. Note that the maximum yield is observed for -500 fs2 and the minimum at 2,400 fsz. We observe a factor of 2.9 in the overall change in the yield for this pathway as a function of chirp. These data are contrasted with a chirp scan obtained under similar conditions with the 624 nm pulses on the yield from three photon excitation of 12 vapor to yield D——>X and D'—>A' fluorescence between 320 - 345 nm. These results are presented in Figure 3.1b and show an opposite trend to that obtained for CH212 dissociation. The minimum yield for the three photon excitation is found at -500 fs2 but increases by ~ 40 % as the magnitude of chirp increases. Wilson and coworkers have investigated the chirp effect on the three-photon absorption yield for 12 at 550 nm, 570 nm and 600 nm.'7sl8 They found that the yields are significantly affected by the chirp (approximately factors of 2 and 3 for 600 nm and 570 nm excitation respectively). Our results on 12 are in very good agreement with their findings. We have also explored the effect of chirp on the multiphoton excitation of CH212 with 312 nm laser pulses. Figure 3.2 presents the yield of the molecular pathway determined by detection of 12 D'—>A' fluorescence intensity as a function of chirp. For 43 1.2 1.0 ’5 0.8 ‘9' T INK» D 4,; _ . ,3, 342m N 0.6 -— 624rm “3 _l_ 5- X >" ' ‘ ' -1000 O 1.6 ‘ I ' E" . D' E Z, .0' 1.4 - 34211.“ ‘ ‘ - .5. 624m L A, W 8 L . J 5 - X 3' 8 1.2 r , fl "5’ - ‘5 w :s Q a W ._,N 1.0 " -1000 ' o 1000 2000 Linear Chirp [fsz] Figure 3.1 (a) Experimental measurement of the yield of the molecular pathway producing 12 from the multiphoton dissociation of CH212 with 624 nm laser pulses as a firnction of chirp. The insert shows the relevant energetics for the reaction. (b) Fluorescence yield as a function of chirp following the multiphoton excitation of 12 vapor. these experiments, both the fluorescence signal and 312 nm laser intensity were recorded and averaged for 500 laser shots at each chirp value. The fluorescence intensity was normalized against the third power of the 312 nm laser intensity (because of the three photon excitation and the inherent variation of 312 nm intensity as a function of chirp). It is clear that increasing the chirp enhances the photodissociation yield significantly. if 44 The molecular pathway enhancement is found to be non-symmetric, favoring positive over negative chirps. This observation implies that the observed enhancements are not due to pulse width effects but rather depend on the magnitude and sign of the linear chirp. The three scans shown in Figure 3.2 were obtained under identical conditions except for the intentional changes in the pulse intensity from 0.8x1012 to 1.6x1012 W/cm2 (calculated for zero chirp pulses). The data is shown normalized to laser pulse energy but is not corrected for variations in peak intensity caused by pulse broadening as a function of chirp. The effect of laser intensity on these control experiments is shown in the insert, where the molecular pathway yield is shown to increase for positive 2400 fs2 chirp by factors of 3 to 25 as the zero-chirp intensity of the laser pulses is increased from 0.8x10’2 to 2.4x1012 W/cmz, see Figure 3.2 insert. We consider the differences between the dependence of the molecular pathway yield on laser pulse chirp for 624 nm and 312 nm excitation to be of great interest. For 624 nm, the yield decreases with absolute chirp, while for 312 nm it increases. We have not taken into account the fact that the temporal pulse width of the pulses increases with linear chirp. Multiphoton transitions are, in general, expected to increase by factors proportional to the peak intensity raised to the nth power (1") where n is the number of photons. Therefore, multiphoton excitation is expected to be maximized when the chirp is zero. For 624 nm excitation, the maximum yield was found for a chirp of -500 fsz; however, for 312 nm the maximum yield in the available chirp range was found for a chirp of 2400 fsz. Based on the pulse width change and the three photon excitation for the 312 nm case, the transition probability, proportional to (1":3), is expected to decrease by a factor of 27 (because the pulse width triples at this chirp value). Therefore, the yield of 12 45 10 . 1 m I f r ' l _ 25 L 12 Yield ‘ o— 1.61012W/cm2 J 15 - / -_ 1.1 1012 W/cm2 8 _ r V— 0.8 11112 WIcm2 _. 5 ' Mr - 1 o .75.. - 0.5 1.0 1.512 2.0 2 2.5 ~ g Laser Intensrty 10 [W/cm ] e 6 _ O 7 1531 \\\\\ - 4‘1 _ IL:— Dr q .5 g . / % 4 — 312 nm 342 nm / /- '- 5‘ l iA' o 0.; l- X I /- .. ._ 2 Dd 2 _ o ‘1 __ V—“v9\"\K~ .ég/ _, 0 . 1 . 1 . 1 . 1 r -1000 0 1000 2000 Linear Chirp [fsz] Figure 3.2 Experimental measurement of the yield of the molecular pathway producing 12 from the multiphoton dissociation of CH212 with 312 nm laser pulses as a function of experimentally available chirp range. The inserts show the relevant energetics for the reaction, as well as a plot of the maximum 12 yield enhancement recorded at 2400 fs2 chirp, as a function of laser peak intensity (measured at zero chirp). production should track with the transition probability instead of showing the observed enhancement. The effects caused by chirp in the excitation pulses reflect characteristics of the potential energy surfaces and the nascent wave packet dynamics. For diatomic 12, Wilson and coworkers have been able to explain their observed chirp effects based on quantum 46 mechanical calculations that show a “wave packet following” effect for positive chirp. Currently, the potential energy surfaces involved for CH212, are not known, preventing us from giving an accurate quantum mechanical description of the effect. In principle a similar wave packet following effect could be responsible, given that the first photon transition is resonant, as in the 12 experiment fi'om Wilson’s group. The order of magnitude changes demonstrated give additional evidence that quantum control of the yield of chemical reactions involving polyatomic molecules is a promising area of research. The chirp of femtosecond pulses can have profound effects on photophysical and photochemical processes. For diatomic iodine, both positive and negative chirps enhance I2 fluorescence signals. These findings are in good agreement with those obtained by Wilson and coworkers. However, for the multiphoton photodissociation of CH212 with 624 nm, the opposite chirping effects, with the maximum 12 yield near chirp zero, are observed. Results obtained for 312 nm multiphoton excitation of CH212 exhibit a minimum 12 yield near chirp zero. The yield of the molecular pathway is found to be asymmetric with respect to the sign of the chirp and enhancements up to a factor of 25 for positive chirps are observed. The amount of enhancement is also found to be highly dependent on the pulse intensities. 47 4. PULSE SEQUENCES AND CHIRP AS A TOOL IN F EMTOSECOND QUANTUM CONTROL OF MOLECULAR EXCITATION AND INTRAMOLECULAR DYNAMICS 4.1. Introduction Work presented in this chapter demonstrates how three-pulse F WM can be used to understand laser excitation, suggests how to predict the outcome from simple pulse sequences, and suggests how to achieve experimental control over the population transfer and coherences between two states. The approach presented here is to use pulse sequences in three-pulse F WM to study closely the interaction between ultrashort laser pulses and model molecular system. A two-level system is the ideal starting point for discussing laser excitation. The probability of excitation from the ground to the excited state in a two-level system is expressed quantum mechanically asl|2 = (e p-E(t) g) , where u is the transition dipole moment and E(t) is the electric field. This expression implies that in order to transfer part of the population from one state to the other, two interactions with the electric field are required, one with E(t) and one with E(t)*. A good example of this kind of population transfer is excitation by any light source as performed routinely in linear spectroscopy methods. Both interactions are with the same field but do not imply that it is a two-photon process. Achieving the population inversion or driving the excitation process itself into desired direction with the laser pulse requires individual 48 manipulation of both of the electric fields involved in the excitation. As will be described in this chapter, full control over the transfer of population between the ground and the excited states could be accomplished.94 Such a technique that allows individual manipulation of the electric fields E(t) and E(t)’I can be obtained if non-linear optical methods are combined with ultrashort excitation sources. The technique has a number of modifications but is commonly referred to as three-pulse FWM. As a version of transient grating (TG) techniques it has been used to investigate solids, liquids and gases almost since the invention of the laser;l55"59 it has been investigated in depth‘50-l62 and textbooks have been written on these subjects.35'37 When used in combination with femtosecond pulses it reveals new possibilities, as it will be demonstrated in this chapter. It was until recently most commonly applied for molecular dynamics studies in condensed phases. 163465 Gas-phase FWM studies have a long history, starting with the observation of photon echoes in SF6 ‘56 and the first observation of molecular rotation by a third-order nonlinear process.‘57 Numerous gas-phase applications of transient grating techniques were pioneered by Fayer,‘59v‘66~I67 especially for the study of flames. These novel probes have been applied to the study of femtosecond dynamics in the gas phase. Zewail and coworkers used degenerate four-wave mixing (DFWM) for probing reaction dynamics.I68 Matemy and coworkers have studied iodine vapor using time-resolved CARS and DFWM.83~‘69"72 By varying the time delay of one of the incident pulses while maintaining the other two incident pulses overlapped in time, they showed that vibrational and rotational dynamics can be observed for both the ground and excited electronic states. Knopp et al. have used two-dimensional time-delay CARS to study high 49 energy ground-state vibrations. ‘73 Brown et al. used off-resonance transient grating methods to study the response of atoms as well as polyatomic molecules in the gas phase and derived a semiclassical expression to analyze the rotational coherences observed in the data.92 Resonance transient grating measurements confirmed the observation of ground and excited states. Our group has explored different laser pulse sequences in three-pulse FWM to control the observation of ground- or excited-state dynamics and has shown that the spectrally dispersed three-pulse FWM method adds an important dimension to the study of non- linear phenomena with transform limited and chirped pulses.”-102 Three-Pulse F WM is a nonlinear spectroscopic method that involves the interaction of three laser pulses in a phase-matched geometry with a sample. It requires a precisely defined time sequence of the pulses. The effect of these sequences in control of molecular dynamics is discussed here. The FWM signal develops from a third-order polarization resulting from the interaction of the three electric fields and is itself a coherent beam forming a fourth electromagnetic wave.87 Its similarity to the pump-probe technique is straightforward; a preparation step is followed after a variable time delay by a probing step. Its advantage lies in the greater degree of control over the preparation and probing processes because various pulse sequences are possible. The technique of quantum control by applying various pulse sequences presented here, is an ideal tool to learn about and manipulate the quantum mechanical processes involved in laser control of chemical reactions and also reveals which ‘knobs’ are available for controlling laser- molecule interactions. 50 In addition to exploring the effect of different pulse sequences, the effect of linearly chirped electric fields on three-pulse FWM measurements is also explored. Chirped pulses affect the phase of the individual contributions from ro-vibrational levels of the molecular system. These effects are followed in time and the data show different dephasing dynamics depending on the magnitude and sign of the chirp. These effects are also observed by spectrally dispersing the three-pulse FWM signal to measure the phase- dependent amplitude of each of the different spectral components. This method, spectrally dispersed three-pulse FWM, is ideal for characterizing and understanding molecular excitation with chirped laser fields.95 The experimental demonstration of laser control of chemical reactions by other methods has been advancing rapidly over last fifteen years following the success of femtosecond laser technology. Achievements in this area of research have been thoroughly reviewed.43-63J37~ Coherent control depends on the quantum mechanical interference between two excitation paths created by two phase-locked lasers. The experiments are carried out by picking two lasers that can reach the same quantum mechanical state by one-, two-, or three-photon transitions. This technique is typically formulated in the frequency domain with phase as the controlling knob. The pump-dump technique involves the transfer of population between two or more states; control is achieved by taking advantage of the dynamics of the molecular system. This technique is typically formulated in the time domain. The time between the pulses is the control knob; therefore, most pump-probe experiments in the femtosecond time scale can be considered examples of this technique. Mode-selective control involves using one pulse to select a vibrational overtone of the molecule and a second pulse to cleave the bond that has been 51 selectively excited. This technique combines frequency and time resolution (usually in nanoseconds). Quantum control usually involves a single laser pulse that is tailored in the time and frequency domains. Typically the control knobs are linear chirp, pulse duration (transform-limited pulses), and pulse intensity. Experimental demonstrations of this technique often involve multiphoton transitions. Optimal control seeks to find the optimal electric field to cause the desired excitation.‘74"76 Theoretically, this search can be carried out by two approaches. The first is to seek a direct solution of the quantum mechanical problem for the optimal field (a problem that is computationally feasible for very simple systems). The second is to use an algorithm that quickly explores different possible fields and contains a feedback on the degree of success that each field attains. Experimentally, groups have demonstrated equipment that can be used to search for the ‘ideal’ laser pulse to achieve a particular outcome.'9”?-"22vI77 This method has been demonstrated recently to control a laser-initiated chemical reaction. It is important to realize that the classification of control approaches should be taken loosely, as the field is relatively new and under constant development. It is clear from this brief overview that laser control of chemical reactions depends on the repeated interaction between the laser fields and molecules in the system. Although there are various formulations of the problem, the fundamental goal is to control population transfer to different quantum states (leading to different products of chemical reactions) by tailoring the electric fields in the time and frequency domains. 52 4.2. Four wave mixing Four-wave mixing techniques provide a powerful platform for combining coherently multiple laser pulses. The results presented demonstrate control over the observed dynamics from ground- and excited-state populations and coherences. The FWM signal results from the polarization of the sample following three different electric field interactions. The effectiveness of these techniques in controlling chemical reactions is explored here by manipulating the time sequence between laser pulses. The phase relationship between the pulses is maintained by detecting the signal in a phase-matching direction. Several different pulse sequences were applied (transient grating (TG), reverse transient grating (RTG), virtual and photon echo (V E and PE)) and their contributions to dynamics of sample molecules were sorted out. For example, the virtual echo sequence is achieved by the interactions of the sample with three consecutive electric fields characterized by exp[i(kx-wt)], exp[-i(kx-at)] and exp[i(kx-ax)]. This sequence allows control over the observed ground or excited state dynamics. With the PE pulse sequence, characterized by interactions with exp[-i(kx-at)], exp[i(kx-at)], and exp[i(kx-at)], we find that control of ground and excited state populations is not achieved. These differences between pulse sequences are shown experimentally. 4.2.1. The role of pulse sequences in controlling ultrafast intramolecular dynamics with FWM This section explores the different four-wave mixing phenomena that can be obtained as a firnction of pulse sequences using degenerate ultrafast pulses. The goal is to reach a more fundamental understanding of these processes by learning how the time 53 order between the pulses can be used to control the different nonlinear optical responses that lead to signal formation. The data from this systematic approach provide the relative intensities from the different nonlinear processes, and reveal pulse sequences that are amenable to control electronic state excitation. The experimental data obtained for different pulse sequences directly demonstrates the need to separate bra and ket interactions for a complete understanding of FWM experiments with non-collinear laser pulses. We have found that under certain conditions, e.g. gas-phase samples and high sample concentration, a cascading process can occur, whereby the free—induction decay emission, stimulated by one of the pulses, can mix with two other beams to produce an additional source of signal, designated here cascaded free-induction decay four-wave mixing (C-FID-FWM). This article provides a large body of experimental data on a model system consisting of isolated molecular iodine that is well understood and can be used to refine present theories of FWM. In the condensed phase, the collective polarizability of the sample contributes to the nonlinear response and it is difficult to separate the intermolecular from the intramolecular processes that cause the observed signal. In dilute gases, it is much easier to analyze in detail the photophysical processes that lead to signal formation. Inhomogeneous broadening is small and the coherence decay is much longer than the laser pulse duration. This latter characteristic leads to long-lived signals, on the order of 100 ps, which in turn allow for very accurate analysis. Gas-phase molecular iodine was chosen as a model molecule for this study for various reasons. (i) The visible X ’Zo.g <—) B ”110+u transition has been well characterized by frequency and time domain spectroscopies. (ii) The vibrational periods of both the X 54 and B states are longer than the duration of our laser pulses, allowing us to impulsively excite wave packets in each electronic state. (iii) The vibrational periods of the X and B states are quite different, 160 fs and 307 fs respectively, making the assignment of the signal relatively easy. (iv) At the wavelength of excitation, molecular iodine can be treated as a two (electronic) level system. The state reached by two-photon excitation is repulsive and does not contribute to the signal discussed in this study. 4.2.2. Formal description of nonlinear optical processes involved in sequences’ control scheme There are numerous publications in the literature on the nature of nonlinear optical processes.”87 Some work involves the density matrix approach,l°4 whereas in other cases the wave packet formalism is used to understand FWM-type processes.'7"I78 We find that the density matrix approach provides a complete framework for interpreting nonlinear optic phenomena. The density matrix approach naturally incorporates all of the initially populated vibrational and rotational states necessary to express the collective nature of a coherent process. However, for some processes the benefits of the intuitive wave packet picture were exploited. Three-pulse FWM involves a sequential interaction of three electric fields with a molecular ensemble. The incident pulses are typically arranged in the forward-box geometry shown in Figure 4.1a (see detailed beam arrangement in Figure 2.6), and the coherent signal is emitted in the direction kg with frequency (05. Both the wave vector and the frequency of the signal are linear combinations of the wave vectors and frequencies of the applied fields. The wave vector kg: ka— kb+ kc satisfies the phase- 55 a) Phase matching ks zka'kb+kc 2'12 T23 2'12 PE t1 t2 (3 tlme c) Virtual Echo sequence 2'12 ,' £723 ) t1 t2 t3 tlme Figure 4.1. (a) Beam arrangement in the forward box configuration for the three-pulse FWM experiment. The three laser fields are applied in a given temporal sequence and overlapped spatially in the sample. The signal is detected in the direction of the wave vector ks that satisfies the phase-matching condition. The dashed line indicates a plane of symmetry that makes fields E8 and Ec equivalent. (b) Temporal sequence of the three laser pulses for a virtual echo measurement. 712 and 123 are the time delays between first and second or second and third pulses, respectively. (c) Temporal sequence of the three laser pulses for a virtual photon echo measurement. Notice that the rephasing occurs at a time (3 + (t2 - t,) for the stimulated photon echo signal and t3 - (t2 — II) for the virtual photon echo signal. matching condition, ensuring conservation of energy and momentum. The source of the signal is the third-order polarization in the direction ks, P513 ’(t), that arises after weak interactions with the applied fields. For homodyne detection, the intensity of the signal i586,87 56 °° 2 [FWM oc flPQ’Kt) dt. (4.1) If the applied electric fields are ultrashort laser pulses, various rotational, vibrational and in some cases, electronic levels can be excited impulsively due to the large bandwidth. In those cases the signal reflects the coherent superposition of molecular polarizations, showing coherent oscillations (or quantum beats). In this way, the optical polarization contains information regarding the electronic and nuclear dynamics manifested in the spectra. The coherent signals in a FWM process involve the entire molecular ensemble that is interrogated by the laser electric fields. The interpretation of the signals requires the density matrix formalism for which the statistical nature of the quantum mechanical ensemble is explicitly taken into account. The density matrix formalism will be just briefly described here. The most detailed description of this formalism is presented by Grimberg et. al.104 The third-order polarization of the molecular ensemble is defined a535'87 13%) = T4132?) (1)] . (4.2) where P is the polarization operator and fig” (t) is the third-order density matrix with a spatial dependence 6”“ 'r . The temporal evolution of the density matrix can be represented diagrammatically using either double-sided Feynman diagrams (DSFD)35'87’I79 or ladder diagrams.180 Each diagram is associated with a sequence of transformations or pathway in a phase space of the density matrix. The total number of diagrams for a given direction ks is 48. However, 57 under the rotating wave approximation (RWA), the number of possible pathways is reduced to eight. When the system includes only two electronic states, the frequency dependence of the slowly rotating terms is exp[i(a)—a)eg )t], where a) is the carrier frequency of the pulses resonant with the transition frequency cocg. Based on the RWA the contribution of slowly rotating terms is much greater, therefore pulse sequences having the first two electric fields interacting with opposite wave vectors are primarily responsible for the FWM signal. As a consequence, the first two pulses of a three-pulse FWM sequence form gratings in the direction i (kl — k2), the third pulse Bragg diffracts (Figure 2.6) from that grating and the direction of the signal is given by kg = k3 i (k1 — k2). The specific pulse sequence indicated by the subindices 1, 2, and 3 determines a specific non-linear process detected (TG, RTG, VE or PE) at the direction kg and the corresponding coherent transient effect probed. When collected at k3 = — k1 + k2 + k3, the signal corresponds to a stimulated photon echo (PE) process.87 This process is similar to Hahn spin echo'31 in NMR.‘32 When the signal is collected at kg = k. — k2 + k3, the signal has been named a virtual photon echo (VE) phenomenon. Alternate names for this non-rephasing FWM signal can be found in the literature. One of the key differences between VB and PE phenomena is the fact that PE phenomena can be regarded as a reversible decay process for which the polarization rephases after the interaction with the third laser pulse. For PE this polarization reaches a maximum at a time (3 + ((2 — t1) (see Figure 4.1b) whereas for VE the polarization reaches a maximum at a time t3 — (t2 — t1) (see Figure 4.1c), hence the name virtual and the lack of observable rephasing. This difference is of critical 58 importance for samples where the inhomogeneous broadening is significant such as in liquids and solids. For PE measurements, the inhomogeneous dephasing is cancelled by the optical rephasing. Each experimental signal obtained with a specific pulse sequence in the direction ks is the coherent sum of the signal corresponding to several third-order pathways in the Liouville space. Therefore, more than one diagram may be associated with a particular three-pulse FWM phenomenon. Each diagram contributes additively to the overall third order polarization and the signal is proportional to the square of that polarization. Diagrams for which the first interaction is on the ket side correspond to VE phenomena, whereas the diagrams that initially show a bra side interaction correspond to PE phenomena. We have chosen the pulse sequence k1 — k2 + k3 (see Figure 4.2) to demonstrate the different density matrix pathways involved in a VB process. In the DSFD (Figure 4.2.a), the bra (right side vertical line) and the ket (left side vertical line) interactions are kept separately. Each electric field interaction is indicated with a short slanted arrow. The convention for the sign of the electric field wave vectors is that positive wave vectors with exp[i(kr-at)] point to the right while negative wave vectors with exp[-i(kr-ax)] point to the left on the diagram. Its relative position and label t., t2 or t3 are used to indicate the time ordering of the laser pulse in a given sequence. The first interaction with the system causes a change on only one side of the Feynman diagram (bra or ket) giving rise to an electronic coherence, designated pmge or p")cg respectively. The second interaction produces a population change, designated pmgg or p‘z’ce, and a ro-vibrational coupling within each electronic state, designated p‘z’g'g or p‘z’ggv in the ground state or 59 pm“, or pm“ in the excited state. The symbols g g' and e e' represent different ro- vibrational levels in the ground and excited states respectively. Note that the DSFD in the left, labeled R4, proceeds through the formation of pmg'g while the one on the right, R1, proceeds through the formation of p(2’ec., The third interaction produces the third order electronic coherence pme'g or p‘J’cg' that gives rise to the signal according to Equations 4.1 and 4.2. a) Double-Sided Feynman Diagrams E SN ES \ E3 pe'g t peg' E3 E2 pg'g 1066' p t2 ’2 p eg t] EST-fig. E2 E1 E1 lg) 80° C. The experiments presented here were carried out at 140° C. Pulses in one of the beams (EC or Eb, depending on the setup) were delayed with respect to the other two by a computer-controlled actuator (variable t 61 delay), while the fixed delay (rm, where n, m=a, b, or c) was adjusted by delaying the pulses from the second beam with respect to the fixed one (beam E3 in all cases) using a micrometer. All the experimental data were obtained in the direction kg = ka — k, + kc. The subindices a, b, and c can take any time order 1, 2 or 3; therefore, for the same phase- matching geometry, different phenomena such PE or VE can be detected. The spatially filtered three-pulse FWM signal was collected by a spectrometer with wide spectral acceptance (8 nm resolution). The transients were taken at 300-400 different time delays (about 200 shown) and averaged for nine scans. At each time delay I, the signal was collected for 10 laser shots. The laser intensity was monitored on a shot-to-shot basis; laser pulses with intensity outside one standard deviation were rejected. The accumulated laser intensity for each given time delay was stored and the final transients were normalized by this intensity. An effort was made to maintain constant laser intensities for all the transients and to present all transients using a consistent intensity scale to allow direct comparisons. Three laser pulses can be arranged into six sequences (abc, acb, bca, bac, cab, cba), each having a distinct pulse ordering. For each pulse sequence, we can define time delays between the first and second pulses (m) and the second and third pulses (223) (see Figure 4.1). We can divide the pulse sequences into those where two pulses overlap in time (Table 4.1) and those where the three pulses are separated (Table 4.2). We can further subdivide the pulse sequences into those where the first time delay is fixed (pulse sequence 1, PS I), and those for which the second time delay is fixed (pulse sequence 11, 62 PS 11). In the experiments presented here, the phase-matching geometry of detection (see Figure 4.1a) makes fields E, and Ec equivalent. The total number of unique combinations and permutations is ten; these are shown in boldface type, with four in Table 4.1 and six in Table 4.2. The ten combinations lead to the observation of different nonlinear optical phenomena, such as transient grating, reverse-transient grating, virtual photon echo, photon echo, and cascaded free-induction decay FWM. Table l. (I) Seqllfinm (11) I ...... I JL ..... I overlapped scanned scanned overlapped “"6 * kl-kl'+k2 TransientGrating 0”” k-k +k' ReverseTG cb,a c,ba a: ' 2 2 ac.b* k.+k,'-k2 ReversePE b,“ * -k1+k2'+k2 PhotonEcho Table 1. Summary of the observed phenomena for the different pulse sequences possible with the three laser pulses when two pulses overlap in time. For the experimental arrangement used here, there are only four distinguishable sequences (shown in bold face). Because beams Ea and Ec are equivalent, two additional (though not unique) sequences are possible. Sequences for which experimental data are presented have been labeled with a star. The phenomena include transient grating (TG), reverse-transient grating (RTG), reverse photon echo (RPE), and photon echo (PE). The experimental data are shown along with the schematic of the pulse sequence and the applicable DSFD. Many of the labels for the DSFD have been omitted in order to simplify the diagrams here. The DSF D corresponding to the formation of an excited state population after the first two interactions are shown on the right side while the lefi-side diagram(s) correspond to formation of a ground state population after the first two interactions. The time at which each field interacts with the system is indicated by a dotted line. For the DSFD shown in figures on following pages (Figures 4.3-4.8,) the 63 distance between these dotted lines indicates the time delay, i. e., the smaller distance depicts the fixed time delay rm and the larger distance depicts the variable time delay I. When the fixed time delay is zero, only the variable time delay is reflected by the dotted lines. The experimental transients show the FWM signal as a function of variable time delay I. All FWM signals are displayed relative to each other, 1'. e., the transients have not been re-scaled. This allows for direct comparisons of signal intensity among the different pulse sequences. Table2. Phenomena (1) Sequences ([1) fixed - - -scanned scanned- - - fixed VirtualPhotonEcho ab,c alt a,bc kl-kz-l-k3 Cm, (VEI) c’ba * (VEIl) Stimulated Photon Echo 1,0,0 * b,ac a1: -k1+k2+k3 baa “’5” b... “’E") C-FlD-FWM ac,b a1: a,cb “+1121, cal, (C-FIDFWMI) c,” * (C-FID-FWMII) Table 2. Summary of the observed phenomena and the different pulse sequences possible with three time-separated laser pulses. For the experimental arrangement used here, there are only six distinguishable sequences (shown in bold face). Because beams Ea and Ec are equivalent, additional (though not unique) sequences are possible. Sequences for which experimental data are presented have been labeled with a star. The phenomena include virtual photon echo (VE), stimulated photon echo (PE), and cascaded free-induction decay FWM (C-FID-FWM). 64 4.2.4. Overlapped Pulses Figure 4.3 displays the dynamics observed when two of the laser pulses are overlapped in time while the third pulse precedes or follows the other two. In the transient grating (TG) measurement (Figure 4.3a), fields Ea and Eb are overlapped in time (rah = 0 fs) and field Ec follows at a variable time delay 1. Three DSFD describe this process and indicate that both ground and excited state dynamics (0(2),;g and pa’cc' respectively) should be observed. In the first two picoseconds of the transient, the dominant oscillations reflect the 307 fs; the vibrational period of the B excited state of iodine. The transient also clearly shows 160 fs oscillations, reflecting the ground state vibrational period of 12. Ground state dynamics are more clearly observed at longer time delays because of wave packet spreading in the more anharrnonic B excited state. The fast Fourier transform (FFT) of this transient confirmed the assignment of the dynamics to the X (208 cm”) and B state (108 cm") vibrations (v" = 2-4 and v' = 6-11).93~94 In the reverse-transient grating (RTG) experiment (Figure 4.3b), field Ec precedes the overlapped fields E, and Eb. In this case, Ec excites an electronic coherence with a relaxation lifetime of ~1 50 ps.97 When Eb and E3 arrive, the grating is formed with Eb and the polarization generated by E... The observed vibrational dynamics are dominated by 307 fs oscillations, corresponding to the excited state of 12. A FFT of the transient confirms that the contribution of the excited state is greater than that of the ground state. The time evolution of the coherence leads to the simultaneous observation of both ground and excited state dynamics for this pulse sequence. 65 a) Transient Grating ab o‘ .r--- -. .- b) Reverse Transient Grating 0 ba b 2 a c o ............................ ° - c) Photon Echo b FWM Signal [arbitrary units] Delay Time 2'[psj Figure 4.3. Experimental FWM data obtained for three different sequences having two pulses overlapped in time. The signal corresponds to: TG, RTG and PE depending on the temporal arrangement of the electric fields. The pulse sequence and the DSFD of the processes contributing to each phenomenon are included in each case. The dominant oscillation period in all cases, 307 fs, corresponds to excited state vibrations. The TG and the RTG transients in (a) and (b) show some additional contribution of ground state vibrations. The RTG transient presents a prominent slow modulation (2 ps) corresponding to rotational dephasing and a large background signal. Notice that the PE data show only excited state vibrations with very little rotational dephasing. 66 In the photon echo (PE) measurement (Figure 4.3c), the pulse sequence is similar to the one responsible for the RTG except that field Eb is applied first, followed by Ea and Ec which are overlapped in time. The data show primarily excited state dynamics as can be observed in the transient. The most important difference between these transients is that field Eb acts first and is then followed by the overlapped fields E8 and EC. Photon echo processes involve a rephasing of the coherence that is lost due to inhomogeneities in the sample. A direct comparison between the reverse-transient grating and photon echo transients reveals that the slow undulation (2 ps) in the RTG measurement as well as the background are smaller in the PE transient. This difference is also apparent in the FFT of the PE and the RTG transients, where the rotational contributions (peaks at frequencies < 20 cm") in the PE FFT are greatly reduced. The predominance of the excited state dynamics in both RTG and PE transients can be attributed to the fact that the wave packet motion in the excited state has a much wider range of internuclear distances. This takes the wave packet in and out of the Franck-Condon region. Therefore, the electronic polarization reflects predominantly excited state dynamics. Only in pulse sequences where pathways involving excited state population transfer can be avoided, the ground state dynamics predominates. Results shown in Figure 4.3. are an example of observation of quantum dynamics, but also of the selectivity. This selectivity in the observation of ground state dynamics can be achieved by choosing the appropriate time delay between the pulses, described in next section. 67 4.2.5. Pulse Sequence 1: Virtual Photon Echo and Stimulated Photon Echo In Figure 4.4, we use Pulse Sequence I (fixed delay followed by variable delay) to observe the dynamics obtained as the fixed delay is changed from 460 fs (Figure 4.41) to 614 fs (Figure 4.4.b) for virtual photon echo (VE I) and stimulated photon echo (PE 1) sequences (Figures 4.4c and d). In both cases, fields Ea and E, are separated by a fixed time delay (rah or rba) and are followed by field Ec after a variable time delay I. There are two DSFD that apply to each of these pulse sequences; one shows pgvgm and the other peeve). In the VB 1 measurement, when Tab is 460 fs (Figure 4a) (3/2 I, where 2', = 21t/cq, for 12 in the B state), the dynamics show 307 fs oscillations, reflecting only an excited state contribution. When Tab is 614 fs (Figure 4.4b) (twice the vibrational period of 12 in the B state), the dynamics show 160 fs oscillations, reflecting predominately a ground state contribution. By changing the fixed time delay between fields E1, and Eb, the observation of excited or ground state dynamics of 12 can be controlled. This type of coherent control should not be confirsed with control experiments without phase locking or phase-matching. In the PE 1 configuration, when rba is 460 fs, the dynamics reflect an excited state contribution with 307 fs oscillations; no ground state contribution is observed in this transient. When rba is 614 fs, the 307 fs oscillations still dominate; however, two picoseconds latter, 160 fs oscillations can be seen. Fourier transforms of these two PE transients confirm that for rba = 614 fs, there is some ground state contribution. The signal contribution due to rotational dynamics (slow 2 ps modulation) that is clearly observed in the VB 1 transients is much smaller in the PE 1 transients. Note that the 68 4 VirtualEchoI 3‘ a) tab=460 fs 2 a b c _-z-_ 1- 111 1 > o ........................... ’ C 4‘ b) tab=6l4 fs c ' b E" 3‘ b g 2-1 ' ‘ I a a S ’ R4 R1 .13 1‘ _Q . E. o .................... . ....... €30 4; Photon Echol 'c'r‘: ‘ c)rba=460fs 3.. E ’ b a c u. 2. . Arr-1 1’ ; C d)rba=614fs c a b a b R3 R2 Delay Time 2' [ps] Figure 4.4. Experimental transients for sequences with fixed time delay between the first two pulses, PS I, with tab: zba=460 fs or Tab: zba=614 fs. The VE signal is obtained when the initial electric field acts on the ket side of the diagram and a PE signal is obtained if the initial electric field acts on the bra side of the diagram. (a) When rab=460 fs, the oscillation period of the VE signal corresponds to the vibrational motion of the excited state molecules. (b) When rab=614 fs, the VE signal oscillates with the ground state vibrational frequency and the rotational dynamics (slow 2 ps modulation) are clearly observed. Notice that the predominant feature in both PE signals (c and d) shows 307 fs oscillations corresponding to the excited state. 69 selection between ground- or excited-state dynamics is much more efficient for the virtual echo setup (Figures 4.4a and 4.4b). The reason for this observation can be deduced from inspecting the DSFD. For VB 1, the DSFD that leads to the observation of ground state has three laser interactions acting on the ket. This leads to high selectivity between the two states. For VB 1, the appearance of ground state dynamics arises from a wave packet being prepared in the excited state, then pumped to the ground state, and finally probed as a function of time, thereby giving a clear and intense ground state signal. For PE 1, the DSFD that leads to ground state dynamics shows that the first two interactions are on the bra while the third interaction is on the ket. This action on an unperturbed ground state by the third pulse leads to loss of selectivity. The reason for the small ground state modulation in the PE 1 data is due to the interference between wave packets generated from different initial states. In Figure 4.5, enlargements of the VE I and PE I transients from Figures 4.4 a and c, for Tab: rba=460 fs are shown. It is clear that the two transients are exactly out of phase with each other; when one is at a maximum, the other is at a minimum. The applicable Feynman diagram for the observation of excited state dynamics is on the right-hand-side in each case of Figure 4.4. In both cases, field Ec must interact with the wave packet formed in the excited state by field Eb. This process is depicted by the action of Ec on the bra for both DSF D responsible for excited-state populations. In Figure 4.5 (bottom), schematics based on calculated wave packets of the excitation process after the first two electric fields are shown. The wave packet resulting from excitation by field Ea is shown as a solid line; the wave packet resulting from excitation by field Eb is shown as a dashed 70 7.3—3“ tab=460fs §2_ VEI El- .971 PEI 2 o 1 2 3 Delay Time t [ps] Figure 4.5. Top: Experimental data for VE (thick line) and PE (thin line) measurements with rab= rba=460 fs. Notice that the signals, corresponding to excited state dynamics, are exactly out of phase. The signal intensity has not been re-scaled, but the VEI signal has been shifted vertically. Note that the photon echo signal peaks at 154 fs. The 180° phase difference between these transients can be understood by looking at the dynamics of the wave packet in the B state of 12 resulting from the interaction with pulse Eb. Bottom: Simulation of the wave packet motion on the B state of 12 under the conditions of the VB and PE signals for PS I when Tab: rba=460 fs. In both cases signal formation depends on de-excitation of ‘I’bm (dotted curve). At time delay F0 fs, which corresponds to the time delay between last two pulses, LPb“) is in the F ranck-Condon region for VB 1, maximizing the transition probability when the third pulse is applied there. However, for case PE 1, LPb") is at the outer turning point of the excited state potential when r=0 fs, minimizing the transition probability when the third pulse is applied in the Franck-Condon region. 71 line. It is important to note that signal formation for both processes requires the de-excitation of the wave packet ‘Pmb by field EC. In the virtual echo case, field Eb interacts with the system second with rab=460 fs. At r=0 fs (time delay for field Ec with respect to the second pulse), the dashed wave packet is at the inner turning point of the potential where the transition probability is maximized. Therefore the interaction results in a maximum signal. However, in the photon echo, field Eb interacts first with the system (still with rba=460 fs). At r=0 fs, the wave packet is at the outer turning point of the potential with negligible transition probability upon interaction with field EC. After half a vibrational period, the wave packet returns to the inner turning point where the transition probability is maximized and the interaction with Ec produces a maximum signal. Therefore, the signals for PE 1 and VE I are exactly out-of-phase. 4.2.6. Pulse Sequence II : Virtual Photon Echo and Stimulated Photon Echo In Figure 4.6, the experimental signal resulting from Pulse Sequence II (variable time delay followed by fixed time delay) is observed. The fixed delay is changed from 460 fs (a) to 614 fs (b). In this case, an electronic coherence is created by the first field (EC) which evolves until fields E, and E2, arrive. The fixed time delay can act as a “filter” rather than a control parameter in selecting the observation of ground or excited state dynamics. There are two applicable DFSD for this configuration. The diagram on the left has all the interactions on the ket while the diagram on the right has the last two interactions on the bra. For rba=460 fs, the transient shows 307 fs oscillations, corresponding to excited-state dynamics. The signal is weaker and shows only a small 72 background. When 2ha=614 fs, the transient is dominated by 307 fs oscillations. The signal is stronger and shows a larger background. Weak 160 fs vibrations are also observed. Fourier transforms have confirmed that the rba = 614 fs transient shows a contribution of the ground state. For a time delay of zba=6l4 fs, the observed background arises from the process depicted by the DSFD on the right. The use of the fixed delay as a filter for the dynamics changes the ground state contributions to the signal slightly but does not give the same degree of control as is observed in the VB 1 case (see Figures 4.4a and 4.4b). 6‘ I V1rtual Echo H 4- _ a) rba — 460 fs E c b a g 2- - - 2'- - _ .>.~ ' All/\AAANMAANW M 13 1‘: 0 . . . . . . a E 1 b _ _ g 4_ b) 2ba=6l4fs a b .9.” ' ' (I) _ C C E 2- R4 R1 LL. 0 1—* v I v v r 0 1 2 3 4 5 Delay Time 1: [ps] Figure 4.6. VE transients obtained for a sequence having the time delay between the last two pulses fixed to rba=460 fs or 21,5614 fs. The vibrational excited-state motion dominates both transients. A very small contribution from the ground state vibrational motion is apparent when 21,,=614fs. A larger rotational dephasing is observed (2 ps modulation) in the 21,5614 fs signal and the signal intensity is higher in this case. The DSFD of the processes that contribute to the signal are included with the pulse sequence. 73 In Figure 4.7, we use Pulse Sequence II to observe the dynamics obtained as the fixed delay is changed 460 fs to 614 fs. In these measurements, field E, precedes fields Ea and EC, setting up a stimulated photon echo (SPE). The coherence generated by the first laser pulse is probed by the second and third pulses. The rotational dynamics, as evidenced by the slow 2 ps modulation, are significantly reduced in the PE 11 transients as compared to those observed in the VE II transients. The transient in Figure 4.7a shows predominately 307 fs dynamics and a small ground state contribution. The transient shown in Figure 4.7b reveals a predominately excited state contribution to the signal. These assignments were confirmed by FFT. Photon Echo II 1 a)rac=460fs i . 1 . 11 E > E o ............................ Te * b)2'ac=6l4fs c a) a £7) 1‘ ' E b b 0 1 2 3 4 5 Delay Time 2'[ps] Figure 4.7. PE transients obtained for PS 11 when rac=460 fs or 23c=614 fs. These transients, mostly dominated by the excited state motion, do not show the selectivity of the molecular dynamics with the selection of the fixed time delay. In addition to the pulse sequence, the DSFD of the processes that contribute to the signal are included. 74 4.2.7. Cascaded F rec-Induction Decay F WM So far we have discussed a number of phenomena and outlined the rules for their diagrammatic representation. The identification of the eight different diagrams that can contribute to the signal within the RWA has been shown in a number of publications. When a particular sequence of resonant fields violates the RWA, the FWM signal should not be observed (it is negligible) as denoted in upper-right comer diagrams in Figure 4.8. An example would be the sequence k1 + k2 - k3 for a two electronic state system. This corresponds to pulses E2, and Ec preceding pulse Eb (see Figure 4.8). We set out to test this hypothesis and detected weak signals corresponding to an interesting FWM phenomena arising under certain conditions described below. In Figure 4.8a, the signal obtained with the pulse sequence Ec followed by Ea and Eb is presented. The signal intensity is about a factor of four lower than the VB 1 data. It is remarkable to see that when 2ab=460 fs, the transient shows 307 fs oscillations, corresponding to excited state dynamics. When rab=614 fs, 160 fs oscillations are evident indicating ground state dynamics. Therefore, we are able to observe ground or excited state dynamics by changing the fixed time delay as it was done in VB 1 (see Figure 4.4a and b). Our explanation for this process is that field (EC) creates a first-order polarization that decays by free-induction decay emission (FIDE) with time T2 > 50 ps. This emission is designated as FIDEC. Beams E2, and Eb form the grating with populations pu’gg and plz’ee', then the FIDEc Bragg diffracts from the grating to form the signal. Because the FIDE is generated in a separate ensemble of molecules we designate this process a cascade, hence the abbreviation C-FID-FWM. This situation is similar to the VE I measurements except that the beam EC is replaced by FIDEc. The observation of 75 C-FID-FWM II kb ’‘2: '5? 1‘ X c .+.:: a)2 Tab — ~460 fs "a c c a a ’51 forbidden a b MAR/MNU‘N‘AAMM 8 E. a E. c ....................... fl . W T T b/Ef 117F113:EC "a ,7, b)2 2ab=614fs . FIDEC E 1- FIDEc c ensemble 121 ensemble2 G ............................ Delay Time 2'[ps] Figure 4.8 Insert: Transient grating and double-sided Feynman diagrams for the case when pulse b is the last to arrive at the sample. The grating formed by pulses a and c (indicated by the diagonal lines) cannot diffract pulse b in the signal direction. Violation of the RWA is identified in the double-sided Feynman diagrams by the crossed arrows. C—FID-FWM transients for rab=460 fs (a) and rab=614 fs (b). When rab=460 fs, the oscillation period of the C-FID-FWM signal corresponds to the vibrational motion of the excited state molecules, whereas when 23b=614 fs the signal oscillates with the ground state vibrational frequency and shows a larger rotational modulation (2 ps). Notice that the data are very similar to those observed for VE I (Figure 4.4a). This implies that the dominant pathway for signal formation can be illustrated using similar DSFD for the two ensembles. selectivity between ground- and excited-state dynamics confirms the interpretation of this phenomenon. Note that the only way to represent this first-order cascading process diagrammatically is by invoking two ensembles, one responsible for the first-order polarization and the other for the FWM process. This explanation does not involve a violation of the RWA. The signal identified here as C-FID-FWM arises when the sample has a relatively 76 long-lived polarization. The lasers are resonant with a transition and the sample is in relatively high concentration. Under these conditions each pulse generates a free- induction decay with an integrated emission intensity that can be comparable with the field that induces it. The FID emission under these circumstances participates in the formation of FWM signal. This type of signal is best observed when other FWM processes are forbidden. Observation of this type of signal can be found in the work of Prior. 1 82 PS I used to explore other C-FID-FWM signals is shown in Figure 4.9. For both transients in this figure, 2ac=460 fs and 2ac=614 fs, the dynamics are dominated by excited state vibrations. F FT confirms this observation. The ground state contribution is slightly higher for the transient with 2ac=614 fs than for 23c=460 fs. These transients can be interpreted as follows: fields Ec and Eb form the grating, then the free-induction decay emission stimulated by Ea, F IDEa, mixes to form the signal, C-FID-FWM 1. Because the fixed time delay is now between the first two pulses, one participating in the population formation and the second in the generation of a F IDEa, selectivity is not expect between ground and excited state dynamics. The time delay difference between the first two pulses is half the excited state vibrational period. This causes the 235460 fs transient to be similar to the rac=6l4 fs transient except for being out-of-phase by half a vibrational period (154 fs). 4.2.8. Conclusions The classification and the comparison of different phenomena that can be observed with femtosecond pulses on gas-phase molecules using specific pulse sequences 77 C-FID-F WM I a) raw = 460 fs 11:2— , FIDE FIDEa b) Z'ac =614 fs aj: c/bCDDEa ’0’” 1 I " , ensemble 1 ensemble 2 o vvvvvvvvvvvvvvvvvvvvvvvvvv FWM Signal [arbitrary units] Delay Time 2'[ps]5 Figure 4.9. Cascaded free-induction decay FWM transients for PS I for 235460 fs (a) and 2ac=614 fs (b). Notice that the signal intensity for these sequences is quite low, especially after the first picosecond. The transients do not show selectivity between the ground and excited state molecular dynamics with the fixed time delay. The pulse sequence and the coupled DSFD are included for each case. are presented. Three laser pulses can generate six different nonlinear optical phenomena: virtual echo, photon echo, transient grating, reverse-transient grating, cascaded free- induction decay four-wave mixing and reverse photon echo. While these phenomena are closely related, it is important to distinguish between them. Some allow the experimenter to control the observation of ground or excited state dynamics. Others allow the control of the phase of the observed molecular vibrations. Finally, some of these techniques achieve cancellation or enhancement of inhomogeneous broadening. One of the important conclusions is the distinction between the bra and ket laser interactions. Quantum mechanically, the transition probability from state g to state e 78 depends on el21El> implying a double interaction. The physical basis for the double interaction comes from the fact that the electric field has two components exp[i(kor- 01)] and exp[-i(kor-at)]. Molecules may interact with either of the components (thus forming a coherence state) or with both (thus forming a population in the original or in the new state). Starting from the ground state and invoking the rotating wave approximation, a ket interaction can only be achieved with exp[i(kr-ax)] and a bra interaction can only be achieved with exp[-i(kr-at)]. When the three incident pulses have distinguishable wave vectors, as for most non-collinear geometries, one is able to define a phase-matching geometry that determines the sign and hence the nature of each electric field interaction. Therefore, a distinction between a bra or a ket interaction can be experimentally achieved. The discussed phenomena correspond to products of four matrix elements (three incident fields and formed signal). However, presented description involves elements up to the second order (pg-(2’). The specific aim is solution of control over the system dynamics. One can envision the presented method as expanded pump- probe experiments, where the first two interactions prepare the system and the third one interrogates it. The time-integrated detected experimental signal is caused by third interaction (00(3)) and carries information about the preparation of the system (with first two fields). Fourth order matrix element would carry the information about the system after the signal is emitted and is not of interest here. Also, the interrogation of the state of the system after third field, and its temporal profile would require additional time gating of the signal. The distinction that arises from the geometric arrangement of the lasers and the 79 possibility to have fixed or variable time delay between them leads to the ten unique pulse sequences described in Tables 1 and 2. The data presented show several differences. For stimulated photon echo measurements, the signal does not contain elements that can be attributed to inhomogeneous broadening while virtual echo measurements do. This difference is best shown in Figure 4.3 by comparing the reverse- transient grating (see Figure 4.3b) and the photon echo signals (see Figure 4.3c). Coherence relaxation measurements on gas-phase iodine as a function of temperature presented in Chapter 5, showed that the photon echo relaxation times were always longer than the corresponding virtual echo ones and were also temperature dependent, as expected. The relaxation measurements for reverse-transient grating measurements were dominated by inhomogeneous contributions such as Doppler broadening and the Boltzmann distribution of states prior to laser excitation. One of the strategies used here to sort among different processes was to delay a pair of pulses by a fixed time interval, chosen to be equal to an integer or a half integer number of vibrational oscillations of the excited state. This reveals a striking difference between VB and PE measurements (Figure 4.4) where the VB 1 signal shows excited or ground state dynamics depending on the delay between the first two pulses. However, we note that for PE measurements this control is not observed. The reason for this observation is that one of the diagrams for VE involves all three interactions on the ket. This places a restriction on the laser-molecule interactions, requiring that the electric field interactions must be in concert with the vibrational dynamics. The PE diagram that corresponds to the formation of a population in the ground state is not as restrictive because only the third pulse acts on the ket. From the point of view of control over ground or excited state dynamics, the sequence k1 — k2 + 80 k3 labeled VB 1 is the most efficient. This phenomenon is easily observed in gas-phase samples; however, in liquids VB is usually not observed because of inhomogeneous broadening. Obtained photon echo results agree with the well-known property of photon echo signals, namely, that the effect of inhomogeneities is cancelled. A technique aimed at suppressing excited-state dynamics was introduced in the nineties by Shank and coworkers.183’184 The mode suppression technique requires fixing the time delay between the first and third pulses while the second pulse is scanned in the pulse sequence kg = - kl + k2 + k3. The method has been found to work only in cases of extremely large inhomogeneous broadening. We have explored the mode suppression method under the same conditions as all of the transients in this article. We found that the apparent suppression of excited—state coherent vibrations is achieved by a combination of Liouville pathways that allows ground- and excited-state contributions, thereby drowning the pure excited-state dynamics that are observed when the time delay between the first and third pulses are out-of-phase with the excite-state oscillations. A new phenomenon in gas-phase FWM experiments, designated here as C—FID- FWM has been presented. This signal is observed for pulse sequences where conventional diagrammatic techniques fail to predict the appearance of signal. The observed signal is not caused by a higher order nonlinear process or from a breakdown of the RWA. As described earlier (and formulated in Appendix 1) the C-FID-FWM signal is caused by an induced polarization that maintains the phase and the vector characteristics of the pumping beam. Because the coherence relaxation in iodine at this temperature has a lifetime of ~10'10 seconds, the free-induction decay emission can act at relatively long 81 times. The participation of an electric field generated by a polarization in a nonlinear optical process has precedents; this phenomenon is called cascading.135,186 More recently, cascading has been implicated in the signal arising in fifth-order nonlinear processes. Upon close inspection the fifth-order signal has been found to result from two third-order processes that have been called sequential CARS or cascade. The emission from one third-order polarization participates in a different third-order polarization.‘87"89 The data being reported here show a different type of FWM cascading. The differences are: (i) resonance excitation leads to long-lived polarizations that can interact with the other laser pulses at very long times; (ii) in the studies from the Albrecht and Fleming groups, the cascading process is a combination of two third-order processes, while in the case presented here, the C-FID-FWM signal results from a first- and a second-order process; (iii) only three pulses were used instead of four or five and (iv) we have observed a C-F ID-F WM in a gas-phase sample. Thus, we have ensured that phase matching occurs only for the third-order process of interest and that the observation of FWM or C-FID-FWM depends on only on the pulse sequence. The C-F ID-F WM transients presented here show two surprising effects: the signal is measurable and the femtosecond time resolution is maintained. Given that at least one of the electric fields is replaced by the free-induction decay emission, one could conclude that the intensity would be much lower and that the time resolution would be lost because of the picosecond lifetime of that emission. The intensity of the C-FID-FWM signal derives from the integral over the long duration of the free-induction decay emission. In fact, we have found that in addition to the concentration dependence the ratio between C— FID-FWM and FWM depends on the ratio (T2)2/(T. 2,,), where T2 is the coherence 82 relaxation time, T1 is the emission lifetime and 2,, is the laser pulse duration.loo The temporal resolution is maintained because the free-induction decay emission arises from a vibronic coherence. Therefore, it is strongly coupled to the coherent vibrational motion in the ground and excited states, being modulated on the femtosecond time scale. Coherent control of chemical reactions depends on the design of an electric field capable of causing a molecular system to yield a specified product. These special fields can be constructed out of coherently coupling multiple laser pulses. This work examined the coherent coupling of three fields using the phase-matching condition and the third- order nonlinear response of the sample as filters of the incoherent contributions. The exploration of the various phenomena allows us to form a ‘toolkit’ that can be used to achieve coherent control. Our understanding shows that with FWM the most efficient control is achieved when the phase-matching geometry and the pulse sequence lead to all laser interactions occurring on the ket. This implies using what we have called the VB 1 pulse sequence. Presented experimental data for a number of different sequences stemming from three degenerate femtosecond laser pulses highlight the need for understanding that laser- molecule interactions take place in a complex frame that can be understood in terms of interaction with exp[-i(kr-at)] or with exp[i(kr-at)]. In the language of quantum mechanics, these interactions correspond to the bra or the ket respectively for the first pulse. Based on this understanding, diagrams can be drawn to illustrate the nature of the measurement, whereas density matrix calculations can then be used to simulate the data. Sorting out all the different phenomena is valuable for understanding different nonlinear 83 optical measurements from different groups. Finally, the observation of the C-FID-FWM transients is an example of nonlinear experiments where the pulses are at least one order of magnitude shorter than the coherence of the sample. 4.3. Control of quantum phase with coherent light Spectrally dispersed FWM is shown to be an ideal tool for studying intramolecular dynamics and this idea is applied to understanding the role of the phase profile of the laser pulse and chirp in controlling molecule-laser interactions. Experimental control and characterization of intramolecular dynamics are demonstrated using chirped femtosecond three-pulse four-wave mixing (FWM). The two-dimensional (spectrally dispersed and time-resolved) three-pulse FWM signal is shown to contain important information about the population and coherence of the electronic and vibrational states of the system. The experiments are carried out on gas-phase I2 and the degenerate laser pulses are resonant with the X (ground) to B (excited) electronic transition. Control over population and coherence transfer in the absence of laser chirp is demonstrated in previous section by selecting specific pulse sequences. When chirped lasers are used to manipulate the optical phases of the pulses, the two-dimensional data demonstrate the transfer of coherence between the ground and excited states. Positive chirps are also shown to enhance the signal intensity, particularly for bluer wavelengths. The calculations based on the multilevel density matrix formalism in the perturbation limit allow prediction of particular pulse sequences that control the final electronic state of the population. In a similar manner, the theory allows us to find critical chirp values that control the transfer of vibrational coherence between the two electronic states. The 84 ability to control population and coherence transfer in molecular systems is of great importance in the quest for controlling the outcome of laser-initiated chemical reactions. 4.3.1. Introduction Experimental demonstrations of laser control of chemical reactions in the past decade have solidified this field of research and garnered great excitement. Emphasis in this section is put on the controlled formation of ground- and excited-state wave packets and the time-dependent polarization that is observed by employing a variation of the femtosecond three-pulse four-wave mixing (F WM) technique, where the signal is spectrally dispersed. This method is very useful to control and characterize the intramolecular dynamics occurring after laser excitation. In this study, we experimentally explore the role of different pulse sequences and optical phase manipulation (linear chirp) in controlling the population and coherence transfer in gas-phase 12. The two-dimensional (time and frequency) data provide phase information that cannot be obtained from similar techniques without spectral dispersion. Presented experimental results and theoretical simulations reveal the role that pulse sequences and laser chirp can play in controlling intramolecular dynamics. The goal of these experiments is to gain a deeper understanding about the role of chirped laser pulses in controlling chemical reactions. The results from this study are consistent with the experimental observation of wave packet modification (for example focusing), and enhancement in the yield of chemical reactions using chirped femtosecond pulses. As in the previous section, gas-phase 12 was chosen as a model molecule for the experimental demonstration of this technique. The visible X ’Zo+g <—) B 3110+u transition 85 has been well characterized by frequency-resolvedl90"95 and time-resolved spectroscopy.l9‘5'I98 The vibrational spacing is quite different for these two states making their identification in the data a simple task. An important step in the control of molecular excitation is the experimental characterization of the system being controlled. Conventional pump probe (PP) and FWM techniques give very high temporal resolution but give little or no spectral information. Some variations of the PP method have been shown to provide frequency resolution; among these are time-gated fluorescence techniques'97 and PP-photoelectron spectroscopy.‘999200 The additional frequency resolution helps one to obtain a more complete characterization of the system. The time-gate used to obtain the spectral information determines the spectral resolution that can be achieved, limited by the uncertainty principle. Here we spectrally disperse the signal from the three—pulse FWM. This signal contains a wealth of spectroscopic information. We have found that this information allows us to characterize the system more fully following different pulse sequences and different chirp values. It is shown that the pulse sequence in three-pulse FWM allows one to form wave packets in the ground and/or excited states. A brief theoretical background is given in terms of the evolution of the density matrix. Calculations are used to predict interesting behavior that can be achieved by particular sequence-chirp combinations. Complete density matrix treatment of non-linear optical processes is beyond the scope of this thesis and can be found in the Reference337~'04. Emphasis is put on the spectroscopic information that is contained in the F WM signal and its connection to the density matrix. Results are presented for various pulse sequences and pulse chirps. Theoretical calculations are found to be in 86 excellent agreement with the experimental data and are presented here insupport of the experimental results. 4.3.2. Theory The density matrix formulation is the preferred theoretical approach for the study of coherent multi-wave mixing experiments.36~87 The diagonal blocks of the density matrix of a multilevel system involving several vibrational levels in the ground and excited electronic states represent properties within each state (see Figure 4.10a). In the diagonal blocks, the diagonal elements (designated as pgg or pee in Figure 4.10a) are the population of each vibrational level, whereas the off-diagonal elements (pgg' or pee.) represent the coherence of the vibrational levels. The elements of the off-diagonal blocks (age or peg) represent the vibronic coherence between the two electronic states. In the weak interaction limit, each field E” interacts linearly with the medium producing a change pl") in the initial density matrix. When n is an odd number, pm contains the changes in the probability amplitude of the electronic couplings, whereas if n is an even number, p("’ represents the changes in the population and the coherence of the vibrational levels within each electronic state. The time evolution of the density matrix for molecules interacting with consecutive applied weak electric fields can be represented by DSFD as described in Section 4.2. The diagrams relevant to our pulse sequence and phase-matching geometry are shown in Figure 4.10b. Only first interactions with electric fields E, and E, separated by a time delay Tab are shown. The interaction with the third electric field EC and the emitted signal E, are not shown. 87 a) Multilevel Density Matrix b) Double-Sided Feynman Diagrams I - P - (2) (2) peg. pgg' pee '08? p” ' Ig> e'l pg'g ‘. E, O 1 a 9: - i" ”"90”? '“ ""1 % Tab: : Tab: 130 ' l ( ' |e> |e> O pe’e A l f l E. E. peg pee pee lg> <81 18> (gI I Figure 4.10. (a) Representation of the density matrix for a multilevel system composed of two electronic states, each one with a manifold of vibrational levels. The indices g, g' range over the number of vibrational levels of the ground electronic state (in the present simulation g, g' = 1, 2). The indices e, e' range over the number of vibrational levels of the excited electronic state (in the present simulation e, e' = 3, 4). The diagonal matrix elements, pgg and Pee, represent the population of the vibrational levels (dark shade). The off-diagonal matrix elements of the diagonal blocks, pgg' and peev, represent the vibrational coherence in the respective electronic state (light shade). The elements of the off-diagonal blocks, pge and peg, represent the vibronic coherence between the electronic states. (b) Double-sided Feynman diagrams for the processes described in the text after the first two electric field interactions only. The complex conjugate diagrams, mirror images of those shown in Figure 4.1b, lead to FWM signal in a different phase-matching geometry and are not shown. The signal is the sum of all possible processes and arises from the contributions of pg:g and peel which correspond to coherence in the ground and excited states, respectively, after the interaction with the first two pulses. Each process depends on the molecular dynamics of the corresponding electronic state; therefore, by enhancing one contribution with respect to the other, it is possible to extract some information about the molecular dynamics of each state. In a FWM process, the first two electric fields interfere and form transient and 88 static gratings in space. The third pulse probes these gratings, inducing a polarization in the medium which emits coherent radiation. When (00, the carrier frequency of the pulses, is resonant with an electronic transition frequency, the interfering fields induce two successive dipole transitions yielding a redistribution of the vibrational populations and coherences of the molecular ensemble, described in pm. In this way the time-dependent matrix elements, plz’gvg (t) and pm“, (t), which oscillate with the vibrational frequencies of the ground and excited states, cog and we respectively, describe the transient grating. The diagonal terms, pmgg and pm“, describe the static grating which decays with T1. Note in Figure 4.10b that the diagram on the right contributes to pmee' while the one on the left contributes to plz’gvg. The transient FWM signal will be a maximum at the times when all molecules radiate in phase as long as the dephasing time of the excitation coherence, T2, is longer than the duration of the pulses. The density matrix formulation for the interpretation of the signal obtained from chirped three-pulse F WM experiments is applied. The theoretical analysis is based on a model that includes two electronic states with two vibrational levels each; the states are labeled as |1>, |2>, |3>, and |4> where g = 1, 2 and e = 3, 4. The coherence relaxation time for 12 in the gas phase is much longer than the times considered here; therefore, no relaxation terms are included. Resonant experiments reduce the electronic manifold to the states that are coupled by the electronic dipole interaction. In order to control the molecular dynamics of such states, the theory should include the parameters that characterize their dynamics (i.e. vibrational frequencies). Therefore, at least two vibrational levels with an energy separation of AEg = haig or AEe = ha), in each electronic 89 state will be needed. Although the effect of anharmonicity cannot be included in the formulation with this reduced model, an analytical solution to the optimal control problem is achieved. Unless otherwise noted, it is assumed that the initial populations are pm)“ = p(°’22 = 0.5 and p(°’33 = p(°’44 = 0. The parameter w = pm)” - p(°’22 is used to indicate a different initial population distribution. The three pulses are of same frequency and designated as Ea, Eb, and EC and the FWM signal is emitted in two directions defined by the wave vectors kyg = k, -— kb + kc and k5 = — k0 + k1, + kc. The former applies to observation of a virtual echo and the latter to the conventional photon echo. Calculations and experimental results shown here apply to detection in the km direction. As described in chapter 4.2 the simultaneous control of the population and vibrational coherence within each electronic state can be achieved by manipulating the time delay between the first two pulses, Tab. After the interaction with two unchirped pulses, the matrix elements of the ground and excited states are (0 Pg) = —A2 COS[&ZTCIP—lcos[(wo i 7g}... " (k, - k;.)' X], (4.3) P1230) = A2 coleBTT‘QJeiiwAH’T”)cos(a)ofab - (kg - kb)' x), (4-4) 0)”! p33) = A2 cosl-iZ—fll cosllwo $—%—l2ab — (kg — k b) xl, and (4.5) 90 a) 2' .02.?)(0 = A2 cos[ ‘2 °° lei’m"(’+’r”” )cos(a)02ab — (kg — k b)- x) (4.6) where A is the area under the pulse and the upper sign corresponds to the first combination of indices and the lower sign to the second. Setting 20;, = (n + l/2) 2,. where 2,. = 2n/we, we obtain a signal that is mostly characterized by the dynamics of the excited state and similarly for 22,), = (n + ’/z)2g with 2,, = 22t/wg. The vibrational coherence in the ground state corresponds to the transformation of the wave packet after an excitation — de-excitation process, whereas the vibrational coherence in the excited state corresponds to the interference of two wave packets after excitation. The probability of each mechanism is represented by the diagonal blocks of the density matrix after two pulses, pm“ and p (2’29! respectively. When the first two pulses are equally chirped, the chirp is manifested in the vibrational coherence matrix elements plz’gga) and pmeea) and is absent in pa)“ and pa)“. This is because the population matrix elements are obtained after a second-order process with two equal and opposite transition paths and the phase shifi of the successive chirped pulses cancels out. When the transition paths are different, as for the coherence matrix elements, a net phase due to the chirped pulses arises. The modifications introduced by the chirped pulses in the spatial average of the density matrix elements are 0C [1 + cos(a)erab )l, (4.7) 91 oc 1+cos(werab)cos ¢ 2" , (4.8) (2) pee < 2> cc [1 + cos(a)grab )l (4.9) n 2 2 a) (2) >01: 1+cos(a)grab)cos ¢ 2g (4.10) pee' < where the brackets denote the average with respect to coordinate 11. ¢" is the linear spectral chirp of the pulses obtained from the phase shift of the chirped electric fields, ¢(a)) -.~: ¢(a)0) + (a) - (any ¢"/2 with ¢" = 6221160)sz Note that these expressions. (Equations 4.7-4.10) are time independent. The modulus square cancels the time- dependent terms in Equations 4.2 and 4.4. After time delay 2', the third pulse in a FWM experiment probes the population and the vibrational coherence generated by the first two pulses. The resulting polarization can be represented by the third-order density matrix [3(3) containing only the matrix elements that oscillate with the transition frequencies (neg = (E, — Eg)/h where g = 1, 2 and e = 3, 4. Therefore, the emitted light contains useful spectroscopic information about 2)” + p(2)E, where the components that the system. The overall signal depends on pm = p‘ lead to signal in the km; direction are separated from those that lead to signal in the k}; direction. The intensity of the emitted light corresponding to each transition in the direction kVE as a function of 2'is given by 92 . , ,. . 2 VE 2)VE o (2)25 (2)15 — <1) (2)VE 1,,g (2) 0C pie +el pee. (2)—pgg +e ' pgrg (2) (4.11) where (I) = ¢"a)gZ/4. Since the diagonal terms of the density matrix p‘z’VEgg and pawn“? are time independent, the time evolution of the spectral line intensity depends on the vibrational coherence matrix elements. Because these matrix elements oscillate with a vibrational frequency cog or (02, Equation (4.11) can be expressed as 1§f(2) 0c 1”,.(2) + 1"E.(2) + 1”“.(2), with (4.12) [1'5 ( (2) 2 g 2) oc lpgg'l [cos(a)g2) + l] and (4.13) z 2 I’Ee(2) 0C [cos(a),,2)+1]. (4.14) 1’58. and [”58 represent the spectral intensity contributions that oscillate with the frequencies mg and (02., respectively. IVEC represents the cross terms that oscillate with a . . . . . . E VE . . lrnear combination of both frequencres; 1fe1ther [V g or I e rs zero, IVEC 1s zero. Figure 4.11 represents the simulated relationship between the amplitude of the averaged density matrix elements, as indicated in Equations (4.7-4.10), and the spectral intensity, as shown in Equation (4.11). The three-dimensional diagrams of the averaged density matrix elements for the four-level system are shown at the top of the figure. The height represents the amplitude of ( l p.112) l 2). The spectral intensities as a function of time delay 2 are shown in the spectrograms. The spectrograms show the contributions of the four transition lines that correspond to the four-level system (141 = 4—21, 131 = 3—>1, 2.42 93 = 4—>2, and 132 = 3——)2). Gaussian functions with a Alt/2 F WHM are used to simulate the experimental spectral resolution where Al = 2.41—2.31 = 142—132. Based on Equations (4.7- 4.10), one can identify the time delay 20b as a useful parameter in the design of pulse sequences that enhance or eliminate specific elements in the density matrix. Two extreme cases with either exclusive ground- or exclusive excited-state dynamics are shown in Figure 4.11. When 2a}, = 400 fs = 52g/2 (Figure 4.11a), the spectral intensities oscillate with the vibrational frequency cog, with 28 = 160 fs, consistent with the vibrational frequency of iodine in the ground state. The signal corresponds to the time evolution of the probability amplitude in the ground state because only the matrix elements pa’gg and plz’gvg( 2) are different from zero (see the averaged density matrix diagram at the top of Figure 4.11a). When 20;, = 460 fs = 32e/2 (Figure 4.11b), the spectral line intensities oscillate with vibrational frequency me, with 2., = 307 fs, consistent with the vibrational frequency of the B 3110.u state of iodine (v' = 7-11). As can be seen from the averaged density matrix diagram (Figure 4.1 lb top), only the elements of the excited state contribute to the signal in this case. Therefore, by changing the time delay between the first two pulses, it is possible to control the population and the vibrational coherence of the system. The use of chirped pulses as an alternative means of controlling the intramolecular dynamics is explored in the simulations shown in Figure 4.12. The spectrograms along with the diagrams of the averaged density matrix elements correspond to time delay 20;, = 460 fs with chirped pulses having a spectral chirp ¢"= 1000 rs.2 or wag/2 = 11/4 (Figure 4.12a) and 25" = 4000 fsz or (25"(082/2 = 1: (Figure 4.12b). 94 a) Tab = 400 fs b) Tab = 460 fs Time delay 2, [ps] A'41 A'31 A'42 A'32 Figure 4.11. Simulations of spectrally dispersed FWM. At the top of the figure are diagrams of the averaged density matrix elements <|p1j(2’|2> calculated with Equations (4.7-4.10). The corresponding spectrograms calculated with Equation (4.11) are shown with wavelength plotted along the horizontal axis and time delay between the second and third pulses (2) plotted along the vertical axis. The darker regions in spectrograms correspond to higher signal intensity. For these calculations, the initial population distribution parameter (w) in the ground state was assumed to be zero and the pulses were unchirped. (a) The delay between first two pulses (2,2,) is 400 fs, corresponding to 5/2 vibrational periods of the ground state, 23. The density matrix diagram for this pulse sequence (top left) shows population and coherence terms only in the upper diagonal block, corresponding to ground-state matrix elements alone. The spectrogram (left) clearly shows 160 fs periods, reflecting the vibration of the ground state (cog). (b) For 22,1, = 460 fs, 3/2 vibrational periods of the excited state 2,, only the matrix elements of the excited state (lower diagonal block) are present in the diagram. In the corresponding spectrogram (right), the excited-state vibrations of 307 fs are evident. 95 The introduction of chirped pulses produces a signal that oscillates with both vibrational frequencies cog and are, confirmed by the presence of coherence vibrational terms belonging to both electronic states in the averaged density matrix diagram. When more than one oscillating frequency is present in the spectrogram, the spectral line intensity is slowly modulated in time with an approximate period of 3 ps. This effect can be attributed to the interference, ICVE( 2), of the vibrational coherence terms from the different electronic states according to Equation (4.12). Comparing the averaged density matrix elements for different values of chirp for 2,), = 460 fs (Figures 4.1 lb top, 4.12a top, and 4.12b top), we can see that the chirp does not affect the population terms but only the vibrational coherence terms, pa’ggfl) and plz’eea), in agreement with previous the discussion. Thus, the time dependence of the spectral signal can be manipulated through chirped electric fields by controlling the coherence vibrational terms of the density matrix after two pulses. Figure 4.12 highlights the importance of spectral dispersion; the intensity of the spectral features varies at different wavelengths due to the different phases involved. This information would not be available if the spectral information was integrated. The density matrix elements after two pulses are the ensemble-averaged products of two time-dependent amplitudes, p‘z’y-(t) = (c,-(t) cj(t)’), of the wave packet components in each electronic state. A change in the magnitude of p‘z’gga) or pmeefl) due to chirped pulses will correspond to a change in the relative phase of the wave packet components. This manipulation shapes the wave packet in each electronic state, allowing the focusing or spreading of the wave packet.‘3~l4,'6~7O 96 Tab = 460 fs a) ¢"=1ooofe:2 b) ¢"= 40002132 Time delay 2, [ps] A'41 A'31 A": A:12 A'41 2‘31 A42 A‘32 Wavelength Wavelength Figure 4.12. Simulations of spectrally dispersed FWM for 22,), = 460 fs, w = 0, and two values of laser pulse chirp. Again, the density matrix diagrams are time independent throughout the region shown in the spectrograms. (a) For a small value of chirp (¢" = 1000 fsz), the changes in the dynamics are obvious (compare to Fig 2b). These differences are expected, as seen in the density matrix diagram by the appearance of vibrational coherences in both the ground and excited states (upper and lower diagonal blocks). (b) It is possible to calculate the appropriate chirp where ground state elements dominate the contribution to the signal even for 22,), = 460 fs where unchirped pulses result in seeing only excited-state dynamics. With ¢" = 4000 fs2 (Figure 3b), 160 fs oscillation periods appear in the simulation, showing primarily ground-state dynamics. 97 When the initial population distribution is not limited to a single vibrational state, the density matrix elements represent the evolution of all levels in the system after successive laser pulses. A coherent coupling arises within each electronic state as a consequence of interaction with phase-matched electric fields. This results in additional coherence terms introduced, for example, by the thermally populated vibrational levels in the sample prior to laser excitation. The density matrix formulation treats these coherences naturally without having to artificially add these contributions using phase coherent or incoherent sums. By taking only four levels into account, analytical expressions are obtained for the spectrally dispersed F WM signal as a function of the control parameters 22,), and (21" using the density matrix formalism. Simulations based on these expressions are in close agreement with the experimental data. 4.3.3. Experimental The experimental setup is identical to one previously described in section 4.2. These experiments were conducted using 60 fs pulses (FWHM when transform-limited) centered at 620 nm with 8 nm (FWHM) spectral width. The chirp was measured by analyzing the pulses with a frequency-resolved optical gating (FROG). The laser was split into three beams, each with energy of ~ 20 11.1, which were arranged in the forward box configuration before being focused in the quartz sample cell containing neat iodine vapor at 140 °C. The time delay between the first two pulses, 20),, was controlled by a manual translator and the time delay between the second and third pulse, 2, was scanned to yield the FWM signal transients. The three-pulse FWM signal was collected by a spectrometer placed at the phase-matching condition km; for different wavelengths within 98 the 605 — 635 nm spectral range. The spectral acceptance of the spectrometer was set to 2 nm. Narrower spectrometer slits did not seem to sharpen the observed spectral features. Both FWM spectra and transients were measured with these same conditions. The spectra were obtained between 600 —- 640 nm and averaged for 10 scans with 200 points per scan and 10 laser shots per point. Each transient was taken at 150 different time delays and averaged for 20 scans. At each time delay 2, the signal was collected for 10 laser shots, where the energy of each laser pulse was required to be within one standard deviation from the mean. These transients, collected at different wavelengths, were then combined to generate the spectrograms shown in the following section. 4.3.4. Spectral information in femtosecond three-pulse FWM Most four-wave mixing experiments measure the intensity of the signal either as a function of excitation wavelength (frequency-resolved FWM) or as a function of time delay between laser pulses (time-resolved FWM). For FWM measurements involving resonant excitation, the polarization can be sufficiently long-lived that it carries relevant spectroscopic information."3 This is particularly the case for gas-phase samples where the polarization can persist for tens to hundreds of picoseconds.100 Spectrally dispersed FWM measurements are not reserved only for long-lived polarization systems; for a radiative decay of 1 ps, one can expect to get 5 cm'1 spectral resolution. The spectral resolution in this case is almost two orders of magnitude better than the bandwidth of the femtosecond lasers used for the experiment. In Figure 4.13, the spectrum of the transform-limited femtosecond laser used for these experiments (b) with the spectra of three-pulse FWM signal for two cases, (a) and 99 Intensity [arb. units] I . —l Tab= 614 fs - T: 614 fs 609 613 617 621 625 629 633 Wavelength 2» [nm] Figure 4.13. Experimental spectra (dots) and fit (lines) of the three-pulse FWM signals and the transform-limited laser. (a) FWM signal observed when all three beams coincide in time (22,), = 0 fs; 2= 0 fs). This spectrum is very broad but shows some spectroscopic transitions. The spectrum is fit with six Gaussian functions. (b) Spectrum of the unchirped femtosecond laser pulses used for these experiments. This profile is fit with a Gaussian function with width 8 nm (FWHM). (c) FWM signal observed when 20,, = 614 fs and 2= 614 fs. This trace shows well-resolved spectroscopic transitions which were fit by six Gaussian firnctions. The spacing between the peaks in spectra (a) and (c) is 4 nm, corresponding to the vibrational spacing of the B state of iodine. (c) is shown. The experimental data (dots) were fit by Gaussian functions (black lines). The laser has a bandwidth of 210 cm". Trace (a) corresponds to the signal observed when all three beams coincide in time. This signal spectrum is very broad but shows some spectroscopic features with an average width of 80 cm". Trace (c) shows well-resolved spectroscopic transitions and corresponds to a three-pulse FWM in which 201, = 614 fs and 2: 614 fs. The average width of these features is 50 cm". The main difference between traces (a) and (c) is that the signal in the former case includes multiple FWM pathways, 100 many involving non-resonant processes that are short-lived,92 and the latter case arises from a long-lived polarization that decays in hundreds of picoseconds.'00 The spectral features are much narrower than the laser but are significantly broadened by the rotational temperature of the sample (spectral congestion) and the spectral resolution of the spectrometer. No significant improvement in the spectral resolution was obtained when much narrower slits were used. The spacing between the features in Figure 4.13c is A?» z 4 nm and corresponds to the vibrational spacing in the B 3110+u state?“ The transitions arise from a polarization involving vibrational levels v' = 7-11 in the B state and vibrational levels v" = 2-4 in the X ’Zo+g state. The vibrational levels with the highest probability of excitation are v' = 9 and 10. Given the very small rotational constant of iodine molecules, there is a great deal of spectral congestion. Furthermore, the coincidental factor of two between the vibrational frequency of the ground-state and that of the B state results in multiple band overlap. The data in Figure 4.13 make it evident that spectral dispersion of the three-pulse FWM signal provides an additional dimension with a wealth of information about the system. In the following section the changes in the spectra as a function of time delay of the pulses involved in the three-pulse FWM process are explored, as well as the effects of chirped-pulse excitation. 4.3.5. Two-dimensional three-pulse FWM measurements As discussed in section 4.3.2, the spectroscopic transitions that are detected here are determined by the system and their wavelengths are time independent. The time- dependent intramolecular dynamics manifest themselves as changes in the intensity of the 101 individual spectroscopic features. Therefore, the modulation of the signal is a manifestation of a time-dependent transition probability of the system. Based on these principles, it is not necessary to take a high-resolution spectrum at each time delay. The two-dimensional data were obtained by taking transients at 5 to 6 different wavelengths and interpolating the rest of the spectroscopic information using multiple Gaussian functions in the same way that they were used in Figure 4.13 to fit the spectroscopic data. In all cases, spectra were taken at different time delays 2 to check the two-dimensional data for wavelength and bandwidth consistency. Given that the data contain temporal as well as spectral information, it should be possible to invert and obtain the real space dynamics of the system in both the ground and the excited states. Figure 4.14 shows spectrally dispersed three-pulse F WM measurements carried out with 20;, = 0 fs. Both plots present the two-dimensional experimental data. Figure 4.14a shows the signal intensity in the third dimension giving a topographic representation. Figure 4.14b shows the same signal in a contour form, apparently more usefiil for quantitative analysis. In the contour plot, darker shades correspond to higher intensity. The contour lines are spaced according to the square root of the signal amplitude. One of the important points to notice is that the spectroscopic features remain fixed in wavelength and one observes their change in intensity as a function of time delay 2. This makes this technique different from wave packet imaging techniques such as those based on photoionization19 or Coulomb explosion.202~203 With those, one obtains the shape of the wave packet and its time evolution. In our measurements, the dynamics of the system are obtained from the change in intensity of every spectral region. The spectroscopy as well as the dynamics reveal the states involved; in the experiments 102 discussed here these would be the X and B states with cog = 208 cm'1 and a), = 108 cm'l (2g = 160 fs and 22 = 307 fs). In Figure 4.14 the observed dynamics for wavelengths longer than 620 nm are a mixture of ground- and excited-state vibrational motion; however, the features with a period of 307 fs, corresponding to the excited state dynamics, are more prevalent especially in the 628 — 632 nm region and after the first 1.5 ps. For wavelengths shorter than 620 nm, the dynamics show a vibrational period of 160 fs that corresponds to the ground state. The observation of ground and excited state dynamics for FWM experiments with 20;, = 0 fs was reported first by Schmitt et al.38,169 Basically, the first two pulses create populations in both the ground and excited states. The third laser pulse probes the dynamics of both populations and the signal contains a mixture of each. Due to the Franck-Condon factors between both of the electronic states involved, probing of the ground state will be enhanced for short wavelengths while probing of the excited state is enhanced for longer wavelengths. These enhancements result in the observed differences in the signal as seen in the experimental spectrograms. With three-pulse four- wave mixing, one is able to prepare different pulse sequences that can be used to control the coherence and population, and therefore the observed dynamics of the system. Pulse sequences have been used in liquid phase FWM experiments to suppress vibrational mode contributions183 and to control the emission of signal by allowing the echo and virtual echo components to interfere.”204 Experimental data and simulation of results for the case when 2a), = 614 fs are shown in Figure 4.15 (a) and Figure 4.15 (b), respectively. 103 Tab=ors ¢"=0fs’ a) Experiment b) Experiment we Time delay 2, [ps] - “flu-rflfiwfidfiefi‘efl! 5"/i‘ffi’/!!§f§};3; t I D ID ‘— 8 612 ' Wavelength [nm] Wavelength [nm] Figure 4.14. Spectrally dispersed experimental data for three-pulse FWM with 202 = 0 fs and transform-limited pulses. Transients taken at six different wavelengths for 3.3 ps are represented as (a) a topographical plot and (b) a contour plot. As before, the darker areas in the contour plot correspond to the higher signal intensity whereas the topographical plot uses lighter shades to show higher intensity. Spectral profiles were approximated by Gaussian profiles (see Figure 5a and c) to generate the experimental spectrograms. Both ground- and excited-state dynamics are observed with these experimental conditions. 104 Tab=s1us ¢=0st a) Experiment b) Theory , 3.3 3.0 2.7 Time delay 2, [ps] Wavelength [nm] Wavelength [nm] Figure 4.15. Spectrally dispersed experimental and theoretical spectrograms of three- pulse FWM for 201, = 614 fs and ¢" = 0 fsz. (a) The spectrogram shows the experimental transients taken at five different wavelengths where the spectral profiles were approximated by Gaussian profiles. For all wavelengths, the dominant feature is the 160 fs vibrational motion reflecting the ground-state dynamics. (b) The initial population distribution was taken to be w = 0.33 for the calculation of the theoretical spectrogram of the experiment shown in (a). The 160 fs vibrations from the ground state dynamics are evident in both the theoretical and experimental spectrograms. This time delay maximizes the contribution from the ground state. The experimental data show fast oscillations with a period that corresponds to the vibrational frequency of the ground state. There is a small contribution from a population in the excited state that is apparent in the 628 nm region. The data show a phase shift of about 80 fs when comparing the maxima of the spectra features for 612 nm to those at 628 nm. This delay is probably due to the anharmonicity of the B state. Notice that the intensity 105 remains high for the bluer wavelengths. The calculations, carried out for w = 0.33, which corresponds to the expected Boltzmann distribution between the two vibrational levels in the ground state at 140 °C, reproduce the observed dynamics showing higher intensities for shorter wavelengths and showing that some of the excited state dynamics are observable for this pulse sequence. The calculations, based on a four-level density matrix, do not have a sufficient number of levels to reproduce the observed phase shift. The wave packet calculations described in the Section 4.2 (Figure 4.5) demonstrate that the observed ground state dynamics arise from a wave packet that evolves in the B state for a time period 201,. During this evolution, the anharmonicity of the B state causes the wave packet to spread. When a pulse sequence is chosen with 20,, = 460 fs, only the excited state is populated (see Figure 4.11). The experimental data, shown in Figure 4.16a, show very clear 307 fs oscillations that correspond to the vibrational period in the excited state for v' = 7-11. In this case there is no evidence of ground state dynamics. This data can be contrasted to the data in Figure 4.15 where ground state dynamics are dominant. The theoretical calculations, shown in Figure 4.16b and carried out for w = 0.33, are in excellent agreement with the experimental data. For this sequence, the intensity distribution across the spectral range is quite symmetric and shows no phase shift. Note that for this 2gb, the calculations are independent of the initial population (w) in the ground state. 106 Tab = 460 fs ¢"= o fsz b) Theory .3 \Jv\,_.4\r/\I 3.0 2.1 2.4 '3‘ 2.1 ,2; 1..“ 1.2 > . 2 - e _ 1.5 U - 0 I 1.2 .E _ 1- : 0.9 ; : 0.5 : ' ' : 0.3 '. l 0 D D 2 225232 =34: Wavelength [nm] Wavelength [nm] Figure 4.16. Spectrally dispersed experimental and theoretical spectrograms of three- pulse FWM for 2a,, = 460 fs and (2" = 0 fsz. (a) The spectrogram generated from the transients taken at five different wavelengths shows only 307 fs oscillations from the excited-state dynamics. There is no evidence of ground state dynamics. (b) The simulation of the experiment again used w = 0.33 for the calculation. The theoretical and experimental spectrograms are in excellent agreement, with the 307 fs oscillations clearly evident in both. 4.3.6. Three-pulse FWM with chirped laser pulses Based on the experimental data shown so far, it is clear that the spectrally dispersed three-pulse FWM data provide a wealth of information about the intramolecular dynamics of a system including information about the population 107 distribution among the active electronic states. The role that chirped laser pulses have on the excitation of molecular systems is explored in this section. Recent experiments where chirped laser pulses have been used to control the excitation of molecules and the yield of chemical reactions provide the motivation for this study. '5'13270273-74 Using the pulse sequence with 20;, = 460 fs and fixing the time delay of the third pulse to 600 fs provides a snapshot of the system in time. While keeping these two parameters fixed, chirp was varied. The experimental data for this measurement are presented in Figure 4.17; data are interpolated between the spectra obtained at the different chirp values. Notice that for zero chirp the spectrum is centered and is quite symmetrical, as discussed earlier. For positive chirps, we notice that the intensity shifts towards shorter wavelengths, while for negative chirps, the intensity shifts towards longer wavelengths. These shifts and their time dependence are discussed below. It is interesting to note that the highest signal intensities are observed for positive chirps. This observation is consistent with the fact that optimal pumping of the B state of iodine requires a positively chirped pulse.‘7’73,74,l ‘0 The experimental and theoretical data for the two-dimensional three-pulse F WM where 2a,, = 460 fs and the laser pulses are chirped to 3300 fs2 are presented in Figure 4.18 (a) and (b) respectively. Comparing the experimental data in Figure 4.18a with the experimental data in Figure 4.16a obtained for the same pulse sequence, it is clear that the chirp has introduced ground state dynamics into the observed signal. While the excited state dynamics are still clearly observable, there are some regions showing 108 600 ' " r , _ 620 , 2800 Wavelength ,1 [11m] Figure 4.17. Three-dimensional representation of the experimental spectra taken using differently chirped pulses for the pulse sequence 2“,, = 460 fs and 2= 600 fs. Chirp values ranged from —2800 fs2 to 2800 fsz. Note the change in intensity towards shorter wavelengths as the chirp values increase from negative to positive values. oscillations with the 160 fs characteristic of the ground state. There are spectral regions, for example 620 nm, where there is a clear doubling of the frequency with every other beat having higher intensity. The theoretical calculations in Figure 4.18b, carried out with w = 0.33 and 25" = 3300 fsz, are in excellent agreement with the experimental data. The calculations reproduce the shift of spectral intensity towards shorter wavelengths and the alternating intensity between the vibrational features. 109 Tab = 460 fs ¢"= 3300 fsz a) Experiment b) Theory 3.0 s: 9 O 2.7 2.4 2.1 3' 1.3 3 1..“ 1.5 >1 2 1.2 g d) 0.9 _§ .— 0.6 7 . I 0.3 ’ ‘ 2 3 § 3 o u- 2 N N a g g ‘9 on 9 '9 10 GD ‘9 O O O (D 0 Wavelength [nm] Wavelength [nm] Figure 4.18. Spectrally dispersed experimental and theoretical spectrograms of three- pulse FWM for 2.2, = 460 fs and ¢" = 3300 fsz. (a) The experimental spectrogram was generated from transients obtained with five different wavelengths as above. The signal observed for wavelengths 624 nm and shorter originates from mixed ground- and excited- state dynamics. It is clear when compared with Figure 8a that the addition of chirp alters the observed signal by introducing ground-state dynamics for shorter wavelengths. (b) The theoretical simulation, where w = 0.33, shows good agreement with the experiment both in the dynamics features and in the intensities of the beats. It is possible to find ‘critical chirp’ values that accomplish certain tasks, for example, eliminateing the ground- or the excited-state dynamics in the signal. Here we would like to focus on two interesting cases. In Figure 4.19 we show simulations for the pulse sequence with 2,, = 614 fs. When a relatively small chirp of 1000 fs2 is used (a), we 110 see that the ground-state contributions to the signal decrease (compare to Figure 4.15b). The data show an apparent slope in which the spectroscopic features on the short wavelength side seem to reach a maximum earlier than those on the long wavelength side. This slope depends on the sign of the laser chirp. For negative chirps, the opposite slope would be observed. The second interesting case, shown in Figure 4.1%, corresponds to the same pulse sequence but with a critical chirp ¢" = (222/20,?) = 4000 fs2 (see Equations 6 and 8). In this case, only 307 fs dynamics are observed. The spectral intensity is still shifted towards shorter wavelengths as expected for w = 0.33 and 20,, = 614 fs. For w < 0, a red shift in spectral intensity would be observed. However, most notable is the phase of the observed oscillations across the different wavelengths. The average density matrix diagrams for both cases are also shown in Figure 4.19 (top). Notice that the laser chirp does not affect the population, but affects the coherence between the states. For the case shown in Figure 4.1%, the ground state coherence is no longer present and therefore not observed. As can be seen in Figure 4.15b, the spectrogram for 20), = 614 fs is dominated by ground-state dynamics. When the critical chirp of 4000 fs2 is introduced for 20), = 614 fs, the ground-state dynamics are completely eliminated (see Figure 4.1%). This result is similar to the one presented in the spectrogram for 20b = 460 fs and no chirp (Figure 4.16b) where only excited-state dynamics is observed. When the critical chirp of 4000 fs2 is introduced, the dynamics are dominated by the ground state (see Figure 4.12b). From a theoretical point of view, changing the chirp (¢") is as effective as changing the pulse sequence (20),) to control the observed molecular dynamics. It is much more complex task experimentally. lll Tab = 6141‘s a) 2521000242 b) (152400011:2 ©(‘) 1:)" '3 o 0 > ”st/7 C2; (filo? (9‘97 Time delay 2, [ps] A11 A'31 A'42 A'32 Wavelength Wavelength Figure 4.19. Diagrams of averaged density matrix elements <|,0y(2)12> and corresponding theoretical spectrograms for time delay 2a), = 614 fs and w = 0.33 with ¢" = 1000 fs2 and (15" = 4000 fsz. (a) For small chirps, the dynamics at shorter wavelengths reflect the ground state while higher wavelengths represent signal predominantly from the excited state. (b) When the pulses are chirped to the critical value, ¢" = (212/082) = 4000 fsz, the signal for all wavelengths reflects excited state dynamics only. Notice that the density matrix diagram in this case shows population but no coherence in the ground state. 4.3.7. Conclusions The signal arising from time-resolved FWM measurements carries valuable information about the intramolecular dynamics of the system. The time delay between the first two pulses is a valuable parameter for controlling the population and coherence transfer. In all cases the theoretical simulations were in good agreement with the experimental observations. The chirped pulse excitation measurements are of current interest as phase-altered pulses are being used to control the excitation process of molecules and the yield of laser- initiated chemical reactions. The data in Figure 4.18, obtained for (15" = 3300 fsz, show a contribution from the ground state that was absent in the dynamics obtained with unchirped pulses in Figure 4.16. It is possible to find ‘critical chirp’ values that can accomplish certain tasks, for example, introduce or eliminate ground state dynamics in the signal. Calculations that include multiple vibrational levels in the excited state would help to reproduce some of the finer features observed, such as the phase delay evident in the data for 22,), = 614 fs in Figure 4.15. From Equations (4.8 and 4.10) in the Theory section, it is clear that the laser chirp (¢") can be used as a parameter to control the coherence, and in a similar way, the time delay (201,) can be used to control the coherence and population transfer. The experimental data shown here illustrate the advantages of spectrally dispersed three-pulse F WM because it yields the following results that could not be obtained from spectrally integrated data: 1. For the case where 20;, = 0 fs, one can see that longer wavelengths are 113 dominated by excited-state dynamics while shorter wavelengths are dominated by ground-state dynamics (see Figure 6). 2. For the case 20,, =614 fs, the data show a phase shift of about 80 fs across the spectrum (see Figure 4.15). This shift is caused by the anharmonicity in the excited state. Therefore, the data contain phase information that is important in understanding the underlying intramolecular dynamics. 3. When chirped pulses are used, important spectral shifts occur. Figure 4.17 shows that for positive chirps the intensity shifts to shorter wavelengths and the signal intensity increases. Negatively chirped pulses give a broader spectrum that is shifted towards longer wavelengths. The above examples demonstrate both experimentally and theoretically that the chirp controls the coherence but not the population transfer in this system. It is important to note that all the laser pulses are equally chirped and only two electronic states participate in the observed dynamics. When the first two laser pulses are chirped differently (magnitude and/or sign), one can control the population transfer. When the system contains a reactive pathway, the coherence can be used to enhance or decrease the probability for the system to follow that pathway. The coherence determines the dynamics of the system. In multidimensional systems, one may want to control the yield of different pathways or affect the rate of intramolecular vibrational relaxation. This type of control, where the phases of the quantum mechanical superpositions are manipulated by the optical phases of the pulses to determine the outcome of chemical reactions, results in true control of final population transfer from reactant to product. Alternatively, 114 the first two pulses can be used to prepare the system in a specific superposition of states and the third pulse can be used to complete the chemical reaction. Work presented here has only involved pulse sequences of linearly chirped pulses. With phase masks one can achieve quadratic and higher order chirps. One can also tailor the pulses into very complex electric fields, which have been used to achieve different product yields in chemical reactions. Based on these observations, we consider three- pulse FWM to be a powerful tool for probing intramolecular dynamics and for achieving laser control of chemical reactions. 115 5. FEMTOSECOND PHOTON ECHO AND VIRTUAL ECHO MEASUREMENTS OF THE VIBRONIC AND VIBRATIONAL COHERENCE RELAXATION TIMES OF IODINE VAPOR 5.1. Introduction Previous chapters described control of intramolecular dynamics by manipulation of vibrational and vibronic coherences as well as pOpulation transfer between two electronic states of gas state molecular iodine. This chapter explores the three-pulse four- wave mixing technique as a tool to sort and measure the processes that contribute to coherence relaxation. All coherent control methods are affected by the rate of decoherence of the sample. Here we show how these rates are measured with FWM techniques. The measurements presented here illustrate how photon echo (PE) measurements yield the homogeneous relaxation rate while the virtual echo (V E) measurements yield the sum of the homogeneous and inhomogeneous relaxation rates. The observed relaxation times for vibronic coherence using photon echo and reverse transient grating measurements are compared at different temperatures to isolate inhomogeneous and homogeneous components. Different pulse sequences are used to select ground- or excited-state vibrational coherences. Measurements of ground- and excited-state wave packet spreading times due to anharmonicity, a process that does not involve energy dissipation or phase relaxation, are also presented. The loss of quantum mechanical coherence, a T2 process, in gas phase samples is 116 caused by collisions and inhomogeneous broadening (for example Doppler broadening). The cross section of these processes depends on the type of coherence. In a molecular system, one can identify electronic, vibrational and rotational coherences. Vibrational and rotational coherences can be found in a selected electronic state (ground or excited), and electronic coherences can include rotational and vibrational components as well. In the present study, three-pulse four-wave mixing (FWM) techniques are applied to measure experimentally the homogeneous, T,’, and inhomogeneous, T; , vibronic coherence relaxation times of iodine vapor. Starting with the first observation of photon echo (PE) in the condensed and gas phase,‘55~156 it was clear that this nonlinear optical method, analogous to spin echo techniques in NMR, was capable of providing coherence relaxation measurements free of inhomogeneous broadening. In the last decade, these measurements have been combined with ultrafast lasers to measure coherence relaxation times in large organic molecules in solution.‘64~205‘207 This method has not been as popular for the study of gas phase samples, and this work illustrates its advantages. There are many studies on the relaxation time of iodine in the gas phase. Among them are measurements of T1 relaxation (loss of population in the excited state) due to self-quenching,208~209 cross section measurements of phase relaxation in the presence of noble gases with femtosecond pump-probe methodsmu“ and T2 measurements by PE techniques on the nanosecond time scale.212 The goal of our measurements is to determine the rate of vibrational coherence relaxation of the B and X states independently as well as the homogenous and inhomogeneous vibronic coherence involving the two 117 electronic states. The theoretical foundation of these measurements is best expressed in terms of the time evolution of the density matrix in Liouville space.“87 This theory as it applies to presented measurements was briefly described in sections 4.2 and 4.3, based on an established formalism.379'0492‘3 In the presence of collisions, two types of relaxation can be observed: T1 processes, which involve the loss of amplitude in the population terms p,-,-. and T 2' processes, which involve the loss of amplitude in the coherence terms pi). Inhomogeneous broadening, T; , does not depend on collisions but also leads to a loss of coherence. Wave packet spreading due to anharmonicity does not involve coherence relaxation, but it leads to dephasing (and rephasing) of the observed vibrational coherence, which depends on the terms pggrand peel. This intramolecular dephasing time is defined here as a T3 process. A set of measurements using three-pulse FWM techniques, presented in this Chapter, sorts out all the contributions to quantum mechanical coherence dephasing in a molecular gas. Four different pulse sequences are used to sort out the different relaxation processes involving the (X)’>:g+ and (B)31'Io,,Jr states of molecular iodine. Homogeneous vibronic coherence relaxation rates between the ground and the excited state were measured using the PE pulse sequence. The full vibronic coherence lifetime was measured with the reverse transient grating (RTG) pulse sequence. The vibrational coherence relaxation in each electronic state was measured with VB pulse sequences as described in Chapter 4. For this method, the timing between the first two pulses was used to separate the contributions from ground and excited states. 118 5.2. Experimental The experimental setup used to carry out the measurements has been described in previous chapters. Briefly, 60 fs transform-limited pulses (FWHM) centered at 620 nm were used. The output was split into three separate arms with ~20 1.11 per pulse, and then combined non-collinearly in a phase matching geometry. The pulses were focused into a 45 mm long quartz cell. Temperature dependence was measured for PE and RTG in a cell that maintained a saturated iodine vapor pressure. The rest of the measurements were made at 200°C in a different cell where the optical density was kept constant at ~ 0.3, a value that is well below the saturated vapor pressure. The two diagonally opposed beams (a and c) and the beam between them (b) define the phase-matched emission in the direction k, = k, - k2, + k. This arrangement predeterrnines the sign for each wave vector. The time ordering of each pulse (the pulse sequence) is changed for each experiment using optical delay lines. The homodyne-detected signal was time (10 ns) and spectrally (16 nm) integrated. For all the pulse sequences, indicated at the top of each figure, the laser intensities were kept constant. The measurements presented here were taken with the same setup and under similar conditions. The only change was in the pulse sequence that selects the process to be measured. 5.3. Results and discussion Homogeneous vibronic coherence relaxation measurements on molecular iodine using the PE sequence are presented in Figure 5.1a as a function of temperature. The first pulse in the sequence generates the vibronic coherence pge, and its dephasing is measured as a function of the time delay 2. The double-sided Feynman diagrams (DSFD) are 119 presented to indicate the nonlinear process involved. In the range from 77 to 108 °C, the signal intensity decays exponentially. Because the signal is homodyne detected, it is proportional to the square of the polarization. Hence, we use T2 = 2 tcxp to calculate the homogeneous relaxation times measured, ranging from 7222:1304 to 310 ps for the lowest and highest temperatures (see Figure 5.1a). In this temperature range, the vapor pressure increases exponentially with temperature?'4 Homogeneous relaxation is caused by collisions; therefore, it is proportional to the iodine concentration in this temperature range, 1/ T,“ = now, where u is the average molecular velocity, n the number density and 0' the phenomenological cross section. Numerical calculation of the homogenous cross section based on the experimental data gives 0' = 11501150 AZ. This implies that long-range interactions are responsible for the collision-induced homogeneous vibronic relaxation. Our value for the homogeneous vibronic dephasing cross section is larger than the vibrational dephasing cross section measurements by Zewail’s research group in the presence of noble gases?‘ 1,212 This is to be expected when comparing self-quenching of iodine molecules to quenching by noble gas atoms such as helium, neon and argon.215 The cross section for coherence relaxation is also larger than those for T, relaxation due to self quenching, as expected. When the RTG pulse sequence is used (see Figure 5.1b top), the vibronic coherence relaxation measurements include both homogeneous and inhomogeneous contributions. The experiments were carried out under the same conditions as the PE experiments. In this case, however, the data no longer follow an exponential decay. Inhomogeneous broadening gives rise to a Gaussian line shape in the frequency domain 120 a) Homogeneous vibronic relaxation Inhomogeneous vibronic relaxation b a a b ca signal __ c badgna] \ -_ -_ a peg peg b p“ -c- -— 2’ time .. b 1' (in: c P110100 EChO Reverse Transient Grating 1.01'"; " " 1.0-rs; - - .4 - - - 2-- - - r- ~r - 1 n" . : x i V," -' : ‘ .' 3 . " 1 I _ D s K. , _ . . a . :i -—': 0.1i o‘oooo 'l '2'. 0.1.1 ': £3 ‘ ' yo 1 a i 1 2. ° a. ., ..- M... a» : z = O O 1 "' ‘ 3 0 ° 0 3 = , . o’ 0 ‘ a 2 ‘ '-' IP13”) 0C exPl-217'Tz) o E lRTdT) 0C CHM-(17'1" 2) ) e 0..., ., 1 5 am. .. : a 78°C,T'2=l302145ps 0' [_ 3o 78°C,T'2 = 229i8ps a 3 'e 88°C,T2= 778il6ps o . 9‘ 1088°C,T'2=221:t5ps o I , .98°C,T§= 460i10ps , ‘498°C,T'2=214tsps ‘ olO8°C, T'2 = 310i 9ps °lO8°C, T’ = l99i3ps 0.001 .................................... 0.002-.-” ,_ n, ...... 0.0 0.1 0.2 0.3 040506 0.70.8 0,0 01 vino: 0,3 0,4 Time Delay 2, [11s] Time Delay 2, [us] Figure 5.1. (a) Pulse sequence, corresponding DSFD, and PE measurements of the homogeneous vibronic relaxation between the X and B electronic states of 12. The time delay between pulses b and a(c) was scanned. The DSFD indicates that the signal depends on the off-diagonal terms pg... The logarithm of the amplitude normalized signal FWM emission is plotted as a function of time delay 2. The solid lines are exponential fits to the experimental data obtained at different temperatures. (b) Pulse sequence, corresponding DSFD, and RTG measurements of the inhomogeneous vibronic relaxation between the X and B electronic states of 12. The time delay between pulses c and a(b) was scanned. The DSFD shows that signal depends on the off-diagonal terms peg. The logarithm of the amplitude-normalized signal FWM emission is plotted as a function of time delay 2. The solid lines are fits to a Gaussian function of the data obtained at different temperatures. The dashed line corresponds to the exponential decay measured for the PE data at 108 °C (in Figure 5.1a). and hence in the time domain. The data in Figure 5.1b, were fit with the formula S = Ae'w '2’ ,where A is an adjustable amplitude. '05 The experimental relaxation times, in the temperature range from 77 to 108 °C, go from T22, = 229 i 8 to 199 i 3 ps (see Figure 1b). Notice that there is only a 13% difference in the relaxation rate, although the number density increased from 3x1023 1113 to 1.7x1024 m'3. In Figure 5.1b, we have 121 included the exponential decay used to fit the PE data at 108° C. Notice that after the first 200 ps, inhomogeneous relaxation is faster than the homogeneous relaxation. The weak dependence on number density and the Gaussian decay pattern indicate that inhomogeneous contributions play a very important role in RTG measurements. Inhomogeneous contributions include Doppler broadening. Using formulas from Siegmartm6 Awnfllflffiflw and T,’=——‘4”’°2, where A019 is the FWHM C A200 linewidth of the Gaussian distribution, M is the molecular weight and a) is the central frequency of the transition, we calculate T; = 1.1 ns for T: 381 K. The value measured is of the same order of magnitude but is five times shorter. One reason for this discrepancy is that the RTG data contain both homogeneous and inhomogeneous relaxation. This explains why the measured rates exceed the expected 7% dependence for inhomogeneous decay. For PE and RTG measurements, the signal drops to zero for long time delays. In Figure 5.2a, we have included an RTG measurement obtained with a cell kept at 200 °C with a low number density of iodine molecules (see experimental section). The signal approaches zero at long times with an exponential decay time of 160 ps. For vibrational relaxation measurements, the transient grating (TG) and VE sequences were used. In both cases, the signals include the diagonal density matrix elements pee and/or pgg, which relax with T1 lifetimes. Under the experimental conditions, T 1 is very long and is manifested experimentally as a constant signal level at long times (see Figure 5.2b). The exponential decay has lifetime 29,”, = 130 ps. The time dependent component of the TG signal includes both p22 and pggr vibrational coherence. 122 8) Vibronic relaxation b) Vibrational relaxation b a __ C c be signal -- '2 ab 2' signal c " a peg peg b m I ’ g’2 b pll's pcc' Lb? :1," c _- c _- 2' lime r a a 0 Reverse Transient Grating ’ ' Transient Grating ? t 5.... [exp = 16012 ps > rexp = 130:1:8ps i ’ . n?" 3: a o a g 1 ~ T=200°C .1 - T=200°C a E‘ 8 m .5 1’: [L3 .5. LD 9‘ 1— . o " 1 l J a L 1 1 0 1' 1 I 2 2 1 1 L 0-0 0-1 0-2 0-3 0-4 0-5 0-6 0.0 0.1 0.2 0.3 0.4 0.5 0.6 Time Delay 2, [nsl Time Delay 2, Ins] Figure 5.2. (a) RTG measurements of the vibronic relaxation between the X and B electronic states of 12. The FWM emission is plotted as a function of time delay 2. The solid line is an exponential fit to the experimental data (dots). (b) The pulse sequence and corresponding DSFD for TG measurements of vibrational coherence in the X and B electronic states of 12. The DSF D shows that the signal depends on the elements in the diagonal blocks pgg, pggv, pee and pan. The FWM emission is shown as a function of time delay 2. The solid line is an exponential fit to the experimental data (dots). The dashed line indicates the long-lived component of this signal, which decays with T1. The data in (a) and (b) were obtained in a special cell at 200° C (see text), and include inhomogeneous and homogenous broadening. In order to separate the ground- and excited-states contributions to the signal, the VE pulse sequences was applied as described by Brown et al.94 (see Figure 5.3 top). When the delay between first two pulses 212 was 460 fs, we measured a coherence lifetime corresponding to the excited state of Texp = 152 ps. When the delay 2, 2 was 614 fs the observed lifetime, corresponding to the ground state, was rcxp = 165 ps (see Figure 5.3). The first tens of picoseconds of these two signals are shown together with their Fourier transforms in Figure 5.4. 123 C Vibrational relaxation " c ‘- a b ’ c signal pf: b ‘31:} lL—‘A , ’ -- b __ T12 T time a a 71209: 21m or 21t(n+%) 3 - o 1 Virtual Echo :2 I . 2,,=460fs.2,,p= 152:1:8 ps —: 2 '- 0 T12 =614fs .Taxp=16519 p5 3‘ ' is. '5 . g .. a _ 3 ’ ’. .5: 1 . l: i , o o'. J oio 0:1 0:2 0:3 0:4 0:5 0.6 Time Delay 2, Ins] Figure 5.3. (a) The sequence and corresponding DSFD for the VE pulse sequences. The DSFD show that the signal depends on the pee' and pgrg elements. (b) VE measurements of vibrational coherence in the B (2,2 =460 fs) and X (2,2 =614 fs) electronic states of 12. The intensity of FWM emission is shown as a function of time delay 2. The solid line is an exponential fit to the experimental data (dots). The data were obtained in a special cell at 200 °C (see text). The excited-state vibrational modulation (307 fs oscillations, 108 cm") is observed for 2,2 = 460 fs, while the ground-state vibrational modulation (160 fs oscillations, 208 cm") is observed for 2, 2 = 614 fs. These vibrational frequencies correspond to wave packet motion in the B state with v' z 7-10 and in the X ground electronic state with v” z 3-4, respectively. The initial wave packet dephasing, dashed line in Figures 5.4a and b, was obtained by eliminating the slow modulation caused by rotational dephasing and then fitting the vibrational modulation amplitude to an exponential. The data in Figure 5.4a show a vibrational revival at 10 ps that is very similar to that observed in pump-probe measurements with a similar preparation pulse.‘96 Because no coherence is lost during wave packet spreading, revivals are typically expected. The wave packet spreading time 124 depends on intramolecular parameters such as anharmonicity and external parameters of the preparation pulse such as pulse duration and chirp. The value measured for wave packet spreading in the B state, 4 ps, is consistent with other femtosecond pump-probe measurements. 2‘0?” The time associated with wave packet spreading of the X state, ~10 ps, is longer indicating a much smaller anharmonicity and a smaller number of vibrational levels involved?“5»217 The measurements on wave packet spreading presented here are in agreement with FWM experiments on gaseous iodine from other groups173218.219 Wave packet dephasing a) T12=460 IS F. b) 712:614fs 31.0 lug” Tarp: 4:1 ps 3°25 Wl llllflll [WI l“. ‘1' Tap: 10:2 ps .2 lllll 111111111111gull‘llllllmlllllHHIUllll’l' l 1,111 g lll'E ’ lull c2.0 ll ‘ Mlll’l’zl will.“ i :30le lllllll‘l’ 8111711,.” llllll g llllllli I'lllllll llllllllll l llll‘i'lli llllllilllilll Wiwillsllmfl E(iv'''55"1'1'ov'n1'5'h'2'0”"2'5 6----é----1,o----1,5--..io.---2,5 Time Delay 2, [ps] ,__, Time Delay 2', [pa] =3 a 30.3 vcxcited state 550.5 ’ V round state §°1l : 1 g 80.1 E . E k t ’G‘TAIA‘ I I I I t I r I I t I “0.0 100 110120130 140150 1601701'180190 200210 220 “0.0 100110120130140 150150 170100190 200 210 220 Wavenumber [cm ’1 Wavennrnber,[cm"] Figure 5.4. Measurements of vibrational wave-packet spreading: (a) excited state (only the first 12 ps), (b) ground state with the VE pulse sequences as in Figure 3. The decay in modulation amplitude caused by wave packet spreading was fit to an exponential function and is included here (dashed lines). The data in (a) show a wave packet revival at 10 ps. The corresponding power FFT is shown. 125 5.4. Conclusions This chapter described the use of different three-pulse F WM sequences that can be effectively used to measure different types of coherence relaxation processes in gas- phase molecules. Experimental results lead to the conclusion that the vibronic coherence relaxation times in iodine vapor are strongly influenced by inhomogeneous broadening. We obtained the homogeneous relaxation time and corresponding cross section using photon echo measurements. The size of the cross section indicates that decoherence is caused by long-range ‘soft’ collisions. In addition, we measured the inhomogeneous vibronic coherence relaxation and vibrational relaxation times following selective coherent excitation in the ground or in the excited state. Times for wave packet spreading in each state are also reported. Using different pulse sequences to select among the relaxation processes, these powerful methods can measure T1 times (typically 10'8 5); T2 times (typically 10'lo s) and T3 times (typically 10'12 5). Future applications will explore decoherence processes in polyatomic molecules, an aspect that is of great interest for efforts in coherent control of molecular dynamics. Of particular interest will be to explore long-range dephasing collisions and wave packet spreading rates in multidimensional systems. These studies have been followed by a series of systematic measurements as a function of density and in the presence of buffer gases.220 The results, obtained by Matt Comstock and Evgeny Sudachenko in our group, are in very good agreement with the dephasing cross section for pure iodine obtained by the initial measurements presented here. 126 6. FEMTOSECOND GROUND-STATE DYNAMICS OF GAS-PHASE DINITROGEN TETROXIDE AND NITROGEN DIOXIDE 6.1. Introduction Nitrogen oxides (NOX) have been the subjects of numerous studies primarily for their roles in both stratospheric and tropospheric chemistry and in particular, their links to ozone photochemical cycles and smog formation. These relatively small molecules are especially interesting due to their complex spectroscopy as evidenced by theoretical and experimental studies of their ground and excited statesm'232 All except NO have multidimensional potential energy surfaces and numerous curve crossings that result in vibronic chaos making these molecules very interesting to dynamicists. The nitrogen dioxide/dinitrogen tetroxide monomer/dimer pair has received particular attention. The dimer is favored at higher pressures and lower temperatures. At one atmosphere and room temperature the equilibrium concentration of NO; is about 16%.226s233 Dinitrogen tetroxide is particularly intriguing because of its unusually long N-N bond (1.776 A, gas phase)23‘lv235 and the fact that the ONO bond length and angle are identical to those of the monomer (see Figure 6.1). Because of its (D2h) symmetry, N204 does not have a permanent electric dipole moment. The dissociation energy of the N-N bond is 0.59 eV and its first electronic resonance is at 340 nm.221 It has been proposed226 that the use of off-resonance excitation should make it possible to directly probe the dimer at room temperature. Most of the vibrational frequencies for these molecules are 127 O 0 Figure 6.1. Sketch of the N204 molecule. The % N-O distance (a) is 1.19 A , the N-N bond (b) \ is 1.776 A, and the angle a is 134.6°, from (X /N __—— N \ reference 234. O b 0 established and assigned,222~2241225 but there is still some disagreement between experiment and theory on several modes. Time-resolved rotational dynamics of NO; were measured by Sarkisov et. al.236 using femtosecond Raman-induced polarization spectroscopy. The photodissociation dynamics of NO; were recently studied using ultrafast methods.237.~238 This chapter exploits the ability to detect low frequency Raman modes of short-lived species and to measure high accuracy rotational constants by femtosecond time-resolved FWM to study the ground-state dynamics of NzO4/NOZ. Non— resonant femtosecond time-resolved four-wave mixing data on gas-phase N02 and N204 are presented. The initial rotational dephasing is observed during the first picosecond after excitation. Fast vibrational dynamics (average value 133 i 1 fs beats) are observed for dinitrogen tetroxide and are assigned to the V3 N-N stretching mode in the ground state ('Ag). Rotational revivals on the picosecond time scale are recorded for both samples. No evidence of a long-lived excited-state participation was found using PE or VE pulse sequences with 800 nm laser pulses, however, NO; photoproducts were observed following N204 excitation. 128 6.2. Experimental The experiments were performed using a regeneratively amplified femtosecond titanium-sapphire laser system (Figures 2.8 and 2.9) that was described in detail in Chapter 2. In short, the output from a continuous wave NszVO4 laser (Millennia — Spectra Physics Laser) pumps the femtosecond titanium-sapphire oscillator (KM Labs), capable of producing <15 fs pulses at 90 MHz. This output was amplified by a regenerative amplifier (Spitfire — Spectra Physics Lasers) pumped by a 1 kHz NszLF laser (Evolution-X — Spectra Physics Lasers). Amplified pulses centered at 805 nm were close to transform limited and had a maximum energy of 0.8 mJ/pulse and time duration of 50 fs. Characterization of the pulses was performed with both non-collinear autocorrelation and frequency resolved optical gating. Both methods confirmed that the pulses were shorter than 50 fs measured at full-width half-maximum and had no residual chirp. The output was split into three beams using dielectric beam splitters and recombined in a forward-box geometry (see Figure 2.6). In all experiments the output beam was attenuated to ~60 u] per pulse measured before splitting. Two beams were overlapped in time and space to form a transient grating in the sample while the third one was time delayed with a computer controlled actuator. This configuration allows TG and RTG experiments described in detail in previous chapters. When the two initial pulses are not overlapped in time, other types of FWM experiments (PE and VE) are possible.104 The event when all three pulses overlap in time is referred to as time-zero. The three 10 mm diameter beams were focused into a 6 inch-long custom-designed quartz cell containing the gaseous sample by a two-inch diameter lens with focal length of 50 cm. Heating the entire cell with heating tape controlled the temperature of the cell. The F WM 129 signal beam was spatially filtered by a set of irises, collimated and sent into a spectrometer (Triax 320, Jobianon). Before reaching the entrance slit of the spectrometer, the signal was attenuated by a neutral density filter of OD=1-2. The signal was then dispersed by the spectrometer’s grating, and was registered by a 2000 x 800 pixel liquid nitrogen cooled CCD. The entrance slit of the spectrometer was 100 micrometers (0.8 nm resolution). The wavelength range on the CCD in this setup (grating dictated) was about 60 nm. In order to preserve disk space, every 20 pixels were binned together. For the shorter scans (~1-2 ps duration) the frequency-dispersed signal was recorded every 10 fs, while for the longer scans the time steps were 20 fs. Signal was registered for each time delay with an integration time of 100 ms. Each data set is the average of several such scans. The sample (99.1% N204) was obtained from Sigma-Aldrich and used without further purification. The sample cell was pumped out to 10'5 Torr on the vacuum line and filled directly from the lecture bottle. There was no experimental evidence of contamination. Experiments were performed at different sample pressures. All the data presented here were obtained with a cell that was loaded to an initial pressure of 400 Torr at room temperature . 6.3. Results and discussion In the gaseous state, the N204 and N02 molecules are in thermodynamic equilibrium (at room temperature K = 8.8 atm").239 The temperature and pressure of the sample cell from 294 K (80% N204) to 363 K (94% N02) were controlled.233 A change 130 in optical density of the sample at 805 nm from 0.045 (294 K) to 0.071 (363 K) was observed and was not considered large enough to affect the measurements qualitatively. Some experiments were carried out with a cell filled to 760 Torr. The overall signal level was somewhat smaller due to the increased OD leading to absorption of the excitation photons. The transients yielded qualitatively identical results, but required a higher temperature to obtain a good NO; signal (383 K), probably due to a change in the equilibrium conditions and some re-absorption of the signal. It is necessary to note that it is impossible to have 100 % pure gaseous N;O4 in the cell because of the aforementioned equilibrium. In Figure 6.2 frequency-dispersed transients are shown that were obtained as a function of time-delay between the two time-overlapped pulses and the third pulse for NO; (6.2a) and N;O4 (6.2c). In the two-dimensional contour plots (a and c) the spectrum of the signal in the region from 780 nm to 830 nm is registered versus delay time every 10 fs. The dark current level of the CCD was subtracted from these data. The intensity scale is logarithmic to highlight the weaker features. In the case of NO; (Fig. 6.2a) there is a broad, featureless peak 160 fs after time zero as well as another at 500 fs. Figure 6.2b shows the frequency—integrated (780-830 nm) transient signal. These features (Figure 6.2a and b) can be assigned to the initial rotational dephasing in the sample and are consistent with results for the off-resonance FWM signal. Measurements for N;O4 were collected in the same manner as described for NO; and are presented in Figure 6.2c and d. These results show, in addition to a broad rotational dephasing envelope, a series of beats with an average period of 133 fs. The beats are detectable up to 3 ps after time zero but with much lower intensity. l3l _ a) E E... (< b) 1045- 1 103 820:— 3 i . ' C) E 810— -t‘ - 2 i E“ 800* 1?: K g (< " - Q O 790_‘ : E. - ‘ : d) 104; ' 103 . . _ o 500 1000 1500 2000 Time Delay [fs] Figure 6.2. Experimental time-resolved FWM signal obtained during the first 2 ps: a) Frequency-dispersed signal obtained as a function of time delay for N02, b) Frequency — integrated NO; data as a function of time delay. c) Frequency-dispersed signal as a function of time delay for N;O4, d) Frequency—integrated N;O4 data as a function of time delay. For all scans, the signal intensity is plotted on a logarithmic scale. 132 Long time scans were performed at 363 K (NO;) or 294 K (N;O4). We present data in the time-delay range from 9 to 42 ps in Figure 6.3. The transients show the frequency-integrated intensity recorded in the 785-825 nm range as a function of time delay. Note that the vertical axis (intensity) is logarithmic. The NO; data presented in Figure 4a show rotational recurrences every 19.680 ps with half recurrences in between (9.840 ps). The transient in Figure 6.3b, taken at 294 K, exhibits similar features with the same periodicity. The recurrences observed at 294 K are about six times lower in intensity than those observed for NO; at 363 K, taking into account the attenuation filter used for the higher temperature transients. The transients for both temperatures were taken under identical laser conditions except for attenuation. This results in slightly different background and noise levels between the two transients. In this range of time delays both data sets exhibit rotational recurrences that can be assigned to NO;. Under close examination some differences are observed in the shape of the rotational recurrences obtained in the two transients. The differences are especially clear at the half recurrence observed in both transients near 29.5 ps (see Figure 6.3). Nitrogen dioxide is an asymmetric top molecule, but the similarity of the B and C constants allows us to view it as a symmetric top with Bave=(B+C)/2. From the observed rotational recurrences of 19.680 ps (9.840 ps half-recurrences) in the long time delay transient of NO; (Figure 6.3a) we can calculate the average rotational constant as Bave=1/4c rm, where c is the speed of light and Tree is the time between rotational recurrences. The obtained value is 0.42343 cm'l i 0.00007 cm". This value compares favorably to the reported experimental240 (0.42202 cm") and theoretical227 (0.423325 cm") values calculated for Bave, respectively. Simulations based on the semiclassical 133 forrnalism,9z1'32 and the average rotational constant for NO; using the literature values are in general agreement with the observed data (see vertical lines in Figure 6.3). The rotational recurrences obtained at longer time delays (r> 40 ps) exhibit a more complex modulation that cannot be simulated with a symmetric top approximation, even though their positions can be exactly reproduced. 104 r l ‘ A Y Y N 9680 fs Intensity [rel.] 103— 1 ._ Fl 1 ll 1 I n :L 1 Id 10000 20000 30000 40000 50000 Time Delay [fs] Figure 6.3. Experimental time-resolved and frequency-integrated FWM signal (a) NO; and (b) N;O4 taken at 363 K and 294 K, respectively. Note the subtle differences in the shape of the rotational recurrences obtained for both cases. The signal intensity is plotted on a logarithmic scale. The vertical lines indicate the position of the rotational recurrences obtained from a simulation using Bave=(B+C)/2. I34 In the case of N;O4, the signal obtained at long time delays is more complex. The small features at 9.840 and 19.680 ps (Figure 6.3b) could be assigned to rotational recurrences originating from the 16% N0; molecules present at room temperature. The positions of these features coincide precisely with the rotational recurrences of NO;, however, the signal intensity is about one order of magnitude stronger than expected based on the observed intensity at 363 K, given the square dependence on number density for these measurements. The signal in TO measurement is proportional to the square of the induced polarization, which depends linearly on susceptibility and therefore on number density.92 If the only source of rotational recurrences at 294 K (Figure 6.3b) were N0; molecules that were present in the sample before the first interaction with laser pulse took place, the signal should be an order of magnitude lower. Furthermore, there are some subtle differences in the shape of the recurrences; for example the half recurrence at 29.5 ps in Figure 6.3 lead us to consider assigning a significant contribution to these features from photoproducts that result from excitation by the first two pulses. The N-N bond dissociation energy is only 0.59 eV while the photon energy of each pulse is 1.54 eV. The principal axis of the NO; moieties is aligned with the N-N bond of N;O4; therefore, the initial rotational alignment would be preserved in the photodissociation process. The product NO; molecules formed by photodissociation would realign and become ‘visible’ to the third pulse at the same time delays as the thermal N0; molecules originally in the sample, but would have a different rotational temperature. There is a possibility that multiphoton excitation yields NO; molecules in the 2B; excited state, which has slightly different (and still not exactly determined) rotational constants. The initial rotational dephasing observed during the first picosecond is 135 vibrationally modulated. The times between first five beats are 158, 125, 131 and 135 fs. Fourier transformation of the signal presented in Figures 6.2 c and d results in a single broad feature centered at 234 cm"(See Figure 6.4). The experimentally obtained value for the N-N stretching mode from gas-phase Raman spectroscopy225 is 254 cm", and from high resolution FTIR241 and Raman spectra on solid242 N;O4 it is 281.3 cm". The only other frequency that could be implicated is the NO; rocking mode v10 = 265.5 cm'l; however, this mode is not Raman active and therefore not observed in our measurements. FFT l‘\ /— -"/' ' ‘-— 1‘ A A n A \T“'r~—l__l—_A ‘h— Irri— 1100‘ A ‘150 ‘ A 200 A A .250 I 300 350 400 wavenumber [cm1 ] Figure 6.4. FFT of the data in Figure 6.3b. The observed peak at 234 cm'I was assigned to the N-N stretching mode. Applying the Maximum Entropy Method (MEM)243 of analysis on the same set of data yielded two frequencies: 238.5 cm'1 and 251.8 cm". These two values were consistently obtained following analysis of different data sets (not shown). The observed wave packet motion involves at least three vibrational levels in the ground state. The difference between the two frequencies obtained by MEM gives the anharmonicity of this Vibrational mode. Presumably, the 251.8 cm'1 value corresponds to the lowest level(s), 136 consistent with the reported gas phase value, and the 238.5 cm'1 value to the higher levels. These findings are consistent with the extremely large anharmonicity of N;O4.244 Only one rotational recurrence feature was observed after the first picosecond that can be assigned to N;O4. At 76.265 ps (Figure 6.5) a small feature is present in the room temperature sample that is clearly absent at higher temperatures. This feature gains in intensity compared to the one at 78.540 ps (NO; recurrence) if the temperature of the sample is lowered to 00 C (data not shown). The lower temperature shifts the equilibrium further towards the dimer. If a semiclassical model and available rotational constants are used, it is possible to simulate this recurrence. The experimental data showed no evidence of a half recurrence at 38 ps. Scans taken for time delays of up to 160 ps showed no other feature that could be distinctly assigned to N;O4. One possible reason for the absence of N;O4 rotational recurrences at long time delays (> 100 ps) is the rate of collision; which we estimate to be r< 0.1 ns at 300 K. 137 10“ :— l\590401s. 78720 fs/ ‘5 '7 : M l a) I B _ m - C 9 l- d E 104 r 1 : l 76265 fs l - Intensity [rel.] 103 _ I 1 L I . I . 1' 60000 70000 80000 Time Delay[fs] Figure 6.5. Experimental data obtained at 363 K (a) and 294 K (b). The main difference between the two transients, after the first two picoseconds, occurs at 76.265 ps and is indicated by an arrow. Both intensity axes are plotted on a logarithmic scale. 6.4. Conclusion The signal observed for both molecules (N O; and N;O4) is primarily non-resonant and therefore depicts ground-state dynamics. Experiments were carried out at both temperatures; photon echo (PE) and virtual echo (VE) pulse sequences were performed but yielded no significant feature other than the time zero signal. If the experiments shown here did involve a resonant one-photon transition to the 2B; state of NO;, the PE and VE signals would have revealed it. The measurements suggest that only ground state NO; contributes to the observed 138 signal based on the following argument. The absorption cross section of NO; at 800 nm is very small.245»246 The reported and calculated values of the B and C rotational constants for the 2B; state are very dissonant, and fall in the range from 0.458 to 0.548 cm'1 and 0.38 to 0.449 cm", respectively. If these transitions were involved, additional rotational recurrences corresponding to these values would have been observed in the time-resolved F WM transients. Close examination of the data did not reveal any other features except for the aforementioned recurrences every 19.680 ps. Given that the homodyne detected FWM signal depends on the transition dipole moment to the eighth power, this process is not likely to be detected by our setup. The average rotational constant measured from our experiments of 0.42343 cm'l reflects the average (B+C)/2 rotational constant of NO; in the ground state. The experiments at room temperature (N ;O4) allowed us to measure the frequency of the N-N stretch of 234 cm'1 and perhaps the anharmonicity of this unusual bond. The next electronic state known is the 3B3u which is reached with photon energies >3 eV and would require two photon absorption to be detected. There is no evidence in our data of single or multiphoton transitions to a long-lived excited state. However, excitation to a dissociative state would be difficult to rule out. Perhaps the negative time delay shoulder found in the N;O4 transient in Figure 6.2d but not found in the NO; transient in Figure 6.2b is evidence of a dissociative pathway. 139 7. SUMMARY AND CONCLUSIONS This work has contributed to the evaluation of different methods of using femtosecond lasers for studying laser control of chemical reactions. The number of ultrafast pulses, the timing between them (sequence), their phase, and chirp are shown to be effective control parameters for the molecular system. All of these parameters play a role in the search for optimal laser fields for control of molecular motion. The experimental demonstration of laser control presented in the previous pages includes the enhancement of the yield of the photodissociation reaction by manipulation of a phase characteristic (chirp) of the applied pulses. Specifically, the formation of I; that followed multiphoton excitation of CH;I; at 312 nm exhibited an asymmetric dependence on the sign of chirp. Enhancements up to a factor of 25 were observed for positive chirps compared to transform-limited pulses and were found to be dependent on pulse intensities. Over last two years this method that involves only one control ‘knob’ has lost popularity despite its success,‘3~‘54844512471248 and has surrendered its place to control schemes using more sophisticated optimal control},'93032’7‘31‘77,249 where the number of control parameters can be virtually unlimited. Very recently, it has been demonstrated that a limited number of parameters is indeed more advantageous.23~250 The work presented in this thesis also demonstrated that femtosecond F WM techniques could be applied in coherent control by combining a discrete number of laser pulses. This also allows us to learn about the role of one or more control parameters in I40 controlling a process. Theoretical as well as some experimental studies in coherent, quantum, and optimal control have found that the optimal fields in many cases correspond to two or three pulses with a particular phase relationship and chirp. Three- pulse FWM methods are shown here to be very useful in the characterization of intramolecular dynamics. Presented measurements show the dynamics of a gas-phase model system (1;) prepared by the first two laser pulses and its control. For the VE sequence, when the pulses are transform limited and the delay between them is 460 fs, only excited-state dynamics are observed, while time delay of 614 fs yields ground-state dynamics. With the VE setup controlling the time between the first two pulses allows to control the observation of ground or excited state dynamics. This type of control is not available with the PE setup. Presumably, under large inhomogeneous broadening, the vibrational coherence is lost leaving only the signal that is independent of the coherence. However, when the laser pulses are chirped, the observed dynamics are quite different. One of the clearest applications of FWM methods to coherent control is in measuring coherence relaxation. It is demonstrated here how the methods work for gas- phase samples and a comparison has been made between photon echo and virtual echo type measurements. The main difference is that for PE the inhomogeneous broadening cancels, giving a more accurate homogeneous coherence relaxation time. The data presented can be used to extrapolate a homogeneous coherence time for iodine in the nanosecond scale, giving the possibility of performing a very large number of coherent interactions with this gas-phase sample at room temperature. Differences between the various phenomena (PE, VE, TG, RTG) have been demonstrated. The results reveal the control schemes of the molecular system within the framework of phase-matching l4l detection. This implies that only the subset of molecules forming a transient grating in the sample is controlled and yields the coherent signal observed. It is also demonstrated how FWM methods can be used to measure coherence life times and to characterize the intramolecular dynamics of the system that is being controlled. Future applications of this form of control could lead to a more general molecular control of quantum mechanical states in the gas phase and perhaps even to realization of molecular computing.24 The non-resonant time resolved FWM technique has been applied to dinitrogen tetroxide and nitrogen oxide in order to study their dynamics. This allowed the precise determination of the average rotational constant and vibrational frequency of the N-N stretch mode. The results suggest that the planarity of N;O4 could be explored with this technique. More detailed exploration of this interesting molecular system requires tunable femtosecond pulses that are resonant with electronic transitions of the molecule. Further explorations would be feasible because such pulses are widely available at the present. 142 8. APPENDIX I Below is a brief theoretical description of C-FID-FWM based on the solution of the Liouville equation. Microscopic expressions for cascading are given in chapter 16 of reference 87. Evolution of the density-matrix elements p1,- in the absence of an external field can be described with the free propagator, under interaction with the vibronic matrix elements V1,. v5“ = Alf'e'ka", where All“ = ‘fislf‘m (8.1) where the label aindicates the interacting pulse a, b or c; i and j are indices for vibrational levels in the excited and ground state respectively; 0),]- and 11,-,- are the transition frequencies and dipole moment on the ijth transition; Sl-a] is the spectral intensity of the corresponding transition induced by the transform-limited pulse a with carrier frequency a)" and wave vector ka; Asa] are the macroscopic space-independent (x) 0] components of V2 . The 1n1t1al condition of the densrty matrix p19) = 11,- 1S glven by a Boltzmann distribution. After pulse c interacts with the sample at time t], the induced polarization causes FID with the wave vector knp = kc, 143 EFID(I) 0C :14 21- “ij[vi[jc]e-i(oij(t-ti)_Vljcl*ei0)ij(t-ti)]nj (82) . . . . no , This field interacts With medium, Vlj / oc ,uleF/D, after pulses a and b at times t; and I 3 causing the third order polarization (3) -' .. - (2) pij (Um-e 1(1),)“ t3)Zj" Pj"j(t3)14ij"zry “no, * f dt'eimij”(t"t3)[e-i[wi'j'(t'-tl)-kFle]-ei[wi'j'(t'-tI)-kFle]]A!f!nj' _' .. - (2) +e “U“ “)2," pii" (101/0021111 Pi’j' [c * floodt'eimi"j(t'-t3)[e'i[mi'j'(t"tl)‘kFIDXI-eilmi'j'(t"tl)‘kFlenA'v'Inj' (83) Only resonance terms can stimulate optical transitions, therefore, terms with wy'zwij in the first sum and a),-~j=a),~j1 in the second sum are kept. The second order . . 2 . . . . denSity matrix elements, pg) in Equation (3), can be calculated With the following formula applicable for well-separated pulses plin)(t) ____ Ie-iwija-t") 1n] (n-l) (n-l) [n] [n] (n-l) ("l *{Zy [Vii'pi'j (tn)'pii' (thi'J n 1+2, [Vij' pj'j (id-Pi)" (tn)Vl3]]}- (8.4) To calculate the C-FID-FWM signal, only non-zero density matrix elements of p33) proportional to exp[-i(ant-(ka-kb+kc)x)] are included, because only these components emit light in the phase-matching direction. The intensity of the time-integrated signal is 144 given by 2 (3) l = Zij 111, where Iij oc uijfidtlpij (t)|2 (5) An expression to calculate the spectral components Ig-FlDlof the C-FID-FWM signal observed for negative rd, = tc-tb can be obtained as follows. The Equation (8.4) is used to calculate psi), and introduce these elements into Equation (8.3). Finally, each element is calculated with Equation (8.5) using rba = tb-ta, to obtain [FID] 2| 10,... 2 [c1 -10” [bl* [al Iij °Cllij 211C " ‘bllij'Aij'nj'Zive ”'"Ai'j' Ai'jn'+ - 2 [c] -° [bl‘ [8| 2 X11 e'm"T°bH1'in'j anj'elm”Tb'A 'Aij'“i’ - (6) i’j The time delay rba is the control parameter that allows us to select between ground and excited state dynamics in both cases. Similarly to VE sequence, the signal is obtained as a function of the time delay 1d,. 145 9. REFERENCES 1. van't Hoff, J. H. Etudes de Dynamiques Chimiques 1884, 114 2. Arrhenius, S. Z. Phys. Chem. 1889, 4, 226 3. Rice, 8.; Zhao, M. Optical Control of Molecular Dynamics; Wiley: New York, 2000. 4. Rabitz, H.; de Vivie-Riedle, R.; Motzkus, M.; Kompa, K. Science 2000, 288, 824 5. Rabitz, H.; Zhu, W. Acc. Chem. Res. 2000, 33, 572 6. Rice, S. R. Nature 2001, 409, 422 7. Henriksen, N. E. Chem. Soc. Rev. 2002, 31, 37 8. Vanderwal, R. L.; Scott, J. L.; Crim, F. F. J. Chem. Phys. 1990, 92, 803 9. Seideman, T. J. Chem. Phys. 1999, 111, 4397 10. Seideman, T. J. Chem. Phys. 1995, 103, 7887 11. Krause, J. L.; Whitnell, R. M.; Wilson, K. R.; Yan, Y.; Mukamel, S. J. Chem. Phys. 1993, 99, 6562 12. Krause, J. L.; Whitnell, R. M.; Wilson, K. R.; Yan, Y. J. Light packet control 146 of wave packet dynamics. In Femtosecond Chemistry; Manz, J., Weste, L., Eds.; VCH Verlagsgesellschaft: Weinheim, 1995. 13. Kohler, B.; Yakovlev, V. V.; Che, J .; Krause, J. L.; Messina, M.; Wilson, K. R.; Schwentener, N.; Whitnell, R. M.; Yan, Y. J. Phys. Rev. Lett. 1995, 74, 3360 14. Kohler, B.; Krause, J. L.; Raksi, F.; Wilson, K. R.; Yakovlev, V. V.; Whitnell, R. M.; Yan, Y. J. Acc. Chem. Res. 1995, 28, 133 15. Bardeen, C. J.; Che, J.; Wilson, K. R.; Yakovlev, V.; Apkarian, V. A.; Martens, C. C.; Zadoyan, R.; Kohler, B.; Messin, M. J. Chem. Phys. 1997, 106, 8486 16. Bardeen, C. J .; Che, J.; Wilson, K. R.; Yakovlev, V. V.; Cong, P.; Kohler, B.; Krause, J. L.; Messina, M. J. Phys. Chem. A 1997, 101, 3815 17. Bardeen, C. J .; Yakovlev, V.; Squier, J. A.; Wilson, K. R. J. Am. Chem. Soc. 1998, 120, 13023 18. Yakovlev, V.; Bardeen, C. J .; Che, J.; Cao, J.; Wilson, K. R. J. Chem. Phys. 1998, 108, 2309 19. Assion, A.; Baumert, T.; Bergt, M.; Brixner, T.; Kiefer, B.; Seyfried, V.; Strehle, M.; Gerber, G. Science 1998, 282, 919 20. Bergt, M.; Brixner, T.; Kiefer, B.; Strehle, M.; Gerber, G. J. Phys. Chem. A 1999,103, 10381 21. Weinacht, T. C.; Bartels, R.; Backus, S.; Bucksbaum, P. H.; Pearson, B.; Geremia, J. M.; Rabitz, H.; Kapteyn, H. C.; Murnane, M. M. Chem. Phys. Lett. 2001, 147 344, 333 22. 23. Levis, R. J .; Menkir, G. M.; Rabitz, H. Science 2001, 292 Herek, J. L.; Wohlleben, W.; Cogdell, R. J .; Zeidler, D.; Motzkus, M. Nature 2002, 417, 533 24. 25. 26. 27. 28. 29. 30. Lozovoy, V. V.; Dantus, M. Chem. Phys. Lett. 2002, 351, 213 Eyring, H.; Polanyi, M. Z. Phys. Chem. 1931, 12, 279 Polanyi, M.; Wigner, E. Z. Phys. Chem. Abt’.A 1928, 139, 439 Eyring, H. J. Phys. Chem. 1935, 3 Dantus, M.; Rosker, M. J .; Zewail, A. H. J. Chem. Phys. 1987, 87, 2395 Dantus, M.; Rosker, M. J .; Zewail, A. H. J. Chem. Phys. 1988, 89, 6128 Dantus, M.; Bowman, R. M.; Baskin, J. S.; Zewail, A. H. Chem. Phys. Lett. 1989, 159, 406 31. Dantus, M.; Bowman, R. M.; Gruebele, M.; Zewail, A. H. J. Chem. Phys. 1989, 91, 7437 32. Bowman, R. M.; Dantus, M.; Zewail, A. H. Chem. Phys. Lett. 1989, 161, 297 33. Zewail, A. H. J. Phys. Chem. 1996, 100, 12701 34. Zewail, A. H. J. Phys. Chem. A 2000, 104, 5660 148 35. Femtochemistry - Ultrafast Dynamics of the Chemical Bond; Zewail, A. H., Ed.; World Scientific: Singapore, 1994; Vol. I and II. 36. F emtosecond Chemistry; Manz, J .; Woste, L., Eds; Verlag Chemie GmbH: Weinheim, 1995; Vol. I and II. 37. Femtochemistry: Ultrafast Chemical and Physical Processes in Molecular Systems; Chergui, M., Ed.; World Scientific: Singapore, 1996. 38. Femtochemistry and Femtobiology; Sundstrtim, V., Ed.; World Scientific: Singapore, 1998. 39. Baskin, J. S.; Zewail, A. H. J. Chem. Edu. 2001, 78, 738 40. Dantus, M. Annu. Rev. Phys. Chem. 2001, 52, 639 41. Baltuska, A.; Wei, Z. Y.; Pshenichnikov, M. S.; Wiersma, D. A. Opt. Lett. 1997, 22, 102 42. Sutter, D. H.; Steinmeyer, G.; Gallmann, L.; Matuschek, N.; Morier-Genoud, F .; Keller, U.; Scheuer, V.; Angelow, G.; Tschudi, T. Opt. Lett. 1999, 24, 631 43. Gordon, R. J .; Rice, S. A. Annu. Rev. Phys. Chem. 1997, 48, 601 44. Stone, J .; Goodman, M. F; Dows, D. A. Chem. Phys. Lett. 1976, 44, 411 45. Mukamel, S.; Jortner, J. Chem. Phys. Lett. 1976, 40, 150 46. Letokhov, V. S.; Makarov, A. A. Appl. Phys. 1978, 16, 47 149 47. Quack, M. J. Chem. Phys. 1978, 69, 1282 48. Quack, M. The influence of intensity and laser energy fluence in unimolecular reactions induced by monochromatic infrared readiation (URIMIR). In Laser Induced Processes in Molecules; Kompa, K., Smith, S. D., Eds.; Springer Verlag: Berlin, 1979; Vol. 6. 49. Robinson, P. J.; Holbrook, K. A. Unimolecular Reactions; Wiley: New York, 1972. 50. Zewail, A. H. Phys. Today 1980, 33, 27 51. Manz, J .; Parmenter, C. S. Chem. Phys. 1989, 139, U1 52. Letokhov, V. S. Science 1973, 180, 451 53. Imre, D. G.; Zhang, J. Chem. Phys. 1989, 139, 89 54. Bar, 1.; Cohen, Y.; David, D.; Rosenwaks, S.; Valentini, J. J. J. Chem. Phys. 1990, 93, 2146 55. Shapiro, M.; Brumer, P. Chem. Phys. Lett. 1993, 208, 193 56. Shapiro, M.; Brumer, P. Int. Rev. Phys. Chem. 1994, 13, 187 57. Chen, C.; Yin, Y. Y.; Elliott, D. S. Phys. Rev. Lett. 1990, 64, 507 58. Zhu, L. C.; Kleiman, V.; Li, X. N.; Lu, S. P.; Trentelman, K.; Gordon, R. J. Science 1995, 270, 77 150 59. 116 60. Baranova, N. B.; Chudinov, A. N.; Zeldovich, B. Y. Opt. Commun. 1990, 79, Muller, H. G.; Bucksbaum, P. H.; Schumacher, D. W.; Zavriyev, A. J. Phys. B-At. Mol. Opt. Phys. 1990, 23, 2761 61. 62. 63. 64. 65. 66. 67. 68. Tannor, D. J.; Rice, S. A. J. Chem. Phys. 1985, 83, 5013 Tannor, D. J .; Kosloff, R.; Rice, S. A. J. Chem. Phys. 1986, 85, 5805 Felker, P. M.; Zewail, A. H. Adv. Chem. Phys. 1988, 70, 265 Fork, R. L.; Greene, B. 1.; Shank, C. V. Appl. Phys. Lett. 1981, 38, 671 Fork, R. L.; Shank, C. V.; Yen, R. T. Appl. Phys. Lett. 1982, 41, 223 Valdmanis, J. A.; Fork, R. L.; Gordon, J. P. Opt. Lett. 1985, 10, 131 Shank, C. V. Science 1986, 233, 1276 Chemical Reactions and Their Control on the Femtosecond Timescale; Gaspard, P.; Burghardt, I., Eds.; Wiley: New York, 1997; Vol. 101. 69. Manz, J. . 1n Femtochemistry and Femtobiology; Sundstrom, V., Ed.; World Scientific: Singapore, 1998. 70. Lozovoy, V. V.; Antipin, S. A.; Gostev, F. E.; Titov, A. A.; Tovbin, D. G.; Sarkisov, O. M.; Vetchinkin, A. S.; Umanskii, S. Y. Chem. Phys. Lett. 1998, 284, 221 71. Ubema, R.; Amitay, Z.; Loomis, R. A.; Leone, S. R. Faraday Discuss. 1999, 151 385 72. Pastirk, 1.; Brown, E. J .; Zhang, Q.; Dantus, M. J. Chem. Phys. 1998, 108, 4375 73. Zadoyan, R.; Schwenter, N.; Apkarian, V. A. Chem. Phys. 1998, 233, 353 74. Buist, A. H.; Muller, M.; Ghauharali, R. 1.; Brakenhoff, G. J .; Squier, J. A.; Bardeen, C. J .; Yakovlev, V. V.; Wilson, K. R. Opt. Lett. 1999, 24, 244 75. Neuhauser, D.; Rabitz, H. Acc. Chem. Res. 1993, 26, 496 76. Rice, S. A.; Manz, J.; Kohler, B.; Mukamel, S.; Brumer, P. W.; Quack, M.; Shapiro, M.; Marcus, R. A.; Schaffner, K.; Rabitz, H.; Flaming, G. R.; Tannor, D. J.; Gerber, G. Perspectives on the control of quantum many-body dynamics: Application to chemical reactions. In Chemical Reactions and Their Control On the F emtosecond Time Scale XXth Solvay Conference On Chemistry, 1997; Vol. 101; 213. 77. Kohler, B.; Krause, J. L.; Raksi, F.; Rosepetruck, C.; Whitnell, R. M.; Wilson, K. R.; Yakovlev, V. V.; Yan, Y. J .; Mukamel, S. J. Phys. Chem. 1993, 97, 12602 78. Yan, Y. J .; Gillilan, R. E.; Whitnell, R. M.; Wilson, K. R.; Mukamel, S. J. Phys. Chem. 1993, 97, 2320 79. Yan, Y. J. J. Chem. Phys. 1994, 100, 1094 80. Weiner, A.; Leaird, D.; Wiederrecht, G.; Nelson, K. Science 1990, 247, 1317 81. Shah, S. P.; Rice, S. A. Faraday Discuss. 1999, 319 152 82. Fujimura, Y.; Gonzalez, L.; Hoki, K.; Manz, J.; Ohtsuki, Y. Chem. Phys. Lett. 1999, 306, 1 83. Scherer, N. F.; Ruggiero, A. J.; Du, M.; Fleming, G. R. J. Chem. Phys. 1990, 93, 856 84. Bergmann, K.; Shore, B. W. . In Molecular dynamics and spectroscopy by stimulated emission pumping; Dai, H.-L., Field, R. W., Eds.; World Scientific: Singapore, 1995; Vol. 4. 85. Shen, Y. R. The Principles of Nonlinear Optics; Wiley: New York, 1984. 86. Boyd, R. W. Nonlinear Optics; Academic Press: San Diego, 1992. 87. Mukamel, S. Principles of Nonlinear Optical Spectroscopy; Oxford University Press: New York, 1995. 88. Schmitt, M.; Knopp, G.; Matemy, A.; Kiefer, W. Chem. Phys. Lett. 1997, 270, 9 89. Schmitt, M.; Kn0pp, (3.; Matemy, A.; Kiefer, W. J. Phys. Chem. A 1998, 102, 4059 90. Pshenichnikov, M. S.; de Boeij, W. P.; Wiersma, D. A. Phys. Rev. Lett. 1996, 76, 4701 91. Brown, E. J. B. Femtosecond gas-phase molecular dynamics: Pump-probe and four-wave mixing experiments. Ph.D. Thesis, Michigan State University, 2001. 153 92. Brown, E. J .; Zhang, Q.; Dantus, M. J. Chem. Phys. 1999, 110, 5772 93. Pastirk, I.; Brown, E. J.; Grimberg, B. 1.; Lozovoy, V. V.; Dantus, M. Faraday Discuss. 1999, 113, 401 94. Brown, E. J.; Pastirk, I.; Grimberg, B. 1.; Lozovoy, V. V.; Dantus, M. J. Chem. Phys. 1999, 111, 3779 95. Pastirk, I.; Lozovoy, V. V.; Grimberg, B. 1.; Brown, E. J.; Dantus, M. J. Phys. Chem. A 1999, 103, 10226 96. Lozovoy, V. V.; Grimberg, B. 1.; Brown, E. J.; Pastirk, I.; Dantus, M. J. Raman Spectrosc. 2000, 31, 41 97. Lozovoy, V.; Brown, E.; Pastirk, 1.; Grimberg, B. 1.; Dantus, M. What role can four-wave mixing techniques play in coherent control? In Quantum control of molecular reaction dynamics; Gordon, R. J., Fujimura, Y., Eds.; World Scientific: Singapore, 2000; Vol. 14; 62. 98. Lozovoy, V. V.; Pastirk, 1.; Brown, E. J .; Grimberg, B. I.; Dantus, M. Int. Rev. Phys. Chem. 2000, 19, 531 99. Lozovoy, V. V.; Pastirk, 1.; Comstock, M.; Dantus, M. Chem. Phys. 2001, 266, 205 100. Pastirk, 1.; Lozovoy, V. V.; Dantus, M. Chem. Phys. Lett. 2001, 333, 76 101. Pastirk, I.; Comstock, M.; Dantus, M. Chem. Phys. Lett. 2001, 349, 71 154 102. Lozovoy, V. V.; Grimberg, B. 1.; Pastirk, 1.; Dantus, M. Chem. Phys. Special Issue on Laser Control of Quantum Dynamics 2001, 267, 99 103. Pastirk, 1.; Comstock, M.; Walowicz, K. A.; Lozovoy, V. V.; Dantus, M. “2D (time-frequency) femtosecond four-wave mixing at 1014 W/cm2: Molecular and electronic response”; Optical Pulse and Beam Propargation III, 2001, San Jose, CA. 104. Grimberg, 1.; Lozovoy, V. V.; Dantus, M.; Mukamel, S. J. Phys. Chem. A 2002, 106, 697 105. Fork, R. L.; Shank, C. V.; Yen, R.; Hirliman, C. A. IEEE J. Quantum Electron. 1983, QE-l9, 500 106. Siegman, A. E. Lasers; University Science Books: , 1986. 107. F emtosecond Laser Pulses: Principles and Experiments; Rulliere, C., Ed.; Springer-Verlag: Berlin, 1998. 108. Link, 8.; Durr, H. A.; Eberhardt, W. J. Phys. Condens. Matter 2001, 13, 7873 109. Cao, J. 8.; Che, J. W.; Wilson, K. R. J. Phys. Chem. A 1998, 102, 4284 1 10. Malinovsky, V. S.; Krause, J. L. Phys. Rev. A 2001, 6304, 3415 111. Dmitriev, V. G.; Gurzadyan, D. N.; Nikogosyna, D. N. Handbook of Non- Linear Optical Crystals; Springer-Verlag: Berlin, 1991. 112. de Silvestri, S.; Laporta, P.; Svetlo, O. J. Quantum Electron. 1984, QE20, 155 533 113. 114. 2463 115. 116. 117. Trebino, R.; Kane, D. J. J. Opt. Soc. Am. A 1993, A10, 1101 DeLong, K. W.; Trebino, R.; White, W. E. J. Opt. Soc. Am. B 1995, 12, Kane, D. J .; Trebino, R. Optics Lett. 1993, 18, 823 Kane, D. J .; Trebino, R. IEEE J. Quantum Electron 1993, 29, 571 Trebino, R.; DeLong, K. W.; Fittinghoff, D. N.; Sweetser, J. N.; Krumbugel, M. A.; Richman, B. A.; Kane, D. J. Rev. Sci. Instrum. 1997, 68, 3277 118. Dantus, M. Femtosecond Transition-State Spectroscopy of Chemical Reactions. Ph.D. Thesis, California Institute of Technology, 1991. 119. Marvet, U. Femtosecond Reaction Dynamics in the Gas Phase. Ph.D. Thesis, Michigan State University, 1998. 120. 18,101 121. 122. 123. 124. Migus, A.; Shank, C. V.; Ippen, E. P.; Fork, R. L. IEEE J. Quant. El. 1982, Fork, R. L.; Martinez, 0. E.; Gordon, J. P. Optics Letters 1984, 9, 150 Valdmanis, J. A.; Fork, R. L. IEEE J. Quantum Electron. 1986, 22, 112 Salin, F.; Brun, A. J. Appl. Phys. 1987, 61, 4736 Spence, D. E.; Kean, P. N.; Sibbett, W. Opt. Lett. 1991, 16, 42 156 125. Huang, C. P.; Kapteyn, H. C.; McIntosh, J. W.; Murnane, M. M. Opt. Lett. 1992, 17, 139 126. Spielman, C.; Krausz, F.; Brabec, T.; Wintner, E.; Schmidt, A. J. Opt. Lett. 1991,16,1180 127. Spielmann, C.; Curley, P. F.; Brabec, T.; Krausz, F. IEEE J. Quant Electron. 1994, 30, 1100 128. Strickland, D.; Mourou, G. Opt. Commun 1985, 55, 447 129. Brumer, P.; Shapiro, M. Acc. Chem. Res. 1989, 22, 407 130. Shapiro, M.; Brumer, P. J. Chem. Phys. 1993, 98, 201 131. Zare, R. N. Science 1998, 279, 1875 132. Dantus, M. Annu. Rev. Phys. Chem. 2001, 52, 639 133. Shapiro, M.; Brumer, P. J. Chem. Phys. 1986, 84, 540 134. Brumer, P.; Shapiro, M. Annu. Rev. Phys. Chem. 1992, 43, 257 135. Zhu, L. C.; Suto, K.; Fiss, J. A.; Wada, R.; Seideman, T.; Gordon, R. J. Phys. Rev. Lett. 1997, 79, 4108 136. Tannor, D. J .; Rice, S. A. Adv. Chem. Phys. 1988, 70, 441 137. Bowman, R. M.; Dantus, M.; Zewail, A. H. Chem. Phys. Lett. 1990, 174, 546 157 138. Baumert, T.; Gerber, G. Isr. J. Chem. 1994, 34, 103 139. Ruhman, S.; Kosloff, R. J. Opt. Soc. Am. B 1990, 7, 1748 140. Shi, S.; Woody, A.; Rabitz, H. J. Chem. Phys. 1988, 88, 6870 141. Broers, B.; van Linden van den Heuvell, H. B.; Noordam, L. D. Phys. Rev. Lett. 1992, 69, 2062 142. Bardeen, C. J .; Wang, Q.; Shank, C. V. Phys. Rev. Lett. 1995, 75, 3410 143. Assion, A.; Baumert, T.; Helbing, J .; Seyfried, V.; Gerber, G. Chem. Phys. Lett. 1996, 259, 488 144. Papanikolas, J. M.; Williams, R. M.; Leone, S. R. J. Chem. Phys. 1997, 107,4172 145. Sarkisov, O. M.; Petrukhin, A. N.; Gostev, F. E.; Titov, A. A. Quantum Electron. 2001, 31, 483 146. Stauffer, H. A.; Ballard, J. B.; Amitay, Z.; S.R.Leone. JChem. Phys. 2002, 116, 946 147. Malinovsky, V. S.; Krause, J. L. Eur. Phys. J. D 2001, 14, 147 148. Okabe, H.; Kawasaki, M.; Tanaka, Y. J. Chem. Phys. 1980, 73, 6162 149. Dyne, P. J.; Style, D. W. G. J. Chem. Soc. 1952, 2122 150. Style, D. W. G.; Ward, J. C. J. Chem. Soc. 1952, 2125 158 151. Fotakis, C.; Martin, M.; Donovan, R. J. J. Chem. Soc. Faraday Trans. 1982, 78, 1363 152. Marvet, U.; Dantus, M. Femtosecond photoassociation spectroscopy: coherent bond formation. In F emtochemistry: Ultrafast Chemical and Physical Processes in Molecular Systems; Chergui, M., Ed.; World Scientific: Singapore, 1996; 138. 153. Zhang, Q.; Marvet, U.; Dantus, M. Farad. Discuss. 1997, 108, 63 154. Marvet, U.; Zhang, Q.; Dantus, M. J. Phys. Chem. 1998, 102, 4111 155. Kumit, N. A.; Abella, I. D.; Hartmann, S. R. Phys. Rev. Lett. 1964, 13, 567 156. Patel, C. K. N.; Slusher, R. E. Phys. Rev. Lett. 1968, 20, 1087 157. Lukasik, J .; Ducuing, J. Phys. Rev. Lett. 1972, 28, 1155 158. Heritage, J. P.; Gustafson, T. K.; Lin, C. H. Phys. Rev. Lett. 1975, 34, 1299 159. Rose, T. S.; Wilson, W. L.; Wackerle, G.; Fayer, M. D. J. Chem. Phys. 1987, 86, 5370 160. Hesselink, W. H.; Wiersma, D. A. Phys. Rev. Lett. 1979, 43, 1991 161. Myers, A. B.; Hochstrasser, R. M. IEEE J. Quantum Electron. 1986, QE- 22, 1482 162. Duppen, K.; Wiersma, D. A. J. Opt. Soc. Am. B 1986, 3, 614 163. Passino, S. A.; Nagasawa, Y.; Fleming, G. R. J. Chem. Phys. 1997, 107, 159 6094 164. de Boeij, W. P.; Pshenichnikov, M. S.; Wiersma, D. A. Annu. Rev. Phys. Chem. 1998, 49, 99 165. 166. 167. 168. 169. 280, 339 170. Yang, M. N.; Fleming, G. R. J. Chem. Phys. 1999, 110, 2983 Fayer, M. D. Annu. Rev. Phys. Chem. 1982, 33, 63 Rose, T. S.; Fayer, M. D. Chem. Phys. Lett. 1985, 117, 12 Motzkus, M.; Pedersen, S.; Zewail, A. H. J. Phys. Chem. 1996, 100, 5620 Schmitt, M.; Kn0pp, G.; Matemy, A.; Kiefer, W. Chem. Phys. Lett. 1997, Meyer, 8.; Schmitt, M.; Matemy, A.; Kiefer, W.; Engel, V. Chem. Phys. Lett. 1997, 281, 332 171. Meyer, 5.; Schmitt, M.; Matemy, A.; Kiefer, W.; Engel, V. Chem. Phys. Lett. 1999, 301, 248 172. Chen, T.; Engel, V.; Heid, M.; Kiefer, W.; Kn0pp, G.; Matemy, A.; Meyer, 8.; Pausch, R.; Schmitt, M.; Schwoerer, H.; Siebert, T. J. Mol. Struct. 1999, 481, 33 173. Kn0pp, G.; Pinkas, 1.; Prior, Y. J. Raman Spectrosc. 2000, 31, 51 174. Judson, R. S.; Rabitz, H. Phys. Rev. Lett. 1992, 68, 1500 175. Amstrup, B.; Toth, G. J .; Szabo, G.; Rabitz, H.; Lorincz, A. J. Phys. Chem. 160 1995, 99, 5206 176. Toth, G. J.; Lorincz, A.; Rabitz, H. Journal of Modern Optics 1997, 44, 2049 177. Brixner, T.; Damrauer, N. H.; Niklaus, P.; Gerber, G. Nature 2001, 414, 57 178. Tannor, D. J.; Rice, S. A.; Weber, P. M. .1. Chem. Phys. 1985, 83, 6158 179. Vaccaro, P. H. Degenerate Four-Wave Mixing (DFWM) Spectroscopy. In Nonlinear Spectroscopy for Molecular Structure Determination; Hirota, E., Field, R. W., Maier, J. P., Tsuchiya, S., Eds.; Blackwell Scientific Publications Ltd.: London, 1997. 180. Lee, D.; Albrecht, A. C. A Unified View of Raman, Resonance Raman, and Fluorescence Spectroscopy (and their Analogues in Two-Photon Absorption). In Advances in Infrared and Raman Spectroscopy; Clark, R. J. H., Hester, R. E., Eds.; Wiley-Heyden: Chichester, 1985; Vol. 12. 181. Hahn, E. L. Phys. Rev. 1950, 80, 580 182. Kinrot, 0.; Prior, Y. Phys. Rev. A 1994, 50, 1999 183. Bardeen, C. J.; Shank, C. V. Chem. Phys. Lett. 1993, 203, 535 184. Schoenlein, R. W.; Mittleman, D. M.; Shiang, J. J.; Alivisatos, A. P.; Shank, C. V. Phys. Rev. Lett. 1993, 70, 1014 185. Meredith, G. R. Phys. Rev. B 1981, 24, 5522 186. Meredith, G. R. J. Chem. Phys. 1982, 77, 5863 161 187. Ulness, D. J .; Kirkwood, J. C.; Albrecht, A. C. J. Chem. Phys. 1998, 108, 3897 188. Kirkwood, J. C.; Ulness, D. J .; Albrecht, A. C.; Stimson, M. J. Chem. Phys. Lett. 1998, 293, 417 189. Blank, D. A.; Kaufman, L. J.; Fleming, G. R. J. Chem. Phys. 1999, 111, 3105 190. Steinfeld, J. I.; Zare, R. N.; Jones, L.; Lesk, M.; Klemperer, W. J. Chem. Phys. 1965,42, 25 191. LeRoy, R. J. J. Chem. Phys. 1970, 52, 2683 192. LeRoy, R. J .; Bernstein, R. B. J. Mol. Spectrosc. 1971, 37, 109 193. Tellinghuisen, J. J. Chem. Phys. 1973, 58, 2821 194. Tellinghuisen, J. Chem. Phys. Lett. 1977, 49, 485 195. Tellinghuisen, J. J. Quant. Spectrosc. Radiat. Transfer 1978, 19, 149 196. Dantus, M.; Bowman, R. M.; Zewail, A. H. Nature 1990, 343, 737 197. Dunn, T. J.; Sweetser, J. N.; Walmsley, I. A. Phys. Rev. Lett. 1993, 70, 3388 198. Gruebele, M.; Roberts, G.; Dantus, M.; Bowman, R. M.; Zewail, A. H. Chem. Phys. Lett. 1990, 166, 459 162 199. Assion, A.; Baumert, T.; Geisler, M.; Seyfried, V.; Gerber, G. Eur. Phys. J. D 1998, 4, 145 200. Blanchet, V.; Stolow, A. J. Chem. Phys. 1998, 108, 4371 201. Barrow, R. F.; Yee, K. K. J. Chem. Soc. Faraday Trans. 11 1973, 69, 684 202. Stapfeldt, H.; Constant, E.; Corkum, P. B. Phys. Rev. Lett. 1995, 74, 3780 203. Stapfeldt, H.; Constant, E.; Sakai, H.; Corkum, P. B. Phys. Rev. A 1998, 58, 426 204. de Boeij, W. P. Ultrafast solvation dynamics explored by nonlinear optical spectroscopy. Ph.D. Thesis, Rijksuniversiteit Groningen, 1997. 205. Fleming, G. R.; Cho, M. H. Annu. Rev. Phys. Chem. 1996, 47, 109 206. de Boeij, W. P.; Pshenichnikov, M. S.; Wiersma, D. A. Chem. Phys. 1998, 233, 287 207. Joo, T.; Albrecht, A. C. Chem. Phys. 1993, 173, 17 208. Capelle, G. A.; Brodia, H. P. J. Chem. Phys 1973, 58, 4212 209. Paisner, J. A.; Wallenstein, R. J. Chem. Phys. 1974, 61, 4317 210. Zewail, A. H.; Dantus, M.; Bowman, R. M.; Mokhtari, A. J. Photochem. Photobiol. A 1992, 62, 301 211. Lienau, C.; Zewail, A. H. J. Phys. Chem. 1996, 100, 18629 163 212. Zewail, A. H.; Orlowski, T. E.; Jones, K. E.; Godar, D. E. Chem. Phys. Lett. 1977, 48, 256 213. Mukamel, S.; Piryatinski, A.; Chemyak, V. J. Chem. Phys. 1999, 110, 1711 214. Tellinghuisen, J. J. Chem. Phys. 1982, 76, 4736 215. Liu, Q.; Wan, C.; Zewail, A. H. J. Phys. Chem. 1996, 100, 18666 216. Vetchinkin, S. 1.; Vetchinkin, A. S.; Eryomin, V. V.; Umanskii, I. M. Chem. Phys. Lett. 1993, 215, 11 217. Gruebele, M.; Zewail, A. H. J. Chem. Phys. 1993, 98, 883 218. Matemy, A.; Chen, T.; Schmitt, M.; Siebert, T.; Vierheilig, A.; Engel, V.; Kiefer, W. Appl. Phys. B 2000, 71, 299 219. Zadoyan, R.; Apkarian, V. A. Chem. Phys. Lett. 2000, 326, l 220. Comstock, M.; Sudachenko, E.; Lozovoy, V. V.; Dantus, M. J. Chem. Phys. 2002, in preparation 221. Hall, T. C.; Blacet, F. E. J. Chem. Phys. 1952, 20, 1745 222. Bibart, C. H.; Ewing, G. E. J. Chem. Phys 1974, 61, 1284 223. Weiner, N. R.; Nixon, E. R. J. Chem. Phys. 1957, 26, 906 224. Luckhaus, D.; Quack, M. Chem. Phys. Lett. 1992, 199, 293 225. Melen, F.; Pokorni, F.; Herman, M. Chem. Phys. Lett. 1992, 194, 181 164 226. Dyer, C.; Hendza, P. J. Chem. Phys. Lett. 1995, 233, 461 227. Kato, T.; Hayashi, S.; Machida, K. J. Molec. Liquids 1995, 65, 405 228. Stirling, A.; Papai, I.; Mink, J.; Salahub, D. R. J. Chem. Phys. 1994, 100, 2910 229. Domenech, J. L.; Andrews, A. M.; Belov, S. P.; Fraser, G. T.; Lafferty, W. J. J. Chem. Phys. 1994, 100, 6993 230. Wesolowski, S. S.; Fermann, J. T.; Crawford, T. D.; Schaefer, H. F. J. Chem. Phys. 1997, 106, 7178 231. Parsons, B. F.; Curry, S. L.; Mueller, J. A.; Ray, P. C.; Butler, L. J. J. Chem. Phys. 1999, 111, 8486 232. Mueller, J. A.; Morton, M. L.; Curry, S. L.; Abbatt, J. P. D.; Butler, J. B. J. Phys. Chem. 2000, 104, 4825 233. Streuli, C.; Averell, P. The analytical chemistry of nitrogen and its compounds Part I; John Wiley & Sons: New York, 1970. 234. Borisenko, K. B.; Lolonits, M.; Rozsondai, B.; Hargittai, I. J. Mol. Struct. 1997,413,121 235. Kvick, A.; McMullan, R. K.; Newton, M. D. J. Chem. Phys. 1982, 76, 3754 236. Sarkisov, O. M.; Tovbin, D. G.; Lozovoy, V. V.; Gostev, F. E.; Titov, A. A.; Antipin, S. A.; Umanskiy, S. Y. Chem. Phys. Lett. 1999, 303, 458 165 237. Davies, J. A.; LeClaire, J. E.; Continetti, R. E.; Hayden, C. C. J. Chem. Phys. 1999, 111, 1 238. Davies, J. A.; Continetti, R. E.; Chandler, D. W.; Hayden , C. C. Phys. Rev. Lett. 2000, 84, 5983 239. Jolly, W. L. The Inorganic Chemistry of Nitrogen; W.A. Benjamin Inc.: New York, 1964. 240. Herzberg, G. Molecular Spectra and Molecular Structure, Reprint ed.; Krieger Publishing Company: Malabar, Florida, 1991. 241. Andrews, B.; Anderson, A. J. Chem. Phys. 1981, 74, 1534 242. Bolduan, F.; Jodi, H. J.; Loewenschuss, A. J. Chem. Phys. 1984, 80, 1739 243. Press, W. H.; Teukolsky, S. A.; Vetterling, W. T.; Flannery, B. P. Numerical Recipes in C; Cambridge University Press: New York, 1992. 244. Hepp, M.; Georges, R.; Herman, M.; Flaud, J .-M.; Lafferty, W. J. J. Molec. Struct. 2000, 517, 171 245. Gillispie, G. D.; Khan, A. U. J. Chem. Phys. 1976, 65, 1624 246. Lievin, J.; Delon, A.; Jost, R. J. Chem. Phys. 1998, 108, 8931 247. Cao, J. S.; Bardeen, C. J .; Wilson, K. R. Phys. Rev. Lett. 1998, 80, 1406 248. Lozovoy, V. V.; Sarkisov, O. M.; Vetchinkin, A. S.; Umanskii, S. Y. Chem. Phys. 1999, 243, 97 166 249. Brixner, T.; Kiefer, B.; Gerber, G. Chem. Phys. 2001, 267, 241 250. Walowicz, K. A.; Pastirk, I.; Lozovoy, V. V.; Dantus, M. J. Phys. Chem. 2002, (in press, September) 167 IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIII 11l11111111111111111111ll