18
IF photodissociation and recombination dynamics in size-selected Ir(COz)n cluster ions John M. Papanikolas, Vasil Vorsa, Mat-la E. Nadal, Paul J. Campagnola, Heinrich K. Buchenau, and W. C. Lineberger Department of Chemistry and Biochemistry and Joint Institute for Laboratory Astrophysics, University of Colorado and National Institute of Standards and Technology Boulder, Colorado 80309 (Received 14 June 1993; accepted2 August 1993) Pump-probe techniquesare used in conjunction with a tandem time-of-flight massspectrometer to investigate the I* * *I- cage recombination dynamics following IF photodissociation in size- selected11 (CO,), cluster ions. The absorption recovery, which reflects the recombination and vibrational relaxation of the photodissociated IT, exhibits a strong cluster size dependence in the range of n= 13-15. Over this limited cluster size range, the absorption recovery time decreases from -40 ps (n<12) to - 10 ps (n) 15). In addition, a recurrence is observed at ~2 ps in the absorption recovery of the larger clusters (II = 14-17). This feature results from coherent 1. * *I- motion following photodissociation. Measurementof the absorption recovery with both parallel and perpendicular pump-probe polarizations demonstrates that the pump and probe transition dipoles lie in the same direction. Analysis of the 1, transition dipole directions shows that the coherent motion takes place on the first two repulsive excited potential surfaces. The two-photon photofragment distribution reflects the solvent cagestructure as a function of pump-probe delay time. I. INTRODUCTION Molecular clusters offer a unique environment in which to study the effectsof increasing solvation on chem- ical systems.1-3 Many of these experiments have utilized laser based techniques both to elucidate “static” cluster properties, such as the geometryk7 and electronic struc- ture,7-‘3 as well as in the study of numerous photoinduced processes.‘3-‘9 In recent years, several groups have ex- tended these techniques to the ultrafast regime in order to investigate cluster fragmentation,“20-22 the competition be- tween fragmentation and ionization,23324 and photoinitiated chemical reactions.‘825-33 Our efforts have focused on using both photofragmentation14”5 and ultrafast pump-probe techniques34’35 to investigate the X; (X=Br, I) photodis- sociation and cagerecombination dynamics in size-selected XT ( C02) n clusters. In a recent communication35 we pre- sented experimental results that demonstrated the exist- ence of coherent I-*-I- motion taking place following the photodissociation of I; within the cluster. This observa- tion provides a detailed picture of the 11 internuclear mo- tion in the fnst few picosecondsafter photodissociation. Contained in this paper is a more detailed account of this unique observation. Coherent nuclear motion can be initiated by photoex- citation to a bound excited state by an ultrafast laser pulse, provided that the coherent spectral bandwidth of the laser pulse exceeds the excited state vibrational spacing. This photoexcitation prepares a coherent superposition of ex- cited state vibrational eigenfunctions, i.e., a wave packet, initially localized at the ground state equilibrium geome- try. The motion of this wave packet on the excited state potential surface, which corresponds classically to coherent vibrational motion of the photoexcited ensemble,can be followed using a second, delayed, ultrafast laser pulse. If the probe laser frequency is tuned such that the absorption of a probe photon can occur only from a restricted set of molecular geometries, then sharp peaks in the transient probe absorption may occur; the maxima correspondto the time(s) at which the wave packet passes through the nar- row Franck-Condon region. These recurrences in the probe absorption furnish a detailed picture of the nuclear motion following photoexcitation. Wave packet motion has been observed in this manner in a number of diatomic molecules such as 12,36-39 IC1,40 NaI,4’ and Na2.42 As the pump-probe delay increases, the recurrencesdisappearas a result of processes that act to destroy the nuclear phase coherence. In the gas phase,nuclear dephasingresults pri- marily from anharmonicity in the potential energy surface, thus the coherent motion can persist for many vibrational periods. In I, (Ref. 37) and Na, (Ref. 42), for example, the vibrational coherence lasts for more than 40 ps, corre- sponding to over 100 vibrational periods. In the condensed phase, collisions with the solvent dominate vibrational dephasing and significantly reduce the amount of time that phase-coherence can be maintained. As a consequence of this rapid dephasing,there have been relatively few examples of phase-coherent motion in solvated environments. Scherer et ~1.~~ report multiple I2 vibrational recurrences that last for only a few picoseconds following photoexcitation of I, in n-hexane.In this system, the dephasing time is due predominantly to population loss via solvent induced predissociation. In addition, Vos et aL” have observed oscillations with periods as long as 2 ps in the transient probe absorption of bacterial reaction centers embedded in low temperature glycerol matrices (T < 100 K). These oscillations, which are attributed to coherent vibrational motion on an excited state surface, are only observedat low temperature. A third example is sup- plied by Ruhman and co-workers,45 who have observed several oscillations in the transient absorption of the 11 J. Chem. Phys. 99 (1 I), 1 December 1993 0021-9606/93/99(11)/8733/18/$6.00 @ 1993 American Institute of Physics 8733 Downloaded 04 Oct 2007 to 128.138.107.158. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

I-2 photodissociation and recombination dynamics in size …jila.colorado.edu/~wcl/wclgroup/Pub downloads/I2... · 2007. 10. 4. · separations) and the delocalized, or molecular,

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: I-2 photodissociation and recombination dynamics in size …jila.colorado.edu/~wcl/wclgroup/Pub downloads/I2... · 2007. 10. 4. · separations) and the delocalized, or molecular,

IF photodissociation and recombination dynamics in size-selected Ir(COz)n cluster ions

John M. Papanikolas, Vasil Vorsa, Mat-la E. Nadal, Paul J. Campagnola, Heinrich K. Buchenau, and W. C. Lineberger Department of Chemistry and Biochemistry and Joint Institute for Laboratory Astrophysics, University of Colorado and National Institute of Standards and Technology Boulder, Colorado 80309

(Received 14 June 1993; accepted 2 August 1993)

Pump-probe techniques are used in conjunction with a tandem time-of-flight mass spectrometer to investigate the I* * *I- cage recombination dynamics following IF photodissociation in size- selected 11 (CO,), cluster ions. The absorption recovery, which reflects the recombination and vibrational relaxation of the photodissociated IT, exhibits a strong cluster size dependence in the range of n= 13-15. Over this limited cluster size range, the absorption recovery time decreases from -40 ps (n<12) to - 10 ps (n) 15). In addition, a recurrence is observed at ~2 ps in the absorption recovery of the larger clusters (II = 14-17). This feature results from coherent 1. * *I- motion following photodissociation. Measurement of the absorption recovery with both parallel and perpendicular pump-probe polarizations demonstrates that the pump and probe transition dipoles lie in the same direction. Analysis of the 1, transition dipole directions shows that the coherent motion takes place on the first two repulsive excited potential surfaces. The two-photon photofragment distribution reflects the solvent cage structure as a function of pump-probe delay time.

I. INTRODUCTION

Molecular clusters offer a unique environment in which to study the effects of increasing solvation on chem- ical systems.1-3 Many of these experiments have utilized laser based techniques both to elucidate “static” cluster properties, such as the geometryk7 and electronic struc- ture,7-‘3 as well as in the study of numerous photoinduced processes.‘3-‘9 In recent years, several groups have ex- tended these techniques to the ultrafast regime in order to investigate cluster fragmentation,“20-22 the competition be- tween fragmentation and ionization,23324 and photoinitiated chemical reactions.‘825-33 Our efforts have focused on using both photofragmentation14”5 and ultrafast pump-probe techniques34’35 to investigate the X; (X=Br, I) photodis- sociation and cage recombination dynamics in size-selected XT ( C02) n clusters. In a recent communication35 we pre- sented experimental results that demonstrated the exist- ence of coherent I-*-I- motion taking place following the photodissociation of I; within the cluster. This observa- tion provides a detailed picture of the 11 internuclear mo- tion in the fnst few picoseconds after photodissociation. Contained in this paper is a more detailed account of this unique observation.

Coherent nuclear motion can be initiated by photoex- citation to a bound excited state by an ultrafast laser pulse, provided that the coherent spectral bandwidth of the laser pulse exceeds the excited state vibrational spacing. This photoexcitation prepares a coherent superposition of ex- cited state vibrational eigenfunctions, i.e., a wave packet, initially localized at the ground state equilibrium geome- try. The motion of this wave packet on the excited state potential surface, which corresponds classically to coherent vibrational motion of the photoexcited ensemble, can be followed using a second, delayed, ultrafast laser pulse. If the probe laser frequency is tuned such that the absorption

of a probe photon can occur only from a restricted set of molecular geometries, then sharp peaks in the transient probe absorption may occur; the maxima correspond to the time(s) at which the wave packet passes through the nar- row Franck-Condon region. These recurrences in the probe absorption furnish a detailed picture of the nuclear motion following photoexcitation. Wave packet motion has been observed in this manner in a number of diatomic molecules such as 12,36-39 IC1,40 NaI,4’ and Na2.42 As the pump-probe delay increases, the recurrences disappear as a result of processes that act to destroy the nuclear phase coherence. In the gas phase, nuclear dephasing results pri- marily from anharmonicity in the potential energy surface, thus the coherent motion can persist for many vibrational periods. In I, (Ref. 37) and Na, (Ref. 42), for example, the vibrational coherence lasts for more than 40 ps, corre- sponding to over 100 vibrational periods. In the condensed phase, collisions with the solvent dominate vibrational dephasing and significantly reduce the amount of time that phase-coherence can be maintained.

As a consequence of this rapid dephasing, there have been relatively few examples of phase-coherent motion in solvated environments. Scherer et ~1.~~ report multiple I2 vibrational recurrences that last for only a few picoseconds following photoexcitation of I, in n-hexane. In this system, the dephasing time is due predominantly to population loss via solvent induced predissociation. In addition, Vos et aL” have observed oscillations with periods as long as 2 ps in the transient probe absorption of bacterial reaction centers embedded in low temperature glycerol matrices (T < 100 K). These oscillations, which are attributed to coherent vibrational motion on an excited state surface, are only observed at low temperature. A third example is sup- plied by Ruhman and co-workers,45 who have observed several oscillations in the transient absorption of the 11

J. Chem. Phys. 99 (1 I), 1 December 1993 0021-9606/93/99(11)/8733/18/$6.00 @ 1993 American Institute of Physics 8733 Downloaded 04 Oct 2007 to 128.138.107.158. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 2: I-2 photodissociation and recombination dynamics in size …jila.colorado.edu/~wcl/wclgroup/Pub downloads/I2... · 2007. 10. 4. · separations) and the delocalized, or molecular,

a734 Papanikolas et aL: I; photodissociation and recombination

photofragment produced by the photodissociation of 1, in ethanol. The authors attribute these oscillations, which last for z 1 ps, to coherent IF vibrational motion in a low vibrational energy level of the ground electronic state. This coherence is similar to that observed in HgI following pho- todissociation of isolated IHgI.& In the examples discussed thus far the phase-coherent motion takes place only along a bound coordinate of the potential energy surface.

If the chromophore is excited to a dksociative poten- tial, then the possibility exists for the wave packet to “re- flect” off the solvent cage, resulting in coherent motion along the dissociation coordinate. This type of coherent motion differs markedly from that observed on bound po- tentials, for in this case the forces that arrest and reverse the motion of the photofragments arise entirely from the surrounding solvent. The experiments described here are a definitive example of this qualitatively different type of co- herence.

The photoexcited I; (CO,),, system provides the means to investigate a number of dynamical processes, in- cluding the mechanisms that lead to the recombination of the dissociating I-*-I- (especially in partially solvated clusters), the electronic surface hopping from the repulsive excited state to the 1, ground state, and the energy transfer from the initially “hot” 1, to the CO2 solvent. Further- more, the presence of the net charge introduces long-range electrostatic forces, whose importance has been demon- strated in the vibrational relaxation of both isolated47 and solvated48-” species. A comprehensive understanding of the caging dynamics must incorporate not only the pres- ence of the ionic charge, but also the charge transfer pro- cess that occurs as the 11 passes between the localized I+I- charge distribution (present at large internuclear separations) and the delocalized, or molecular, 11 distri- bution. One consequence of this process is highlighted by molecular dynamics simulations on 1; ( C02) ,, clusters,52 which indicate that there is substantial enhancement of the IF relaxation rate coincident with this change in the ionic charge distribution.

In this paper, we report the results of time-resolved pump-probe experiments on 12 ( C02) n (n = 8-22) cluster ions. The primary experimental observations are as fol- lows. (i) We observe a strong cluster size dependence in the absorption recovery for the cluster sizes in the range of n= 12-15. This abrupt change in absorption recovery be- havior over such a small cluster size range suggests that the IF cage recombination dynamics are qualitatively different in the small (n<12) and large (n)15) size regimes. (ii) There is a “peak” at ~2 ps in the absorption recoveries of the larger clusters (n = 14-22), which is attributed to co- herent I- * +I- vibrational motion following the photoexci- tation of 1~ to a dissociutiue electronic state. (iii) We present pump-probe photofragmentation experiments which suggest that substantial “break up” of the solvent cage occurs in the first 4 ps following photodissociation. Together, these experiments provide insight into, not only the dynamics of the 1, solute, but the CO2 solvent as well.

The organization of this paper is as follows. In Sec. II we summarize the conclusions of previous investigations of

1.0 2iL9~~~31k0a~ Sep2ation, f&O)

6.0

FIG. 1. Scmiempirical 11 potential curves derived by Chen and Went- worth (Ref. 53). Each electronic state is identified with the Hund’s case (c) symmetry label. The term symbol in parenthesis is the symmetry of the corresponding case (a) electronic state. The arrow labelled Ppump indicates the $ s(2fls,In) -4 ,(‘2:) transition used to photodissociate the I,. The arrow labeled Pz indicates the transition assigned to the transient probe absorption at z-2 ps in the 11 (CO,), (n> 14) absorption recoveries.

11 and 1, ( C02), cluster ions. A detailed description of the pump-probe experiment appears in Sec. III, and our results are presented and discussed in Sec. IV. Finally, we review our major findings and present prospects for future experiments in Sec. V.

II. PREVIOUS INVESTIGATIONS OF I,- AND l;(C02), CLUSTER IONS

In our earlier experiments on 11 (C02) n cluster ions we employed nanosecond and picosecond techniques to study the effects of stepwise solvation on the 1, cage re- combination dynamics. Nanosecond photofragmentation experiments probedI “static” aspects of the 11 caging pro- cess, while time-resolved pump-probe techniques were used to investigate34”5 the 1, recombination and vibra- tional relaxation dynamics. A. 1; electronic structure

The 1, gas phase potential energy curves are displayed in Fig. 1. These semiempirical potential curves were deter- mined by Chen and Wentworth’ based on matrix absorp- tion spectra, dissociative electron attachment data, elec- tron scattering, and self-consistent field (SCF ) calculations. The ground state potential is characterized by a fundamental vibrational frequency54 and bond dissocia- tion energy54’55 of - 115 cm-’ and - 1.1 eV, respectively. There is a large quantitative uncertainty in these potential curves, especially at internuclear separations significantly different than r=re.56 Nevertheless, they provide useful, qualitative information concerning the electronic structure of IF. The arrow labelled PpUmp in Fig. 1 indicates the 4 J 2111,1,2) -4 ,( ‘2: ) transition employed to photodisso- ciate the 1, within the cluster. Photoexcitation of this tran- sition at 720 nm (the wavelength used in these experi- ments) produces I and I- with a kinetic energy release of ~0.5 eV,

J. Chem. Phys., Vol. 99, No. 11, 1 December 1993 Downloaded 04 Oct 2007 to 128.138.107.158. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 3: I-2 photodissociation and recombination dynamics in size …jila.colorado.edu/~wcl/wclgroup/Pub downloads/I2... · 2007. 10. 4. · separations) and the delocalized, or molecular,

Because of the extensive spin-orbit mixing that occurs inI;, these potential curves are labeled with the appropri- ate Hund’s case (c) symmetry labels. Spin-orbit coupling in 1; alters both the magnitude and direction of the elec- tronic transition dipole moments. As we will see later, the determination of the transition dipole directions in 11 plays a central role in assigning the probe transition at early pump-probe delay times. This point is addressed in more detail in Appendix B.

One concern in these experiments is the effect of CO2 solvation on the I; electronic structure, which can be ad- dressed by comparing the isolated 11 and 1; ( COz), ab- sorption spectra. The absorption spectrum for the $ J211s,tn) -f J2Zz) transition in isolated 11 is a broad bandI peaking at 750 nm. The I; ( COZ), cross sections at 720 nm (Ref. 14) do not depend strongly upon cluster size. From a purely pragmatic standpoint, this result indicates that the CO2 solvent evaporation does not obscure the IF dynamics in the absorption recovery as probed through this transition. Additional spectroscopic measurements are required, however, before the effect of solvation on the electronic structure of I; is understood in detail.

B. I;(CO,), photofragmentation experiments

Photodissociation of I; inside the cluster results in either the escape of an I atom, or solvent induced recom- bination of the dissociating I-*-I-. These two product channels lead to the formation of I--based (uncaged) and IF-based (caged) ionic photofragments, respectively. The I; recombination probability increases smoothly from 0 at 1; ( C02)s to unity when the first solvation shell is com- pleted’* at 1; ( C02) i6; for n> 16 IF recombination occurs with 100% efficiency.

Photoexcited cluster ions cool by the evaporation of CO2 monomers from the cluster. The total number of CO, molecules ejected depends upon the amount of energy de- posited into the cluster, the binding energy of the COZ to the cluster, and the kinetic energy of the departing mono- mers. Photofragmentation experiments on 11 (CO,), clus- ter ionsI indicate that each CO, monomer evaporated re- moves ~2250 meV of energy from the cluster. For caged photofragments all of the photon energy is available for CO, evaporation and as a result ~7 COz molecules are evaporated following the absorption of a 720 nm photon. For uncaged fragments the 11 bond dissociation energy is not available to the cluster, and only 3-4 CO2 molecules are lost.

The fact that I- * *I- recombination is always observed in the larger clusters suggests that the IF is embedded within the CO2 solvent cage. This is consistent with the 11 ( C02), cluster ion minimum energy structures deter- mined by both molecular dynamics5’ (MD) and Monte Carlo14 techniques. These simulations also indicate that the strong electrostatic interactions present in this system re- sult in rigid geometries at internal temperatures below - 100 K. Since the internal temperature of the 11 ( C02), clusters is estimated to be in the range of 20-60 K,14 they are best thought of as initially being “solid.”

Papanikolas et a/.: I; photodissociation and recombination 8735

-> ChA:ltrO.

FIG. 2. Schematic of the cluster ion source and tandem time-of-flight mass spectrometer.

C. I;(COP)n time-resolved experiments Early pump-probe experiments34 (6 ps resolution) on

11 (C0219 and IF (CO,) 16 measured the absorption recov- ery time at 720 nm to be - 10 ps for 12 (CO,) i6 and - 30 ps for 11 (CO*) 9, indicating a strong cluster size depen- dence on the absorption recovery. Subsequent improve- ments in our experimental apparatus provide data which allow a more detailed analysis. In a recent communica- tion3’ we reported the observation of a “peak” at ~2 ps in the 720 nm absorption recoveries for clusters with n) 14, a feature that is attributed to coherent 1-e -I- “vibrational” motion following 11 photodissociation. In this paper we present the detailed arguments which lead to this conclu- sion. This work is complemented by the experiments of Barbara and co-worker?’ who have recently performed pump-probe experiments on IF emersed in several polar solvents. Data obtained for 1; solvated in H,O, methanol, ethanol, and propanol also exhibit the rapid recovery (5-10 ps). In addition, the IF/ethanol transient data ob- tained with jlpump= 770 nm and jlprobe= 700-740 nm ex- hibit a “shelf” at -2 ps. It is nuclear whether this “shelf” results from the same dynamical process that gives rise to the “peak” in the 12 ( C02) n absorption recoveries or some other process.

III. EXPERIMENTAL

The 11 (COz), caging dynamics are investigated by coupling the cluster ion source and tandem time-of-flight (TOF) mass spectrometer used in our previous experi- ments with an ultrafast laser system. Since a detailed de- scription of the cluster ion apparatus has appeared else- where,2*14>‘5 only a brief description will be given here. A detailed description of the ultrafast laser system and pump-probe experimental procedure follows.

A schematic diagram of the cluster ion apparatus is shown in Fig. 2. IF (CO*) ,, cluster ions are created in the source upon crossing a pulsed supersonic expansion of CO, ( - 1% I,) with a 1 keV continous electron beam. As the unskimmed expansion drifts, cluster ion growth occurs via CO, condensation around the ion cores. Approximately 20 cm downstream from the nozzle orifice a pulsed electric field extracts the ions present in the expansion, without significant collisional excitation, into a tandem TOF mass spectrometer. At the spatial focus (SF1 ) of the first TOF

J. Chem. Phys., Vol. 99, No. 11, 1 December 1993 Downloaded 04 Oct 2007 to 128.138.107.158. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 4: I-2 photodissociation and recombination dynamics in size …jila.colorado.edu/~wcl/wclgroup/Pub downloads/I2... · 2007. 10. 4. · separations) and the delocalized, or molecular,

8736 Papanikolas et a/.: I; photodissociation and recombination

( i . . . . . . . . . --, j , A, j Optic-1 Del*g

j Line

A i 10, \w v i Laser Interaction

Region

mass spectrometer the ion sample is intersected with a pulsed laser. Adjustment of the time between the extrac- tion pulse and the laser pulse provides for mass selective photoexcitation. A pulsed mass gate placed before the laser interaction region reduces the amount of noise on the par- ticle detectors by deflecting all of the ions, except the one mass of interest, from the beam path. Approximately 10 ps after laser interaction, the parent ions, along with the ionic and neutral photodissociation products, enter a single field reflectron TOF mass analyzer. Adjustment of the electric field within the reflectron allows for the refocusing of either parent ions or photofragment ions at the second spatial focus (SF2) located at the microchannel plate (MCP) de- tector designated Dl. The front of the MCP detector is biased such that it provides the ions with an additional z 3 keV of kinetic energy prior to striking the, detector, thus increasing the MCP detection efficiency. At low bias volt- ages (1200 V) the MCP operates in an analog mode and can be used to measure the entire parent ion beam. If the bias voltage is increased to 2000 V, then the MCP can detect single ions, and thus can be used in a particle count- ing mode. When the ionic photofragments are refocused at Dl, the reflectron field is insufficient for the reversal of the parent ion trajectory. As a result, the parent ions, along with any neutral photodissociation products, traverse the reflectron and strike the particle multiplier designated D2. A. Ultrafast laser system

A schematic diagram of the ultrafast laser system used in these experiments is shown in Fig. 3. The 76 MHz out- put of a Quantronix 416 mode-locked Nd:YAG laser was frequency doubled in KTP producing laser pulses at 532 nm with a 70 ps pulsewidth. This 532 nm pulse train (15 nJ/pulse) was then used to synchronously pump a Coher- ent 702 dye laser whose output was cavity dumped at 3.8 MHz in order to obtain higher pulse energies. Wavelength selection was achieved with a two-plate birefringent filter. This two-jet laser, operated with pyridine 2 (LDS722) in the gain jet and a saturable absorber dye (DDI) in the

PIG. 3. Schematic of the picosecond laser system. Picosecond laser pulses at 720 nm are generated using a synchronously pumped dye laser, and amplified to =: 1 mJ of energy and z-1 ps duration. Identical pump and probe pulses are created by split- ting the 720 rmr pulse with a 50/50 beam- splitter. The two pulses traverse dierent paths to the mass spectrometer where they are intersected with the ion beam. An op- tical delay line sets the pulse timing and a half-wave plate allows the adjustment of the angle between the pump and probe po- lariz.ation vectors.

second jet, produced 720 nm laser pulses with an energy content of z 10 nJ. The temporal profile of the dye laser output was monitored continuously by directing the 10% reflection off the surface of a pellicle into an Inrad 5-14 quasi-real-time autocorrelator. The 720 nm laser pulse is characterized by an autocorrelation FWHM of z 1.5 ps and a spectral bandwidth of 17 cm-‘, implying that the temporal width of the laser pulse (- 1 ps) is ~20% greater than the transform limit.

Amplification of the Coherent 702 dye laser output was accomplished using a Spectra Physics PDA pulsed dye amplifier pumped at 30 Hz with the second harmonic out- put (330 mJ/pulse) of a Quanta Ray GCR3 Q-switched Nd:YAG laser. Four amplification stages each using LD700 as its gain medium were employed; the first three were contained within the commercial PDA, and the fourth was added externally. The tlrst two cells were side- pumped, each receiving 8% of the initial pump beam en- ergy, and the last two cells were end-pumped with 42% of the total pump beam energy. Amplified spontaneous emis- sion (ASE) was minimized by spatial lilters placed after the first two amplification stages and a saturable filter (RG715) placed after the fourth stage. Typical amplified pulse energies were - 1 mJ at 720 nm, with ASE levels limited to less than 0.5% of the total pulse energy.

Pump and probe pulses were generated by splitting the PDA output into two identical pulses using a 50% beam- splitter. The two beams traveled along different paths to the mass spectrometer where they were focused to a spot size of 5-7 mm diameter and crossed at an angle of ~5” at the laser interaction region. Adjustment of the pump- probe delay was accomplished with the use of a computer controlled optical delay line. A micrometer gauge with - 5 pm resolution was used to measure the distance traveled by the stage, and thus determines the pump-probe delay to within - 15 fs. The optical delay line has a minimum step size of -50 fs. A half-wave plate placed in one beam al-

J. Chem. Phys., Vol. 99, No. 11, 1 December 1993 Downloaded 04 Oct 2007 to 128.138.107.158. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 5: I-2 photodissociation and recombination dynamics in size …jila.colorado.edu/~wcl/wclgroup/Pub downloads/I2... · 2007. 10. 4. · separations) and the delocalized, or molecular,

1.1 h

.t: ; 0.9

3

2 0.7

[ 0.5

x a 0.3

g 0 aJ 0.1

VI

-0.1 -3 -2 Pu-Lip-ProL!

De& 2 3

(ps)

FIG. 4. Autocorrelation trace of the amplified picosecond and femto- second laser pulses obtained using a noncollinear SHG autocorrelation technique. The solid line is the autocorrelation of a laser pulse with a se& temporal profile. The F W H M of the picosecond autocolrelation is - 1.1 ps, while that of the femtosecond aytocorrelation is -280 fs.

lows the adjustment of the angle between the pump and probe polarization vectors.

The temporal proflle of the amp lified picosecond laser pulse was determined using a noncoll inear autocorrelation technique. In this measurement, z 1% of the amp lified pump and probe pulses were focused into a 1 m m thick KDP crystal cut at 48.33 and the intensity of the second harmonic light generated when the pump and probe pulses over lapped spatially and temporally inside the crystal was detected by a photomultiplier tube. The width of the laser pulse autocorrelation provides a direct measurement of the pump-probe instrument response time. A typical autocor- relation trace of the amp lified picosecond laser pulse is shown in F ig. 4 (solid circles). The single points in the figure represent the second harmonic intensity obtained as a function of pump-probe delay time. The solid curve through the data points is the autocorrelation function for a laser pulse with a sech2 temporal profile. This autocor- relation has a full width at half maximum (FWHM) of 1.1 ps. Continuous mon itoring of the amp lified laser pulse au- tocorrelation is not possible with this experimental config- uration. However, based on laser pulse autocorrelations obtained during the collection of this data, we lim it the instrument response time to less than 1.25 ps FWHM.

For several of the measurements presented here, the output of the Coherent dye laser was compressed using standard pulse compression techniques.58-M) The dye laser output was coupled into a 40 cm long, 3.6 pm diameter, single-mode, polarization preserving optical fiber. The chirped pulses that emerged from the fiber were ~4.5 ps in duration, with a spectral bandwidth of ~280 cm-‘. Com- pression of the chirped pulse, which was achieved by dou- ble passing the fiber output through a SF10 prism pair spaced 152 cm apart, yielded a nearly transform-limited laser pulse ~80 fs in duration with an energy content of ~6 nJ. The compressed pulse was amp lified to -600 pJ/ pulse using the PDA described previously with two mod- ifications. A telescope and optical isolator were inserted

before the PDA to collimate the beam, and to prevent reflected light and backward traveling ASE from damaging the fiber, respectively. The autocorrelation of the amp lified femtosecond laser pulse (open circles) is shown along with the picosecond autocorrelation in F ig. 4. This autocorrela- tion has a FWHM of 280 fs, representing an improvement in the instrument response time by a factor of 4-5.

B. Description of pump-probe experiment

The manner in which this apparatus is used in a pump- probe experiment is as follows. A size-selected IF (CO,), cluster ion absorbs a 720 nm photon from an ultrafast pump pulse, dissociating the 1, inside the cluster. Imme- diately following the photodissociation of IF, the cluster ion is transparent to 720 nm radiation. On ly when the 11 has reached a region of the potential energy surface with an appreciable absorption cross section at 720 nm can the photoexcited cluster ion absorb a second photon from a probe pulse of the same color. The most obvious region where such a transition can take place corresponds to vi- brationally relaxed IT. Absorption of two 720 mn photons results in a narrow fragment distribution at masses which correspond to the evaporation of as many as 12-13 CO2 mo lecules. Since the absorption of a single photon by the cluster results in the loss of only 7 CO2 mo lecules, the two-photon photofragments are easily separated from the one-photon photoproducts by the secondary mass ana- lyzer. The absorption recovery is mapped out by measuring the total number of two-photon products generated as a function of delay between the pump and probe pulses. The fact that we observe the appearance of these two-photon product ions against a nearly zero background is essential to the success of this experiment, making it possible to perform experiments on mass-selected samples that con- tain only lo3 to lo4 ions.

Absorption recovery data for IF ( C02), , n = 8, 9, and 12-17 are displayed in F igs. 5 and 6. Near zero pump- probe delay the probe absorption is weak. As the delay is increased the amount of probe absorption also increases, reaching an asymptotic lim it at long pump-probe delay times. The symmetry in the absorption recovery about zero pump-probe delay is a consequence of the fact that the pump and probe pulses have the same wavelength, and that we detect the absorption of two photons via the appearance of photoproducts. This symmetry provides a basic test of the experimental apparatus.

The transient bleach created by the pump pulse is mon itored by measuring the number of two-photon ionic photofragments generated as a function of pump-probe delay. To quantify the dependence of the two-photon ion yield on pump-probe delay time, we define the following two-photon cross section, a2h,,( At) :

’ dh’(At) %hv(ht)=f iz, ;par+? G (1)

i I where n?““(At) is the number of two-photon product ions observed at pump-probe delay At, Np and @ i are the number of parent ions and’laser flux measured on the ith

Papanikolas et a/.: I; photodissociation and recombination 8737

J. Chem. Phys., Vol. 99, No. 11, 1 December 1993 Downloaded 04 Oct 2007 to 128.138.107.158. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 6: I-2 photodissociation and recombination dynamics in size …jila.colorado.edu/~wcl/wclgroup/Pub downloads/I2... · 2007. 10. 4. · separations) and the delocalized, or molecular,

8738 Papanikolas et a/.: I; photodissociation and recombination

1.0

8 Z&O.6 k $ 0.6

$ E 0.4 a

0.2

0.0 ., -10 -5 0 PLnp’-oPrdL &ay25(ps;

35 40

FIG. 5. I,(CO,), (n=8, 9, 12, and 13) absorption recovery data ob- tained with 720 nm pump and probe pulses with parallel polarization. Since the pump and probe pulses are identical, these curves should be symmetric about zero ps delay. The smoothing algorithm used to remove high frequency noise from the data is described in the text.

laser pulse, respectively, and J is the total number of laser shots over which the cross section is averaged, for the ex- periments discussed here J=500. Included in G are a number of (presumed slowly varying) instrumental geom- etry factors, such as the laser-ion beam spatial overlap, the homogeneity of the laser beam profile, and the detection efficiencies of the particle multipliers. In this experiment the fluxes of the pump and probe beams are nearly identi- cal, thus maximizing the two-photon product signal.

At long pump-probe delay times, where the two- photon signal is largest, the production of a single two- photon ion requires interaction of the pump and probe pulses with - 50 000 parent ions: only one two-photon ion is produced for every five laser shots ( 103-lo4 parent ions per valve opening). In this low signal limit, rather than

G(C02L

“....o==*= .-a*.

. Xl4 d

7

I $ $6 ,G e0

‘r /.. an.m=~**-~

5 \. J

g r

. .

l . . . l l

-g 0 c J.” n=15 .

7% ,: I ; 2;--” 1

110 Lnp-Kobe %lay 3;ps)

40

FIG. 6. IT(COz)n (n=14-17) absorption recovery data obtained with 720 nm pump and probe pulses with parallel polarization.

1 40 ps Pump-Probe Delay

1 i

G(CWk I’ I I I 51 I 17cfM, 4’ I I I 61 I

I 1 0 I I I 1 I I I 21 22 23 24 25 26 27 26 29 30 31

Flight Time (ps)

FIG. 7. Two-photon photofragment distribution of 1, (CO*) r6 observed following the sequential absorption of two 720 nm photons separated by 40 ps. This mass spectrum was obtained by averaging the two-photon signal over 3000 laser shots.

“measuring” the two-photon cross section on every laser shot, it is better to sum the two-photon product signal and remaining normalization terms (with larger signals) over a large number of laser shots. Thus, we approximate the two-photon cross section in Eq. ( 1) as

Z;FI n;*“( At) Nd At) u2hv(At) =: 2;zI NT”@; G=2;F, N;“@; G9 (2)

where N2hY( At) is the total number of two-photon product ions observed during 500 laser shots at pump-probe delay At.

Shown in Fig. 7 is the two-photon photofragment mass spectrum observed after the sequential absorption of two 720 nm photons (At=40 ps) by 1~ ( C02) i6. This frag- ment distribution is obtained by adjusting the reflectron electric field to refocus the two-photon product ions at Dl. The two-photon products for 1~ (CO,) 16 arrive in four in- tense “peaks” spread out over ~5 ys. Each “peak” is ac- tually a doublet consisting of an IF ( C02) n caged fragment and an I-(C&)n+3 uncaged fragment. The other cluster sizes exhibit product distributions that are qualitatively similar, but which occur within a different mass range. In the absorption recovery experiment, the two-photon prod- uct signal, N2h,,( At), is determined by counting the total number of ions that arrive within the four time windows indicated in Fig. 7 by brackets. The manner in which the two-photon signal is determined from experimental mea- surements depends on the magnitude and physical origin of several sources of “noise” counts in the two-photon prod- uct channel. Appendix A is presented for those readers who wish to know precisely the details surrounding the measurement of this two-photon cross section, including the nature and magnitude of various background correc- tions to the two-photon signal.

In principle, one could map out the absorption recov- ery by measuring 022h,,(At) at a number of different delay times. However, the presence of the geometrical factor, G, in Eqs. ( 1) and (21, which depends sensitively on a num-

J. Chem. Phys., Vol. 99, No. 11, 1 December 1993 Downloaded 04 Oct 2007 to 128.138.107.158. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 7: I-2 photodissociation and recombination dynamics in size …jila.colorado.edu/~wcl/wclgroup/Pub downloads/I2... · 2007. 10. 4. · separations) and the delocalized, or molecular,

ber of experimental conditions that are not easily con- trolled, makes it difficult to obtain reproducible measure- ments of azn,,(At) over extended periods of time. It is possible, however, to make reproducible measurements by the repeated normalization of CQ,,( At) to the cross section obtained at a long delay time At,,, e.g.,

a,.JAt) = Oihv(At) aid be,) ’

able to absorb a second 720 nm photon. As the pump- probe delay is increased, more time is allowed for this ensemble to populate regions of the potential surface where a second 720 nm photon can be absorbed, resulting in the increase in probe absorption. The most obvious region where a second photon can be absorbed corresponds to vibrationally relaxed I,, and thus the simplest interpreta-

(3) tion of the absorption recovery is that it reflects the relax- ation of the photoexcited 1, ensemble to low vibrational levels of the ground electronic state. Earlier data have shown that at 40 ps the absorption re-

covery is essentially complete, and for the data presented here At, was set to this delay time. Since G varies slowly with time this normalization procedure provides reproduc- ible measurements of a,i(At). By definition the cross sec- tion at 40 ps has a value of 1.0.

Two tlnal operations were performed on these data. Fit, the absorption recoveries for n = 15 and n = 16 were adjusted to take into account the two-photon ion signal that was not included in the four ion “peaks” monitored in the measurement of N2hY( At) (see Fig. 7). The required correction was estimated from the two-photon photofrag- ment mass spectra obtained at a number of different pump-probe delay times. This correction corresponds to an approximately 10% modification in Q( At) for pump- probe delays near 0 ps, decreasing to zero at ~5 ps. Cor- rections to the absorption recoveries of the other cluster ions were not necessary. Finally, the absorption recovery data obtained with the picosecond laser pulse were smoothed. Near zero delay the data points were separated by 0.5 ps, corresponding to & of our time resolution. Hence, in this region a three-point filter ( 1:2:1 weights) was employed. At long pump-probe delay times a five- point running average was used to remove high frequency noise from the absorption recovery. Unlike the picosecond data, the femtosecond data were not smoothed.

IV. RESULTS AND DISCUSSION

These experiments investigate the dynamics of both the 11 solute and CO2 solvent cage following the photodisso- ciation of 1~ within an Ic( C02) n cluster. The 1~ absorp- tion recovery discussed in Sec. IV A probes the recombi- nation, vibrational relaxation, and rotational motion of the 11 chromophore following its photodissociation within the cluster. Section IV B addresses a complementary aspect of the cluster caging dynamics-the destruction of the solvent cage that occurs simultaneously with the 11 relaxation. The time dependence of this process is inferred from the evolution of the two-photon ionic photofragment distribu- tion with pump-probe delay. A. 1; dynamics-absorption recovery experiments

The absorption recoveries for 11 ( C02) ,, (n = 8, 9, and 12-17) clusters are shown in Figs. 5 and 6. In addition, we have obtained the absorption recovery for 11 (CO,),, and found it to be similar to those observed for n= 16 and 17. At short pump-probe delay times there is essentially no probe absorption due to the fact that immediately follow- ing photodissociation the photoexcited 11 ensemble is un-

Papanikolas ef a/.: I; photodissociation and recombination 8739

Several observations, however, indicate that the 1. * *I- dynamics following photodissociation are much more com- plicated. First, a “peak” is observed at ~2 ps in the ab- sorption recoveries of the larger clusters (n) 14). Its pres- ence is an indication that the absorption recovery is not soZeZy due to the formation of vibrationally relaxed 11. Much of the discussion below focuses on the physical ori- gin of this “peak.” It is ultimately attributed to coherent I** .I- internuclear motion that is initiated on the f g(*&/*) excited state, survives an electronic surface hopping, and continues on the 5 ,J211g,s,z) excited state. Second, the absorption recovery can not be fit by a single exponential, indicating the presence of multiple time scales in the 11 relaxation process. Finally, the absorption recov- ery depends strongly on cluster size in the range n= 13-15. This pronounced cluster size dependence suggests that the 12 relaxation dynamics are qualitatively different in the small (n< 13) and large (n> 14) cluster size regimes.

The data shown in Figs. 5 and 6 were obtained with parallel pump and probe laser polarizations, and thus re- flect 11 rotational motion as well as the I* * *I- motion on the internuclear potential surface. Photodissociation of the cluster ion ensemble with a plane polarized pump pulse results in an anisotropic distribution of the probe transition dipoles. As a consequence, the magnitude of the two- photon cross section depends on whether the ensemble is probed with light polarized parallel or perpendicular to that of the pump pulse. We can thus obtain two cross sections, oII (At) and al (At); the difference between them reflects the degree of anisotropy in the probe absorption. Several processes can lead to the decay of this anisotropy, including simple cluster rotation. In order to ascertain the relative contribution of the internuclear and rotational mo- tions to the absorption recovery we have performed addi- tional experiments with perpendicular pump-probe polar- izations for n = 12, 14, and 16.

Shown in Fig. 8 are olI (At) and al (At) for 11 (CO*) i6 and 11 ( C02) 12. At early pump-probe delay times there is a substantial anisotropy in the 11 (CO,) i6 probe absorption, which vanishes after approximately 15 ps. The decay of the probe anisotropy occurs on a time- scale similar to that of the absorption recovery, signifying the presence of substantial orientational effects in cl1 (At) and al (At). The presence of the “bump” in both the par- allel and perpendicular absorption recoveries indicates that it is nof an orientational effect. The decay in the probe absorption anisotropy for n= 12 and n= 14 is faster than that observed for n = 16. For 11 ( C02) 14 [a1 (At) is not shown], there is a significant anisotropy at early times

J. Chem. Phys., Vol. 99, No. 11, 1 December 1993 Downloaded 04 Oct 2007 to 128.138.107.158. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 8: I-2 photodissociation and recombination dynamics in size …jila.colorado.edu/~wcl/wclgroup/Pub downloads/I2... · 2007. 10. 4. · separations) and the delocalized, or molecular,

8740 Papanikolas et a/.: I; photodissociation and recombination

G(CO2)lS 1.0 - e •"~Oo-o

z l q* Q .eme

go.0 . 0 - .O g . . 0 2 0.6 - . l .O

. .T?O d .

2 0.4 - a 3

< ’ II 0.2 -

0.0

<O L . && I. l ,m.. ,”

FtqTT O ’ (a)

I

. 1.0

2 50.8

8 3 0.6 I G(CO2)12

Q a- m%.*’

abc?z

‘:;?oo~~~%~~ , , :! ~b) -10 -5 0

Am,'"Pr,'~e &ay2~ps) 30 35 40

FIG. 8. Absorption recoveries for (a) I,(CO,),, and (b) IF (C02)12 obtained with perpendicular polarizations (open circles). These absorp- tion recovery curves are superimposed on the parallel polarization data (solid circles) reproduced from Figs. 5 and 6.

which disappears within - 6 ps. In contrast to 1, ( COZ) 16 and Ii-(c& 011 ( At), and al (At) for n = 12 show no detectable anisotropy at early pump-probe delay times. However, the small two-photon signal for At < 10 ps makes the definitive detection of any anisotropy for this cluster very difficult.

Each of these observations is discussed in detail in the following subsections. In Sec. IV A 1, the rotational and internuclear contributions to the 1, ( C02) 16 absorption re- covery are separated by transforming cl1 (At) and o, (At) for this cluster into a “magic angle absorption recovery” and a “probe anisotropy decay”; the former reflecting only the 1.. *I- motion on the internuclear potential surface; the later reflecting the 1, rotational motion. The discussion of the 1.. .I- internuclear motion, contained in Sets. IV A 2 and IV A 3, is divided into the short time (At<3 ps) and the long time (At)3 ps) dynamics. The cluster size depen- dence of the absorption recovery is examined in Sec. IV A 4. Finally, the 12 rotational motion is addressed in Sec. IV A 5.

1. Separation of the internuclear and rotational contributions to the absorption recovery

Orientational effects can be removed from the absorp- tion recovery by measuring the two-photon cross sections with the angle between the pump and probe laser polariza- tions set to the magic angle ( 54.71”).61d3 This absorption recovery is denoted bus and can be either measured directly or calculated from 011 (At) and al (At) :

1.0

g 0 3 0.8 “Ml? 4x p 0.8 2 4 2a ZIP & 0.4

a

0.2

l ..e.* . . . . . l

. .

(4

z x

_ 0.3

2

l 0.2

..-I

d 0.1

x z a 0.0

1 . -0.1 I I , I 1 I I I 0 5 10 Pu&Prcze

tkflty (,"s", 35 40

FIG. 9. (a) The “magic angle” absorption recovery for Ij-(COz),s de- termined from (~1, (t) and a, (t) using Eq. (4). This absorption recovery reflects the I...I- motion on the internuclear potential surface free of orientational effects. (b) Anisotropy [EQ. (5)] in the probe absorption for IF (C02)16 (individual points) displayed as a function of pump-probe delay. The solid line is the probe anisotropy determined by modeling the rotational motion of the cluster (see text).

a,,(AO=i tall (Af)+20, (At)]. (4) The IF (CO,) 16 magic angle absorption recovery deter- mined from oil (At) and a, (At) is displayed in Fig. 9 (a). The presence of the “peak” in the magic angle configura- tion establishes that this feature is not an orientational effect. The physical origin of this feature is discussed in depth in Sec. IV A 2.

The degree of orientational anisotropy in the total probe absorption can be quantified by calculating an effec- tive anisotropy parameter, r(At),61P64 defined in terms of oII (At) and al (At) as follows:

r( At) = qI (W--o, (AtI olI (At)+201 (At) ’ (5)

The aI’dSOtrOpy in the probe absorption for 11 (c02)16 [Fig. 9(b)] is observed to exceed 0.3 at delay times less than 3 ps and decays smoothly to zero after approximately 15 ps. The value observed for r( At) depends not only on the degree of anisotropy present, but also on the relative orientation of the pump and probe transition dipoles with respect to the internuclear axis. If the pump and probe transition dipoles lie in the same direction, then the limit- ing value for r( At) is 0.4, whereas if they are perpendicular to each other then the limiting value is -0.2 (Ref. 61). Based upon the fact that r( At) approaches its maximum of

J. Chem. Phys., Vol. 99, No. 11, 1 December 1993

Downloaded 04 Oct 2007 to 128.138.107.158. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 9: I-2 photodissociation and recombination dynamics in size …jila.colorado.edu/~wcl/wclgroup/Pub downloads/I2... · 2007. 10. 4. · separations) and the delocalized, or molecular,

I;wM16 . . 00

. 0 0 -0 00

90 o 9 0 . y

ooTo~ . .; ocD 0 . %0Qw 0 0 8 l Pa , 0 fs

-0.1’ @ ’ ’ I ’ ’ ’ ’ I ’ t ’ -2 -1 0 1 2 3 4 5 6 7 6 9 10

Pump-Probe Delay (ps)

FIG. 10. Expanded view of the 11 ( C02) 16 absorption recovery obtained using the femtosecond laser apparatus. Superimposed on the data is the parallel absorption recovery obtained using the picosecond laser system and a typical autocorrelation trace of the amplified femtosecond laser pulse. Unlike the picosecond data, the femtosecond data have not been smoothed.

0.4 at delay times less than 3 ps, we conclude that at the time of the “bump” the pump and probe transition dipoles lie in the same direction. This is an essential piece of in- formation used in the assignment of the probe absorption spectrum at early pump-probe delay times.

2. Coherent 1. - .I- motion: Internuclear dynamics in the first 3 ps

The “bump” at ~2 ps in the absorption recovery of the larger clusters (Fig. 6) is a particularly prominent fea- ture of the transient data that has a number of possible origins. Its appearance at both positive and negative delay times eliminates a multitude of potential experimental ar- tifacts. This feature, which occurs on a time scale close to the instrument response time, could arise from temporal structure in the picosecond laser pulse. To address this concern the 11 (COr) i6 absorption recovery was also ob- tained using the femtosecond laser system described ear- lier. This absorption recovery is shown in Fig. 10, along with the picosecond data and the autocorrelation of the amplified femtosecond laser pulse. The presence of the “bump” in the femtosecond absorption recovery signifies that it is not a result of temporal structure in the laser pulse. In addition, the improved time resolution enables us to determine the width of the “bump” to be ~2 ps F?VHM. The question still remains, however, as to the physical origin of this feature.

The exact form of the absorption recovery reflects both the time-dependent relaxation of the 11 ensemble, and the location of those regions on the potential energy surface with an appreciable absorption cross section at 720 nm. The “peak” in the absorption recovery at ~2 ps implies the existence of an excited region on the potential surface where the absorption of a second 720 nm photon is possi- ble, and at the same time rules out vibrationally relaxed IF as the chromophore responsible for the “bump.” Thus, the absorption recovery must consist of (at least) two compo-

Papanikolas et a/.: I; photodissociation and recombination 8741

nents: a transient 720 nm absorption at early delay times (resulting in the “peak” at 2 ps) and an increasing probe absorption at longer delays. The long-time component, which is discussed in the following section, is attributed to the recombination and vibrational relaxation of IF. Our immediate interest is in the origin of the transient absorp- tion observed at early delay times.

This Franck-Condon region could manifest itself as a “peak” in the absorption recovery in one of several ways. For example, if 720 nm transitions could occur from a discrete band of ground state vibrational levels, possibly located near the 1, dissociation limit, then a transient in- crease in the probe absorption would occur as the 1, en- semble vibrationally relaxed through these states. As we will see later, however, the relevant transitions have tran- sition dipole directions that are inconsistent with the pump-probe polarization data at early delay times. Instead the transient absorption at =: 2 ps is attributed to a fraction of the photoexcited ensemble undergoing coherent I* **I- motion through this excited Franck-Condon region. Knowing the location of this Franck-Condon region pro- vides a detailed description of the I* * *I- internuclear mo- tion in the first few picoseconds after photodissociation.

One possible reason for the successful observation of coherent 1. *aI- nuclear motion in the 11 (CO,), cluster ions stems from the strong electrostatic 1~ * * *CO2 forces that result in a large CO* binding energy ( - 150 meV/ C02). Both MD (Ref. 57) and Monte Carlo (Ref. 14) calculations predict that at the temperatures reached in these experiments ( -40 K) the clusters are initially solid. A combination of the strong ion-neutral forces and low internal temperature results in an ensemble of cluster ions with a restricted set of initial geometries. The number of initial isomers is most likely reduced even further for the case of 1~ (CO*) i6, where the closing of the first solvation shell is observed to occur.14 These conditions explain why the 1.. *I- motion is able to maintain nuclear coherence during its first “vibrational” period following photodisso- ciation. In addition, preliminary results from MD simula- tions on IF ( CO1) ,, cluster ions suggest52 that strong inter- molecular forces give the first solvation shell some of the character of a soft “net” that can stretch in order to ac- commodate the dissociating motion of the massive I atoms. This property of the solvent cage is the main reason why the 2 ps recurrence time is longer than one would expect for dissociation in a more constrictive liquid environment.

a. Assignment of the probe absorption spectrum. In order to determine on which surface(s) this coherent mo- tion takes place, one must assign the probe absorption spectrum at pump-probe delay times less than 3 ps. This task is hampered by the fact that little is known about the electronic structure of 12, and even less is known about that of 1, ( C02) ,, clusters. Identification of the probe tran- sition based solely on energetic arguments applied to the 1~ potential surfaces is impossible due to the quantitative uncertainty in these curves. Instead, it must be based on factors that do not depend on the details of the 1, potential energy surfaces. Our approach is to consider all of the

J. Chem. Phys., Vol. 99, No. 11, 1 December 1993 Downloaded 04 Oct 2007 to 128.138.107.158. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 10: I-2 photodissociation and recombination dynamics in size …jila.colorado.edu/~wcl/wclgroup/Pub downloads/I2... · 2007. 10. 4. · separations) and the delocalized, or molecular,

8742 Papanikolas et a/.: I; photodissociation and recombination

possible transitions, and eliminate those that are inconsis- tent with other experimental observations.

The qualitative potential energy curves in Fig. 1 show the six 11 electronic states present in this manifold. In principle, the probe transition could occur between any two of these electronic states, provided that the transition is not symmetry forbidden. The only rigorous selection rule in isolated 11 arises from the inversion symmetry. Consideration of all six electronic states gives rise to eight symmetry-allowed transitions, including the pump transi- tion. Of these eight transitions, the total probe absorption at 2 ps is ultimately assigned to have a substantial compo- nent of the f U(211U,3,2) -3 g(211g,3,2) transition. The de- tailed arguments that lead to this conclusion are as follows.

Four of the eight transitions terminate on either the f J22J) or f U(211U,1,2) excited states, which differ from the other electronic states in that they correlate with I- + I*( 2P,,2) at large internuclear separations. As long as the I* retains its electronic excitation it will be unable to recombine with I- to form IF. Since I* stores z 1 eV of energy, its ejection from the cluster would lead to the evap- oration of 4-5 fewer CO2 molecules. These two-photon ionic fragments would have masses, and flight times, that are similar to the more intense one-photon fragment ions, and as a consequence they would not affect the absorption recovery. On the other hand, if the I* is returned to the lower spin-orbit manifold through some subpicosecond ra- diationless process, then all of the photon energy will be available for CO2 evaporation. The collisional deactivation of I* by CO2 is extremely inefficient, requiring z lo6 col- lisions.65 This suggests that photoexcitation of 1~ to either the f J2zg+ 1 or f u(2&,1j2) excited state would result ex- clusively in the ejection of I* from the cluster. A fast ra- diationless process returning the I- * * *I* to the lower spin- orbit manifold can not be completely ruled out. However, the large energy gap between the two spin-orbit manifolds should make any such process slow. Therefore, photoexci- tation of the 11 to either the f J2Xi) or $ U(211U,1,2) ex- cited state will most likely result in the escape of I* from the cluster. We conclude that the transient probe absorp- tion does not involve transitions that terminate on either of these states.

Since transitions to the upper spin-orbit manifold are excluded, we focus on the transitions among the ground and three lowest excited states of I; shown in Fig. 11. The solid arrows labeled Ppump, P, , P2, and P3 depict the sym- metry allowed transitions between the four lowest elec- tronic states. The removal of the inversion symmetry, as is the case for 12 embedded within a cluster, gives rise to additional (presumed weak) transitions. These transitions, which are indicated in Fig. 11 by dashed arrows, are des- ignated P4 and P5. The polarization data provide an essen- tial piece of information needed to determine which of these five transitions is responsible for the transient absorp- tion at early pump-probe delay times.

Since r(At) for I; (Co2)i6 approaches its maximum possible value of 0.4 at short pump-probe delay times, the pump and probe transition dipoles at the time of the re- currence must lie in the same direction. In a Hund’s case

~_1_ yYI!i-

Pdl) f I :1- PZU) P,(L) j

P pump

k(2r-L,3,2)

&J2Q,1,2)

gg( 2Q,3,2>

) &l(“C) PI(l) P5s-I

PIG. Il. Schematic energy diagram of the ground state and three lowest excited states of I;. The solid lines indicate the symmetry allowed tran- sitions that can occur in isolated 11. If the inversion symmetry is broken then the dotted transitions could also occur. The direction of the transi- tion dipole with respect to the internuclear axis is also indicated in pa- rentheses. The P pump transition is anticipated to possess both parallel and perpendicular components; however, experimental evidence suggests that its parallel component is substantially larger than the perpendicular com- ponent (see Appendix B) .

(a) molecule the transition dipole is classified as either a pure parallel (AA =O) or a pure perpendicular (AA = f 1) transition. Because of the extensive spin-orbit mix- ing that takes place in IT, all of the fi = f states are ad- mixtures of Z ( A=O) and II (A= 1) states, resulting in electronic transition dipoles that can have both parallel and perpendicular components. In a separate experiment described in Appendix B, we observed the pump transition dipole to be dominated by its parallel component. There- fore, the probe transition excited at early pump-probe de- lay times must also have a predominantly parallel transi- tion dipole.

We have analyzed the 11 electronic structure in the Hund’s case (c) coupling limit in order to determine the presence of parallel and/or perpendicular components for the transitions displayed in Fig. 11. Details of this analysis are outlined in Appendix B. Since we cannot determine a priori the relative magnitude of the parallel and perpendi- cular components, any transition that has a parallel com- ponent could in principle be responsible for the recurrence in the absorption recovery. As indicated in Fig. 11, the P, , P3, P4, and PS transitions are all anticipated to be purely perpendicular transitions. The P2 transition, which is pre- dicted to be purely parallel, is the only transition consistent with the polarization data at early pump-probe delay times. The location of the corresponding Franck-Condon region at 720 nm is suggested by the I; potential curves to be at an internuclear separation of ~3 A. At this distance it appears energetically possible for a 720 mu transition between the z.J211g,3,2) and 1 U(211U,s,2) excited states to occur. The location of this 720 nm Franck-Condon region is indicated in Fig. 1 by the arrow labeled P2.

b. 1. * *I- Dynamics immediately following photodikso- cia tion. Based upon the aforementioned analysis, the I* * *I- motion in the first 3-4 ps following photodissocia-

J. Chem. Phys., Vol. 99, No. 11, 1 December 1993

Downloaded 04 Oct 2007 to 128.138.107.158. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 11: I-2 photodissociation and recombination dynamics in size …jila.colorado.edu/~wcl/wclgroup/Pub downloads/I2... · 2007. 10. 4. · separations) and the delocalized, or molecular,

tion can be described as follows. About 1 ps after photo- dissociation the separating I * * *I- motion is arrested by the solvent, leaving a substantial fraction of the photoexcited 11 ensemble on the i 8(2118,3,2) excited state at large inter- nuclear separations, perhaps 5-6 A. This portion of the photoexcited ensemble moves toward smaller internuclear separations, reaching the Franck-Condon region on the i J2$3,2) excited state ~2 ps after photodissociation, at which lime an increase in the probe absorption occurs. The subsequent decrease in probe absorption between 2 and 3 ps results from the I** *I- leaving this Franck-Condon region.

Coherent nuclear motion, both in isolated and solvated environments, has been limited largely to motion on single, bound electronic potential energy surfaces. The 12 (CO,) n system differs in two respects. First, the nuclear coherence involves motion that occurs on two different potential sur- faces. Second, both of these electronic states are dissocia- tive, and as a consequence, the forces that arrest and re- verse the dissociating I* * *I- motion must arise entirely from the CO2 solvent. Another possible example of coher- ent motion following photodissociation is provided by Nel- son and co-workers.66 They report fluctuations in the tran- sient absorption of Mn( CO), following photocleavage of the Mn-Mn bond in s( CO)Mn-Mn( CO), solvated in eth- anol. Although the origin of these fluctuations is not en- tirely understood, one possibility is that they result from Mn-Mn “vibration” following photoexcitation to a disso- ciative potential.

3. I,- relaxation processes: Internuclear dynamics after 3 ps

The analysis of the IT ( C02) i6 absorption recovery in the previous section indicates that for a fraction of the photoexcited ensemble, the 1. *aI- dynamics in the first two picoseconds takes place on the lowest two excited states of 1~. Eventually the 11 must recombine on the ground electronic state and undergo vibrational relax- ation.14 Whether the observed absorption recovery in the first 40 ps arises from 11 recombination and vibrational relaxation, or from the continued I** *I- motion on the lowest excited state, cannot be ascertained solely on the basis of the absorption recovery obtained at a single probe wavelength. Other experimental observations, however, can answer this question. Pump-probe photofragmentation experiments described later show that substantial “disinte- gration” of the solvent cage occurs in -4 ps. This process makes I atom escape more probable the longer the 1, re- mains on the repulsive i J 211,3,2) surface. Because 1. * .I- recombination is observed to occur in every 11 (CO,) i6 cluster that is photodissociated at 720 nm, it seems unlikely that 12 remains on this excited state for an extended period of time,67 implying that the absorption recovery between 3 and 40 ps reflects the recombination and vibrational relax- ation of IF.

The exact form of the absorption recovery is deter- mined not only by the time dependence of the 1~ relax- ation, but also by the location of the 720 nm Franck- Condon regions. If the only 720 nm Franck-Condon

region on the ground electronic state is located near r,, then the absorption recovery at delay times greater than 3 ps will reflect solely the appearance of vibrationally relaxed IF. This would imply that 11 vibrational relaxation in 11 (CO,) i6 is nearly complete within 10 ps after photodis- sociation. In order to confirm this, however, one must show that the absorption spectrum of the photoexcited en- semble is time independent, requiring data at multiple probe wavelengths. While our data are limited to a single probe wavelength, Barbara and co-workers5’ have con- ducted such experiments on 11 in a number of polar sol- vents. They demonstrate that 11 relaxation occurs in -5 ps in ethanol, a time scale that is comparable to that ob- served for 1, ( C02) 16.

Our results in combination with the I,/alcohol studies indicate that the rate for vibrational relaxation of solvated 1; is significantly faster than for I2 .68 This could be a result of both the net charge, which introduces long-range elec- trostatic interactions between the ion and the solvent, and the fact that the lower IF vibrational frequency provides a better match to the solvent phonon modes than does I,. Benjamin and Whitnell have used MD techniques to elu- cidate the relative importance of these properties in the vibrational relaxation of 11 solvated in water and ethanol. This study probes the 1, vibrational relaxation in the bot- tom half of the ground state potential well, where it is presumed that the excess electron charge is always equally distributed over both iodine atoms. These calculations in- dicate that the increased vibrational relaxation rate for 1~ is due to both the lower IF vibrational frequency and the presence of the ionic charge. In addition, MD simula- tions52 of 1~ ( C02) ,, clusters, which include the I + I- -+ 1~ charge transfer, suggest that there is a substantial enhance- ment of the 1~ relaxation rate coincident with the passage of the 11 from the localized to the delocalized charge dis- tribution.

4. Cluster size dependence

Figure 12 shows the 80% absorption recovery time plotted as a function of cluster ion size. Two different ab- sorption recovery behaviors are observed with a distinct transition occurring between n = 12 and n = 15. The ab- sorption recoveries for n = 8, 9, and 12 are nearly indistin- guishable (Fig. 5), suggesting that similar 1, recombina- tion and vibrational relaxation dynamics occur in these clusters. For 12<n<15, the absorption recovery times are strongly dependent on cluster size, while similar absorp- tion recovery times are observed for n = 15-22. The de- crease in the absorption recovery times is accompanied by the emergence of the recurrence at ~2 ps. This sudden change in the absorption recovery behavior, which occurs for cluster sizes near the completion of the first solvation shell, suggests that the IF caging dynamics are qualita- tively different in these two regimes.

One possible explanation for this behavior is suggested by the MD simulations carried out by Amar and Perera.57’69 Their calculations indicate that in the smaller clusters (n=8-12) the photodissociated 11 is able to un- dergo large amplitude motion before recombination. As

Papanikolas et al.: 12 photodissociation and recombination 8743

J. Chem. Phys., Vol. 99, No. 11, 1 December 1993 Downloaded 04 Oct 2007 to 128.138.107.158. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 12: I-2 photodissociation and recombination dynamics in size …jila.colorado.edu/~wcl/wclgroup/Pub downloads/I2... · 2007. 10. 4. · separations) and the delocalized, or molecular,

8744

cl” ’ * ” ’ ” * ’ ” a ’ * ’ * ’ n= 6 6 10 12 14 16 16 20 22 24 Precursor Ii(CO&

FIG. 12. Absorption recovery time displayed as a function of cluster ion size. The absorption recovery time is detined as the pump-probe delay time at which the parallel absorption recovery (Figs. 5 and 6) reaches 80% of its asymptotic value.

the I and I- separate, a CO2 molecule could fit between them forming a “solvent-separated” pair, which would de- lay IF recombination and destroy any 1. * *I- phase coher- ence. As more CO2 molecules are added to the solvent cage, the I - * -I- motion becomes increasingly restricted, forcing the 11 to remain a “contact pair.” Much more information is required before any picture of this type can be confirmed.

5. 1; rotational motion Several processes can contribute to the decay of total

probe absorption anisotropy. The most obvious is the re- orientation of 1, between the absorption of the pump and probe photons,” due to the photoexcitation of a rotating cluster. In addition to cluster rotation, dynamical processes arising after photoexcitation could also contribute to this decay. For example, if the forces exerted by the solvent on the dissociating iodine atoms do not lie along the 1~ bond, then an internal rotation of the 1, will occur, hastening the decay of the probe anisotropy. One might anticipate this process to be more important in the smaller clusters where the solvation shell is only partially filled. An anisotropy decay that is significantly faster than what could be ac- counted for by cluster rotation alone would indicate the presence of an additional decay mechanism such as the one just described.

In order to estimate the contribution of overall cluster rotation to the decay in the probe anisotropy, we have simulated r( At) for a rigid spherical top with a point tran- sition dipole embedded in the center. The strong electro- static interactions, and the relative sixes of the 11 and CO2 , suggest that the 11 is held rigidly within the solvent cage prior to its photodissociation. The rotational constant, B, of the sphere was chosen to approximate that of IF ( COa) i6; its value was estimated from minimum energy structures of Amar and Perera” to be ~0.0024 cm-‘. The rotational temperature for these simulations was chosen to be 50 K. The result of this model, displayed in Fig. 9(b) as

Papanikolas et al.: I; photodissociation and recombination

a solid line, shows that r( At) for the spherical top decays on a timescale similar to that observed for 1, (CO,) i6. This similarity in timescales indicates that the photoexci- tation of a rotating cluster is a major contributor to the decay of the probe absorption anisotropy. Felker and Ze- wai161 have calculated similar curves for rigid symmetric tops and have shown that the decay time scales inversely with @ and @, thus depending only weakly on the (somewhat uncertain) values chosen for T and B. The general form of the observed and simulated decays are qualitatively different at late times. The r(At) calculated for the rigid top shows an anisotropy that grows in and persists at long pump-probe delay times, in contrast to the experiment which shows no detectable anisotropy after 15 ps. The absence of this long time anisotropy is not partic- ularly surprising since the cluster has substantially disinte- grated by the time (30 ps) this residual anisotropy might become visible.

The anisotropy in the probe absorption for n = 12 and n= 14 decays much faster than one would expect from overall cluster rotation alone. For 11 ( COz) i4 [ol (At) not shown] the anisotropy vanishes after -6 ps, which is sig- nillcantly faster than the - 15 ps decay time predicted by the rotational model. Furthermore, the parallel and per- pendicular absorption recoveries for 1, (CO*) i2 [Fig. 8(b)] show no detectable anisotropy at any delay time. However, weak probe absorption precludes a quantitative measurement of r( At) at delay times below 5 ps, and large errors in this quantity persist out to - 10 ps, making the detection of any anisotropy difficult. Nevertheless, the po- larization data for n = 12 and n = 14 imply the existence of an additional anisotropy decay mechanism for clusters with partially filled solvation shells (n < 16). Additional experiments and more sophisticated modeling of the disso- ciation dynamics is required before the origin of this addi- tional decay mechanism can be determined.

B. Two-photon photofragmentation experiments The pump-probe experiments described above focus

on the relaxation of the 12 moiety. As the 1, relaxes, it transfers energy to the COz. solvent, ultimately leading to solvent cage “disintegration,” which in this context refers to COz evaporation and/or structural changes resulting from cluster “melting.” The experiments described in this section probe this disintegration, and thus provide a com- plementary view of the photodissociation dynamics.

To motivate how this is accomplished, first consider the sequential absorption of two photons by I,(C0,)i6 separated in time by 10 ps (the time between laser inter- action and photofragment mass analysis). With this delay the caged photofragments are allowed sufficient time to cool evaporatively to an internal temperature similar to that of the parent ion before the absorption of the second photon. As a result, the known one-photon fragment dis- tributions (see Table I in Ref. 14) can be used to predict the two-photon caging fraction. For example, the most probable one-photon photofragment ion resulting from the 720 nm photodissociation of 12 ( CO1) 16 is n =9, and the

J. Chem. Phys., Vol. 99, No. 11, 1 December 1993

Downloaded 04 Oct 2007 to 128.138.107.158. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 13: I-2 photodissociation and recombination dynamics in size …jila.colorado.edu/~wcl/wclgroup/Pub downloads/I2... · 2007. 10. 4. · separations) and the delocalized, or molecular,

Papanikolas et a/.: I; photodissociation and recombination 8745

6 2

40 ps I

(b) 63% Caged , ,

I;(co& I’ ’ ’ ’ 51 1 I-(co*), 4’ ’ ’ ’ 8’ 1

FIG. 13. Two-photon photofragment distribution observed following the squential absorption of two 720 nm photons separated by (a) 3OC0 ps, (b) 40 ps, and (c) -1 ps.

one-photon caging fraction for this photofragment cluster is ~50%. Therefore, one would expect the fraction of caged (IF based) products in the IF ( C02) 16 two-photon photofragment distribution to be =: 50%. This caging frac- tion reflects the degree of solvent cage “disintegration” that has occurred during this time.

The photofragment mass spectra resulting from the sequential absorption of two 720 nm photons by 1; (CO;?) ,6 at pump-probe delays of 3000, 40, and - 1 ps are shown in Fig. 13. Since the photoproducts are mass analyzed z 10 ~CLS after interaction with the pump and probe pulses, there is a sufficient amount of time for exten- sive CO, evaporation to occur. Thus, the most intense caged ionic photofragment corresponds to the evaporation of 13 CO, molecules from Ii ( COZ) 16, nearly double the number lost when a single photon is absorbed. The phot- ofragment distributions at 3000 and 40 ps are qualitatively similar. The 1 ps mass spectrum shows a substantial reduc- tion in the fraction of I--based (uncaged) photofragments.

Similar mass spectra have been obtained at a number of other delay times, and their caging fractions are dis- played in Fig. 14. The caging fraction is essentially con- stant at -60% for pump-probe delays between 3000 and 4 ps, and increases to -75% near zero delay. Between 0 and 4 ps there is a sharp peak in the caging fraction reach- ing 90% at z 1 ps. This peak occurs slightly earlier than the recurrence peak observed in the It (CO,) 16 absorption recovery (see Fig. 10). The increase in the caging fraction at short pump-probe delays is accompanied by a simulta- neous shift in the photofragment distribution of z 1 CO2

100 I;( CO& 1

50 ’ : , I I I ,, 1 I

-6 -4 -2 Pump-l’:obe 0 4Dela; 40 3000 (ps)

FIG. 14. The two-photon caging fractions for 11 ( COz) 16 displayed as a function of pump-probe delay time. Both pump and probe photons have a wavelength of 720 nm and 280 fs pulse width.

monomer to higher mass (Fig. 13). Because the photon energy is partitioned between the COZ binding energy and the kinetic energy of the ejected monomers, the evapora- tion of fewer CO2 molecules at early delay times indicates that a larger portion of the probe photon energy is parti- tioned into the kinetic energy of the departing CO2 mole- cules. The behavior for n = 14 and n = 15 is qualitatively similar to that of IF( COz) 16. The two-photon mass spectra for n = 12 also exhibit an increase in the caging fraction near zero delay, but the weak two-photon signals at short delay times prohibit a quantitative evaluation.

At the simplest level, the mass spectra show that the caging fraction for At>4 ps is not significantly different from that calculated from the one-photon fragmentation data, suggesting that the breakup of the solvent cage by the first photon is complete within -4 ps. This conclusion requires no real modeling of the cluster dynamics. In order to determine whether the peak in the transient caging frac- tion data results from the P2 photoexcitation (see Fig. 1) to the + J ‘II,& excited state at early delay times, coher- ent motion of the solvent cage, or some other dynamical process will require more sophisticated modeling of the cluster dynamics. In the discussion that follows we employ a simple model to look at the energy flow from the initially hot I; to the CO, solvent in the first four picoseconds.

Employing an analysis similar to that used to predict the two-photon fragment distribution at 10 ps for 1, (CO,) 16 we can construct a picture of the effective sol- vent cage size following 1~ photodissociation. For exam- ple, the observed two-photon caging fraction for n = 16 at 2 ps is ~80-85 %. The one-photon caging fractions suggest that at this time the solvent cage resembles that of n= 13- 14, implying that 2-3 CO2 monomers have effectively de- parted the cluster. This number increases to -4-5 COZ molecules at 4 ps, and 5-6 by 3000 ps.

Using this model the number of COz molecules ejected from the cluster can be used to determine the amount of the photon energy that remains at a given time after 1, photoexcitation by the pump pulse. This energy is parti- tioned between the 11 and the solvent cage and, hence,

J. Chem. Phys., Vol. 99, No. 11, 1 December 1993 Downloaded 04 Oct 2007 to 128.138.107.158. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 14: I-2 photodissociation and recombination dynamics in size …jila.colorado.edu/~wcl/wclgroup/Pub downloads/I2... · 2007. 10. 4. · separations) and the delocalized, or molecular,

8746 Papanikolas et a/.: I; photodissociation and recombination

represents only an upper limit to the amount of energy localized in the IT. Since the evaporation of each CO? monomer removes on average ~250 meV of energy from the cluster, the loss of 2-3 CO1 molecules in the first 2 ps following 11 photodissociation indicates that z 1.2 eV of energy remains. At this delay time it is likely that much of this remaining energy is localized in the I,. This places the I; at the time of the recurrence (2 ps) in a band of ener- gies near its dissociation limit, an observation which is consistent with the P2 probe transition assignment and the proposed Franck-Condon region on the 4 s(211k3,2) ex- cited state.

TABLE I. Representative two-photon signal levels observed under our typical operating conditions. These signal levels are based upon actual 1, (CO*) i6 data, but for clarity they do not contain the shot noise inher- ent in any single measurement. They are representative of the interaction of the 0.5 cm2 pump and probe pulses ( -500 d/pulse) with -2X lo6 12 (COs) i6 ions, the total number of parent ions produced by 500 valve pulses. The signals are essentially the same for both the 1.25 ps and 280 fs pulse durations. The pump-probe delay for the on/on configuration was 40 ps, corresponding to the maximum two-photon signal.

V. CONCLUSIONS

Ion beam

OE

on

Pump laser

On Off On Off

Probe laser On Off

0 0 0 0

150 25 25 20

We have carried out pump-probe experiments on size- selected IF (CO,) ,, cluster ions at 720 run. In these experi- ments the IF is photodissociated within the cluster with a 720 nm ultrafast pump pulse. The recovery of the 720 mn absorption is then probed with a second laser pulse of the same color. A “peak” is observed at ~2 ps in the absorp- tion recovery of the larger cluster ions, n = 14-17. Absorp- tion recovery curves obtained with both parallel and per- pendicular pump-probe polarization allow us to determine the I* * *I- internuclear motion free of orientational effects. This analysis indicates that this feature is not the result of cluster rotational motion; it is instead attributed to coher- ent I* * *I- nuclear motion occurring on an excited region of the 1, potential surface. The pump-probe polarization data, together with the observation of a parallel transition dipole for the pump transition, indicate that the probe transition also has a substantial parallel component. Of the possible probe transitions considered, only the f J2K,3,2) + 2 Z B( 2II9,s,2) transition possesses a parallel transition dipole. The potential curves suggest that the 720 nm Franck-Condon region is located on the f s(211~,3,2) excited state at an internuclear separation of -3 A. In addition, the disintegration of the CO2 solvent cage follow- ing 1, photodissociation is inferred from the evolution of the two-photon photofragment distribution with pump- probe delay time. These experiments suggest that substan- tial CO, evaporation occurs in the first 5 ps following pho- todissociation.

ferent heavy atom losses and will certainly add to our knowledge of cage recombination dynamics in cluster en- vironments.

ACKNOWLEDGMENTS

We wish to acknowledge a number of stimulating dis- cussions with Robert Parson, Paul Barbara, Joe Alfano, Gustav Gerber, Nancy Levinger, and Douglas Ray. Partial support for M. E. N. was provided by the Ford foundation. H. K. B. was supported, in part, by a Feodor-Lynen fel- lowship of the Alexander von Humboldt-Stiftung, Ger- many. This research is supported by the National Science Foundation Grants Nos. PHY 90-12244 and CHE 88- 19444.

APPENDIX A: DATA AQUISITION METHODOLOGY

1. Measurement of experimental observables a. Two-photon product signal

The presence of “noise” counts in the two-photon product channel requires us to perform two measurements to determine N2h,,( At) :

There is a pronounced dependence of the absorption recovery on cluster ion size. The clusters with n = 8, 9, and 12 have indistinguishable absorption recoveries which re- quire ~30 ps to reach 80% of their asymptotic value. The larger cluster ions (n= 15-22) exhibit faster absorption recoveries, obtaining 80% of the asymptotic value in 8-10 ps. The cluster size range between n= 12 and it = 15 cor- responds to the transition region in which the recurrence emerges. This change in the absorption recovery behavior suggests that the 1, recombination and vibrational relax- ation dynamics are qualitatively different in these two size regimes.

N2dAr) =Nodo,W -Nb, (Al)

where N ,,(Ar) is the number of two-photon ions gen- erated by the interaction of the pump and probe beams with the ion sample, and Nb is the number of “two-photon background” ions. Nb is composed of those ions arriving at Dl at times corresponding to the masses of the two-photon product ions, but which do nor result from the sequential absorption of one photon from the pump pulse and one photon from the probe pulse. Whereas experimental mea- surement of N On/O,,( At) is straightforward, the background measurement requires further consideration.

Future experiments will use two-color pump-probe techniques to probe other regions of the potential surface and establish the timescale required for vibrational relax- ation. Cluster ions with heteronuclear chromophores, such as ICl- ( C02) ,, or IBr- ( C02) ,, provide possibilities of dif-

Shown in Table I are the representative count rates (total number of ions per 500 laser shots) observed for the four different pump/probe laser combinations (on/on, on/ off, off/on, and off/off) in the presence and absence of the ion beam. These count rates have been chosen for illustra- tive purposes to be free of the shot noise inherent in any real measurement. In the absence of precursor ions no ion counts are observed. Thus, we need consider only the ion

J. Chem. Phys., Vol. 99, No. 11, 1 December 1993 Downloaded 04 Oct 2007 to 128.138.107.158. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 15: I-2 photodissociation and recombination dynamics in size …jila.colorado.edu/~wcl/wclgroup/Pub downloads/I2... · 2007. 10. 4. · separations) and the delocalized, or molecular,

counts generated by the interaction of the parent ions with the various pump/probe laser combinations. The designa- tions %dmW)t Nodoff, No~lon, and No~lof, cor=spond to the interaction of the 0.5 cm2 pump and/or probe beams (-500 d/pulse) with -2 x lo6 precursors, the total number of I; (CO,) 16 parent ions produced by 500 valve pulses. They are the same for both the 280 fs and 1 ps laser pulse durations. The “on/on” signal presented in Table I is that observed at At=40 ps, a delay time corresponding to the maximum two-photon signal. At shorter pump-probe delay times Nodo, is substantially smaller and, hence, the correct determination of Nb at these delay times is much more critical. The count rates in Table I indicate that (i) the largest contribution to Nb corresponds to the signal that arises directly from our ion beam with both lasers gated “off’ ( Noff/off , ) and (ii) the presence of a weak laser induced background.

N6 is defined in terms of the experimental measure- ments Non/o~p Noff/on, and Nowo~ a~

Nb=Notf/oir+kr,

where

(A21

r=f[ Won/or-Notd + Wowon--oworr) I- The quantity F is the number of two-photon ions produced by a single laser pulse, and k is a parameter that ranges from 2 to 4, depending on the source of this laser induced background.

In order to choose the correct value for k in Eq. (A2), we must first determine the source of the laser induced background. The possible sources can be broken into three categories: (1) Single photon absorption followed by

extensive fragmentation, or vice versa

(2) Direct absorption of two photons:

I Overlapped pump and probe pulses Well-separated pump and probe pulses

(3 ) Sequential absorption of two photons:

1

(a) which does involve ASE (b) which does nor involve ASE

(k=2).

(k=4) (k=2).

(k=4)

(k=2);

where ASE denotes amplified spontaneous emission. Since each category corresponds to a different value for k, it is necessary that the relative importance of these processes be determined. The discussion that follows shows that the origin of the laser induced background in this experiment is process ( 1) and, hence, k=2.

If the direct absorption of two photons, process (2), were important, then one would expect substantially more two-photon products to be produced by a 1 ps pulse then by a 6 ns pulse, provided that the two pulses have the same energy. This is not observed. The interaction of 12 (CO,) 16

with a 6 ns laser pulse generates in excess of 200 two- photon product ions in 500 laser pulses, whereas the two- photon ion yield produced by a 1 ps laser pulse is smaller by almost 2 orders of magnitude (I’- 5-10). In addition, no increase in I’ is observed upon decreasing the pulse- width from 1 ps to 280 fs. These observations indicate that direct two-photon absorption does not contribute signifi- cantly to the laser induced background.

Sequential two-photon absorption from within a single laser pulse is classified according to whether the photons are absorbed from the picosecond portion of the pulse or the broad ASE background. The most likely contribution involving ASE arises from the absorption of one photon from the picosecond portion of the pulse and one photon from the ASE background, process 3 (a). Since the ASE is limited to less than 0.5% of the total pulse energy, the number of two-photon ions generated in this manner should be no more than 1 ion count per 500 laser pulses. This suggests that the laser induced background does not arise from sequential two-photon absorption involving ASE. The sequential absorption of two photons from the ultrafast portion of the laser pulse, process 3 (b), is also possible. The number of two-photon products produced in this manner depends upon the I; absorption recovery time in comparison to the laser pulse duration, increasing for longer pulse width. The fact that I’ does not depend upon whether a 1 ps or 280 fs laser pulse is used indicates that this process is not an important contributor to the laser induced background.

Based on the previous discussion, we anticipate that the most important process contributing to the laser in- duced background is single photon absorption coupled with extensive fragmentation [process ( 1 )], perhaps as the result of collisions of the parent ion with background gas or ion-optic apertures. Under these circumstances, the cor- rect form of Eq. (A2) is obtained with k=2. The validity of this definition for Nb is further established by the fact that, as expected, N2,J At = 0) =: 0.

Since the two-photon ions have well defined arrival times at Dl, one can reject the single ions that strike the detector during the time between the ion “peaks” by indi- vidually gating around the four most intense doublets in the mass spectrum. This gating reduces Nb by a factor of 2-3, resulting in a signal to noise ratio of ~5-10 at long pump-probe delay times. The time windows used to obtain the two-photon ion signal in the 12 (CO,) 16 absorption recovery experiments are indicated in Fig. 7 by brackets.

Measurement of this two-photon product signal was accomplished by operating the MCP detector in a particle counting mode with the output going to a 300 MHz dis- criminator and gated counter. The pulse pair resolution of this system is ~5 ns and is limited by the pulsewidth of the single ion peaks produced by the MCP detector. Under our experimental conditions N2,,,,(Ar) was observed to scale quadratically with the laser power, and linearly with the number of parent ions.

Papanikolas et al.: 1; photodissociation and recombination 8747

J. Chem. Phys., Vol. 99, No. 11, 1 December 1993 Downloaded 04 Oct 2007 to 128.138.107.158. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 16: I-2 photodissociation and recombination dynamics in size …jila.colorado.edu/~wcl/wclgroup/Pub downloads/I2... · 2007. 10. 4. · separations) and the delocalized, or molecular,

8748 Papanikolas et a/.: I; photodissociation and recombination

b. Parent ion signal

When the reflectron is adjusted to reflect the two- photon products onto Dl, both the one-photon ions and unphotodissociated parent ions have enough energy to traverse the reflectron and strike D2. Since the number of one-photon ions is notably smaller than the number of parent ions, the ion signal measured on D2 is essentially that of the parent ions. The parent ion signal, NYU, was measured on D2 concurrently with the detection of the two-photon product ions on D 1. The particle multiplier D2 was observed to have linear response to the number of ions impinging on the detector.

states. Consequently, the transition dipole moment in a case (c) molecule need not lie along a molecular symmetry axis. Here we express the case (c) states as a linear com- bination of case (a) basis states. The transition moment integral, ( $f I c1 I $i) 9 is then examined for each transition. These integrals typically contain several “case (a)” terms; the nonzero terms are then classified as contributing to either the parallel ( AA=O) or perpendicular (AA= f 1) component of the total transition dipole.

c. Laser power

Spin-orbit coupling mixes the case (a) electronic states with the same inversion symmetry and the same total angular momentum a.” If only the six 1~ electronic states in this manifold are considered, then we can express the relevant case (c) electronic states in terms of the case (a) states as

Measurement of the laser flux @i was accomplished by monitoring a fraction of the amplified picosecond pulse transmitted through a mirror on a photodiode. A teflon sheet was used to homogenize, as well as attenuate the beam. A second lens collected the light and focused it onto a photodiode. This arrangement provided a linear photo- diode response with input laser power.

1; J21-&,3,2)) = I 2Q,3/2>, (Bl)

I~g(2~g,1/2))=bl12~g,*,2)+~212~Cgf),

2. Data collection algorithm I4 u(2%3,2) > = I 2K4,w2)- The data collection algorithm used in this experiment

is structured around the measurement of two quantities, the two-photon background, Nb, and the two-photon cross section, azhy(Ar). First NodoE, N,,,r,0n, and NOs,,,s, were measured, and from these Nb was calculated as outlined earlier. The second step involved nine measurements of N on/on ( At) and corresponding normalization factors (ZNf%$) for Ar=Ar,,,, Ar,,...,Ar,, At,,,. The two- photon cross sections, oZ,J At), were calculated for each of the nine N on/on( At) values, using Nb determined in the first step and the nine normalization factors. The oZih,,(Ar,,f) values at the beginning and end of the delay “set” were used to judge the quality of the data. If the percent change in the cross sections measured at At,,, exceeded 25%, then those data were discarded. If N,,,(Ar) was significantly greater than Nb, as is observed at long delay times, then the second step was repeated several times before the two- photon background was measured again. On the other hand, near zero delay, where Non/,,* zNb, a more frequent two-photon background measurement is required. At these delay times, the first and second steps were alternated, providing a more frequent background measurement. At long delay times 10-15 % of the data collection time was devoted to the measurement of Nb, while near zero delay the fraction of time spent measuring Nb was increased to 30-40 %.

All of the n=f states are admixtures of Z and I1 states. ‘r-he 2Q~~2 and 2k,~~2 case (a) states do not spin-orbit mix with any other states in this manifold, and thus remain pure II states. The coefficients, a,, a2, bt , and b2 are func- tions of the spin-orbit matrix elements, (lcti I I?” j qj), and thus depend on the 11 internuclear separation. The direc- tion of the electronic transition dipoles is determined by examining its transition moment integral. 1. Pump transition [ i o(21TI,,,2)+ B u(28u+)]

732 (4 J2ng,1i2 ) I p I i U ( 28,f ) ) transition moment inte- gral is written in terms of the case (a) states as

=c1(2~,1,21~12~,+)+c2(2~g+ IP12~u,,,2>

+~~~2~,fI~12~uf~+~~~2~g,l/21~12~*,~,2~. WI

Since the first two terms correspond to perpendicular case (a) transitions, and the last two correspond to parallel case (a) transitions, it is evident that this transition dipole will possess both parallel and perpendicular components. Be- cause we have no knowledge of the electronic wave func- tions, we cannot determine a priori the relative magnitude of the parallel and perpendicular components.

APPENDIX B: TRANSITION DIPOLE DIRECTIONS IN I<

If 11 were a Hund’s case (a) diatomic molecule then the photoexcitation between two electronic states would be classified as either a pure parallel (AA =0) or a pure per- pendicular (AA= f 1) transition. However, in a case (c) molecule A is not a good quantum number due to the extensive spin-orbit mixing of the case (a) electronic

We can, however, determine experimentally72 the di- rection of this transition dipole. Photodissociation of the 1, molecular ion with 720 nm light polarized parallel to the ion beam results in two distinct I atom arrival times at D2. The observation of this doublet signifies that the phot- ofragments are ejected along the polarization vector of the laser, indicating that the 4 8(2118,1,2) -4 U(2X,+) transition dipole lies predominately parallel to the internuclear axis. (The doublet disappears if the I; is photodissociated with light polarized perpendicular to the ion beam.) The dou-

J. Chem. Phys., Vol. 99, No. 11, 1 December 1993 Downloaded 04 Oct 2007 to 128.138.107.158. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 17: I-2 photodissociation and recombination dynamics in size …jila.colorado.edu/~wcl/wclgroup/Pub downloads/I2... · 2007. 10. 4. · separations) and the delocalized, or molecular,

blet splitting, which is related to the I* * *I- kinetic energy release, qualitatively confirms the - 1.1 eV I; bond disso- ciation energy. The observation of a large parallel compo- nent for this transition is not surprising, given the presence of the (‘2,’ 1~ 1 22,‘> term in the above equation. This term corresponds to the contribution from the strong charge- transfer transition, and is responsible for both the large absorption cross section and the parallel transition dipole observed for this transition. This result is consistent with the spectroscopic measurements of 11 embedded in a KI matrix performed by Delbecq er al.73

2. Probe transitions

All of the possible probe transitions are analyzed in a similar manner. The PI transition moment integral con- tains two terms:

(Q2&,3,2) IPI; ut2Z))

=c1(2~g,3/21P12~ u,1/2~+~~~2~g,3/21~12~u+~. (B3) The (211g,3,2~~(L211,,1,2) term in expression (B3) is zero due to the fact that it corresponds to a forbidden case (a) transition. The remaining Z - II term makes it a pure per- pendicular transition. Similar arguments establish that P3 is also a perpendicular transition. Because the 4 8( ‘rIs3/2) and i a(2rIy,3,2) states remain pure II states, the transition designated P2 is expected to be a pure parallel transition. Thus, P2 is the only transition that is consistent with the polarization data and represents the extra absorption asso- ciated with the peak at -2 ps in the absorption recovery data.

Strictly speaking the I; g-u symmetry is broken inside the cluster. As a result, all of the Q=i states undergo spin-orbit mixing, adding additional terms to the transition dipole moment integrals. The removal of the 11 inversion symmetry requires us to also consider the P4 and Ps tran- sitions indicated in Fig. 11 in addition to those already discussed. However, all of the case (a) terms that could give rise to parallel components are of the form (2r13/2 1~ I 2111,2) and, hence, are equal to zero. Therefore, in the absence of the It inversion symmetry the P, , P3, P4, and Ps transitions are all still expected to have transition dipoles perpendicular to the 1, internuclear axis.

‘J. M. Farrar, in Current Topics in Ion Chemistry and Physics, edited by C. Y. Ng and I. Powis (Wiley, New York, 1992).

‘M. A. Johnson and W. C. Lmeberger, in Techniques@ the Study of Zen Molecule Reactions, edited by J. M. Farrar and W. Saunders, Jr. (Wiley, New York, 1988), pp. 591-635.

‘A. W. Castleman, Jr., in Clusters ofAtoms and Molecules, edited by H. Haberland ( Springer-Verlag, New York, 1992).

‘L. I. Yeh, M. Okumura, J. D. Myers, J. M. Price, and Y. T. Lee, J. Chem. Phys. 91, 7319 (1989); M. Okumura, L. I. Yeh, and Y. T. Lee, ibid. 88,79 ( 1988); M. Okumura, L. I. Yeh, J. D. Myers, and Y. T. Lee, ibid. 85, 2328 (1986).

‘U. Buck, X. Gu, R. Krohne, C. Lauenstein, H. Linnartz, and A. Ru- dolph, J. Chem. Phys. 94,23 (1991); U. Buck, X. Gu, C. Lauenstein, A. Rudolph, J. Phys. Chem. 92, 5561 (1988).

6T. J. Selegue and J. M. Lisy, J. Phys. Chem. 96, 4143 (1992); J. A. Draves, 2. Luthey-Schulten, W.-L. Liu, and J. M. Lisy, J. Chem. Phys. 93, 4589 ( 1990).

‘C, S. Yeh, K. F. Willey, D. L. Robbins, J. S. Pilgrim, and M. A.

Duncan, J. Chem. Phys. 98, 1867 (1993); K. F. Willey, C. S. Yeh, D. L. Robbins, J. S. Pilgrim, and M. A. Duncan, ibid. 97, 8886 (1992).

*P J Campagnola, D. J. Lavrich, M. J. Deluca, and M. A. Johnson, J. . . Chem. Phys. 94, 5240 (1991); M. J. DeLuca, B. Niu, and M. A. Johnson, ibid. 88, 5857 (1988).

‘M. H. Shen and J. M. Farrar, J. Chem. Phys. 94,3322 (1991); 93,4386 (1989).

‘OH. Haberland, B. V. Issendorf, T. Kolar, H. Kommeier, C. Ludewigt, and A. Risch, Phys. Rev. Lett. 67, 3290 (1991).

“Z. Y. Chen, C. D. Cogley, J. H. Hendricks, B. D. May, and A. W. Castleman, Jr., J. Chem. Phys. 93, 3215 (1990).

‘*E. C. Honea, M. L. Homer, J. L. Persson, and R. L. Whetten, Chem. Phys. Lett. 171, 147 (1990); M. Y. Hahn and R. L. Whetten, Phys. Rev. Lett. 61, 1190 (1988).

“N. E. Levinger, D. Ray, M. L. Alexander, and W. C. Lineberger, J. Chem. Phys. 89, 5654 (1988); N. E. Levinger, D. Ray, K. K. Murray, A. S. Mullin, C. P. Schulz, and W. C. Lineberger, ibid. 89, 71 (1988).

“5. M. Papanikolas, J. R. Gord, N. E. Levinger, D. Ray, V. Vorsa, and W. C. Lmeberger, J. Phys. Chem. 95, 8028 ( 1991).

I’M. L. Alexander, N. E. Levinger, M. A. Johnson, D. Ray, and W. C. Lineberger, J. Chem. Phys. 88, 6200 (1988).

16M. L. Alexander, M. A. Johnson, N. E. Levinger, and W. C. Line- berger, Phys. Rev. Lett. 57, 976 (1986); M. L. Alexander, M. A. Johnson, and W. C. Lineberger, J. Chem. Phys. 82, 5288 (1985).

“P. J. Campagnola, L. A. Posey, and M. A. Johnson, J. Chem. Phys. 95, 7998 (1991); L. A. Posey and M. A. Johnson, ibid. 89, 4807 (1988).

‘*R. Knochenmuss and S. Leutwyler, J. Chem. Phys. 91, 1268 (1989); 0. Cheshnovsky and S. Leutwyler, ibid. 88, 4127 (1988); Chem. Phys. Lett. 121, 1 (1985); R. Knochenmuss, 0. Cheshnovsky, and S. Leut- wyler, ibid. 144, 317 (1988).

“0. Echt, P. D. Dao, S. Morgan, and A. W. Castleman, Jr., J. Chem. Phys. 82, 4076 (1985).

*‘T. Baumert, C. Riittgermann, C. Rothenfusser, T. Thalweiser, V. Weiss, and G. Gerber, Phys. Rev. Lett. 69, 1512 (1992).

*‘J. J. Breen, D. M. Willberg, M. Gutmann, and A. H. Zewail, J. Chem. Phys. 93, 9180 (1990).

**C. A. Schmuttenmaer, J. Qian, S. G. Donnelly, M. J. DeLuca, D. F. Varley, L. A. DeLouise, R. J. D. Miller, and J. M. Farrar, J. Phys. Chem. 97, 3077 (1993).

*‘J Steadman, E. W. Foumier, and J. A. Syage, Appl. Opt. 29, 4962 (1990).

*‘S. Wei, J. Pumell, S. Buzza, R. J. Stanley, and A. W. Castleman, Jr., J. Chem. Phys. 97, 9480 ( 1992).

“1. R. Sims, M. Gruebele, E. D. Potter, and A. H. Zewail, J. Chem. Phys. 97, 4127 (1992).

26M Gruebele, I. R. Sims, E. D. Potter, and A. H. Zewail, J. Chem. Phys. 95; 7763 (1991).

*‘N. F. Scherer, L. R. Khundkar, R. B. Bernstein, and A. H. Zewail, J. Chem. Phys. 87, 1451 (1987).

**S. I. Ionov, G. A. Brucker, C. Jaques, L. Valachovic, and C. Wittig, J. Chem. Phys. 97, 9486 (1992).

*‘J. J. Breen, L. W. Peng, D. M. Willberg, A. Heikal, P. Gong, and A. H. Zewail, J. Chem. Phys. 92, 805 (1990).

30M. F. Hineman, G. A. Brucker, D. F. Kelly, and E. R. Bernstein, J. Chem. Phys. 97, 3341 (1992).

“J. Steadman and J. A. Syage, J. Phys. Chem. 95, 10 326 (1991). ‘*J. Steadman and J. A. Syage, J. Chem. Phys. 92, 4630 ( 1990). “3. A. Syage, Chem. Phys. Lett. 202, 227 (1993). “D. Ray, N. E. Levinger, J. M. Papanikolas, and W. C. Lineberger, J.

Chem. Phys. 91, 6533 (1989). “J. M. Papanikolas, V. Vorsa, M. E. Nadal, P. J. Campagnola, and W. C.

Lineberger, J. Chem. Phys. 97, 7002 ( 1992). 36N. F. Scherer, R. J. Carlson, A. Matro, M. Du, A. J. Ruggiero, V.

Romero-Rochin, J. A. Cina, G. R. Fleming, and S. A. Rice, J. Chem. Phys. 95,1487 (1991); N. F. Scherer, A. J. Ruggiero, M. Du, and G. R. Fleming, ibid. 93, 856 ( 1990).

“R. M. Bowman, M. Dantus, and A. H. Zewail, Chem. Phys. Lett. 161, 297 (1989).

38A. H. Zewail, M. Dantus, R. M. Bowman, and A. Mokhtari, J. Pho- tochem. Photobiol. A 62, 301 (1992).

“Y Yan, R. M. Whitnell, K. R. Wilson, and A. H. Zewail, Chem. Phys. Lktt. 193, 402 (1992).

4oM. H. M. Janssen, R. M. Bowman, and A. H. Zewail, Chem. Phys. Lett. 172, 99 ( 1990).

Papanikolas et a/.: I; photodissociation and recombination 8749

J. Chem. Phys., Vol. 99, No. 11, 1 December 1993 Downloaded 04 Oct 2007 to 128.138.107.158. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 18: I-2 photodissociation and recombination dynamics in size …jila.colorado.edu/~wcl/wclgroup/Pub downloads/I2... · 2007. 10. 4. · separations) and the delocalized, or molecular,

8750 Papanikolas ef al.: I; photodissociation and recombination

“P. Cong, A. Mokhtari, and A. H. Zewail, Chem. Phys. Lett. 172, 109 (1990); T. S. Rose, M. J. Rosker, and A. H. Zewail, J. Chem. Phys. 88, 6672 (1988); M. J. Rosker, T. S. Rose, and A. H. Zewail, Chem. Phys. Lett. 146, 175 (1988).

“T. Baumert, V. Engel, C. Riittgermann, W. T. Strunx, and G. Gerber, Chem. Phys. Lett. 191, 639 ( 1992).

“N. F. Scherer, L. D. Ziegler, and G. R. Fleming, J. Chem. Phys. 96, 5544 (1992); N. F. Scherer, D. M. Jonas, and G. R. Fleming, ibid. 99, 153 (1993).

&M. H. Vos, J. Lambry, S. J. Robles, D. C. Youvan, J. Breton, and J. Martin, Proc. Natl. Acad. Sci. 88, 8885 (1991).

‘sU. Banin and S. R&man, J. Chem. Phys. 98, 4391 (1993); U. Banin, A. Waldman, and S. R&man, ibid. 96, 2416 (1992).

&R. M. Bowman, M. Dantus, and A. H. Zewail, Chem. Phys. Lett. 156, 131 (1989).

“E E. Ferguson, J. Phys. Chem. 98, 731 ( 1986), and references therein. ‘sR1 M. Whitnell, K. R. Wilson, and J. T. Hynes, J. Chem. Phys. 96,5354

(1992); J. Phys. Chem. 94, 8625 (1990). 491. Benjamin and R. M. Whitnell, Chem. Phys. Lett. 284, 45 ( 1993). %D. A. V. Kliner, J. C. Alfano, and P. F. Barbara, J. Chem. Phys. 98,

5375 (1993); A. E. Johnson, N. E. Levinger, P. F. Barbara, J. Phys. Chem. 96, 7841 (1992); J. C. Alfano, Y. Kimura, P. K. Walhout, and P. F. Barbara, Chem. Phys. 175, 147 (1993).

“J. C. Owrutsky, Y. R. Kim, M. Li, M. J. Sarisky, and R. M. Hochstrasser, Chem. Phys. Lett. 184, 368 (1991); E. J. Heilweil, F. E. Doany, R. Moore, and R. M. Hochstrasser, J. Chem. Phys. 76, 5632 (1982).

“J. M. Papanikolas, P. Maslen, and R. Parson (unpublished). “E. C. M. Chen and W. E. Wentworth, J. Phys. Chem. 89,4099 (1985). %G. N. R. Tripathi, R. H. Schuler, and R. W. Fessenden, Chem. Phys.

L&t. 113, 563 (1985). “L. Bar&es and A. G. UreiIa, Chem. Phys. Lett. 176, 178 (1991). 56R. Axria, R. Abouaf, and D. Teillet-Billy, J. Phys. B 21, L213 (1988). 57F. G. Amar and L. Perera, Z. Phys. D 20, 173 (1991). ‘*W. J. Tomlinson, R. H. Stolen, and C. V. Shank, J. Opt. Sot. Am. B 1,

139 (1984).

59J. D. Kafa and T. Baer, Opt. Lett. 12, 401 (1987). 6oM. Kakuxawa, T. Nakashima, H. Kubota, and S. Seikai, J. Opt. Sot.

Am. B 5, 215 (1988). 6’ P. M. Felker and A. H. Zewail, J. Chem. Phys. 86, 2460 (1987). 62A. J. Cross, D. H. Waldeck, and G. R. Fleming, J. Chem. Phys. 78,

6455 (1983); 79, 3173(E) (1983). 63G. R. Fleming, Chemical Applications of Utmfast Spectroscopy (Ox-

ford, New York, 1986), pp. 124-130. 6”Conventionally the anisotropy parameter r(t) is used to describe the

anisotropy present in absorption or emission dipoles and b, which is equal to 2r(t), is used to describe the anisotropy present in photofrag- ment and photoelectron distributions.

65D. H. Burde and R. A. McFarlane, J. Chem. Phys. 64, 1850 (1976). %A. G. Joly, S. Ruhman, B. Kohler, and K. A. Nelson, Springer Series in

Chemical Physics: Ultmfast Phenomena VI, edited by T. Yakajima, K. Yoshihara, C. B. Harris, and S. Shionoya (Springer-Verlag, Berlin, 1988), Vol. 48, pp. 506-510.

67The exact nature of the 12 potential curves is unknown. If the ~s(211r,3,2) state is actually bound, then trapping for longer periods could occur.

68A. L. Harris, J. K. Brown, and C. B. Harris, Ann. Rev. Phys. Chem. 39, 341 (1988), and references therein.

69L. Perera and F. G. Amar, J. Chem. Phys. 98, 7354 (1989). “Another possible decay mechanism involves the contribution from per-

pendicular transitions to the total probe absorption at later pump-probe delay times. If such transitions exist, they will serve to “dilute” the anisotropy, and will lead to its premature decay. There is no experi- mental evidence, however, that suggests the existence of this phenom- enon.

“H. Lefebvre-Brion and R. W. Field, Perturbations in the Spectra of Diatomic Molecules (Academic, New York, 1986), pp. 81-89.

‘*M. E. Nadal, V. Vorsa, J. Smith, P. J. Campagnola, H. K. Buchenan, P. Kleiber, J. M. Papanikolas, and W. C. Lineberger (unpublished).

73C J Delbecq, W. Hayes, and P. H. Yuster, Phys. Rev. 121, l&3.( 1961).

J. Chem. Phys., Vol. 99, No. 11, 1 December 1993

Downloaded 04 Oct 2007 to 128.138.107.158. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp