Eigenstate resolved infrared/infrared double resonance spectroscopy of the 3~~ overtone band of I-propyne: Intramolecular vibrational energy redistribution into a Coriolis-coupled bath Joan E. Gambogi, Erik R. Th. Kerstel, a) Kevin K. Lehmann, and Giacinto Stoles Department of Chemistry, Princeton University, Princeton, New Jersqv 08542 (Received 3 September 1993; accepted 5 November 1993) Sequentialinfrared/infrared double resonanceexcitation of an optothermally detectedmolecular beam has been used to obtain the eigenstate resolved spectrum of the second GH stretch overtone in propyne near 9700 cm-’ . The high resolution and sensitivity of this technique allows for extraction of detailed information about the dynamics of intramolecular vibrational energy redistribution from this highly fractionated spectrum. The analysis suggestsa coupling mechanism consisting of anharmonic coupling out of the bright state through a doorway state or first tier, followed by subsequentcoupling to a strongly Coriolis mixed bath. The lifetime of the bright state, which is determined by the first step and is not dependent on the total angular momentum, is measuredto be about 320 ps for the K=O clumps and about 210 ps for the K= 1 clumps. The root mean squarecoupling matrix element determined for the J’ =0 clump is 0.008 cm-’ and decreaseswith increasingJ’. According to the level spacing and Heller’s F statistics, the spectrum shows evidenceindicating that the underlying dynamic behavior is chaotic. I. INTRODUCTION The combined use of continuous-wavelasers and molecular beam techniques has made possible the study of intramolecular vibrational redistribution (IVR) in the fundamental and, in a few cases,the first overtone vibrations of C-H stretches with sub-Doppler resolution which allows the exploration of intramolecular dynamics in a time window from a few picosecondsto tens of nanoseconds.‘” These eigenstate-resolved, frequency-domain measurements have provided detailed information about time scalesand mechanismsof energy relaxation for a substantial number of polyatomic molecules.The molecules studied have total density of background states ranging from 0.1 to lo9 vibrational states per cm-r. Naturally, there is ample motivation for pushing to higher energies,so as to achieve a level of excitation comparable to the activation barrier of important chemical reactions. In earlier work, Reilly and co-workers’have used similar methods to study the spectroscopy of the third overtone bands of a number of polyatomic molecules. However, because of the extremely small cross sections of third overtone bands, these studies were only successfulfor small molecules for which the oscillator strength is diluted into at most a small number of transitions. Transitions to higher vibrational levels have also been observed using the techniques of long-path absorption,’ photoacoustic spectroscopy,g intracavity laser absorption,” stimulated emission pumping (SEP), r1 visiblemicrowave double resonance spectroscopy,” infrared (IR)-visible double resonancespectroscopy,l3 and Raman shifting of a pulsed visible laser with optothermal detection of a molecular beam.14For the vast majority of these studies, however, the resolution has been limited to at most 1 GHz, due to both Doppler broadeningand the linewidth of %hrrent address: Laboratorio Europeo di Spettroswpie Non-linear4 (LENS), Largo E. Fermi no. 2 (Arcetri), 50125 Firenze, Italy. typical pulsed lasers. In addition, the first three methods, which are the most sensitive, are also limited by the spectral congestion of room temperature (or at best modestly cooled”) samples. At 1 GHz spectral resolution, a maximum of about 10 lines/cm-’ can be resolved, while from recent2>3 and not so recentr6t8 studies in the region of the fundamentals, there is ample evidence that a density of states of 10-100 states/cm-’ is typically required for statistical IVR of the excitation energy over all the isoenergetic states of the molecule to occur. A systematic exploration of the same density of states region for smaller molecules at higher energiesrequires an improvement in the resolution by one to two orders of magnitude. In the present paper, we report the extension of our sub-Doppler resolution, molecular beam technique to the energy region of three quanta of the C-H stretch by use of sequential IIUIR double resonancespectroscopy. Double resonancetechniques have the advantage over single photon experiments of providing state-selected spectra, in which all lines are unequivocally assigned the quantum numbers of the transition pumped by the first laser, while retaining the high resolution characteristics of continuouswave one photon experiments.In our experiments,the first laser is used to pump a transition of the y1 band, the acetylenic C-H stretch fundamental vibration in propyne, and the secondlaser is scannedthrough the u = 3 + 1 transitions of the same mode. The energy deposited in this way into three quanta of vibration is -9700 cm-’ or 1.2 eV. This energy is well below the barrier to isomerization of propyne to allene (-22 750 cm-r) or to cyclopropene (20 650 cm-‘) .I9 Propyne was chosen for this study becauseits acetylenic C-H stretch has been well characterized at the U= 1 and u=2 levels in this and other laboratories.20-23 It is also convenient that in propyne, the fundamental or spectrum is not fractionated by accidental resonant perturbations, allowing the energy to remain localized in the bright state.21 J. Chem. Phys. 100 (4), 15 February 1994 2612 0021-9606/94/100(4)/2612/11/$6.00 @ 1994 American Institute of Physics Downloaded 18 Mar 2002 to 128.112.83.42. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp Gambogi et al.: The 3v, overtone band of 1-propyne In addition, the transitions are easily located since the 3~~ band of propyne has been previously studied at lower resolution by Herzberg in 1938 (Ref. 24) and again in 1984in the thesis of Hall using Fourier-transform infrared (FTIR) spectroscopy.25Both studies, conducted at room temperature, resolved the J but not the K structure of the P and R branches of the spectrum of this prolate symmetric top, parallel transition. Our particular interest is to seewhat characteristics of the intramolecular dynamics can be extrapolated from the behavior at lower energiesand to compare molecules with similar density of states but very different levels of excitation. The fundamental Y* spectrum of propyne, for which the calculated density ofA, states Pcalc=0.3/cm-r, is characterized by nonresonant perturbations to the zero order spectrum which causeband origin shifts in the different K components, but no fractionation of lines2* The nonresonant perturbations are deduced to be due to, at least, two states acting through a mechanism of anharmonic coupling of the Al symmetry bright state to background states that are themselves strongly mixed by z-axis Coriolis coupling. The first overtone spectrum ( p$, = 1 l/cm-‘) shows evidence of one or more nonresonant doorway states which anharmonically mix with the bright state and subsequently couple mainly by z-axis Coriolis interactions to the isoenergetic bath states.22Go and Perry have observedan additional state, shifted 0.1 cm-’ lower in energy than the 2v, state, which is likely to be one of these doorway states.23 For terminal acetylenic C-H stretches at low energies, anharmonic coupling typically dominates1-3v5,6while strong Coriolis coupling has been implicated in the study of the fundamental C-H stretch of trans-ethanolz6 At higher energies, the SEP spectrum of H,CO is dominated by Coriolis couplin2’ while the overtone spectrum of NH3 shows largely anharmonic couplings.” Anharmonic coupling is recognizable in a spectrum with the interaction strength showing no systematic dependenceon rotational quantum numbers. Z-axis Coriolis coupling, in symmetric tops, is distinguished as having an interaction strength that scales with K, is nonexistent at K=O, and follows the selection rules AJ=O, AK=O; while x,y-axis Coriolis coupling scales as dJ(J+ 1) -K(K& 1) with the selection rules AJ=O,AK= i 1. For the second overtone excitation of the acetylenic C-H stretch in propyne, the calculated density of A, vibrational states is 150/cm-‘. This is expected to be well into the “intermediate” regime of IVR, with considerable mixing of the 3~~bright state with the isoenergetic bath states made up of vibrational states with energy distributed among the 15 normal modes of vibration. ,This energy region, around 9700 cmT1, has not previously been studied at this level of resolution ( 10 MHz), which is what is neededto fully resolve the spectrum. Highly mixed spectra are thought to be a consequenceof deterministic chaos, with the intramolecular dynamics producing something akin to a microcanonical distribution of the vibrational energy over all the degrees of freedom. There is strong evidence that classical chaos in a quantum molecular system results in the quantum spectrum having statistical 2613 properties similar to those of a Gaussian orthogonal ensemble (GOE) of random matrices.29’30In this paper, the observed 3~~ spectrum of propyne is analyzed with the typical measuresof highly mixed spectra such as dilution factors, 16-r8Heller’s F statistics,31level spacing statistics,32 and distribution of coupling matrix elements33to conclude that the underlying dynamic behavior is indeed chaotic. II. EXPERIMENT The description of our optothermal spectrometer can be found in Ref. 5. Additional experimental details pertaining to this study and the double resonance methodology are listed here. Propyne (Farchan Laboratories) was diluted to 1% in helium and expanded through a 50 ym nozzle with a backing pressure of 4 atm. This results in a very cold expansion (4-6 K) . Initial excitation of a single rovibrational state (J,K) is provided by a Burleigh FCG20 color center laser pumped by a Krf ion laser. In the region of the acetylenic C-H stretch the laser produces -20 mW of single mode power which is enough to saturate the v1 transitions in propyne. The 3vr+~~ step of the sequential excitation is obtained with a Burleigh FCG120 color center laser which has been converted to an FCL-220 laser by replacing the original crystal with a NaCI:OH-; crystal. The laser presently tunes in the range 1.475-1.725,um with peak power of - 180 mW at 1.6 ym. The laser is pumped by 4 W of 1.06pm radiation from a mode locked Nd:YAG and 5-10 mW of 0.532 pm green (F-band) light generated by passing the 1.06 pm light through a BBO doubling crystal. Due to the long excited state lifetime of the F-center crystals, the mode locking of the pump laser is inconsequential to the single mode, cw operation of the FCL-220 laser and is implemented in order to generate sufficient F-band, green light and becauseunder these conditions the stability of the pump laser is improved. Two key elements for obtaining high quality double resonance spectra are a power level sufficient to saturate transitions and suppression of noise. Both lasers are crossed about 40 times with the molecular beam by means of a pair of parallel mirrors separatedby precision spacers. The fundamental propyne transitions are saturated with -5-10 mW of power and so there is ample power to saturate the first step. The overtone laser has been found to be not far from saturating the 2~~~0 transitions in HCN and the 3v,+ lvr transitions are expected to be about three times stronger. However, as the spectra become fractionated, the transition dipole is diluted over many eigenstates and consequently more power is required for saturation. The second concern is suppressionof noise. Bolometer detection is sensitive to changes in the heat content of the molecular beam due to absorption of radiation. Therefore, beam flux, the size of the photon, laser power, and the molecules’partition function, all determine the magnitude of the signal. In a single laser experiment, the bolometer noise is determined by the stability of the molecular beam flux and by intrinsic bolometer (l/f ) noise. However in a double resonanceexperiment where the first laser is kept on top of a transition and a signal observed as the additional energy input due to the second laser, the background J. Chem. Phys., Vol. subject 100, No. to 4, AIP 15 February 1994 Downloaded 18 Mar 2002 to 128.112.83.42. Redistribution license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp 2614 Gambogi ef al.: The 3q overtone band of I-propyne noise level may be dominated by instabilities in the frequency or amplitude of the pump laser. A typical 3 pm transition has a full width of 8 MHz and our color center laser has frequency fluctuations on the order of a few MHz, thus leading to a sizable increase in the noise when sitting on the pumped transition. We have reduced this source of noise by frequency stabilizing the 3 pm laser. This was done by locking the laser to the side of a fringe produced from a high finesse &talon (Burleigh) which has a 150 MHz free spectral range. The differencesignal from a thermoelectrically cooled InAs detector (Electra-Optical Systems) monitoring the etalon transmission and a reference voltage corresponding to 30% of the &talon transmission peak voltage was amplified with an adjustable gain and bandwidth and added to the voltage that controls the PZT of the laser cavity length. The noise of the double resonance experiments was increasedin comparison to the one photon experiments to -0.4% of the 3 pm signal. From this, one can estimate the fluctuations of the 3 pm laser being held on top of the transition to be on the order of 300 kHz. In order to lock the &talon onto the pump transition, the &talon was dithered (which dithered the pump laser by the first feedback loop), and the demodulated bolometer response used to provide feedback to the &talon cavity length. Over a period of time, the lock loops were not able to eliminate all dc drifts and these were corrected by hand, through monitoring the dc level of the &talon feedback signal on an oscilloscope. It was possible in this manner to lock onto a transition for several hours. A homebuilt wavemeter was used to monitor the 1.6 pm laser fi-equency.34The wavemeter has a resolution of -0.001 cm-‘. The absolute frequency was calibrated against known transitions in acetylene3’and resulted in a 6.7 ppm correction to the output of the wavemeter. Relative frequency measurementsfor the recorded spectra were determined from the transmission of a second 150 MHz &talon used in connection with the 1.6 pm laser. III. RESULTS Propyne is a symmetric top molecule of C,, symmetry with only Al and E modes.‘t The 3vt normal mode is of Al symmetry and undergoes a parallel transition. A series of individual clumps, characterized by the final J and K quantum numbers, of the 3vt mode in propyne have been recorded and a few shown in Figs. 1 and 2. Line fractionation increasesdramatically with J’ and equally as dramatically from the K=O to the K= 1 clumps. Only clumps with J’<4 and K=O, have signal to noise ratios sufficiently high to allow confidencethat most of the lines have been observed. A detailed spectroscopic analysis will therefore be limited to these clumps. Listed in Table I are the number of assignedtransitions for each rotational level and the sequenceused to arrive at the transition, while Table II lists the transitions for the observedeigenstatesfor J’<4 and K=O. A wavemeter was used to measure the frequency of the 1.6 pm laser used in the v= 3 c 1 transitions and this was added to the frequency of the fundamental transitions determined previou~ly.~l Rotational constants for the 3v1 band of propyne I- (0) 9702.333 9702.393 WAVENUMBERS (cm-l 9702.453 ) (b) ~~~~ 9702.322 L&T& 9702.382 WAVENUMBERS (cm-l 9702.442 ) (cl I .I’‘@4#v L 9702.280 9702.340 WAVENUMBERS (cm-l 9702.400 ) FIG. 1. Portions of the 3v1 spectrum of 1-propyne, K’=O, observed by WIR double resonance and bolometer detection of a skimmed molecular beam (a) J’=O, K’=O; (b) J’=2, K’=O; and (c) J’=4, K’=O. were previously determined by Hall25 andno improvement in the constants can be extracted from our data mainly becauseof the low signal to noise above Jr=4 and above K’=o. An extremely large splitting between the K=O and 1 clumps was measured,with the K= 1 clumps lower in energy. This splitting appears to slightly decrease with J’. Some error is inherent in this measurement because the small S/N ratio for the K=l clumps makes the center of gravity determination uncertain due to themtransitions which are not accounted for. However, the 3 GHz splitting measured is much greater (and in the opposite direction!) than that expectedbased on the analogous splitting in the fundamental band (260 MHz) (Ref. 21) and that in the first overtone (576 MHz) .22The difference in the subband origins found in v1 was attributed to nonresonant pertur- J. Chem. Phys., Vol. 100, No. 4, 15 February ‘l994 Downloaded 18 Mar 2002 to 128.112.83.42. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp Gambogi et al.: The 3v, overtone band of I-propyne 9702.216 9702.276 WAVENUMBERS (cm-l this C,, molecule there are no A, modes. Therefore such coupling must occur indirectly through anharmonic mixing of Al with another Al state containing multiple quanta in E modes resulting in A1+A2 pairs. The line splittings in the K=O and K= 1 clumps for this spectrum may have the same origin, however. with only two K subbranches measured, there is not enough evidence to determine the mechanism conclusively. The anomalous behavior could be explained equally well by the presenceof a near resonant anharmonic state which tunes in and out of resonancewith the bright state. Background states are expected to tune very rapidly as a function of K and more slowly as a function of J (since K tunes as f CAK and c can vary significantly between perturbing states). Perpendicular or x,y-axis Coriolis coupling can be ruled out as the primary causefor this perturbation in the K lines since the splitting is only a very weak function of J. 9702.336 ) r 2615 (b) A. Density of states determination from the spectrum Density of states were calculated from the spectrum by counting the number of transitions and dividing by the energy window according to the formula 9702.172 9702.232 WAVENUMBERS (cm-1 9702.292 FIG. 2. Portions of the 3~~ spectrum for K’= 1; (a) J’=2, J’=4, K’=l. K’= 1 and (b) bations through a parallel or z-axis Coriolis interaction. In the 2vt spectrum parallel Coriolis coupling was also thought to play an important role since the matrix elements scaled with K and there was no J dependence.In both of those cases,as well as in the present spectrum, the vi transition is of A, symmetry and cannot interact directly with other modes through z-axis Coriolis coupling. Z-axis Coriolis coupling mixes Al and A2 normal modes and for TABLE I. Summary of experimental results from the 3~~ spectrum of propyne. Sequence (3 pm, 1.5 pm) 00 2 R(O),P(l) R(l), P(2) 7, R(2), R(l), R(4), R(5), R(4), R(5), 11 R(l), P(2) 30 40 50 60 21 31 41 Level lo-11 %I-21 30-31 40-41 exp= (N- l)/AE. (1) There are many methods to extract experimental density of states from a spectrum, but in the case of a highly fractionated spectrum it is most direct to just count states. In the analysis of 2vt in propyne,** and also for 1v16in butyne,6 an alternative approach was used which involved the strength of coupling for each eigenstateand the finite S/N of the spectrum. However, this method has been shown to be accurate only when the product of the density of states times the root mean squared coupling matrix element is -0.5 or less, which is not valid for this strongly fractionated propyne 3~~ spectrum. On the other hand, formula ( 1) underestimatesthe true density of states by a factor of 1 over the fraction of eigenstatesobservedin the spectrum. P ) P(3) R(2) P(5) P(6) R(5) R(6) R(2), pl3) R(l), R(2) R(4), f’(5) Ntimber of assigned eigenstates pexp (/cm-‘) 8 17 25 32 47 24 27 29 18 41 44 17 67 126 249 408 485 225 359 482 258 285 404 223 Total energy, center of gravity [cm-‘) . .. 9702.389 9702.388 9705.220 9702.348 9702.325 9708.577 9709.683 9702.275 9702.286 9705. IO7 9702.247 Energy difference (cm-‘) 0.114 0.102 0.113 0.101 J. Chem. Phys., Vol. 100, No. 4, 15 February 1994 * Downloaded 18 Mar 2002 to 128.112.83.42. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp Gambogi et a/.: The 3v, overtone band of 1-propyne 2616 TABLE II. Observed transitions from the 3v, band in propyne for 0<3’<4 and K'=O. The energy reflects the total energy deposited into the molecule between the two lasers. Wave number (cm-‘) 00 10 20 30 Wave number (cm-‘) Relative intensity 0.061 0.206 0,019 0.520 0.068 0.017 0.056 0.053 0.023 0.094 0.013 0.048 0.017 0.066 0.061 0.189 0.143 0.141 0.045 0.015 0.029 0.051 0.020 0.025 0.019 0.011 0.015 0.035 0.019 0.015 0.029 0.055 0.047 0.016 0.012 0.262 0.019 0.023 0.015 0.021 0.011 0.125 0.087 0.072 0.014 0.024 0.012 0.032 0.017 0.012 0.019 0.018 0.011 0.031 0.024 0.009 0.045 0.072 0.014 0.079 0,130 0.006 0.030 0.006 0.014 9702.4453 9702.4092 9702.4082 9702.3877 9702.3818 9702.3760 9702.3564 9702.3398 9702.4531 9702.4150 9702.4141 9702.4053 9702.4014 9702.3955 9702.3936 9702.3906 9702.3896 9702.3857 9702.3838 9702.3701 9702.3643 9702.3613 9702.3496 9702.3457 9702.3262 9702.4307 9702.4268 9702.4258 9702.4111 9702.4092 9702.4053 9702.4033 9702.4004 9702.3975 9702.3945 9702.3945 9702.3906 9702.3877 9702.3857 9702.3857 9702.3818 9702.3789 9702.3760 9702.3730 9702.3711 9702.3584 9702.3516 9702.3477 9702.3418 9702.3340 9705.2607 9705.2559 9705.2529 9705.2490 9705.2480 9705.2461 9705.2344 9705.2305 9705.2285 9705.2275 9705.2256 9705.2236 9705.2227 9705.2207 9705.2188 J.Chem. Phys.,Vol. 9705.2178 9705.2158 9705.2139 9705.2129 9705.2090 9705.2080 9705.2070 9705.205 1 9705.2051 9705.2020 9705.2011 9705.1992 9705.1973 9705.1953 9705.1895 9705.1885 9705.1846 9702.3994 9702.3965 9702.3867 9702.3818 9702.3809 9702.3750 9702.3721 9702.3691 9702.3662 9702.3643 9702.3633 9702.3633 9702.3604 9702.3594 9702.3564 9702.3555 9702.3545 9702.3535 9702.3525 9702.3516 9702.3506 9702.3506 9702.3496 9702.3477 9702.3477 9702.3457 9702.3438 9702.3428 9702.3418 9702.3398 9702.3389 9702.3369 9702.3359 9702.3350 9702.3320 9702.3311 9702.3301 9702.329 1 9702.3262 9702.3252 9702.3203 9702.3184 9702.3154 9702.3135 9702.3125 9702.3086 9702.3047 40 100, No.4,15 Relative intensity 0.047 0.012 0.025 0.017 0.133 0.048 0.026 0.029 0.040 0.013 0.009 0.014 0.011 0.014 0.020 0.023 0.011 0.013 0.007 0.006 0.011 0.021 0.005 0.019 0.007 0.023 0.013 0.010 0.021 0.015 0.032 0.007 0.054 0.024 0.021 0.100 0.014 0.018 0.006 0.084 0.075 0.009 0.035 0.028 0.009 0.013 0.032 0.013 0.008 0.011 0.024 0.012 0.011 0.054 0.017 0.008 0.008 0.034 0.008 0.019 0.011 0.012 0.006 0.012 February1994 Downloaded 18 Mar 2002 to 128.112.83.42. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp Gambogi et a/.: The 3~; overtone band of 1-propyne B. Time evolution p,Q 360 (2J' -t 1) FIG. 3. Plot of experimental density of states as a function of J’. Only 5’==0-4. K'=O clumps are shown. The calculated density of vibrational states is 150/cm-‘. For J/-O, pexp is 67/cm-’ which can be compared with the result of a direct count of A1 symmetry states which gives 150/cm-‘. The density of states has been calculated including the diagonal anharmonicities compiled by McIlroy and Nesbitt.20 With a S/N of -70 for the strongest peak in this clump, it is unlikely that any states except the very weakly coupled ones have been missed. Therefore, we can state that at J’=0 the 3~~bright state is coupled to -50% of the symmetry allowed background states suggestingthat the dynamics only explore half of the available vibrational phase space. As J’ increases peXPincreasesdramatically to 485/cm-’ at J’ =4 (Table I). As shown in,Fig. 3, this increase withJ’ in the range O<J’<4 is approximately proportional to .(2J’-I- 1). Above J’ =4 the decreasing signal to noise, caused mainly by the increasedfractionation of the spectrum, strongly distorts the statistics and does not allow for the drawing of clear cut conclusions. Upon cooling in a supersonic expansion, conservation of nuclear spin symmetry is expected, and as a result the room temperature distribution of 50% ortho (K= 3N) and 50% para (K=3N& 1) will be preserved, with half of the population in K=O levels and half in K= 1. In the fundamental spectrum, the K= 1 lines in the R and P branches are slightly stronger than the K=O lines, due to the fact that the J=K=O level gives the K=O population one more state to distribute its- population over compared to K= 1. Similar behavior also occurs for the first overtone band. For the 3vi band, however, the total peak intensity measured, for J’ = 1 and 2, in the K= 1 spectrum is a factor of 5 lower than for K-O. While the number of lines observed in the K= 1 clumps did not increase by the same factor, this most likely reflects the fact that a large fraction of lines go undetected due to insufficient signal to noise ratio of the K= 1 spectrum. The actual densities of states for the K- 1 clumps are probably considerably larger than those calculated here. 2617 of the bright state Information about the intramolecular dynamics of the molecule can be obtained from the survival probability, P&t>, which gives the probability that the molecule will be in the bright state at time t, given that it was prepared in that state at time equal to zero.36In the case of spectra showing an intermediate level of IVR (i.e., where the spectral fractionation is high but the eigenstatesare still resolvable), such as that considered here, the time evolution of Pa, is given by an initial decay followed at later time by quantum beats which occur as the energy flows back into and out again of the initially prepared state. P,,(t) can be directly computed from an observed spectrum by XCOS[(Ej-Ej>t/h] xexp(--yJ), (2) where I Cjl 2 are the normalized intensities measuredin the spectrum arising from a single bright state (i.e., all inhomogeneous contributions to the spectrum have already been separated out) and correspond to the amount of bright state character in a particular eigenstate;E,.‘s are the transition frequencies and l/y, is the radiative lifetime of the upper states which is negligibly long. Calculating lifetimes by this formula corresponds to taking a Fourier transform of the autocorrelation of the spectrum. The time evolution curves calculated for the different clumps are shown in Fig. 4. For the K=O lines at low J’s there is an initial decay of intensity followed by the population beating back into the initial state. For the higher J’s the amplitude of early time recurrencesare greatly reduced and thus the population is distributed over a larger volume of phase space.In spite of the different qualitative behavior of the time evolution with J, the lifetimes (estimated from the curve at the point where l/e of the initial intensity has decayed) are all approximately equal to 320 ps. For the K=l clumps, which are notably poorer in S/N and have many transitions below the noise level, the l/e lifetimes calculated from the decay curves are also all approximately equal and correspond to a lifetime of about 210 ps. Given the large number of lines which are likely to be missing from the K=l spectrum (especially in the wings of the clump), the true decay time is likely to be shorter than 210 ps. C. Coupling matrix elements The spectrum was deconvoluted to obtain the positions of the bright and bath states using a modified version37of the Lawrance-Knight deconvolution method.38 Assumptions necessaryto use this procedure are the existence of a single bright state and the lack of direct coupling between the bath states. As will be discussed below, the spectrum indicates that the bath states are in fact strongly coupled among each other. The deconvolution procedure is still valid however if one understands that the “bath” states that are determined are not near normal mode basis functions (as the bright state), but the linear combinations of those states that “prediagonlize” the Hamiltonian neglect- J. Chem. Phys., Vol.subject 100, No. Februaryor copyright, 1994 Downloaded 18 Mar 2002 to 128.112.83.42. Redistribution to 4, AIP15license see http://ojps.aip.org/jcpo/jcpcr.jsp Gambogi et al.: The 3v, overtone band of 1-propyne 2618 --.O.&O0 0.0008 _ 0.0016 MATRIX TIME 0 1000 TIME 3000 4000 (psec) FIG. 4. Tie evolution of the fluorescence intensity calculated using Eq. (2); (a) K’=O, 00~6; (b) K’=l, l<P<4. Thecorrespondingsymbols _ J’zl, -.-J’z2; - . .._ J’z3; ___J’z4. and curves are . . . J’zo; ing coupling to the bright state. Table III lists the average and rms average coupling strengths determined for each clump. For a Gaussian distribution of elements,the ratio of these two averagesis 1.25. Most of the levels have a ratio TABLE J+.. 00 lo 20 3, 40 5, 60 70 11 21 31 41 I i I .-.-___ __ . O.Ob32<:0040 ELEMENT (cm-l 0.0048 0.0056 ) FIG. S. Distribution of coupling matrix elements obtained from a modified Lawrance-Knight deconvolution for the J’=4, K’=O clump. Spectra showing chaotic dynamics are expected to have a Gaussian distribution of elements. (psec) 2000 _-_. 0.0024 very near this value, indeed the histogram for J’=4, K’ =0, shown in Fig. 5, appearsGaussian considering that there are only 47 elements plotted. As a function of J’, the average coupling matrix element as well as.the sum of the squares of the matrix elements decreasewith increasing J’. This is a natural consequence of the fact that the density of lines in each clump increases with J’, while the decay rate of P,,(t), which reflects the overall width of the clump, is largely independent of J’. The width of each clump basically reflects the product of the density of states times the mean square matrix element as in the Fermi golden rule expression. The linear increase in the density of states with ( 2J’+ 1) strongly suggeststhat the coupling of vibration and rotation has completely broken down the rigid rotor type motion and thus each eigenstate has contributions from basis functions spanning all K values allowed by symmetry. This type of strong intramolecular vibrationrotation energy transfer has been discussedpreviously, and reflects the presenceof strong Coriolis or centrifugal coupling in the molecule. The independenceof the decay rates with J, for the K=O bright state spectra, is consistent only III. Coupling matrix elements calculated from a modified Lawrance-Knight tv> (cm-‘Y 0.006785 0.004336 0.003398 0.002634 0.002292 0.003 695 0.002540 0.0023cKJ 0.003092 0.003 683 0.003444 0.003 650 V,, (cm-‘)b 0.008 193 0.005048 0.003 818 0.002955 0.002574 0.004147 0.003306 0.002601 o.OQ3733 0.004303 0.003 896 0.004105 deconvolution. v,nl,/( v) Z)Vl” (VZ) x Pexp 1.21 1.16 I.12 1.12 1.12 1.12 1.30 1.13 1.21 1.17 1.13 1.12 o.Gm470 0.000408 o.oOO35o 0.000271 o.cm305 0.000396 0.000284 0.000189 o.oOo571 o.Ooo741 O.COO653 O.ooO270 0.00450 0.00321 0.00363 0.003 56 0.00321 0.003 87 0.00392 0.00326 0.003 60 0.00528 0.006 13 0.00376 .~ “( v) is a straight average of the matrix elements &V~/Nbath. b V,, refers to root mean square matrix element calculated from the formula ( ZiG/Nbafh) 1’z. J. Chem. Phys., Vol. 100, No. 4, 15 February 1994 Downloaded 18 Mar 2002 to 128.112.83.42. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp Gambogi et al.: The 3v1 overtone band of I-propyne with a model where the coupling between the bright state and the bath states is dominated by anharmonic interactions. Perpendicular or xg Coriolis coupling would result in a J(J+ 1) dependenceof the decay rate. Since the anharmonic interaction will only couple the K=O bright states to other K=O basis states, the decreasing mean squared matrix element reflects the decreasingK=O character of the “bath states” in the spectrum. If each eigenstate is a random mixture of all allowed K values, one would expect the mean square coupling matrix element to be inversely proportional to (2J’-I- 1), which is approximately born out by the results listed in Table I. The strong dependenceof the decay rate on K would appear to point to a strong contribution of z-axis Coriolis coupling to the decay rate for K> 0, but as discussedin the paper on the 2vi spectrum,22 there should be no z-axis Coriolis force operating on the atoms while undergoing motion along the v1 normal mode, and thus this K dependence must reflect the presence of a doorway state that increasesthe coupling of the bright state to the bath. This doorway state may be n-axis Coriolis coupled with the bath states as we believe occurs in the 2vt spectrum, or the doorway state is just tuned closer to resonance at K= 1 than K-O. With our limited data, we cannot rule out either of these explanations, and in fact both could be operating simultaneously. The observed strong shift of. the K= 1 levels from the K=O energy cannot continue unabated at higher K or the room temperature spectrum would not be expectedto show resolved Jstructure. Therefore, we believe that the tuning into closer resonanceof a specific doorway state is the most likely explanation for the increased coupling of the K= 1 states. If our current attempts to increasethe 1.6 pm laser intensity prove successful, we plan to re-examine the propyne spectrum to observe higher K states. D. Energy level spacing statistics A well resolved intermediate case IVR spectrum, such as this, is an ideal candidate for a statistical analysis. Spectra that have regular classical dynamics are generically expected to show a Poisson’s distribution of spacings.39Chaotic spectra are expected to closely resemble the GOE spectrum, due to the energy level repulsion of each level with its neighbors.40In order to remove the effect of the presenceof the bright state, which causesan artificial enhancement of ‘level density close to it, the statistics are calculated using a deconvoluted spectrum, i.e. only the positions of the bath states are considered. Comparison of the level spacing statistics from a LawranceKnight deconvolution [Table IV(a)] with the original eigenstates(not listed) produces slight differences for each J’. The normalized moments are 5%-10% larger for the molecular eigenstatesindicating greater clustering of levels, as expected,by inclusion of the bright state. It is expected that the clumps with fewer amount of levels would be affected the most, however the differencesappear to be uncorrelated to the number of levels. 2619 TABLE IV. (a) Normalized moments for the level spacing statistics calculated from the spectrum of 3vr in propyne. The spacings are calculated using the bath states energy levels obtained from a Lawrancs Knight deconvolution. (b) Normalized moments for a fractional GOE level spacing allowing for missing states.a The quantityfis the fraction of states measured in the spectrum. (a) xv 00 10 20 30 40 50 60 70 1, 21 31 41 V)/W b-?/M3 (s4)/W4 1.33 2.22 1.57 1.70 1.59 1.85 *2.13 1.42 1.36 2.69 1.48 1.59 2.09 7.63 3.49 4.78 3.78 4.63 6.85 2.51 2.23 13.89 3.39 3.91 3.65 31.03 9.28 18.18 12.15 13.04 26.15 5.01 4.09 89.33 10.94 12.25 2.00 2.01 1.93 1.83 1.72 1.62 1.53 1.45 1.38 1.33 1.27 6.00 6.51 6.04 5.28 4.48 3.79 3.23 2.78 2.44 2.17 1.91 24.00 30.07 27.46 22.00 16.48 12.07 8.89 6.69 5.18 4.15 3.24 (b) f 0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 1.00 &The normalized moments accounting for missing levels are calculated based on Ref. 48. Considering first the S =0 clump of 8 assignedtransitions, the level spacings correspond to a Wigner distribution with f =0.9 or only 90% of the levels observed. This does not agree with the 50% of levels coupled from the count of eigenstatesin the spectrum however with only 8 states to consider, level spacing statistics are subject to some error. In spite of this, the clump appears to exhibit close to Wigner statistics indicating chaotic dynamics. For J’ =04, where the signal to noise ratio is high enough for other analysis, such as density of states or lifetimes, to show clear trends, the level spacing statistics do not. Certain clumps actually show normalized moments larger than 2 which is indicative of extreme level clustering. This behavior, GOE like statistics for the J=O spectrum, but Poison or even more highly clustered statistics for the rovibrational spectrum, despite an apparent breakdown of the K selection rules and thus believed strong K mixing, has been observed previously in the optical spectrum of NOz.41 Steve Coy has done extensive simulations of what he calls the K-breakdown model where he adds weak K mixing interactions to a strongly anharmonically coupled system.j2 He has found that he can reproduce statistical properties similar to the NO2 spectrum, and thus the propyne 3vi spectrum as well, if the K-mixing matrix elements are of intermediate strength that is, strong Downloaded 18 Mar 2002 to 128.112.83.42.J.Redistribution subject AIP or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp Chem. Phys., Vol. 100, toNo. 4, license 15 February 1994 Gambogi et al.: The 3v, overtone band of 1 -propyne 2620 TABLE V. Statistics calculated from the intensity information in the spectrum. Shown below are the dilution factors, I&, the number of effectively coupled states, q5; ‘, and the fraction of phase space explored by the dynamics of the system, F. s,, N 0, 8 2 +d 0.321 0.103 0.110 0.062 0.041 0.052 0.049 0.045 0.079 0.032 0.029 0.086 2: 2 50 60 70 1, 21 31 4, 32 47 24 27 29 18 41 44 17 9G’ 3.05 9.68 9.11 16.09 24.43 19.29 20.42 22.22 12.65 31.72 34.57 11.64 F 0.20 0.17 0.13 0.21 0.11 0.15 0.15 0.17 0.43 0.30 0.3 1 0.12 enough to mix the eigenstates allowing most to be observed, but not strong enough to drive the rovibrational statistics into the GOE limit. The apparent clustering of levels even beyond the Poisson predictions can then be understood as reflecting experimental detection bias. If a rovibrational level is particularly close to one that already has transition strength, then high order K-mixing interactions will give this other state more chance to be sufficiently strong in the spectrum and to be observed. An interesting dynamic effect of this model is that if the K-mixing matrix elementsare weaker than the anharmonic matrix elements, then the time scale for the relaxation of molecular orientation will be slower than the rate of IVR. This has potential experimental consequencesand could in principle be probed by IR-UV double resonance experiments where the anisotropy of the UV absorption is measured as a function of time following overtone excitation.“3 The prediction is that rate of decay of such anisotropy would be slower than the IVR decay rate of the 3~~state. Since Crim and co-workers have observedelectronic spectra after pumping acetylene to u= 3, this experiment is likely feasible with existing laser technology.M E. Intensity statistics Other information can be gained from the spectra by considering the intensities corresponding to the different eigenstates.Quantities usually calculated are dilution factors, as introduced by McDonald,i6 and Heller’s F.31 The dilution factor, a,, =is calculated with the formula ud= xiff/(zJf) ‘2 (3) where l/cd is referred to as the effective number of coupled levels to the bright state (N,,) . As pointed out by Perry, od values are artificially large at linite S/N becauseof the loss of low intensity liies.6 The dilution factors, listed in Table V, shows a definite decreasewith Jr which- in reality is probably even larger since the S/N also decreaseswith J’. McDonald and co-workers have calculated dilution factors for a great number of molecules using total and dispersed infrared laser induced fluorescenceand exciting the fundamental C-H stretch.““’ They have found a rough correlation between dilution factors and density of rovibrational states where there is a sharp falloff in dilution factors at around 70 states/cm- 1.17,18 The measurements made here for J’ =0 through 4 all fall within their plot of measuredvalues (Fig. 10 of Ref. 18). It is interesting that the threefold increase in vibrational energy observedhere does not appear to change the order of magnitude of the dilution factor. Minton and McDonald describe a model basedon statistical coupling of the rovibrational statesthat was used to predict dilution factors for different J levels in norbornadiene.18The model assumed equal coupling between bath and bright states and has the averageenergy over which the bright state character is spread as the model’s one adjustable parameter. Setting the coupling width equal to 0.05 cm-‘, they were able to approximately obtain .the dilution factor’s found experimentally. A coupling width from our spectrum was calculated by lb&p. (4) For the case of J’ =0, the number of effectively coupled states (a; ‘) is 3 divided by 67 states/cm-‘, corresponding to a coupling width of approximately 0.05 cm-‘. Indeed, the coupling widths at each J’ fluctuate around a value of 0.05 cm-‘. Since the coupling matrix elements calculated by deconvoluting the spectrum indicate a Gaussian distribution of values, a statistical coupling model is perhaps quite descriptive of this system. McDonald’s measurements were all taken at low resolution and therefore the derived quantities are highly averaged.The present results indicate however, for highly mixed spectra, that quantities such as dilution factors can be deduced at high and loti resolution with comparable results. Heller’s F is a quantity that measures the fraction of available phase space accessibleto the system.3’F is deflned as F=u;‘/N, (5) where N is the total number of phase space cells available and a;’ is the actual number explored by the dynamics. Classically ergodic systems would have F= 1, while a pure chaotic quantum state has a maximum F= l/3.“’ Twothirds of the phase spaceis unaccesableto the dynamics of a state due to a quantum interference effect, basically becausea quantum state must have nodes. The value of F for a quantum state deviates less from one as the purity of the initial state degrades.The quantity N was estimated from the calculated density of rovibrational states [ 15O/cm-’ X (W+ l)] multiplied by the energy window observed in the spectra at each J’ and a; ’ as defined in Eq. (3). For the J’ =0 level, F is calculated to be 0.20 which is close to the chaotic limit of l/3. In fact, all of the values calculated for K’ -0 fall around 0.2. This is further support that there appearsto be no residue of K as a good quantum number. The F values calculated for the K’ = 1 spectra fluctuate alot more and this can be attributed to the low signal to noise ratio. Decreasing the signal to noise will tend to increase the value of F since it reduces the dynamic range of the observed spectral lines. J. Chem. Phys., Vol. 100, No. 4, 15 February 1994 Downloaded 18 Mar 2002 to 128.112.83.42. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp Gambogi et al.: The 3v, overtone band of 1-propyne IV. DISCUSSION The analysis of the data of the 3~~spectrum of propyne strongly indicates a two step coupling mechanism in which the bright state is anharmonically coupled to a first tier of all states which have the same rotational quantum numbers and vibrational symmetry. These states, in turn, are mixed to other tiers of rovibrational states having different values for K, but the same overall rovibrational symmetry. The couplings between these further tiers are expected to occur largely via x,y Coriolis interactions (which mix states with AK= * 1), although some mixing via centrifugal distortions (which allow up to AK= =.t2 t6 occur directly) may occur. This mechanism is supported by the almost linear increase in density of coupled states with J’ which implies that the bright state can mix with almost any symmetry allowed rovibrational state in the bath. Each state of the prediagonlized bath (which diagonlizes the couplings betweenthe latter tiers) must contain significant population in every K value. As such the eigenstatesare predicted to have almost none of the aligriment expected for a rigid symmetric top eigenfunction. The almost constant lifetimes calculated at each J, for the same K, imply that the initial coupling out of the bright state is similar at each J and thus likely anharmonic in nature. The model is further supported by the calculated coupling matrix elements, which decreasewith J’ as expectedfor strong bathbath Coriolis interactions. This strong bath coupling has the effect of diluting the bright to bath vibrational state coupling by the rotational population factor which describes how much population of each eigenstate has the correct K value for an anharmonic interaction (which should obey the AK=0 selection rule). The term “keyhole model” state was introduced by Go et al. to describe such rovibrational dynamics in the spectrum of the aliphatic C-H stretch in trans-ethanol.26In their analogy the bright state is the keyhole which allows one to look into the dynamics in the room behind the door, which in the case of trans-ethanol, like the present case, is a Coriolis or centrifugally mixed bath. Go et al. attributed the different behavior of the bath in ethanol, as compared with butyne where the effective bath density did not systematically increase with J, as due to the presenceof the OH torsional mode. Li, Ezra, and Philips were able to explain the increased density of states (over the calculated) in the spectrum of 24luoroethanol as due to the presenceof low frequency, large amplitude torsional modes inducing the breakdown of rotational selection rules.46 They indicated that this effective K mixing may be expected for asymmetric top molecules when the moment of inertia value and directions are a strong function of the torsion angle. Propyne, however, is a symmetric top with no internal rotational modes and so torsional modes cannot be implicated as the cause of the complete K mixing seen in this spectrum. Analysis of the data support that the spectrum is highly mixed and shows evidence of quantum chaos. At J’ -0, both the level spacing statistics and Heller’s F statistic are consistent with a GOE model which should describe the statistical properties of a chaotic spectrum. The 2621 distribution of coupling matrix elements appear to be GaussianJurther supporting similarity to the GOE model. At higher J, the level spacing statistics do not follow a GOE model. This is similar to the caseof NO2 where it has been establishecl-thatthe vibronic levels follow GOE statistics while the modest J (- 10) rovibrational states do not follow GOE statistics despite the strong breakdown in the spectrum of the K selection rules.“’ A few other molecules have been analyzed using statistical methods searching for the signature of chaos, however the results are often ambiguous. The spectrum of highly vibrationally excited acetylene obtained using SEP, around 15 000 cm-l, has been studied and showed some signs of spectral rigidity indicative of chaos.47Both NO2 and acetylene, even at these very high energies,have relatively small density of states (l/cm-’ and 0.5/cm-‘, respectively). In contrast, the dynamics of larger molecules at relatively low energiesbut higher densities of states have also been analyzed. Evidence was presentedin the form of level spacingsand distribution of coupling matrix elements that the spectrum of the first overtone of the acetylenic C-H stretch of propyne (pcalc= 1l/cm-‘) supports regular behavior22while that of the first overtone of tritlouropropyne (pcalc= 1000/cm-‘)3 shows evidence of chaotic dynamics. The onset of chaotic behavior in propyne, at least for initial excitation in the acetylenic C-H stretch, probably falls somewhere between the 3v1 state which shows signs of chaotic dynamics beginning at J’ -0, where It is interestP exp=67/cm-1, and 219 where p= 1 l/cm-‘. ing to note that this onset coincides with that determined by McDonald of 70 rovibrational states/cm-l for the onset of complex state mixing. 17r1* V. CONCLUSIONS We now have a detailed picture of a single molecule, in which the same vibrational mode is excited at three energy levels, u= 1,2, and 3. At U= 1 where the density of states is in the sparse limit only nonresonant perturbations to the K=O and K= 1 subband origins are observed.‘l A predicted nonresonant perturbation”” was observed by Perry at v=2 -0.1 cm-’ lower in energy from the main clump.23 Evidence of anomalous K=O and 1 behavior is very strong at u=3 as demonstrated by the large splitting, 0.1 cm-‘, between the clumps. While z-axis Coriolis coupling was evident in the coupling mechanismsat both U- 1 and 2, no indications of x,y-Coriolis were present. At v= 3, z-axis Coriolis may possibly be responsible for the large K===O/l splitting, however there is defmitely evidence for x,y-Coriolis mixing among the bath states. It is interesting that this perpendicular coupling occurs, since it was not seen at U= 1 or 2 of propyne nor in many other terminal acetylene molecules that have been studied at U= 1 in this laboratory. We have recently observed the 2~6 band in our laboratory as well as the ( v1+ 2~6) band in double resonance. Given the rich display of dynamics already observed and the rapidly increasing set of bands analyzed by eigenstate resolved methods, this molecule is becoming an excellent J. Chem. Phys., Vol. subject 100, No.to4,AIP 15 license Februaryor1994 Downloaded 18 Mar 2002 to 128.112.83.42. Redistribution copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp 2622 Gambogi et al.: The 3v, overtone band of 1-propyne testing ground for the fundamental theories of intramolecc ular dynamics. ACKNOWLEDGMENTS We are extremely grateful to D. Komanini for building the wavemeter. This work was funded by the NSF. ‘A. M. de Souza, D. Kaur, and D. S. Perry, J. Chem. Phys. 88, 4569 (1988). ‘A. McIlroy and D. J. Nesbitt, J. Chem. Phys. 92, 2229 (1990). 3B. H. Pate, K. K. Lehmann, and G. S&es, J. Chem. Phys. 95, 3891 (1991). 4C. L. Brummel, S. W. Mork, and L. A. Phillips, J. Chem. Phys. 95, 7041 (1991). ‘E. R. Th. Kerstel, K. K. Lehmann, T. F. Mentel, B. H. Pate, and G. Stoles, J. Chem. Phys. 95, 8282 (1991). 6G. A. Bethardy, and D. S. Perry, J. Chem. Phys. 98, 6651 (1993). ’ (a) C. Douketis and J. P. Reilly, J. Chem. Phys. 91, 5239 (1989); (b) 96,343l ( 1992); (c) K. Borass, D. F. DeBoer, Z. Lin, and 3. P. Reilly (preprint). ‘Extensive early work done by this method is summarized in G. Herzberg, Molecular Spectra and Molecular Structure Vol. II, Infrared and Raman Spectm of PoIyatomic Molecules (Van Nostrand, Princeton, 1945). As an example of modern work, see A. M. Smith, S. L. Coy, W. Klemperer, and K. K. Lehmann, J. Mol. Spectrosc. 134, 134 (1989). 9(a) K. V. Reddy, D. F. Heller, and M. 5. Berry, J. Chem. Phys. 76, 2814 (1982); (b) K. K. Lehmann, G. J. Scherer, and W. Klemperer, ibid. 77, 2853 (1982). ‘OF. Stoeckel, M.-A. Melieres, and M. Chenevier, J. Chem. Phys. 76,219l (1982). ‘I (a) D. E. Reissner, R. W. Field, 3. L. Kmsey, and H.-L. Dai, J. Chem. Phys. 78, 2817 (1983); (b) C. Kiltrell, E. Abramson, J. L. Rmsey, S. A. MacDonald, D. E. Reisner, R. W. Field, and D. H. Katayama, ibid. 75, 2056 (1981). t*(a) K. K. Lehmann and S. L. Coy, J. Chem. Phys. 81, 3744 (1984); (b) 83, 3290 (1985). I3 X. Luo and T. R. Rizzo, J. Chem. Phys. 93, 8620 (1990). 14M Scotoni, A. Boschetti, N. Oberhofer, and D. Bassi, J. Chem. Phys. 94,971 (1991). I5 (a) G. J. Scherer, K. K. Lehmann, and W. Klemperer, J. Chem. Phys. 81, 5379 (1984); (b) M. W. Crofton, C. G. Stevens, D. Klenermen, J. H. Gutow, and R. N. Zare, ibid. 89, 7100 (1988). 16G. M. Stewart, and J. D. McDonald, J. Chem. Phys. 78, 3907 (1983). “H L. Kim, T. J. Kulp, and J. D. McDonald, J. Chem. Phys. 87, 4;76 (1987): ‘*T. K. Minton and J. D. McDonald, Ber. Bunsenges. Phys. Chem. 92, 350 (1988). Iv Hydrocarbon Thermal Isomerizations, Organic Chemistry A Series of Monographs, edited by J. J. Gajewski (Academic, New York, 1981), Vol. 44, p. 26. ‘“A..McIlroy, and D. J. Nesbitt, J. Chem. Phys. 91, 104 (1989). ‘I E. R. Th. Kerstel, K. K. Lehmann, B. H. Pate, and G. Stoles J. Chem. Phys. 100, 2588 ‘( 1994). “A McIlroy, D. J. Nesbitt, E. R. Th. Kerstel, K. K. Lehmann, B. H. Pate, and G. Stoles, J. Chem. Phys. 100, 2596 (1994). 23(a) J. Go and D. S. Perry, J. Chem. Phys. 97, 6994 (1992); (b) J. Go, T. J. Cronin, and D. S. Perry, Chem. Phys. 175, 127 (1993). 24G.,Herzberg, F. Patat, and H. Verleger, J. Phys. Chem. 41, 123 (1937). z5R. R. Hall, Ph.D. thesis, Rice University, 1984. ‘6.T. Go, G. A. Bethardy, and D. S. Perry, J. Phys. Chem. 94, 6153 (1990). “H. L. Dai, D. L. Korpa, J. L. Kinsey, and R. W. Field, J. Chem. Phys. 82, 1688 (1985). 28K. K. Lehmann and S. L. Coy, J. Chem. Sot. Faraday Trans. 2 84, 1389 (1988). 29T. A. Brody, J. Flores, J. B. French, P. A. Mello, A. Pandey, and S. S. M. Wang, Rev. Mod. Phys. 53, 385 (1981). 30M. L. Mehta, Random Matrices and the Statistical Theoty of Energy Levels (Academic, New York, 1967). 3’E. J. Heller, Faraday Discuss. Chem. Sot. 75, 141 ( 1983). “(a) 0. Bohiges and M. J. Giannoni, in Mathematical and Computational Methods in Nuclear Physics, Lecture Notes in Physics 209, edited by J. S. Dehesa et al. (Springer, Berlin, 1984); (b) G. Persch, E. Mehdizadeh, W. Demtroder, Th. Zimmermann, H. Koppel, and L. S. Cederbaum, Ber. Bunsenges. Phys. Chem. 92, 312 (1988). 33R. D. Levine, Ber. Bunsengea. Phys. Chem. 92, 222 (1988). 34The wavemeter was built by D. Romanini based on a design by H. Lew, N. Marmet, M. D. Marshall, A. R. McKellar, and G. W. Nichols, Appl. Phys. B 42, 5 (1987). 35W. J. Lafferty and A. S. Pine, J. Mol. Spectrosc. 141, 223 (1990). 36K. F. Freed, and A. Nitzan, J. Chem. Phys. 73, 4765 (1980). 37K. K. Lehmann, J. Phys. Chem. 95, 7556 (1991). ‘*W. D. Lawrance and A. E. W. Knight, J. Phys. Chem. 89,917 (1985). “M. V. Berry and M. Tabor, Proc. R. Sot. London Ser. A 356, 375 (1977). 4oE. P. Wigner, Phys. Rev. 48, 749 (1932). 4’K. K. Lehmann and S. L. Coy, Ber. Bunsenges. Phys. Chem. 92, 306 (1988). 4zS. L. Coy (private communication). 43K. K. Lehmann, J. Chem. Phys. 96,7402 (1992). 44E. M. Carresquillo, A. L. Utz, and F. F. Crim, J. Chem. Phys. 88, 5976 (19881, 45E. B. Stechel and E. J. Heller, Annu. Rev. Phys. Chem. 35,563 (1984). 46H. Li, G. S. Ezra, and L. A. Philips, J. Phys. Chem. 97, 5956 ( 1992). 47D. M. Jonas, Ph.D. thesis, MIT, 1992. 48K. K. Lehmann and S. L. Coy, J. Chem. Phys. 87, 5415 (1987). J. Chem. Phys., Vol. 100, No. 4, 15 February 1994 Downloaded 18 Mar 2002 to 128.112.83.42. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp I ;