chapter 5 excited state intramolecular hydrogen transfer...

24
RINI KUMMROW DREYER NIBBERING ELSAESSER INTRAMOLECULAR HYDROGEN TRANSFER Chapter 5 Excited State Intramolecular Hydrogen Transfer: Ultrafast Mid-Infrared Spectroscopy probing Reaction Dynamics and Vibrational Relaxation Processes H N S O HBT-enol H N S O HBT-enol hydrogen transfer hydrogen transfer HBT-keto O H N S Published as: M. Rini, A. Kummrow, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Faraday Discussions 122, 27-40 (2003). M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical Physics Letters 374 (1), 13-19 (2003). Structure: - electronic state coupling - hydrogen bond reaction coordinate - anharmonic coupling between vibrational modes Dynamics: - intramolecular hydrogen transfer - intramolecular vibrational redistribution - vibrational energy dissipation

Upload: others

Post on 29-May-2020

13 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

RINI ■ KUMMROW ■ DREYER ■ NIBBERING ■ ELSAESSER INTRAMOLECULAR HYDROGEN TRANSFER

Chapter 5

Excited State Intramolecular Hydrogen Transfer:

Ultrafast Mid-Infrared Spectroscopy probing Reaction Dynamics and Vibrational Relaxation Processes

HN

S

O

HBT-enol

HN

S

O

HBT-enol

hydrogentransfer

hydrogentransfer

HBT-keto

OHN

S

Published as: M. Rini, A. Kummrow, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Faraday Discussions 122, 27-40

(2003). M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical Physics Letters 374 (1), 13-19

(2003).

Structure: - electronic state coupling - hydrogen bond reaction coordinate - anharmonic coupling between vibrational modes Dynamics: - intramolecular hydrogen transfer - intramolecular vibrational redistribution - vibrational energy dissipation

Page 2: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

130 RINI ■ KUMMROW ■ DREYER ■ NIBBERING ■ ELSAESSER

Page 3: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

INTRAMOLECULAR HYDROGEN TRANSFER 131

122/3

Femtosecond mid-infrared spectroscopy of condensed phase

hydrogen-bonded systems as a probe of structural dynamics

Matteo Rini, Andreas Kummrow, Jens Dreyer, Erik T. J. Nibbering* and

Thomas Elsaesser

Max Born Institut fur Nichtlineare Optik und Kurzzeitspektroskopie, Max Born

Strasse 2A, D-12489 Berlin, Germany. E-mail: [email protected]

Received 29th January 2002, Accepted 12th April 2002

We report the first time-resolved site-specific mid-infrared study of the photo-inducedexcited state hydrogen transfer reaction in 2-(20-hydroxyphenyl)benzothiazole (HBT) with130 fs time resolution. The transient absorption of the C=O stretching band marking theketo*-S1-state appears delayed on a time scale of 30–50 fs after electronic excitation to theenol*-S1-state. Its line center subsequently shifts up by about 3–5 cm�1 after excitation,depending on the excitation wavelength tuned between 315 and 349 nm. This effect isattributed to intramolecular vibrational energy redistribution (IVR) and vibrational energyrelaxation (VER) processes. We observe for the first time the coherent effects ofanharmonic coupling of low frequency modes (�60 cm�1, �120 cm�1), on the C=O modemarking the product state. We ascribe the 120 cm�1 mode to a Raman-active in-planedeformation mode that is coherently excited by the UV-pump pulse. We tentatively explainthe coherent excitation of the infrared active 60 cm�1 out-of-plane deformation mode bynonradiative processes within the excited enol state after electronic excitation.

I. Introduction: Ultrafast structural dynamics in the condensed phase

When dealing with ultrafast chemical reaction dynamics in the condensed phase, the object ofresearch is not only to decipher the underlying molecular parameters determining femtosecondchemical events of bond breaking and formation (known as femtochemistry1,2), but also to gain aclear picture of the roles that nearby solvent molecules have in the chemical reaction dynamics.Solvent motions occur with a broad range of time scales, with ultrafast components at femtosecondand picosecond time scales having a major impact on the outcome of these reactions due toprocesses such as molecular conformation (de-)stabilization, electronic and vibrational energyexchange, electronic and vibrational coherence decay and the cage effect in molecular dissociation.3

The ideal method for time-resolved studies of chemical reaction dynamics would be a techniquewhere the spatial parameters of the structural dynamics are resolved for the molecular probe understudy as well as for the nearby solvent molecules. We focus on femtosecond mid-infrared pump–probe and four-wave mixing techniques (photon echoes) as an alternative method to probeultrafast structural dynamics, since methods such as ultrafast X-ray diffraction,4–7 ultrafast X-rayspectroscopy8,9 or ultrafast electron diffraction10,11 are still in their infancy. It is well known thatwith time-resolved vibrational spectroscopy one can grasp the dynamics of specific chemical bondsin the case where the probed vibrations can be regarded as local modes. In contrast, by probing

DOI: 10.1039/b201056a Faraday Discuss., 2002, 122, 000–000 1

This journal is # The Royal Society of Chemistry 2002

Page 4: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

132 RINI ■ KUMMROW ■ DREYER ■ NIBBERING ■ ELSAESSER

electronic states with optical frequencies additional arguments are necessary to determine whethercertain chemical bonds are involved in the reaction pathways.

Following a description of several major advantages, well-known from stationary vibrationalspectroscopy, that enable one to gain detailed structural information on molecular geometries andinteractions in section I, we review ultrafast spectroscopic work on hydrogen bonding dynamicsand hydrogen transfer reactions in section II. We then give a short overview in section III of earlierultrafast mid-infrared spectroscopic studies on the hydrogen bond dynamics of Coumarin 102. Wepresent our new results on ultrafast site-specific mid-infrared spectroscopy on excited-state intra-molecular hydrogen transfer in 2-(20-hydroxyphenyl)benzothiazole in section IV, and concludewith some prospects for our method in section V.

A. Site-specific information from inspection of vibrational bands

In the case where a molecular vibration is coupled to the reaction coordinate, leading to markedfrequency shifts and intensity changes, this mode can be used as a spectator for the state of thechemically reactive bond. For instance, we have demonstrated the potential of probing site-specifichydrogen bond cleavage and rearrangements in the case of hydrogen-bonded complexes of Cou-marin 102, a standard probe in solvation dynamics studies.12–17

B. Geometric information from vibrational band patterns

Vibrational marker modes can appear, shift or disappear, indicating the occurrence of a rear-rangement of nuclear coordinates when crossing occurs from the reactant state towards the pro-duct state. Moreover, comparison of experimental vibrational patterns with results of quantumchemical calculations leads to determination of geometric structures of transient states, asdemonstrated in the case of the locally excited and charge-transfer states of 4-(dimethylamino)-benzonitrile.17–24

C. Structural information from anharmonic coupling between vibrational modes

Direct geometric information can be obtained from mid-IR pump–mid-IR probe experiments onvibrational bands that exhibit features of anharmonic coupling between vibrational modes. Byinspection of one specific mode not only the dynamics of this mode can be followed, but also thedynamics of a second mode is monitored due to anharmonic coupling between these modes.In principle, due to the implicit geometrical dependences of the anharmonic couplings betweenvibrational normal modes the potential of the method lies in resolving information on the mole-cular structure. In addition, due to the inherent time-resolution, structural dynamics of evolvingmolecular species can be grasped.

As an illustration of the potential of IR spectroscopy of probing anharmonic coupling we notehere recent results obtained on intramolecular hydrogen bonds of organic molecules in the elec-tronic ground state. A modulation of the IR pump–probe signal in phthalic acid monomethylester(PMME) with a period of 330 fs has been observed, indicating a coherent motion of the hydrogenbond.25 The coherent motions have been ascribed to anharmonic coupling between the O–Dstretching mode and a low-frequency out-of-plane motion of the two sub-groups that are con-nected to each other through the hydrogen bond. Similar features have been obtained for 2-(20-hydroxyphenyl)benzothiazole (HBT), where the hydrogen bond distance is modulated by a 120cm�1 in-plane deformation mode.26

Determination of molecular structure is also pursued by investigation of excitonic couplingbetween vibrational modes with comparable transition frequencies by use of two-dimensional IRspectroscopy.27 Examples on which structural information has been obtained include small pep-tides,28–33 N-methylacetamide34,35 and Rh(CO)2(C5H7O2).

36,37 Two-dimensional IR spectroscopycan be extended towards resolution of transient structures by means of an additional optical triggerpulse.

2 Faraday Discuss., 2002, 122, 000–000 122/3

Page 5: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

INTRAMOLECULAR HYDROGEN TRANSFER 133

II. Hydrogen bonding dynamics and hydrogen transfer reactions

The structural dynamics of protic solvents such as water and of biomolecular polymers such asproteins and DNA are largely determined by the special properties of hydrogen bonds. Thedynamics of the structural properties of hydrogen bonding are crucial to understanding suchsystems on a microscopic level. In addition proton exchange along hydrogen bonds occurs fre-quently in hydrogen-bonded networks such as water (proton hops, Grotthuss mechanism).38–40

Proton pumps through membranes also operate by proton transfer along ‘‘water wires ’’ tomaintain a certain pH gradient along membranes, e.g. in the case of photosynthetic reactioncentres.41 The dynamics of hydrogen bonds occur on ultrafast time scales, mainly set by motions ofthe hydrogen donor and acceptor groups, both in thermal equilibrium and for non-equilibriumexcitations. In thermal equilibrium, i.e. in the electronic ground state, vibrational and translationalmotions of the nuclei comprising the hydrogen-bonded system cause fluctuations in the geometricalparameters of the hydrogen bond, leading to e.g. proton transfer. In addition, fluctuating solventconfigurations may alter the relative energies of the donor and acceptor moieties of the hydrogen-bonded complex, thereby stabilising or de-stabilising the newly formed configuration after protontransfer.42,43

Static and dynamical properties of hydrogen bonds are intensively studied using a multitude ofspectroscopic techniques, varying from X-ray diffraction, electronic and vibrational spectroscopy,and nuclear magnetic resonance. In particular, it has been shown that steady-state infrared spec-troscopy of the O–H/O–D stretching mode has the potential to reveal the strength of the hydrogenbond due to a direct relationship to the red-shift of the vibrational frequency of the O–H stretchingband.44 The same applies for N–H/N–D or F–H/F–D stretching vibrations. In addition similarcorrelations have been found for C=O, C=N, C–O and C–N vibrations when these groups are partof a hydrogen-bonded system.The ultrafast dynamical properties of intermolecular hydrogen bonds, in the electronic ground

state, in water, alcohols and acids have been studied intensively during the last 5 years aftertriggering a non-equilibrium vibrational excitation of an O–H vibration with an ultrafast mid-IRpump, after which the non-equilibrium dynamics are determined by transmission changes of a mid-IR pulse or by anti-Stokes Raman emission. Important information has been obtained on vibra-tional lifetimes with typical values in the subpicosecond time range,45–48 vibrational dephasing andspectral diffusion,49,50 energy redistribution,51–54 energy transfer to neighbouring molecules,55 bandsubstructures56,57 and band shifts due to the dynamical Stokes shift53,58–61 or due to local heating,62

and orientational dynamics.63,64

III. Hydrogen bond dynamics in solvation studies

Solute–solvent interaction in liquids has so far mainly been studied by monitoring transient red-shifts of electronic emission bands or the electronic coherence decay of the chromophore in orderto characterize the time scales of solvent response. Such solvation experiments provide ensembleaveraged time-correlation functions for liquid motion, but give very limited information onchanges of microscopic solvent structure. An important issue to be inspected is how a hydrogen-bonded complex in the condensed phase responds to electronic excitation. In this solvation sce-nario, changing local hydrogen bonds lowers the energy of the transition dipole. A prototypicalcase is hydrogen-bonded complexes embedded in a nonpolar aprotic solvent, as the relative con-centrations of the species forming the complex and, thus, the site geometry and molecular inter-actions can be varied over a wide range. With this the problem is simplified as compared to the caseof a solute hydrogen-bonded to protic solvents like water, where the staggering magnitude of thehydrogen-bonded networks makes the interpretation a formidable task. New and highly specificinsight has been obtained into the ultrafast structural dynamics of hydrogen-bonded complexesbetween an organic chromophore, Coumarin 102 (C102), serving as hydrogen acceptor, andhydrogen-donating phenol molecules dissolved in the nonpolar solvent tetrachloroethene, C2Cl4 .The advantage of these complexes as compared to the case of a hydrogen-donating solvent likechloroform (CHCl3) is that one can study the dynamics on both the acceptor and donor sides.

122/3 Faraday Discuss., 2002, 122, 000–000 3

Page 6: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

134 RINI ■ KUMMROW ■ DREYER ■ NIBBERING ■ ELSAESSER

From femtosecond vibrational spectroscopic experiments, where C102 is excited electronicallyand the spectator carbonyl-stretching mode of C102 and the spectator O–H-stretching mode ofphenol are probed, one has learnt that:13,14,16,17

(1) The hydrogen bond between C102 and (phenol)1,2 breaks within 200 fs (time resolution of theexperiments), similar to C102–CHCl3 complexes.12,15

(2) The time-dependent features in the O–H region show in addition that: (a) the solute dipolechange induces strong changes of vibrational transition moments; (b) the released (phenol)2-moietyreorganizes itself to its equilibrium configuration with an 800 fs time constant.

(3) In the reported studies, dynamics is described involving the phenol-unit not directly bondedto C102. With this result a demonstration is given of the ability of femtosecond vibrationalspectroscopy for probing dynamics in that part of the hydrogen-bonding network that is notdirectly linked to the initially excited acceptor. This is a result that by no means can be directlydetermined with techniques where only electronic transitions are probed, such as photon echoes ortime-resolved Stokes-shift measurements.

Since hydrogen bonding occurs in a wide range of solvents, the above-mentioned results havefar-reaching consequences for interpretation of solvation studies.

IV. Ultrafast mid-infrared study of the excited state intramolecularhydrogen transfer

The transfer of hydrogen atoms or protons represents one of the most elementary structuralchanges occurring in chemical and biological processes.65–71 In general, hydrogen transfer betweenexcited singlet states displays ultrafast dynamics and, thus, spectroscopy with ultrashort lightpulses represents a powerful technique to monitor and analyze such reactions. So far, most workhas concentrated on dynamics in the condensed phase, where both intramolecular hydrogentransfer and the release of protons from so-called photoacids into the surrounding solvent havebeen studied. In pump–probe experiments, the reaction has been initiated by pico- or femtosecondexcitation to a higher electronic state and the subsequent time evolution of the system has beenmonitored through changes of electronic absorption or emission spectra. Time-resolved studies ofphotoacids have shown proton transfer dynamics on time scales of tens to hundreds of picose-conds.72–76 In contrast, early picosecond studies of intramolecular hydrogen transfer gave an upperlimit for the transfer time of the order of a few picoseconds, even at low temperatures of thesamples.77–81 The first femtosecond study of intramolecular hydrogen transfer was performed onthe enol*–keto* transformation of 2-(20-hydroxyphenyl)benzothiazole (HBT) and gave a rise timeof the red part of the keto*-emission of 170 fs.82 After this work, a number of similar studies with atime resolution of about 100 fs have been reported for other intramolecular hydrogen transferreactions, giving rise times of the product species between 50 and 200 fs.83–87 Recent experimentswith a time resolution of about 20 fs allow the observation of vibrational coherences occurringduring and after excited state hydrogen transfer. This has been demonstrated for the first time for2-(20-hydroxy-50-methylphenyl)benzotriazole where two low-frequency modes at 299 and 470 cm�1

are elongated upon electronic excitation and give rise to oscillatory features in the rise of productemission.88 An analysis of the microscopic motions connected with such modes, based on reso-nance Raman studies and theoretical calculations,89 led to a qualitative microscopic picture ofhydrogen transfer in which motions along low-frequency modes determine the pathway of thereaction on the excited state potential energy surface. Though such experiments have providedsubstantial insight into ultrafast hydrogen transfer, the study of electronic spectra does not allowfor a direct characterization of transient molecular structures during and after the reaction. Here,ultrafast vibrational spectroscopy monitoring local changes of molecular geometries provides muchmore information.

The object of our studies is the hydrogen transfer reaction of 2-(20-hydroxyphenyl)benzothiazole(HBT) dissolved in tetrachloroethene (see Fig. 1). After excitation to the enol*-S1-state, the keto*product structure is formed in its first excited state Already 15 years ago it was shown withpicosecond infrared spectroscopy that a clear signature of the keto*-S1-state population density ofHBT is the mid-infrared (MIR) absorption band around 1535 cm�1 corresponding to the C=O

4 Faraday Discuss., 2002, 122, 000–000 122/3

Page 7: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

INTRAMOLECULAR HYDROGEN TRANSFER 135

double bond stretching mode.90 In such measurements, the transfer reaction was not temporallyresolved. Femtosecond pump–probe studies of stimulated emission in the red part of the keto S1–S0emission spectrum gave a rise time 170 fs for the emission which was attributed to hydrogentransfer.82,83 Very recently, this process has been studied with a substantially improved timeresolution of 30 fs and over a spectral range covering the full keto*-emission band.91,92 At shortwavelengths, emission rise times of about 60 fs were found, whereas a substantially slower rise wasdetected at long wavelengths, in quantitative agreement with the data of Laermer et al.82 Thedelayed onset of stimulated emission was explained by a ballistic reaction on the excited statepotential surface leading to delayed population of the keto*-S1-state of HBT. The 60 fs rise timeof emission at short wavelengths was attributed to the hydrogen transfer. The occurrence ofpronounced oscillatory features on these UV/pump–probe measurements has been ascribed tocoherent motion of four low-frequency modes induced by the nuclear rearrangement that occursduring the hydrogen transfer process. It should be noted, however, that such modes are stronglyelongated upon excitation of the enol*-S1-state, as they couple strongly to the electronic transition.

Fig. 1 Molecular structures of HBT in the enol and keto-tautomer configurations. Hydrogen transfer (HT)occurs after electronic excitation of the enol*-S1-state. The keto*-S1-state decays on subnanosecond time scalesto the keto-S0 state by photo-emission. HBT relaxes then to the enol-S0 state by back hydrogen transfer (backHT). The optical spectra show the large Stokes-shift between absorption and emission bands that typifies theexcited state hydrogen transfer reaction.

122/3 Faraday Discuss., 2002, 122, 000–000 5

Page 8: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

136 RINI ■ KUMMROW ■ DREYER ■ NIBBERING ■ ELSAESSER

In the following we show that with ultrafast vibrational spectroscopy one can monitor changes inthe vibrational spectra of relevant functional groups, in this way providing direct site-specific accessto molecular structure and nuclear motions. In particular, we monitor the dynamics of the C=Ocarbonyl stretching mode formed by hydrogen transfer, and, thus, representing a direct probe ofthe formation of the keto species. In this way, the ambiguity resulting from the wavelengthdependent rise time of keto*-emission is avoided. In addition, the shape of the band is potentiallysensitive to vibrational energy redistribution after hydrogen transfer, either when the C=Ostretching mode is vibrationally excited, or by anharmonic coupling to other intramolecular modesthat are on their own transiently excited.

A. Experimental procedure

The pump–probe set-up was similar to the one used earlier for femtosecond mid-IR spectroscopyon hydrogen bond dynamics of Coumarin 102 complexes13 and intramolecular charge transfer of4-(dimethylamino)benzonitrile20 with three modifications used to improve time resolution, tun-ability and signal to noise ratio. First, electronic excitation was performed using near-UV pulsesgenerated by sum frequency mixing of the fundamental of a 1 kHz amplified Ti : sapphire laser andvisible pulses generated by a noncollinear optical parametric amplifier as described elsewhere.93 Inthe experiments the excitation wavelength was tuned from 315 nm to 350 nm. The excitation pulseenergy was 2–4 mJ and the pulse duration around 40 fs. The pump pulses could be variably delayedand were focused on the sample with a concave mirror with a beam diameter of approximately150 mm.

Second, the mid-infrared pulses were generated using double-pass collinear optical parametricamplification followed by difference frequency mixing of signal and idler.94 The center frequencywas tuned to 1550 cm�1 and the output energy was around 400 nJ. Probe and reference pulses werederived using reflections from a BaF2 wedge, and focused in the sample with off-axis parabolicmirrors (focal diameter 100 mm). The whole pump–probe set-up was purged with nitrogen gas toavoid spectral and temporal reshaping of the MIR pulse by the absorption of water vapor and CO2

in air.Third, probe and reference pulses were dispersed in a polychromator and complete spectra were

recorded simultaneously for each shot using a liquid nitrogen cooled 2� 31 HgCdTe detectorarray. Normalizing probe and reference signal on a single shot basis provides highly reliablespectra. The polychromator was not tuned during measurements to avoid the repositioning error(�2 cm�1). Synchronous chopping of the UV-pump pulses was used to eliminate long term drifteffects. Experimental curves shown here represent an average of 100–200 delay time traces eachtaking 100 shots average per delay step.

HBT was dissolved in C2Cl4 and pumped through a free streaming jet (nominal thickness 100mm). The group velocity mismatch between UV-pump pulse and the MIR-pulses was measured tobe 880 fs mm�1 in the neat solvent, the UV-pulse traveling more slowly. In order to minimize theeffect of temporal walk-off between the pulses a relatively high concentration of 4 g l�1 was chosen,leading to an effective jet thickness of less than 30 mm. Nevertheless the high extinction of the UVpulse does not lead to substantial temporal reshaping, since excitation was performed around thepeak of the broad absorption band, which leads to spectrally uniform attenuation.

For reference purposes the jet was replaced by a polished ZnSe window to determine the zerodelay point (including the chirp of the infrared pulse) and to control the time resolution of the crosscorrelation (FWHM 120–140 fs).

B. Transient spectra obtained on the ESIPT reaction of HBT

We investigated the mid-IR spectral region between 1400 and 1600 cm�1, where the C=O stretchingabsorption of the keto*-state can be found. We have also inspected the spectral region of the O–Hstretching band of the enol-form and the N–H spectral region of the keto*-form. However, sincethe broad O–H and N–H stretching bands overlap significantly with implicit complications in theinterpretation of the experimental data, we focus here on our results obtained on the C=Ostretching band.

6 Faraday Discuss., 2002, 122, 000–000 122/3

Page 9: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

INTRAMOLECULAR HYDROGEN TRANSFER 137

Typical experimental transient UV-pump–mid-IR-probe data of HBT in C2Cl4 are shown inFig. 2 and 3. In this case the UV excitation pulse was tuned at 335 nm. In Fig. 2 we presenttransient infrared spectra, i.e. the change of vibrational absorption recorded between 1410 and1590 cm�1 together with the infrared spectrum of the enol ground state. Negative contributions tothe absorbance change DOD (bleach) can be observed already at negative delay times at thosefrequencies where ground state bands are located. These signals result from perturbed freeinduction decay contributions in the pump–probe signal due to bleaching of ground state modesupon UV-excitation.95,96 In addition, around zero delay, a contribution of the solvent occurs, ofwhich the response is identical at all detected frequencies of the probe pulse.A prominent new band is formed at 1530 cm�1, representing the C=O stretching band of the

keto*-species. The strength and spectral position of this band agree with earlier picosecond mea-surements.90 The new C=O band builds up with a delay of 30–50 fs (Fig. 3A), representing theformation time of keto*–HBT.

Fig. 2 Three-dimensional plot of the spectrally and time-resolved pump–probe signals of HBT with electronicexcitation at 335 nm (upper graph). The data show typically perturbed free induction decay contributions atnegative delays, solvent-signals around zero delay, and bleaches and induced absorptions of transient vibra-tioanl bands at positive delays. Transient spectra at several pulse delays are shown in the lower graph (a),indicating the rapid rise of the C=O stretching band and subsequent up-shifting at picosecond time scales. Theground state infrared spectrum of HBT is shown in (b).

122/3 Faraday Discuss., 2002, 122, 000–000 7

Page 10: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

138 RINI ■ KUMMROW ■ DREYER ■ NIBBERING ■ ELSAESSER

We performed Gaussian fitting of the C=O band to derive the temporal development of thestrength, width and line centre, of which results for excitation at 335 nm are shown in Fig. 4. Weobserve that the strength and width of the C=O band does not change at longer delay times. Fromthis we deduce that the C=O mode is populated in its v ¼ 0 state, since a feeding of the v ¼ 1 statewould lead to both transient stimulated emission and absorption contributions to the pump–probesignal for the following reasons. For a harmonic oscillator one would not expect to observe anychanges in absorbance of the probe signals, since the excited-state v ¼ 1! v ¼ 2 absorption has across section which represents twice that of stimulated emission from the v ¼ 1! v ¼ 0 transition.Any changes in absorbance do not occur when the v ¼ 1 state decays to the v ¼ 0 state bypopulation relaxation. For an anharmonic oscillator similar ratios in transition moments occur formoderate anharmonicities.97 However, for an anharmonic oscillator the anharmonic shift betweenthe v ¼ 0! v ¼ 1 and the v ¼ 1! v ¼ 2 leads to spectrally displaced excited-state absorption andstimulated emission bands. For the C=O stretching mode with an estimated anharmonic shift of15–20 cm�127 an initial feeding of the v ¼ 1 state and subsequent decay should thus be clearlyobservable.

Moreover, a second even stronger argument that the C=O stretching mode is created in the v ¼ 0state can be derived from the observed spectral width of 12.5� 0.5 cm�1 of the C=O stretchingband directly after generation of the keto*-state. Such a value for the width of a vibrational band

Fig. 3 Transient signals measured at 1530 cm�1 with excitation at 335 nm. In (a) the initial delayed dynamicsof HBT (dots) are compared with a cross-correlation signal measured in ZnSe (solid line). In (b) the dynamicsare shown over a larger time range, with the measured signal (solid line) compared to that of HBT only aftersubstraction of the solvent contribution (dashed line). In (c) a blow-up is presented of the signal due to HBTand solvent indicating the oscillatory contribution due to the anharmonically coupled low-frequency modes.

8 Faraday Discuss., 2002, 122, 000–000 122/3

Page 11: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

INTRAMOLECULAR HYDROGEN TRANSFER 139

indicates that even for the extreme case of a line shape completely determined by populationrelaxation, the v ¼ 1 state of the C=O transition has a lower limit of the population lifetime T1 of424 fs. More likely the T1-value will be larger, since pure dephasing processes will also contribute tothe line width.

C. Energy redistribution and dissipation

The centre frequency of the C=O band blue-shifts by about 2.5–5.5 cm�1 after the hydrogentransfer reaction is finished (Fig. 2 and 4). Biexponential fitting shows a temporal behaviour with a0.5 ps component (relative weight 30%) and picosecond component (70%) with a time constantdepending on excitation wavelength. Blue-shifts of vibrational bands after a photo-induced che-mical reaction have been observed before in the picosecond time regime in the case of for instancetrans-stilbene98 and azobenzene.99 In these studies it was assumed that at picosecond time scales theexcess energies are thermally distributed amongst the vibrational modes. Specific spectator modesappear shifted due to anharmonic coupling with low-frequency modes that are highly populated.Usually a red-shift results upon excitation since most anharmonic coupling terms have negativevalues. However, the same red-shift behaviour would occur when only a limited set of highlypopulated normal modes takes place. This may be the case for HBT, since a full equilibration ofexcess energy is unlikely on the time scale of 50 fs. We expect that only a limited number ofaccepting normal modes strongly coupled to the hydrogen bond coordinate will initially redis-tribute the excess energy between the enol* and keto* excited states.100,101

The subsequent up-shift of the C=O transition frequency of HBT is understood to be a con-sequence of ongoing IVR among intramolecular modes and vibrational relaxation of the highlypopulated intramolecular modes towards thermal equilibrium through dissipation of the excessenergy towards the surrounding solvent (vibrational energy relaxation, VER). We tend to ascribethe initial subpicosecond component to IVR, and the longer picosecond dynamics of the blue-shiftas an indicator of cooling of the excited molecule towards ambient temperatures. For medium-sizedmolecules typical values for time scales of cooling are found to lie in the picosecond range.102–105

Fig. 4 Results of the fitting procedure of the measured C=O band (excitation at 335 nm) with a Gaussian lineshape. The width (a) and strength (b) of the C=O band do not change at longer delays. At early times the fittingroutine is affected by the additional solvent contribution. In (c) the line centre of the C=O band (dots) shows afrequency up-shift with components of 0.5 ps and 5 ps. The coherent modulation of the line centre is clearlyvisible in (d), where the residue (dots) is shown after substraction of the biexponential fit on the data in (c). Thesolid line in (d) shows a function consisting of two oscillating contributions with frequencies of 60 and 120 cm�1

and a decaying constant of 1 ps.

122/3 Faraday Discuss., 2002, 122, 000–000 9

Page 12: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

140 RINI ■ KUMMROW ■ DREYER ■ NIBBERING ■ ELSAESSER

The hydrogen transfer reaction of HBT occurs on such a fast time scale that we expect thatenergy equilibration over all intramolecular modes is highly unlikely to occur on the same timespan of 50 fs. It would be very interesting to identify the vibrational modes that accept the excessenergy generated when HBT converts from the enol* to the keto* state. These accepting modes areextremely effective in redistributing the energy in such a way that the reverse ESIPT reaction fromketo* to enol* state is prevented. We note that transient anti-Stokes Raman spectroscopy isselective for vibrationally excited modes, and with that technique non-equilibrium populations ofvibrational (Raman-active) modes can be grasped.98,106,107 This method is thus, in principle, idealfor identifying the energy flow in reactive systems. In the case of HBT, however, one would have torely on coherent techniques such as femtosecond CARS108 with sufficient time resolution.109

Subpicosecond time-resolved detection of transient incoherent resonance Raman signals impliesloss of frequency resolution. In addition, due to small Raman cross-sections, the latter form oftransient Raman spectroscopy often fails for fluorescing molecules.

D. Coherence phenomena in photoinduced chemical reactions

Inspection of the transient recorded at 1530 cm�1 after excitation at 335 nm shows that oscillatorycomponents are present in the response of the C=O mode (Fig. 3). The Gaussian fitting procedureof the transient C=O stretching mode band shape reveals that the temporal behaviour of the linecentre exhibits these oscillatory features superimposed on the dynamical blue-shift in a morepronounced way (Fig. 4). Fourier analysis of the transient line centre recorded with UV excitationat 335 nm reveals two components with frequencies of about 60 and 120 cm�1. In contrast, withUV excitation at 349 nm only the 120 cm�1 component is observed.

Oscillatory features on vibrational bands have been observed before in IR-pump–IR-probestudies on O–H and O–D stretching modes in intramolecular hydrogen bonds in the electronicground state.25,26 In these works the oscillatory parts of the pump–probe signals are caused bycoherent modulations of the O–H/O–D stretching band positions due to anharmonic couplingwith coherently excited low-frequency modes. In the present study we ascribe our observation thatthe frequency position of the C=O stretching mode is modulated by anharmonic coupling to one ortwo low-frequency modes, respectively. We thus detect, for the first time, such a spectral mod-ulation by coherently excited low-frequency modes of a vibrational band marking the product stateafter a chemical reaction.

The assignment of these low-frequency modes in the keto*-product state can be made bycomparison to the well-known vibrational mode spectrum in the enol-ground state89,110 (see alsoFig. 5). Although there is no guarantee that the vibrational mode spectrum is similar, ultrafastelectronic pump–probe spectroscopic data have not shown substantial differences for the low-frequency Raman-active modes.91,92 One could anticipate that this also applies for other (infrared-active) modes. With this in mind one can correlate the 120 cm�1 vibrational mode to an in-planebending motion of the two ring systems modulating the hydrogen bond length.89,110 This mode hasbeen observed in electronic resonance Raman spectra indicating a strong displacement between thetwo enol potential energy surfaces. An out-of-plane twisting motion of the ring systems can beascribed to the 60 cm�1 mode, also modulating the hydrogen bond length. This mode has anextremely low electronic resonance Raman cross section, in contrast it is known to be infraredactive.89 A third mode is present at the low-end of the vibrational spectrum. This mode howeverinvolves an out-of-plane tilting motion of the two ring systems without strong motions of the atomsthat constitute the hydrogen bond. We tentatively discard the involvement of this mode in thedynamics of photo-excited HBT in the following discussion.

At this point we want to make a clear distinction between ultrafast UV-pump–VIS-probe andUV-pump–IR-probe spectroscopy in relation to what one can expect in terms of coherently excitedwave packet motions in the system under study. The ultrashort UV-pump coherently excitesvibrational coherences on the enol-excited and enol-ground states of those modes that are stronglyRaman active, and whose frequency lies within the bandwidth of the UV excitation pulse.1,111–118

These light-field driven coherences then evolve on the respective potential energy surfaces, and mayeven survive level crossings if the modes are not strongly coupled to the reaction coordinate.Another option is that vibrational wave packets are generated by rapid nonradiative processesfrom the initially excited reactant state to a product state.119–123 In the latter case the surface

10 Faraday Discuss., 2002, 122, 000–000 122/3

Page 13: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

INTRAMOLECULAR HYDROGEN TRANSFER 141

crossing along the reaction coordinate is accompanied by the creation of vibrational coherencesalong the coordinate of the vibrational mode that, in contrast to the previously mentioned case, isstrongly coupled to the reaction coordinate. In the case of probing the product state the transientabsorption or emission of a UV/VIS probe pulse is modulated by coherences in Raman-activevibrational modes. If no other knowledge than the Franck–Condon factors of the electronictransition induced by the UV-pump pulse exists on the particular molecular systems, one cannotdecide whether the coherences are induced by the applied light pulse or by the subsequent chemicalreaction. With regard to HBT for instance the Franck–Condon factors of the enol-S0 to enol*-S1state transition are well known,89,110 however this is not the case for the keto*-S1- to keto-S0 state.Only when a coherence in the keto*-S1-state is observed that cannot be correlated to one of themodes that are Raman-active in the enol! enol* transition, can one safely conclude that thecoherence is driven by nonradiative processes.In the case of our ultrafast UV-pump–IR-probe experiment we probe the keto*-S1-state by

inspection of a C=O stretching vibrational marker mode. Vibrational coherences in low-frequencymodes are observed if the anharmonic coupling constants to the C=O stretching mode are suffi-ciently large. The factors determining the magnitude of coherent modulation of the infrared-signalsby anharmonic coupling are different from those for coupling of vibrational modes to electronictransitions, and it should be no surprise that coherences in both infrared- and Raman-active modesare visible.With regard to our observation of the frequency and damping time of the 120 cm�1 mode we

confirm within experimental error earlier UV-pump–VIS-probe observations.91,92 Higher frequencyoscillations, most notably the 254 cm�1 mode, are not detectable in our experiment due to limitedtime resolution in comparison to the UV-pump–VIS-probe experiment. The observation of the60 cm�1 mode, however, is a surprising result. The fact that this mode is not observed in the UV-pump–VIS-probe experiment suggests that the mode is infrared active (not Raman active) in theketo*! keto transition. Moreover it is known that the mode cannot be coherently excited by theUV-pump pulse as it has a low Franck–Condon factor in the enol! enol* transition. Excitation ofthis mode is thus driven by nonradiative processes after preparation of the excited enol*-state. Onecould regard this fact as an indication that the mode is impulsively excited by the hydrogen transferreaction, where the reaction time is only one tenth of the oscillation period of the mode. From theobservations that (a) the 60 cm�1 mode is not observable at 350 nm where the electronic origin (0–0)transition is located, whereas (b) the reaction time appears to be independent of the excitationenergy, we are led to the tentative explanation that the coherent excitation of the infrared-active 60cm�1 out-of-plane deformation mode is driven by an IVR mechanism in the enol excited state.

Fig. 5 Schematic representation of the low-frequency in-plane bending, out-of-plane twisting and out-of-plane tilting modes. These modes are similar in the enol- and keto-tautomers.

122/3 Faraday Discuss., 2002, 122, 000–000 11

Page 14: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

142 RINI ■ KUMMROW ■ DREYER ■ NIBBERING ■ ELSAESSER

V. Conclusions and prospects

We have studied the excited state intramolecular hydrogen transfer reaction of 2-(20-hydro-xyphenyl)benzothiazole (HBT) with femtosecond site-specific UV-pump–IR-probe spectroscopy.For the first time we observe a delayed onset of the C=O stretching mode band, indicating that thetime scale of the actual hydrogen transfer process lies in the range of 30–50 fs. After hydrogentransfer we observe that the spectral width and intensity of the C=O band does not change withinexperimental error, indicating that the C=O vibrational mode is generated in the v ¼ 0 state. Theline centre of the C=O stretching band blue-shifts after excitation, from which we deduce asubpicosecond component assigned to intramolecular vibrational redistribution (thermalization)within the HBT-molecule, and a picosecond component that is correlated to cooling by vibrationalenergy relaxation to the solvent. We also observe oscillatory contributions to the position of theC=O stretching band that are due to anharmonic coupling with two coherently excited modes: a120 cm�1 Raman active mode and a 60 cm�1 infrared active mode. The coherence of the lattermode, that is only observed when the electronic transition of HBT to the excited enol*-state occurswith significant excess energies, cannot directly be coherently driven by the UV-pump light field,and should thus be impulsively generated by nonradiative processes during the hydrogen transferreaction. We tentatively assign a mechanism of coherent feeding by IVR within the excited enol*-state as the origin of this phenomenon.

The experimental results on excited-state intramolecular hydrogen transfer obtained with fem-tosecond mid-infrared spectroscopy demonstrate the potential of this method. We aim to extendthis method to investigate excited state intermolecular proton transfer of photoacids.

Acknowledgement

Our progress has benefited from the financial support of the Deutsche Forschungsgemeinschaftthrough the Schwerpunktprogramm ‘‘Femtosekunden-Spektroskopie elementarer Anregungen inAtomen, Molekulen und Clustern ’’. The expertise of Dr. P. Hamm on mid-IR generation anddetection has been extremely helpful in the pursuit of success in this project.

References

1 A. H. Zewail, Science, 1988, 242, 1645.2 A. H. Zewail, J. Phys. Chem. A, 2000, 104, 5660.3 G. R. Fleming, Chemical Applications of Ultrafast Spectroscopy, Oxford University Press, Oxford, 1986.4 C. Rischel, A. Rousse, I. Uschmann, P.-A. Albouy, J.-P. Geindre, P. Audebert, J.-C. Gauthier, E. Froster,

J.-L. Martin and A. Antonetti, Nature, 1997, 390, 490.5 C. Rose-Petruck, R. Jimenez, T. Guo, A. Cavalleri, C. W. Siders, F. Raksi, J. A. Squier, B.C. Walker,

K. R. Wilson and C. P. J. Barty, Nature, 1999, 398, 310.6 C. W. Siders, A. Cavalleri, K. Sokolowski-Tinten, C. Toth, T. Guo, M. Kammler, M. Horn von Hoegen,

K. R. Wilson, D. von der Linde and C. P. J. Barty, Science, 1999, 286, 1340.7 A. Rousse, C. Rischel, S. Fourmaux, I. Uschmann, S. Sebban, G. Grillon, P. Balcou, E. Forster, J. P.

Geindre, P. Audebert, J. C. Gauthier and D. Hulin, Nature, 2001, 410, 65.8 F. Raksi, K. R. Wilson, J. Zhimig, A. Ikhlef, C. Y. Cote and J. C. Kieffer, J. Chem. Phys., 1996, 104, 6066.9 M. Bauer, C. Lei, K. Read, T. R. Tobey, J. Gland, M. M. Murnane and H. C. Kapteyn, Phys. Rev.Lett.,

2001, 87, 25 501.10 J. C. Williamson, C. Jianming, I. Hyotcherl, H. Frey and A. H. Zewail, Nature, 1997, 386, 159.11 H. Ihee, V. A. Lobastov, U. M. Gomez, B. M. Goodson, R. Srinivasan, C.-Y. Ruan and A. H. Zewail,

Science, 2001, 291, 458.12 C. Chudoba, E. T. J. Nibbering and T. Elsaesser, Phys. Rev. Lett., 1998, 81, 3010.13 E. T. J. Nibbering, C. Chudoba and T. Elsaesser, Isr. J. Chem., 1999, 39, 333.14 C. Chudoba, E.T. J. Nibbering and T. Elsaesser, J. Phys. Chem. A, 1999, 103, 5625.15 E. T. J. Nibbering and T. Elsaesser, Appl. Phys. B, 2000, 71, 439.16 E. T. J. Nibbering, F. Tschirschwitz, C. Chudoba and T. Elsaesser, J. Phys. Chem. A, 2000, 104, 4236.17 E. T. J. Nibbering and J. Dreyer, in Femtochemistry, ed. F. C. de Schryver, S. de Feyter and G.

Schweitzer, Wiley-VCH, Weinheim, Germany, 2001, p. 345.18 C. Chudoba, A. Kummrow, J. Dreyer, J. Stenger, E. T. J. Nibbering, T. Elsaesser and K. A. Zachariasse,

Chem. Phys. Lett., 1999, 309, 357.19 J. Dreyer and A. Kummrow, J. Am. Chem. Soc., 2000, 122, 2577.

12 Faraday Discuss., 2002, 122, 000–000 122/3

Page 15: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

INTRAMOLECULAR HYDROGEN TRANSFER 143

20 A. Kummrow, J. Dreyer, C. Chudoba, J. Stenger, E. T. J. Nibbering and T. Elsaesser, J. Chin. Chem.Soc., 2000, 47, 721.

21 H. Okamoto, J. Phys. Chem. A, 2000, 104, 4182.22 W. M. Kwok, C. Ma, P. Matousek, A. W. Parker, D. Phillips, W. T. Toner and M. Towrie, Chem. Phys.

Lett., 2000, 322, 395.23 W. M. Kwok, C. Ma, D. Phillips, P. Matousek, A. W. Parker and M. Towrie, J. Phys. Chem. A, 2000,

104, 4188.24 W. M. Kwok, C. Ma, P. Matousek, A. W. Parker, D. Phillips, W. T. Toner, M. Towrie and S. Umapathy,

J. Phys. Chem. A, 2001, 105, 984.25 J. Stenger, D. Madsen, J. Dreyer, E. T. J. Nibbering, P. Hamm and T. Elsaesser, J. Phys. Chem. A, 2001,

105, 2929.26 D. Madsen, J. Stenger, J. Dreyer, E. T. J. Nibbering, P. Hamm and T. Elsaesser, Chem. Phys. Lett., 2001,

341, 56.27 P. Hamm and R. M. Hochstrasser, inUltrafast Infrared and Raman spectroscopy, ed. M. D. Fayer, Marcel

Dekker, New York, 2001, p. 273.28 P. Hamm, M. Lim, W. F. DeGrado and R. M. Hochstrasser, J. Chem. Phys., 2000, 112, 1907.29 S. Woutersen and P. Hamm, J. Chem. Phys., 2001, 114, 2727.30 S. Woutersen and P. Hamm, J. Chem. Phys., 2001, 115, 7737.31 M. T. Zanni, S. Gnanakaran, J. Stenger and R. M. Hochstrasser, J. Phys. Chem. B, 2001, 105, 6520.32 S. Woutersen, Y. Mu, G. Stock and P. Hamm, Proc.Natl. Acad. Sci. USA, 2001, 98, 11 254.33 M. T. Zanni, N.-H. Ge, Y. S. Kim and R. M. Hochstrasser, Proc. Natl. Acad. Sci. USA, 2001, 98, 11 265.34 M. T. Zanni, M. C. Asplund and R. M. Hochstrasser, J. Chem. Phys., 2001, 114, 4579.35 S. Woutersen, Y. Mu, G. Stock and P. Hamm, Chem. Phys., 2001, 266, 137.36 O. Golonzka, M. Khalil, N. Demirdoven and A. Tokmakoff, J. Chem. Phys., 2001, 115, 10 814.37 N. Demirdoven, M. Khalil, O. Golonzka and A. Tokmakoff, J. Phys. Chem. A, 2001, 105, 8025.38 R. Pomes and B. Roux, J. Phys. Chem., 1996, 100, 2519.39 M. E. Tuckerman, D. Marx, M. L. Klein and M. Parrinello, Science, 1997, 275, 179.40 P. L. Geissler, C. Dellago, D. Chandler, J. Hutter and M. Parrinello, Science, 2001, 291, 2121.41 R. A. Mathies, S. W. Lin, J. B. Ames and W. T. Pollard, Annu. Rev. Biophys. Biophys. Chem., 1991,

20, 491.42 D. Borgis and J. T. Hynes, J. Chem. Phys., 1991, 94, 3619.43 D. Borgis and J. T. Hynes, Chem. Phys., 1993, 170, 315.44 D. Hadzi and S. Bratos, in The Hydrogen Bond: Recent Developments in Theory and Experiments, ed.

P. Schuster, G. Zundel, and C. Sandorfy, North Holland, Amsterdam, 1976, vol. II, p. 565.45 S. Woutersen, U. Emmerichs and H. J. Bakker, Science, 1997, 278, 658.46 S. Woutersen, U. Emmerichs, H.-K. Nienhuys and H. J. Bakker, Phys. Rev. Lett., 1998, 81, 1106.47 H. K. Nienhuys, S. Woutersen, R. A. van Santen and H. J. Bakker, J. Chem. Phys., 1999, 111, 1494.48 G. M. Gale, G. Gallot and N. Lascoux, Chem. Phys. Lett., 1999, 311, 123.49 J. Stenger, D. Madsen, P. Hamm, E. T. J. Nibbering and T. Elsaesser, Phys. Rev. Lett., 2001, 87, 27 401.50 J. Stenger, D. Madsen, P. Hamm, E. T. J. Nibbering and T. Elsaesser, J. Phys. Chem. A, 2002, in press.51 R. Laenen, C. Rauscher and A. Laubereau, Chem. Phys. Lett., 1998, 283, 7.52 L. K. Iwaki and D. D. Dlott, J. Phys. Chem. A, 2000, 104, 9101.53 J. C. Deak, S. T. Rhea, L. K. Iwaki and D. D. Dlott, J. Phys. Chem. A, 2000, 104, 4866.54 L. K. Iwaki and D. D. Dlott, Chem. Phys. Lett., 2000, 321, 419.55 S. Woutersen and H. J. Bakker, Nature, 1999, 402, 507.56 R. Laenen, C. Rauscher and A. Laubereau, Phys. Rev. Lett., 1998, 80, 2622.57 R. Laenen, C. Rauscher and A. Laubereau, J. Phys. Chem. B, 1998, 102, 9304.58 G. M. Gale, G. Gallot, F. Hache, N. Lascoux, S. Bratos and J. C. Leicknam, Phys. Rev. Lett., 1999,

82, 1068.59 S. Woutersen and H. J. Bakker, Phys. Rev. Lett., 1999, 83, 2077.60 S. Bratos, G. M. Gale, G. Gallot, F. Hache, N. Lascoux and J. C. Leicknam, Phys. Rev. E, 2000, 61, 5211.61 G. Gallot, N. Lascoux, G. M. Gale, J. Leicknam, S. Bratos and S. Pommeret, Chem. Phys. Lett., 2001,

341, 535.62 A. J. Lock, S. Woutersen and H. J. Bakker, J. Phys. Chem. A, 2001, 105, 1238.63 H. K. Nienhuys, R. A. van Santen and H. J. Bakker, J. Chem. Phys., 2000, 112, 8487.64 H. J. Bakker, S. Woutersen and H. K. Nienhuys, Chem. Phys., 2000, 258, 233.65 Proton-transfer Reactions, ed. E. Caldin and V. Gold, Chapman and Hall, London, 1975.66 A. Weller, Die Naturwissenschaften, 1955, 42, 175.67 A. Weller, Progr. React. Kinet., 1961, 1, 187.68 M. Eigen, Angew. Chem. Int. Ed. Engl., 1964, 3, 1.69 E. Vander Donckt, Progr. React. Kinet., 1970, 5, 273.70 W. Klopffer, Adv. Photochem., 1977, 10, 311.71 M. Kasha, J. Chem. Soc., FaradayTrans. 2, 1986, 82, 2379.72 E. M. Kosower and D. Huppert, Annu. Rev. Phys. Chem., 1986, 37, 127.73 E. Pines, B. Z. Magnes, M. J. Lang and G. R. Fleming, Chem. Phys. Lett., 1997, 281, 413.

122/3 Faraday Discuss., 2002, 122, 000–000 13

Page 16: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

144 RINI ■ KUMMROW ■ DREYER ■ NIBBERING ■ ELSAESSER

74 D. Huppert, L. M. Tolbert and S. Linares-Samaniego, J. Phys. Chem. A, 1997, 101, 4602.75 L. Genosar, B. Cohen and D. Huppert, J. Phys. Chem. A, 2000, 104, 6689.76 T. H. Tran-Thi, T. Gustavsson, C. Prayer, S. Pommeret and J. T. Hynes, Chem. Phys. Lett., 2000,

329, 421.77 P. F. Barbara, L. E. Brus and P. M. Rentzepis, J. Am. Chem. Soc., 1980, 102, 5631.78 K. Ding, S. J. Courtney, A. J. Strandjord, S. Flom, D. Friedrich and P. F. Barbara, J. Phys. Chem., 1983,

87, 1184.79 S. R. Flom and P.F. Barbara, Chem. Phys. Lett., 1983, 94, 488.80 D. B. O’Connor, G. W. Scott, D. R. Coulter, A. Gupta, S. P. Webb, S. W. Yeh and J. H. Clark, Chem.

Phys. Lett., 1985, 121, 417.81 M. Lee, Y. J.T. and R. M. Hochstrasser, J. Phys. Chem., 1987, 91, 4621.82 F. Laermer, T. Elsaesser and W. Kaiser, Chem. Phys. Lett., 1988, 148, 119.83 W. Frey, F. Laermer and T. Elsaesser, J. Phys. Chem., 1991, 95, 10 391.84 B. J. Schwartz, L. A. Peteanu and C. B. Harris, J. Phys. Chem., 1992, 96, 3591.85 T. Elsaesser, in Femtosecond Chemistry, ed. J. Manz and L. Woste, Wiley-VCH, Weinheim, Germany,

1995, vol. 2, p. 563.86 T. Fiebig, M. Chachisvilis, M. Manger, A. H. Zewail, A. Douhal, I. Garcia-Ochoa and A. de La Hoz

Ayuso, J. Phys. Chem. A, 1999, 103, 7419.87 S. Ameer-Beg, S. M. Ormson, R. G. Brown, P. Matousek, M. Towrie, E. T. J. Nibbering, P. Foggi and

F. V. R. Neuwahl, J. Phys. Chem. A, 2001, 105, 3709.88 C. Chudoba, E. Riedle, M. Pfeiffer and T. Elsaesser, Chem. Phys. Lett., 1996, 263, 622.89 M. Pfeiffer, K. Lenz, A. Lau, T. Elsaesser and T. Steinke, J. Raman Spectrosc., 1997, 28, 61.90 T. Elsaesser and W. Kaiser, Chem. Phys. Lett., 1986, 128, 231.91 S. Lochbrunner, A. J. Wurzer and E. Riedle, J. Chem. Phys., 2000, 112, 10 699.92 A. J. Wurzer, S. Lochbrunner and E. Riedle, Appl. Phys. B, 2000, 71, 405.93 A. Kummrow, M. Wittmann, F. Tschirschwitz, G. Korn and E. T. J. Nibbering, Appl. Phys. B, 2000,

71, 885.94 P. Hamm, R. A. Kaindl and J. Stenger, Opt. Lett., 2000, 25, 1798.95 K. Wynne and R. M. Hochstrasser, Chem. Phys., 1995, 193, 211.96 P. Hamm, Chem. Phys., 1995, 200, 415.97 E. E. Nikitin, C. Noda and R. N. Zare, J. Chem. Phys., 1993, 98, 46.98 K. Iwata and H. Hamaguchi, J. Phys. Chem. A, 1997, 101, 632.99 P. Hamm, S. M. Ohline and W. Zinth, J. Chem. Phys., 1997, 106, 519.

100 P.-T. Chou, S. L. Studer and M. L. Martinez, Chem. Phys. Lett., 1991, 178, 393.101 M. A. Rıos and M. C. Rıos, J. Phys. Chem. A, 1998, 102, 1560.102 A. Laubereau and W. Kaiser, Rev. Mod. Phys., 1976, 50, 607.103 A. Seilmeier and W. Kaiser, in Ultrashort Laser Pulses. Generation and Applications, ed. W. Kaiser,

Springer, Berlin, 2nd edn., 1993, p. 279.104 T. Elsaesser and W. Kaiser, Annu. Rev. Phys. Chem., 1991, 42, 83.105 J. C. Owrutsky, D. Raftery and R. M. Hochstrasser, Annu. Rev. Phys. Chem., 1994, 45, 519.106 S. Hogiu, W. Werncke, M. Pfeiffer and T. Elsaesser, Chem. Phys. Lett., 1999, 312, 407.107 S. Hogiu, W. Werncke, M. Pfeiffer, J. Dreyer and T. Elsaesser, J. Chem. Phys., 2000, 113, 1587.108 A. P. Shkurinov, N. I. Koroteev, G. Jonusauskas and C. Rulliere, Chem. Phys. Lett., 1994, 223, 573.109 A. Vierheilig, T. Chen, P. Waltner, W. Kiefer, A. Materny and A. H. Zewail, Chem. Phys. Lett., 1999,

312, 349.110 M. Pfeiffer, K. Lenz, A. Lau and T. Elsaesser, J. Raman Spectrosc., 1995, 26, 607.111 M. J. Rosker, F. W. Wise and C. L. Tang, Phys. Rev. Lett., 1986, 57, 321.112 S. Ruhman, A. G. Joly and K. A. Nelson, J. Chem. Phys., 1987, 86, 6563.113 Y.-X. Yan and K. A. Nelson, J. Chem. Phys., 1987, 87, 6240.114 Y.-X. Yan and K. A. Nelson, J. Chem. Phys., 1987, 87, 6257.115 H. L. Fragnito, J. Y. Bigot, P. C. Becker and C. V. Shank, Chem. Phys. Lett., 1989, 160, 101.116 W. T. Pollard and R. A. Mathies, Annu. Rev. Phys. Chem., 1992, 43, 497.117 M. H. Vos, F. Rappaport, J.-C. Lambry, J. Breton and J.-L. Martin, Nature, 1993, 363, 320.118 T. S. Yang, M. S. Chang, R. Chang, M. Hayashi, S. H. Lin, P. Vohringer, W. Dietz and N. F. Scherer,

J. Chem. Phys., 1999, 110, 12 070.119 R. M. Bowman, M. Dantus and A. H. Zewail, Chem. Phys. Lett., 1989, 156, 131.120 U. Banin, A. Waldman and S. Ruhman, J. Chem. Phys., 1992, 96, 2416.121 N. Pugliano, D. K. Palit, A. Z. Szarka and R. M. Hochstrasser, J. Chem. Phys., 1993, 99, 7273.122 Q. Wang, R. W. Schoenlein, L. A. Peteanu, R. A. Mathies and C. V. Shank, Science, 1994, 266, 422.123 T. Kuhne and P. Vohringer, J. Phys. Chem. A, 1998, 102, 4177.

14 Faraday Discuss., 2002, 122, 000–000 122/3

Page 17: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

INTRAMOLECULAR HYDROGEN TRANSFER 145

Ultrafast vibrational relaxation processes induced byintramolecular excited state hydrogen transfer

Matteo Rini, Jens Dreyer, Erik T.J. Nibbering, Thomas Elsaesser *

Max-Born-Institut f€uur Nichtlineare Optik und Kurzzeitspektroskopie, Max-Born-Str. 2 A, D-12489 Berlin, Germany

Received 21 February 2003; in final form 10 April 2003

Abstract

Vibrational spectra between 1000 and 1700 cm�1 are studied after ultrafast transfer of a hydrogen atom in the

excited state of 2-(20-hydroxyphenyl)benzothiazole. Femtosecond pump–probe experiments reveal new vibrational

bands of the keto-S1 state, including the carbonyl stretching band formed by hydrogen transfer. Such bands display a

negligible spectral reshaping but blue-shift by up to 7 cm�1 following biexponential kinetics with time constants of

700 fs and 15 ps. The blue-shift is attributed to the anharmonic coupling of the fingerprint vibrations to Raman-active

low-frequency modes that are excited upon electronic excitation and depopulated by intramolecular redistribution and

cooling to the solvent.

� 2003 Elsevier Science B.V. All rights reserved.

1. Introduction

Photoinduced transfer of a hydrogen atom or a

proton in electronically excited states plays a key

role for the function of a variety of molecular sys-tems, among them ultraviolet stabilizers of poly-

mers, photoacids in liquid and/or protein

environment, as well as reactive biological systems

[1]. Intramolecular hydrogen transfer in electroni-

cally excited states, where a hydrogen atom is

moving from its initial binding site to another

functional group in the same molecule, has received

substantial interest for studying the basic micro-scopic mechanisms behind such reactions. In mol-

ecules like 2-(20-hydroxyphenyl)benzothiazole (HBT),

ultraviolet photoexcitation of the initial enol spe-

cies (I) results in a femtosecond hydrogen transfer

creating a new keto-type geometry (II). This

structural change is evident from the carbonylstretching band of the keto tautomer as observed in

picosecond experiments probing the infrared spec-

tra of the keto-S1 state after ultraviolet excitation

of the enol tautomer [2]. Recently, a formation time

of the carbonyl band and – thus – the keto tautomer

of 30–50 fs was determined in femtosecond ultra-

violet pump/mid-infrared probe studies [3].

Chemical Physics Letters 374 (2003) 13–19

www.elsevier.com/locate/cplett

* Corresponding author. Fax: +49-30-6392-1409.

E-mail address: [email protected] (T. Elsaesser).

0009-2614/03/$ - see front matter � 2003 Elsevier Science B.V. All rights reserved.

doi:10.1016/S0009-2614(03)00650-X

Page 18: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

146 RINI ■ KUMMROW ■ DREYER ■ NIBBERING ■ ELSAESSER

The formation kinetics of the keto species have

been investigated in a less direct way by probing

femtosecond changes of electronic spectra, in par-

ticular the rise of stimulated emission from the keto

tautomer [4–6]. Such experiments show an emis-

sion rise on a 60 to 170 fs time scale depending onthe spectral position within the keto emission band.

Superimposed on this incoherent kinetics are pro-

nounced oscillations with frequency components of

118, 254, 289, and 529 cm�1 [5,6]. Such oscillations

which have been observed before in other hydrogen

transfer systems [7], persist for 2–3 ps and reflect

coherent wavepacket motions along several Ra-

man-active low-frequency modes which couplestrongly to the enol S0–S1 transition and are elon-

gated upon electronic excitation. Based on such

findings and theoretical ab initio calculations of

potential energy surfaces, microscopic pictures for

�barrierless� hydrogen transfer have been developedin which coherent motions along one or several

low-frequency modes promote the reaction.

The range of the excited state potential popu-lated by the excitation on the enol S0–S1 transition

has an excess energy of several thousand cm�1 with

respect to the keto minimum. This excess energy is

redistributed during and after hydrogen transfer

and a vibrationally relaxed keto product species is

formed. So far, the time scales and pathways of

this redistribution process which may play a key

role for the irreversible character of hydrogentransfer, are not understood. In particular, the

extent to which vibrations of the carbonyl and the

N–H groups formed by hydrogen transfer accept

excess energy, has remained unknown.

In this Letter, we study vibrational redistribu-

tion processes during and after excited state hy-

drogen transfer in HBT by femtosecond infrared

spectroscopy. After electronic excitation by a sub-40 fs ultraviolet pulse, transient infrared spectra of

HBT in the keto-S1 state are measured in the fin-

gerprint region between 1000 and 1700 cm�1. The

different vibrational bands display a blue-shift by

2–7 cm�1 to their final spectral position which is

reached within 50–60 ps. The blue-shift follows

nonexponential kinetics with a fast subpicosecond

component followed by slower kinetics. The blue-shift is caused by transient populations of Raman-

active low-frequency modes which couple anhar-

monically to the modes in the fingerprint region.

During the first few picoseconds, such low-fre-

quency modes contain the major fraction of vi-

brational excess energy which subsequently flows

into a multitude of modes cooling down on a 50–

60 ps time scale.

2. Experimental

The experiments are based on an ultraviolet

pump/mid-infrared probe scheme where the pump

pulse initiates the intramolecular hydrogen transfer

by exciting HBT molecules to the S1 state. The re-sulting change of vibrational absorption is moni-

tored by weak mid-infrared probe pulses which are

spectrally dispersed after interaction with the

sample. The experimental setup which has been

described in detail previously [3], is based on a

home-built amplified Ti:sapphire laser. Ultraviolet

pulses tunable from 310 to 350 nm were generated

by sum-frequency mixing of the fundamental of theTi:sapphire laser with visible pulses obtained by

noncollinear optical parametric generation [8]. The

pulse energy was 2–3 lJ and the pulse duration, asdetermined by self-diffraction (third-order) auto-

correlation measurements was below 40 fs (repeti-

tion rate 1 kHz). The pump pulses excited about

10% of the HBT molecules in a sample volume of

180 lm diameter. Mid-infrared pulses tunable be-tween 1000 and 2000 cm�1 were generated by

phase-matched difference frequency mixing in

GaSe of near-infrared signal and idler pulses from

an optical parametric amplifier [9]. Probe and ref-

erence pulses, the latter derived at a BaF2 wedge

and passing an unexcited area of the sample, were

focused onto the sample with a spot diameter of 150

lm. After interaction with the sample, probe andreference pulses were dispersed in a grating spec-

trometer and detected by a cooled double HgCdTe

detector array with 2� 32 elements (spectral reso-lution 4 cm�1). Normalization of probe and refer-

ence signals and synchronous chopping of the

pump allowed for measuring absorbance changes

as small as DA ¼ 0:2 mOD. The delay zero, the timeresolution and the chirp of the infrared pulses weredetermined from the frequency resolved cross-cor-

relation function of pump and probemeasured with

14 M. Rini et al. / Chemical Physics Letters 374 (2003) 13–19

Page 19: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

INTRAMOLECULAR HYDROGEN TRANSFER 147

a thin ZnSe crystal. The time resolution varied be-

tween 100 and 150 fs with increasing values from

high to low frequencies.

HBTwas dissolved in the nonpolar solventC2Cl4with a concentration of 0.02 M and pumped

through a free streaming jet of a nominal thicknessof 100 lm. Under such conditions, the enol species(I) with an intramolecular hydrogen bond repre-

sents the predominant ground state species [10].

Measurements in the neat solvent were performed

under the same experimental conditions. The

pump–probe set-upwas purgedwith nitrogen gas to

avoid spectral and temporal reshaping of the mid-

infrared pulses due to water and CO2 absorption.

3. Results

The present study concentrates on the finger-

print region of the vibrational spectrum from 1000

to 1700 cm�1. The structure and the vibrational

spectra of HBT in the enol ground state were

calculated using the B3LYP/6-31G(d) level of

theory implemented in GAUSSIANAUSSIAN 98 [11]. The

calculated bands are listed in Table 1, Fig. 1 shows

a comparison of the calculated and the measured

infrared spectrum of HBT in C2Cl4. The very goodagreement allows for the assignment of essentially

all the lines. The following bands are relevant in

terms of hydrogen transfer forming the keto tau-

tomer: m(phOH-19b) + m(C–O) + d(OH)+ m(C@N)¼ 1483 cm�1, m(phBT-19a) + m(phOH-19a) + m(C @N)+ d(OH)+ m(C–O)¼ 1460 cm�1, d(phOH:CH-3)+d(OH)¼ 1406 cm�1, m(phOH-14)+ m(C–O)¼ 1318cm�1 and m(C–O, 13)¼ 1272 cm�1.In Fig. 2a, we present transient infrared spectra

of HBT recorded at different time delays between

the pump pulses at 330 nm and the infrared probe

pulses. The measured change of absorbance

DAðmÞ ¼ � logf½T ðmÞ � T0ðmÞ=T0ðmÞg is plotted as afunction of probe frequency, where T ðmÞ and T0ðmÞare the infrared transmissions of the sample with

Table 1

Experimental (in C2Cl4) and calculated (B3LYP/6- 31G(d), scaled by 0.965) infrared bands for HBT in the enol ground state in the

fingerprint region (in cm�1)

Experimental Calculation Assignmenta

1624 1622 m(phOH-8a)+ d(OH)1593 m(phBT-8a)

1588 1581 m(phOH-8b)+ d(OH)1561 1553 m(phBT-8b)1504 –

1492 1508 m(C@N)+ d(OH)1483 1488 m(phOH-19b) + m(C–O) + d(OH) + m(C@N)1460 1456 m(phBT-19a) + m(phOH-19a) + m(C@N) + d(OH) + m(C–O)

1448 m(phBT-19a)) m(phOH-19a) + m(C@N) + d(OH) + m(C–O)1441 1433 m(phBT-19b)+ m(C@N)1406 1412 d(phOH: CH-3) + d(OH)1318 1323+ m(phOH-14)+ m(C–O)

1318 m(phBT-14)1272 1275 m(C–O, 13)1254 1250 d(phBT:CH-3)1242 1234 d(phBT:CH-9a)+ m(phBT–N)1217 1205 d(OH)+ m(phOH–CS@N)1155 1151 d(phOH:CH-9a)1128 1117 d(phBT:CH-18a)1122 1112 d(phOH:CH-18a)1071 1047 d(phOH-12) + m(phBT-S)1035 1029 d(phOH:CH-18b)+ m(phOH-1)1017 1007 d(phBT:CH-18b) + m(phBT-1)

aVarsanyi nomenclature [22], m¼ stretching mode, d¼ in-plane deformation mode, phOH¼ phenol ring, phBZ¼benzene ring in thebenzothiazole fragment.

M. Rini et al. / Chemical Physics Letters 374 (2003) 13–19 15

Page 20: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

148 RINI ■ KUMMROW ■ DREYER ■ NIBBERING ■ ELSAESSER

and without excitation, respectively. Fig. 2b shows

DAðmÞ between 1350 and 1580 cm�1 on an extended

scale. The data represent an average of 10–20 time

traces each taking 512 shots per delay step. The

negative contributions to DAðmÞ originate from thebleaching of ground state modes upon ultravioletexcitation. Such bleaching signals occur already at

negative time delays, as is evident from time-re-

solved measurements at a fixed probe frequency

(Fig. 3a). The absorbance change at negative delay

times is due to the perturbed free induction decay

(PFID) of the coherent polarization created on the

vibrational transition by the probe and disturbed

at later times by the pump pulse [12]. At positivedelays, the bleaching signals are nearly constant

for tens of picoseconds. 1

The positive signals in Fig. 2 represent the vi-

brational spectrum of the keto-S1 state. The band

at 1530 cm�1 is caused by the C@O stretching

absorption of the keto tautomer. With respect to a

free C@O vibration, the frequency is lower becausethe carbonyl group is part of a hydrogen bond andconjugated to a C@C bond [3,13]. Recently, a

formation time of this band and – thus – of the

keto tautomer of 30–50 fs was determined [3]. In

Fig. 3b, the absorption change at this spectral

position is plotted up to picosecond delay times.Apart from a slight broadening by 2–3 cm�1 at

early delay times, the spectral envelopes of the new

bands are time-independent. However, all new

bands occur initially at spectral positions which

are 5–7 cm�1 down-shifted from the positions

reached after 50–60 ps. To quantify such spectral

shifts, Gaussian envelopes were fitted to the mea-

sured bands and the line centres determined in this

Fig. 1. Ground state absorption spectrum of HBT in C2Cl4compared with results of B3LYP/6-31G(d) calculations in the

fingerprint region. The calculated lines are listed in Table 1.

Fig. 2. (a) Transient infrared spectra of HBT measured at delay

times of )1 ps (dotted line), 300 fs (dash-dotted line) and 50 ps(solid line) after electronic excitation. The change of infrared

absorbance DAðmÞ ¼ � logf½T ðmÞ � T0ðmÞ=T0ðmÞg is plotted as afunction of infrared probe frequency (T ðmÞ; T0ðmÞ: transmissionwith and without excitation). (b) Transient spectra measured at

300 fs (dotted line), 3 ps (dash-dotted line), and 50 ps (solid line)

after electronic excitation on an extended frequency scale from

1350 to 1550 cm�1.

1 The peak around delay zero is observed at all probe

frequencies and due to the nonlinearity of the solvent C2Cl4.

16 M. Rini et al. / Chemical Physics Letters 374 (2003) 13–19

Page 21: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

INTRAMOLECULAR HYDROGEN TRANSFER 149

way are plotted as a function of delay time (Figs.3c and 4). The blue-shift of all bands displays

nonexponential kinetics with a fast subpicosecond

and a slower kinetics. The observed time traces are

well accounted for by fitting biexponential kinetics

to the data with time constants of 600–750 fs and

15 ps. Such time constants are substantially

shorter than the lifetime of the keto-S1 state of

approximately 300 ps [2].

4. Discussion

We first address the decrease of absorption in

the transient infrared spectra which is due to the

bleaching of vibrational bands of the enol ground

state of HBT (cf. Fig. 2). The bands at 1272,

1318, 1460 and 1483 cm�1 possess, according tothe calculations, some C–O single bond character

which should disappear upon formation of the

C@O double bond in the keto tautomer. At po-sitive time delays, the analysis is complicated by

the fact that new keto-S1 bands are created in the

same spectral region (Fig. 2a). At negative time

delays, however, the time-resolved data measured

at these frequencies show typical PFID features,i.e., a slow rise of bleaching with time, confirming

that the original bands bleach after ultraviolet

excitation (Fig. 2a, dotted line). These observa-

tions together with the formation of the C@Ostretching band around 1530 cm�1 give direct

evidence of the formation of the keto geometry of

HBT.

The new (positive) vibrational bands in thespectra of Fig. 2 mark the keto-S1 state. All new

bands display blue-shifts with time without un-

dergoing a significant reshaping and/or broaden-

ing of their spectral envelopes. We attribute such

new bands to the v ¼ 0! 1 transition of the re-

spective vibration. Transient populations of vP 1

levels of such modes would lead to stimulated

emission and/or absorption contributions to thepump–probe signals. Attributing the spectral

width of the measured lines to (homogeneous)

lifetime broadening, a lower limit to the popula-

tion lifetimes of vP 1 levels of 200–400 fs is es-

timated, i.e., populations of vP 1 levels could

easily be detected in our experiment. As all modes

in the fingerprint region display a finite anhar-

monicity on the order of 10–30 cm�1 [14], stim-ulated emission and/or absorption from vP 1

levels would give rise to additional spectral

Fig. 3. Temporal evolution of the absorption changes at spec-

tral positions of (a) 1483 cm�1 (bleaching of a ground state

vibration) and (b) 1530 cm�1 (formation of the carbonyl

stretching vibration) as a function of the delay time between

pump and probe. (c) Line centre of the C@O band as a functionof time delay (symbols) and biexponential fit (solid line) with

the following time constants and relative amplitudes: s1 ¼610� 80 fs, A1 ¼ 0:28; s2 ¼ 14:5� 1 ps, A2 ¼ 0:72.

Fig. 4. Long time behaviour of the line centres (symbols) of the

transient bands at (a) 1530, (b) 1396, and (c) 1301 cm�1 as a

function of delay time. The data follow biexponential kinet-

ics (solid lines) with the following time constants si andrelative amplitudes Ai ði ¼ 1; 2Þ: (a) s1 ¼ 610� 80 fs, A1 ¼0:28; s2 ¼ 14:5� 1 ps, A2 ¼ 0:72, (b) s1 ¼ 750� 80 fs,

A1 ¼ 0:35; s2 ¼ 15:6� 0:9 ps, A2 ¼ 0:65, (c) s1 ¼ 590� 100 fs,A1 ¼ 0:13; s2 ¼ 14:9� 1:1 ps, A2 ¼ 0:87.

M. Rini et al. / Chemical Physics Letters 374 (2003) 13–19 17

Page 22: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

150 RINI ■ KUMMROW ■ DREYER ■ NIBBERING ■ ELSAESSER

components, different from the v ¼ 0! 1 transi-

tion. We conclude from the absence of such fea-

tures that nonequilibrium populations of vP 1

levels are negligible. All such modes are created

in their v ¼ 0 ground state, i.e., are not elongatedand do not accept vibrational excess energy uponhydrogen transfer.

The blue-shift of the central frequencies of all

new bands by 2–7 cm�1 within 50–60 ps is due to

transient populations in other modes that couple

anharmonically to the fingerprint vibrations of the

keto tautomer. This behaviour has also been ob-

served in azobenzene [15] and trans-stilbene [16]. It

was rationalized by a model describing the spectralposition of vibrational lines as a function of vi-

brational populations of anharmonically coupled

modes [15]. For negative values of the off-diagonal

anharmonic coupling constants, which is the case

for most vibrations, this model predicts a red-

shifted v ¼ 0! 1 transition of a mode k for high

excess populations of modes i ði 6¼ kÞ coupling tothe mode k, and a subsequent blue-shift to thesteady-state spectral position with decreasing

population.

The transient spectra in the fingerprint region

give no direct insight into which anharmonically

coupled modes underlie the blue-shift of the dif-

ferent bands. There is, however, detailed infor-

mation from steady-state resonance Raman

spectra of enol-HBT and from femtosecond stud-ies of electronic transitions: (i) The resonance

Raman spectra give evidence of several modes at

frequencies below 700 cm�1 that are strongly

coupled to the S0–S1 transition of the enol tauto-

mer and display substantial anharmonicities giving

rise to combination bands in the Raman spectrum

[17,18]. Upon femtosecond excitation, such modes

are elongated, i.e., acquire a substantial nonequi-librium population. Some of them, the vibrations

at 118, 254, 289, and 530 cm�1, display coherent

oscillations within the first few picoseconds [5,6].

Such oscillations represent the coherent motion of

vibrational wavepackets which are made up of a

coherent superposition of at least the v ¼ 1 andv ¼ 2 eigenstates of the respective oscillator, i.e.,contain at least two vibrational quanta of the re-spective mode. This gives a minimum excess en-

ergy in such modes of about 2200 cm�1 which is a

major fraction of the total excess energy of �3000cm�1 released upon hydrogen transfer [19]. These

findings strongly suggest that the blue-shift of the

vibrational bands in the fingerprint region is

mainly due to excess populations of such Raman

active low-frequency vibrations.Coherent wavepacket motions in the anhar-

monically coupled low-frequency modes result in a

periodic modulation of the spectral position of

high-frequency modes. For the O–H stretching

mode in the enol ground state of HBT, oscillations

with an amplitude on the order of 10�4 OD have

been observed [20]. In the measured blue-shift of

the fingerprint vibrations (Figs. 3c and 4), oscil-lations due to Raman-active modes are absent

because of the limited spectral resolution and

sensitivity of the present experiment which in fact

averages over the spectral modulation. Conse-

quently, we observe a continuous blue-shift of the

bands with time constants of 700 fs and 15 ps,

reflecting the decreasing excess populations of the

low-frequency modes. The fast component is mostprobably due to intramolecular redistribution

processes by which the excess populations are

randomized within the excited HBT molecules

[21]. It is interesting to note that the damping time

T2 ¼ 1–2 ps of the coherent oscillations observedin [5,6] is about twice the redistribution time

T1 � 700 fs found here, as is expected for a deph-asing dominated by the population relaxation ofexcited vibrational levels. The slower component

in the blue-shift is attributed to a flow of excess

energy from the HBT molecule in the S1 state into

the liquid surroundings [21].

In conclusion, we have studied vibrational re-

laxation processes after excited state intramolecu-

lar hydrogen transfer in HBT by measuring

transient vibrational spectra in the fingerprint re-gion between 1000 and 1700 cm�1. The measured

vibrational bands of the keto-S1 state display a

transient blue-shift following nonexponential ki-

netics. The absence of any significant spectral re-

shaping of these bands indicates the absence of

strong nonequilibrium excess populations in any

of the fingerprint vibrations which play a minor

role for energy redistribution. The blue-shift ismainly caused by transient populations of Raman-

active low-frequency modes which couple anhar-

18 M. Rini et al. / Chemical Physics Letters 374 (2003) 13–19

Page 23: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

INTRAMOLECULAR HYDROGEN TRANSFER 151

monically to the fingerprint modes. We attribute

the fast component of the blue-shift to the intra-

molecular redistribution of vibrational excess en-

ergy, followed by a slower picosecond transfer of

excess energy to the solvent.

Acknowledgements

We gratefully acknowledge financial support by

the Deutsche Forschungsgemeinschaft (project NI

492/2-2).

References

[1] A. Douhal, S.K. Kim, A.H. Zewail, Nature 378 (1995)

260.

[2] T. Elsaesser, W. Kaiser, Chem. Phys. Lett. 128 (1986)

231.

[3] M. Rini, A. Kummrow, J. Dreyer, E.T.J. Nibbering, T.

Elsaesser, Faraday Discuss. 122 (2003) 27.

[4] F. Laermer, T. Elsaesser, W. Kaiser, Chem. Phys. Lett. 148

(1988) 119.

[5] S. Lochbrunner, A.J. Wurzer, E. Riedle, J. Chem. Phys.

112 (2000) 10699.

[6] E. Riedle, S. Lochbrunner, A.J. Wurzer, V. de Waele, R. de

Vivie- Riedle, in: T. Elsaesser et al. (Eds.), Ultrafast

Phenomena XII, Springer, Berlin, 2001, p. 645.

[7] C. Chudoba, E. Riedle, M. Pfeiffer, T. Elsaesser, Chem.

Phys. Lett. 263 (1996) 622.

[8] A. Kummrow, M. Whitmann, F. Tschirchwitz, G. Korn,

E.T.J. Nibbering, Appl. Phys. B 71 (2000) 885.

[9] R.A. Kaindl, M. Wurm, K. Reimann, P. Hamm, A.M.

Weiner, M. Woerner, J. Opt. Soc. Am. B 17 (2000) 2086.

[10] T. Elsaesser, B. Schmetzer, Chem. Phys. Lett. 140 (1987)

293.

[11] M.J. Frisch et al., GAUSSIANAUSSIAN 98 (Revision A.2), Gaussian

Inc., Pittsburgh, PA.

[12] K. Wynne, R.M. Hochstrasser, Chem. Phys. 193 (1995)

211.

[13] K. Hafner, H.E.A. Kramer, H. Musso, G. Ploss, G.

Schulz, Chem. Ber. 97 (1964) 2066.

[14] P. Hamm, R.M. Hochstrasser, in: M. Fayer (Ed.), Ultra-

fast Infrared and Raman Spectroscopy, Marcel Dekker,

New York, 2001, p. 273.

[15] P. Hamm, S.M. Ohline, W. Zinth, J. Chem. Phys. 106

(1997) 519.

[16] K. Iwata, H. Hamaguchi, J. Phys. Chem. A 101 (1997) 632.

[17] M. Pfeiffer, K. Lenz, A. Lau, T. Elsaesser, J. Raman

Spectrosc. 26 (1995) 607.

[18] M. Pfeiffer, K. Lenz, A. Lau, T. Elsaesser, T. Steinke,

J. Raman Spectrosc. 28 (1997) 61.

[19] P.T. Chou, S.L. Studer, M.L. Martinez, Chem. Phys. Lett.

178 (1991) 393.

[20] D. Madsen, J. Stenger, J. Dreyer, E.T.J. Nibbering, P.

Hamm, T. Elsaesser, Chem. Phys. Lett. 341 (2001) 56.

[21] T. Elsaesser, W. Kaiser, Annu. Rev. Phys. Chem. 42 (1991)

83.

[22] G. Varsanyi, Vibrational Spectra of Benzene Derivatives,

Academic Press, New York, London, 1969.

M. Rini et al. / Chemical Physics Letters 374 (2003) 13–19 19

Page 24: Chapter 5 Excited State Intramolecular Hydrogen Transfer ...staff.mbi-berlin.de/nibberin/duitse_tijd_berlin/... · M. Rini, J. Dreyer, E. T. J. Nibbering and T. Elsaesser, Chemical

152 RINI ■ KUMMROW ■ DREYER ■ NIBBERING ■ ELSAESSER