femtosecond transient infrared and stimulated raman … · the journal of chemical physics 142,...

12
Femtosecond transient infrared and stimulated Raman spectroscopy shed light on the relaxation mechanisms of photo-excited peridinin Mariangela Di Donato, Elena Ragnoni, Andrea Lapini, Paolo Foggi, Roger G. Hiller, and Roberto Righini Citation: The Journal of Chemical Physics 142, 212409 (2015); doi: 10.1063/1.4915072 View online: http://dx.doi.org/10.1063/1.4915072 View Table of Contents: http://scitation.aip.org/content/aip/journal/jcp/142/21?ver=pdfcov Published by the AIP Publishing Articles you may be interested in Fragmentation mechanism of UV-excited peptides in the gas phase J. Chem. Phys. 141, 154309 (2014); 10.1063/1.4897158 Selective bond breaking mediated by state specific vibrational excitation in model HOD molecule through optimized femtosecond IR pulse: A simulated annealing based approach J. Chem. Phys. 139, 034310 (2013); 10.1063/1.4813127 Photodissociation of CH 3 Cl , C 2 H 5 Cl , and C 6 H 5 Cl on the Ag(111) surface: Ab initio embedded cluster and configuration interaction study J. Chem. Phys. 132, 074707 (2010); 10.1063/1.3322289 Femtosecond time-resolved absorption anisotropy spectroscopy on 9 , 9 ′ -bianthryl: Detection of partial intramolecular charge transfer in polar and nonpolar solvents J. Chem. Phys. 130, 014501 (2009); 10.1063/1.3043368 Time resolved infrared absorption studies of geminate recombination and vibrational relaxation in OClO photochemistry J. Chem. Phys. 121, 4795 (2004); 10.1063/1.1778373 Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 137.111.13.200 On: Tue, 31 May 2016 00:48:46

Upload: others

Post on 09-Oct-2020

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Femtosecond transient infrared and stimulated Raman … · THE JOURNAL OF CHEMICAL PHYSICS 142, 212409 (2015) Femtosecond transient infrared and stimulated Raman spectroscopy shed

Femtosecond transient infrared and stimulated Raman spectroscopy shed light onthe relaxation mechanisms of photo-excited peridininMariangela Di Donato, Elena Ragnoni, Andrea Lapini, Paolo Foggi, Roger G. Hiller, and Roberto Righini Citation: The Journal of Chemical Physics 142, 212409 (2015); doi: 10.1063/1.4915072 View online: http://dx.doi.org/10.1063/1.4915072 View Table of Contents: http://scitation.aip.org/content/aip/journal/jcp/142/21?ver=pdfcov Published by the AIP Publishing Articles you may be interested in Fragmentation mechanism of UV-excited peptides in the gas phase J. Chem. Phys. 141, 154309 (2014); 10.1063/1.4897158 Selective bond breaking mediated by state specific vibrational excitation in model HOD molecule throughoptimized femtosecond IR pulse: A simulated annealing based approach J. Chem. Phys. 139, 034310 (2013); 10.1063/1.4813127 Photodissociation of CH 3 Cl , C 2 H 5 Cl , and C 6 H 5 Cl on the Ag(111) surface: Ab initio embedded clusterand configuration interaction study J. Chem. Phys. 132, 074707 (2010); 10.1063/1.3322289 Femtosecond time-resolved absorption anisotropy spectroscopy on 9 , 9 ′ -bianthryl: Detection of partialintramolecular charge transfer in polar and nonpolar solvents J. Chem. Phys. 130, 014501 (2009); 10.1063/1.3043368 Time resolved infrared absorption studies of geminate recombination and vibrational relaxation in OClOphotochemistry J. Chem. Phys. 121, 4795 (2004); 10.1063/1.1778373

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 137.111.13.200 On: Tue, 31

May 2016 00:48:46

Page 2: Femtosecond transient infrared and stimulated Raman … · THE JOURNAL OF CHEMICAL PHYSICS 142, 212409 (2015) Femtosecond transient infrared and stimulated Raman spectroscopy shed

THE JOURNAL OF CHEMICAL PHYSICS 142, 212409 (2015)

Femtosecond transient infrared and stimulated Raman spectroscopy shedlight on the relaxation mechanisms of photo-excited peridinin

Mariangela Di Donato,1,2,3,a) Elena Ragnoni,1,2,a) Andrea Lapini,1,2,3 Paolo Foggi,1,2,4

Roger G. Hiller,5 and Roberto Righini1,2,31LENS (European Laboratory for Non-Linear Spectroscopy), via N. Carrara 1, 50019 Sesto Fiorentino, Italy2INO (Istituto Nazionale di Ottica), Largo Fermi 6, 50125 Firenze, Italy3Dipartimento di Chimica “Ugo Schiff,” Università di Firenze, via della Lastruccia 13,50019 Sesto Fiorentino, Italy4Dipartimento di Chimica, Università di Perugia, via Elce di Sotto 8, 06100 Perugia, Italy5Department of Biology, Faculty of Science and Engineering, Macquarie University, Sidney 2109, Australia

(Received 7 January 2015; accepted 4 March 2015; published online 27 March 2015)

By means of one- and two-dimensional transient infrared spectroscopy and femtosecond stimulatedRaman spectroscopy, we investigated the excited state dynamics of peridinin, a carbonyl carotenoidoccurring in natural light harvesting complexes. The presence of singly and doubly excited states, aswell as of an intramolecular charge transfer (ICT) state, makes the behavior of carbonyl carotenoids inthe excited state very complex. In this work, we investigated by time resolved spectroscopy the relax-ation of photo-excited peridinin in solvents of different polarities and as a function of the excitationwavelength. Our experimental results show that a characteristic pattern of one- and two-dimensionalinfrared bands in the C==C stretching region allows monitoring the relaxation pathway. In polarsolvents, moderate distortions of the molecular geometry cause a variation of the single/double carbonbond character, so that the partially ionic ICT state is largely stabilized by the solvent reorganization.After vertical photoexcitation at 400 nm of the S2 state, the off-equilibrium population moves to theS1 state with ca. 175 fs time constant; from there, in less than 5 ps, the non-Franck Condon ICT stateis reached, and finally, the ground state is recovered in 70 ps. That the relevant excited state dynamicstakes place far from the Franck Condon region is demonstrated by its noticeable dependence on theexcitation wavelength. C 2015 AIP Publishing LLC. [http://dx.doi.org/10.1063/1.4915072]

I. INTRODUCTION

Peridinin is a highly substituted carbonyl carotenoid oc-curring in the light harvesting complexes of some dinofla-gellates.1,2 These organisms have evolved an efficient light-harvesting system, able to absorb sunlight mostly in the blue/green spectral region, with the red-most part of the solar spec-trum being highly absorbed by water.3,4 One of the best charac-terized peridinin containing antenna systems is the water solu-ble peridinin-chlorophyll-protein (PCP), whose crystal struc-ture has been solved at 2.0 Å resolution.2 Numerous experi-mental studies on PCP and similar antennas reveal that carot-enoid is the main light harvesting pigment, which efficientlytransfers energy to chlorophyll a. This situation differs fromthat in higher plants, where sunlight is principally absorbed bychlorophyll b and chlorophyll a. Furthermore, peridinin hasan important photo-protective role, since it can either quicklyquench chlorophyll triplet states or directly react with singletoxygen.3,5,6 The peculiar excited state characteristics of peri-dinin prompted many experimental studies about its photo-physics.7–11

The elucidation of the light induced processes occurringin this molecule is a crucial step in the understanding of themechanism operating in the natural antennas. Furthermore,

a)M. Di Donato and E. Ragnoni contributed equally to this work.

it assumes a great importance in view of current researchefforts aiming at developing molecular systems capable of effi-ciently harvesting light in all the regions of solar spectrum andtransferring energy among different molecular components, tobe used in artificial photosynthetic devices.12 The additionalability of peridinin to generate triplet states with high yields,6,13

and promptly react with potentially harmful radical species,14

further increases the interest about this molecule and its ana-logues, since their utilization in artificial devices would allowthe feedback control of the operation of the light absorbingmachinery.

It is now widely accepted that the efficient light harvest-ing and subsequent energy transfer to chlorophylls in natu-ral antennas are strongly related to the structural characteris-tics of peridinin, and, in particular, to the presence of a low-lying excited state with intramolecular charge transfer (ICT)character.9,11,15–19 Peridinin is a highly substituted carotenoid,with an unusually short carbon skeleton (C37 instead of C40)compared to the majority of carotenoids occurring in photo-synthetic systems, whose spectroscopic properties are influ-enced by the presence of carbonyl groups and, possibly, ofother substituents also present in its polyene chain, like anallene group. The excited state properties of carotenoids are infact mainly determined by the structure and functionalizationof their conjugated polyene chain. Their electronic states areusually classified on the basis of the C2h point group symmetry

0021-9606/2015/142(21)/212409/10/$30.00 142, 212409-1 © 2015 AIP Publishing LLC

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 137.111.13.200 On: Tue, 31

May 2016 00:48:46

Page 3: Femtosecond transient infrared and stimulated Raman … · THE JOURNAL OF CHEMICAL PHYSICS 142, 212409 (2015) Femtosecond transient infrared and stimulated Raman spectroscopy shed

212409-2 Di Donato et al. J. Chem. Phys. 142, 212409 (2015)

of the all-trans, un-substituted polyene; in this approxima-tion, both S0 and S1 are Ag

− states, while S2 has Bu+ symme-

try. The intense green visible absorption observed for carot-enoids is assigned to the S0 → S2 electronic transition, withthe S0 → S1 transition forbidden by symmetry.19 Numerousstudies have shown that the molecular structure of peridininleads to a substantial breaking of the idealized C2h symmetry,as evidenced by its unusually strong fluorescence comparedto other carotenoids.7,11,20 Furthermore, time resolved inves-tigations revealed the strong dependence on the solvent po-larity of the S1 excited state lifetime, which decreases from160 ps in non-polar media to 10 ps in highly polar solventslike methanol.7,8,10,16,21 Although it is commonly accepted thatthe shortening of peridinin excited state lifetime is ascribableto the occurrence of the ICT state, the exact electronic natureand formation mechanism of ICT are still debated.9,10,15,22,23

The presence of the carbonyl group, primarily responsiblefor setting up and stabilizing the charge transferred electronicstructure, has a great influence on both the stationary andtransient absorption spectra of peridinin. In the ground stateabsorption spectrum, the clear vibronic progression observedin non-polar solvents is smeared out in polar media, wherethe absorption band broadens and shifts to the red.24 As forthe excited state spectral properties, the transient absorptionspectra of peridinin measured in non-polar solvents are domi-nated by the narrow S1-Sn transition, occurring around 510 nm.In polar media, new red-shifted broad bands develop, extend-ing up to 700 nm, whose intensities increase with the polarity ofthe solvents. These bands, together with a stimulated emissionband observed at 950 nm, are considered markers for the ICTstate formation.4,9,11,22

Various models have been proposed to explain the exper-imental data and to account for the nature of the ICT state.However, it is still unclear whether S1 and ICT are a single elec-tronic state, often indicated as S1/ICT,10,11,19 or two distinct butpossibly coupled electronic states.8,9 Recently, further insighton the photophysics of peridinin has been obtained throughthe study of a series of peridinin analogues, differing for thelengths of the polyene chain.18,25 These studies revealed thatthe formation of the ICT state is related to the chain length andis enhanced in short chain analogues; in addition, the amountof polyene conjugation also regulates the energy differencebetween S2 and S1 states. In recent papers,26,27 we analyzedthe deactivation pathway of a different and structurally simplercarbonyl carotenoid, trans-β-apo-8′-carotenal, by means ofvisible pump/Mid-IR probe spectroscopy, transient absorptionspectroscopy, and transient 2DIR spectroscopy. Initial datawere recorded in two solvents with different polarities, thenonpolar cyclohexane and the slightly polar chloroform, andthe measurements were subsequently extended to a num-ber of solvents, differing both in polarity and polarizability.Based on a global analysis of the time resolved data and onCASPT2/CASSCF (and TD-DFT) quantum chemical calcula-tions of energies, vibrational frequencies, and possible decaypaths, we proposed a kinetic scheme for the polarity-dependentrelaxation dynamics of this carbonyl carotenoid and investi-gated the nature of its ICT state. According to our picture,the major contributor to the ICT state is the bright 1Bu

+ state,which, in polar solvents, is dynamically stabilized through

molecular distortion and solvent relaxation. Furthermore, wedemonstrated that solvent polarizability has an important roleon the dynamical stabilization of the ICT state.27 The use ofnon-linear time-resolved spectroscopic methods in the mid-IRregion and, in particular, of transient two-dimensional infraredspectroscopy in the excited state (T2D-IR),28 was particularlyinformative for the characterization of the dynamics of the ICTstate formation. Based on these previous results, here, we applythe same methods to peridinin, with the aim of clarifying therelation between the photophysical properties of carotenoidsand peculiar structural properties, namely, chain length andnature of substituents. We measured the visible pump/Mid-IRprobe (T1D-IR) spectra of peridinin in cyclohexane and chlo-roform, using three different excitation wavelengths, 400 nm,480 nm, and 510 nm (530 nm in chloroform). Furthermore,we collected non-equilibrium 2D-IR spectra in chloroformexciting the sample at both 400 nm and 530 nm. Furthermore,in order to better identify the contribution of C==C stretchingin the excited state IR spectrum of peridinin, we also appliedFemtosecond Stimulated Raman Spectroscopy (FSRS). Thelatter technique has been successfully applied to study theexcited state relaxation of β-carotene, allowing the unambigu-ous assignement of the contribution of the symmetric C==Cstretching of both S2 and S1 states.29 Our experiments revealan interesting spectral evolution with striking similarities towhat observed in the case of trans-β-apo-8′-carotenal, and arehighly suggestive of a similar interpretation about the natureof the ICT state.

II. MATERIALS AND METHODS

A. Sample preparation

Peridinin was extracted from PCP protein.2 Samples forinfrared measurements were prepared by dissolving the mole-cule in different solvents; the concentration was adjusted for anoptical density of about 0.2 OD for a 50 µm optical path at theexcitation wavelength. To avoid photodegradation, sample wasmanipulated under a nitrogen atmosphere in a sealed dry box.For transient infrared measurements (T1D-IR and T2D-IR),the sample cell consisted of two CaF2 windows (2 mm thick),separated by a 50 µm Teflon spacer. For the stimulated Ramanexperiments, the sample was held in a 2 mm quartz cuvette. Allthe solvents (HPLC purity grade) were purchased from SigmaAldrich and used without further purification.

B. Experimental setup

The setup used to record visible pump/Mid-IR probespectra (T1D-IR) and transient 2DIR (T2D-IR) spectra waspreviously described.30 Briefly, for the T1D-IR experiment, theoutput of a Ti:Sapphire regenerative amplifier (Legend Elite,Coherent), operating at 800 nm, at 1 kHz of repetition rateand with <50 fs pulse duration, was split to pump a home-built two-stage optical parametric amplifier (OPA) equippedwith a difference frequency generator and a commercial opticalparametric amplifier (TOPAS, Light Conversion). The first wasused to generate broad (∼200 cm−1) mid-IR probe pulses in the5-7 µm region and the second to generate the visible excitation

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 137.111.13.200 On: Tue, 31

May 2016 00:48:46

Page 4: Femtosecond transient infrared and stimulated Raman … · THE JOURNAL OF CHEMICAL PHYSICS 142, 212409 (2015) Femtosecond transient infrared and stimulated Raman spectroscopy shed

212409-3 Di Donato et al. J. Chem. Phys. 142, 212409 (2015)

pulses by frequency mixing the signal output of TOPAS withthe residual fundamental at 800 nm in a BBO crystal. Theexcitation wavelength generated with the TOPAS was set at480 nm, 510 nm, and 530 nm in the different experimentsperformed. The visible pump pulses were attenuated to provide100-200 nJ at the sample and focused to a spot of ∼150 µm indiameter. A moveable delay line made it possible to controlthe time-of-arrival-difference of the pump and probe beams upto 1.8 ns. The pump-probe relative polarization was set to themagic angle by means of a λ/2 plate; the measurements weretaken in two partially overlapping spectral windows, coveringthe spectral region 1450-1800 cm−1.

We collected the 2D-IR spectra in the dynamic hole-burning configuration. In this scheme, the frequency of thenarrow-band pump pulse is scanned throughout the spectralregion of interest: it determines one of the frequency axes (thepump axis). The broad-band probe pulse, once the frequencyis dispersed by a monochromator placed after the sample,provides the second frequency axis (the probe axis), thusgenerating the two-dimensional spectral map.31 Narrow bandinfrared pump pulses were obtained by passing the outputof the TOPAS OPA through a Fabry-Perot element, whichprovides pulses with a spectral width of about 15 cm−1. Inthe non-equilibrium extension of the 2D-IR measurement, anadditional visible pump pulse, centered at 400 or 530 nm,prepares the system in the S2 electronically excited state. Byscanning the tvis time delay between the visible and infraredpump pulses (see Scheme 1 for the pulse sequence employed),it is possible to collect 2D-IR snap-shots of the system whilethe electronic excited state relaxes.

All-parallel polarization geometry was adopted both forground state equilibrium 2D-IR spectra and for excited statenon-equilibrium T2D-IR experiments. The sample was movedwith a homebuilt scanner to refresh the solution and avoidphotodegradation. A fraction of the probe pulse was split offbefore the sample and was used as a reference. After thesample, both probe and reference were spectrally dispersedin a spectrometer (TRIAX 180, HORIBA JobinYvon) andimaged separately on a 32 channels double array HgCdTe(MCT) detector (InfraRed Associated Inc., Florida, USA) witha sampling resolution of 6 cm−1.

For FSRS39 measurements, the 800 nm fundamental wassplit to generate three beams: (i) the actinic pump at 400 nm,obtained by frequency doubling the fundamental in a typeI BBO crystal, with <100 fs pulse duration; (ii) the Ramanpump at 800 nm, narrowed to 1-2 nm bandwidth by a 4fgrating compressor with a mask in the Fourier plane; and(iii) the super-continuum probe, generated by focusing thefundamental in a sapphire plate. A fourth beam, obtained by

SCHEME 1. Pulse sequence used to record the T2D-IR spectra.

splitting the probe beam, acted as a reference beam. The actinicand Raman pump powers were adjusted to provide 100 nJ at400 nm and 2 µJ at 800 nm at the sample. The three beams werefocused in a 80 µm spot by a spherical mirror placed beforethe sample. Finally, the probe beam was spatially selected andsent, together with the reference beam, to a 25 cm Jobin Yvonmonochromator (HR 250), equipped with a homemade CCDsignal analyzer, based on a S8380-512Q double array fromHamamatsu.

All measurements were performed at room temperature.The integrity of the sample was checked by FTIR (BrukerAlpha-T) and visible absorption (Perkin Elmer LAMBDA950) before and after the time resolved measurements.

Transient infrared data were analyzed by applying a globalanalysis procedure,32 using a sequential decay scheme withincreasing lifetimes and employing the software packagingGLOTARAN.33

III. RESULTS

A. Ground state FTIR and Raman spectra

The vibrational spectra of peridinin in different solventshave been reported and discussed by several authors.17–20 Themolecular structure is sketched in Figure 1.

In Figure 2, we report the FTIR spectrum of peridininmeasured in methanol (where the solubility is higher) andthe stimulated Raman spectra recorded in cyclohexane andchloroform.

Several bands can be distinguished and assigned to thespecific vibrations of the different substituents. In agreementwith the previous reports, in the FTIR spectrum, we distin-guish two bands in the carbonyl stretching region at 1738and 1756 cm−1. A single band at 1743 cm−1 is observed inthe Raman spectrum in chloroform and, possibly, in cyclo-hexane: in the latter solvent, the noise is much higher, dueto the low solubility of the molecule. Peridinin contains twocarbonyl groups: the lactone C==O, strongly conjugated withthe polyene chain, and the more isolated ester C==O. Thestretching vibrations of both groups are expected to appear inthe 1680-1760 cm−1 spectral region. Assignments have beenproposed in the literature, also based on quantum chemicalcomputations.34,35 Vanishing intensity is expected for the esterC==O in the Raman spectrum, while it should be visible inthe FTIR spectrum; detectable Raman activity is predicted forthe lactonic C==O because of the conjugation with the polyenechain.

In a recent paper,35 Kish et al. reported on the presenceof doublets in the carbonyl region of the Raman spectrumof peridinin in methanol and other solvents (acetonitrile andcyclohexane). While the presence of a doublet in methanolcould be interpreted as arising from distinct populations ofH-bonded and non H-bonded molecules, this explanation isnot valid in case of aprotic solvents. In Ref. 35, the presenceof two peaks in the Raman spectrum around 1770 cm−1 innon-protic media was attributed to the lactone C==O, whichsplits by effect of the Fermi resonance with the overtone of theCH wagging mode of the lactone ring. In the FTIR spectrumregistered in methanol and reported in Figure 2, the observed

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 137.111.13.200 On: Tue, 31

May 2016 00:48:46

Page 5: Femtosecond transient infrared and stimulated Raman … · THE JOURNAL OF CHEMICAL PHYSICS 142, 212409 (2015) Femtosecond transient infrared and stimulated Raman spectroscopy shed

212409-4 Di Donato et al. J. Chem. Phys. 142, 212409 (2015)

FIG. 1. Structure of peridinin.

doublet could be either due to the coexistence of H-bondedand non H-bonded populations of the lactone C==O or due tothe quasi coincident vibrations of the ester and lactone C==Ogroups. Measurements of 2DIR spectra in protic and non-protic solvents are in progress to clarify this point.

At lower frequencies, in the 1600-1400 cm−1 region,absorption of C==C stretching modes is expected, and accord-ingly, several bands are observed both in the FTIR and Ra-man spectra, which can be assigned to normal modes of thepolyene chain. In particular, an intense band is observed inthe Raman spectrum, peaked at 1516 cm−1 in chloroform andat 1534 cm−1 in cyclohexane, which is commonly attributedto a symmetric C==C stretching mode.35 A similar band hasbeen reported also for other carotenoids and polyenes.36,37

This band is also observed in the FTIR spectrum, althoughwith lower intensity, together with another weak absorptionat 1611 cm−1 in chloroform, which can also be attributed

FIG. 2. (a) Stimulated Raman spectrum of peridinin measured in cyclo-hexane (black line) and chloroform (red line). The lower panel highlightsthe weak bands observed in the 1600-2000 cm−1 region. Bands at 1266and 1444 cm−1 are due to cyclohexane and (b) FTIR spectrum of peridininregistered in methanol.

to a chain mode. At lower frequency, where contributionsdue to chain bending or C—C stretching are expected, a lowintensity band appears in the Raman spectrum at 1330 cm−1 inchloroform (1319 cm−1 in cyclohexane); in the same spectralregion, a relatively more intense absorption is present in theFTIR at 1390 cm−1, possibly due to a mode of the same nature.Finally, the Raman peak at ca. 1934 cm−1 (also distinguishablein the FTIR spectrum) is assigned to the allene stretchingvibration.34

B. Transient infrared measurements

Transient infrared spectra were collected in the 1450-1800 cm−1 region. Different excitation wavelengths were used:400 nm, on the blue side of the static absorption spectrum ofperidinin; 480 nm, close to the maximum of the absorptionpeak; and 515-530 nm, on the red side of the absorption band.As already noticed, the character of the ICT state stronglydepends on the ambient polarity; to get a better insight of itsformation dynamics and evolution, we repeated the measure-ments in two solvents: the non-polar cyclohexane and theslightly polar chloroform.27 Representative transient spectra atdifferent pump-probe delays for all the excitation wavelengthsare reported in Figure 3.

Visual inspection of the spectra shows notable differencesbetween the two solutions. The spectra recorded in cyclo-hexane are dominated by two positive bands in the 1700-1800 cm−1 region. The most intense excited state absorption

FIG. 3. Normalized transient infrared spectra recorded in cyclohexane andchloroform after 400 nm excitation, (top panel), 480 nm excitation (interme-diate panel), and 515 nm excitation (lower panel). The absorption spectrumof peridinin in cyclohexane and chloroform is reported in the inset.

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 137.111.13.200 On: Tue, 31

May 2016 00:48:46

Page 6: Femtosecond transient infrared and stimulated Raman … · THE JOURNAL OF CHEMICAL PHYSICS 142, 212409 (2015) Femtosecond transient infrared and stimulated Raman spectroscopy shed

212409-5 Di Donato et al. J. Chem. Phys. 142, 212409 (2015)

FIG. 4. Normalized kinetic traces recorded at 1630 cm−1 (left panel) and 1660 cm−1 (right panel) after exciting the sample at 400, 480, and 530 nm in chloroform.Experimental data are shown as scattered points; the fit obtained by global analysis is shown as a continuous line.

peak is at 1705 cm−1 and the second intense band is at1743 cm−1. At lower frequency, bands are much less intense.At all the excitation wavelengths, spectral evolution is limitedand negative signals due to ground state depletion are barelydetectable over the positive signals due to excited states ab-sorption. In chloroform, bands are much broader and threemajor excited state absorptions can be distinguished, peaking,respectively, at 1533 cm−1, 1630 cm−1, and 1720 cm−1. In thissolvent, the excitation wavelength effect is evident on bothspectral shape and dynamics. When the sample is excited at400 nm, the band peaked at 1630 cm−1 is initially very broad,and presents a shoulder on its blue side, at 1660 cm−1. At longerpump-probe delay, the intensity of the shoulder decreases,and the band both sharpens and red shifts.

The band measured at 1630 cm−1 upon 400 nm excitationmoves to 1620 cm−1 when the sample is excited at 530 nm.Furthermore, the evolution of the band shape observed whenthe sample is excited on the blue side of its absorption ismissing when the excitation is set on its red side. The kineticbehavior of the intensities of the band at 1630 cm−1 and its highfrequency shoulder at 1660 cm−1 in chloroform is shown inFigure 4, as a function of the excitation wavelength. In the samefigure, the fit obtained by applying a global analysis procedureis also shown. The 1630 cm−1 trace rises much more slowlywhen the sample is excited at 400 and 480 nm, relatively towhat happens when the excitation is at 530 nm.

To investigate the relaxation dynamics in the two solvents,we globally analyzed the time-resolved spectra imposing asequential decay scheme with increasing lifetimes. A singu-lar value decomposition (SVD) procedure38 isolated at leastthree components to satisfactorly fit the data. The evolution-associated decay spectra (EADS) obtained from global anal-ysis are reported in Figure 5.

In both solvents, the first decay component, with a lifetimeof 100-200 fs, can be assigned to the decay of the initiallyexcited S2 state. The subsequent evolution, occurring in 3-5 ps, is associated with different spectral changes in the twosolvents. Finally, the longer lifetime extracted from the fitcorresponds to the recovery of the ground state. The associatedkinetics reflect the well-known dependence of the peridininexcited state lifetime on the solvent polarity: in agreementwith previous reports,10,19 we find, upon 400 nm excitation,

a lifetime of about 160 ps in cyclohexane and of 70 ps inchloroform. Moving the excitation wavelength towards the redhas the effect of slightly shortening the excited state lifetime inboth the solvents.

In cyclohexane, within the instrumental time resolution, abroad absorption develops on the high frequency side (peakingat about 1700-1710 cm−1); this feature is mostly evident whenthe sample is excited at 400 nm (see first EADS, black line inFigure 5). The high frequency band around 1710 cm−1 slightlyblue-shifts, gains intensity, and narrows on a sub-ps timescale(second EADS in Figure 5). Finally, on a 3-4 ps timescale,a further blue shift and narrowing of the excited state bandsis observed. The spectral evolution is almost independent onthe excitation wavelength, except for a slight shortening of thelong lifetime component, which decreases from 163 ps, whenthe excitation is at 400 nm, to 149 ps, when the excitationis at 530 nm. Based on these observations, the excited statedynamics of peridinin in non-polar media can be safely inter-preted in the frame of the well-known excited decay schemegenerally applied for carotenoids not containing any carbonylfunction. The fastest process (black-red evolution in Figure 5)represents the decay of the initially excited state S2 into the

FIG. 5. EADS obtained by global analysis in cyclohexane (left panel) andchloroform (right panel) at three excitation wavelengths: 400 nm (top),480 nm (intermediate layer), and 530 nm (bottom).

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 137.111.13.200 On: Tue, 31

May 2016 00:48:46

Page 7: Femtosecond transient infrared and stimulated Raman … · THE JOURNAL OF CHEMICAL PHYSICS 142, 212409 (2015) Femtosecond transient infrared and stimulated Raman spectroscopy shed

212409-6 Di Donato et al. J. Chem. Phys. 142, 212409 (2015)

low-lying S1 excited state. Population of S1 is signaled by therise of the narrow excited state bands located at 1710 and1740 cm−1. The following evolution, occurring on a few ps timescale, can be interpreted as vibrational cooling in S1, while thelong-time evolution represents the decay of S1 and the recoveryof the ground state.

The excited state decay in chloroform is more complexthan in cyclohexane. As shown in Fig. 5, significant spectraland dynamical differences are observed on varying the exci-tation wavelength. Clear bleaching signals are not visible inany EADS, although dips in the overall positive signal canbe noticed both in the 1760-1780 cm−1 region and around1550 cm−1. According to the assignment of the FTIR spectrumreported in Figure 2, the high frequency dips correspond toground state carbonyl vibrations, while that at lower frequencypertains to a chain stretching mode. Two broad excited stateabsorption bands at 1620 and 1720 cm−1 are observed, besidesanother broad absorption band at lower frequency. In about150 fs, we observed a 15 cm−1 shift towards higher frequenciesof the 1620 cm−1 band and the development of a shoulder on itsblue side, at 1660 cm−1 (see second EADS in chloroform, exci-tation at 400 nm). Furthermore, the absorption at 1525 cm−1

increases in intensity. Finally, on a 5 ps timescale, the broad1630-1660 cm−1 band blue shifts and becomes significantlysharper, losing intensity on the 1660 cm−1 side. Ground staterecovery is faster than in cyclohexane and occurs on a 70 pstimescale.

When the visible excitation is at lower frequency, thisspectral evolution is less evident, even in chloroform. Follow-ing 530 nm excitation, the excited state absorption bands aresignificantly narrower. In particular, the spectral evolutionin the 1630-1660 cm−1 region is absent, and only minorband shifts occur. The 1660 cm−1 shoulder is not evident,and already on the short timescale, the spectral shape in the1630 cm−1 region is similar to that observed on a long timescaleafter 400 nm excitation. The ground state recovery is slightlyfaster. The effect of the excitation wavelength on the bandshapes observed in chloroform at short pump-probe delays isshown in Figure 6, where the time resolved spectra, measured

FIG. 6. Normalized transient infrared spectra of peridinin in chloroformrecorded at 1 ps pump-probe delay after exciting the sample at 400 nm (blackline), 480 nm (red line), and 530 nm (blue line).

at 400, 480, and 530 nm excitation at the pump-probe delay of1 ps, are compared.

C. Femtosecond stimulated Raman spectrum

To gain further insights about the shifts of the vibra-tional modes in peridinin upon photoexcitation, we measuredthe FSRS spectrum of the molecule in cyclohexane, usingan actinic pump centered at 400 nm and a Raman pump at800 nm. The spectrum registered 10 ps after actinic excitationis reported in Figure 7 and compared with the ground stateRaman spectrum of peridinin.

The most striking difference between the Raman spectrain the excited and ground states is the appearance, followingthe excitation at 400 nm, of two positive peaks at ca. 1700and 1750 cm−1. Their frequencies almost coincide with thoseof the bands observed in the excited state T1D-IR spectrum(1710 and 1740 cm−1) in cyclohexane. Since the spectrumrefers to a delay of 10 ps, the observed excited state bandscan be attributed to the S1 state. We tentatively assign the twobands to the symmetric C==C stretching (1700 cm−1), which,according to previous measurements on different carotenoidsand polyenes, upshifts upon excitation,26,39 and to the lactoneC==O (1750 cm−1), whose vibration acquires some Ramanactivity because of the conjugation with the polyene chain, andis expected to downshift in the excited state.34

The strong negative peaks in the Raman spectrum inFig. 7 correspond to ground state vibrations (those at 1270 and1450 cm−1 are solvent bands).

We performed the FSRS experiment also in chloroform,but in this case, the excited state bands could not be clearlyresolved. It is possible that in the polar solvent, the excited stateenergy changes enough to move the resonance far from the800 nm Raman pump, so that the pre-resonance amplificationeffect is canceled and the Raman activity becomes negligiblysmall.

FIG. 7. Stimulated Raman spectrum of peridinin in cyclohexane registered10 ps after the excitation at 400 nm (red line). The ground state spectrumis reported on top for comparison (black dashed line). Excited state Ramanbands are evident at 1700 and 1750 cm−1. The two bands at 1270 and1450 cm−1 are solvent bands.

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 137.111.13.200 On: Tue, 31

May 2016 00:48:46

Page 8: Femtosecond transient infrared and stimulated Raman … · THE JOURNAL OF CHEMICAL PHYSICS 142, 212409 (2015) Femtosecond transient infrared and stimulated Raman spectroscopy shed

212409-7 Di Donato et al. J. Chem. Phys. 142, 212409 (2015)

FIG. 8. Transient 2D-IR spectra measured for peridinin at 400 nm excitation in chloroform. Spectra registered at different Vispump-IRprobe delays are reported.For all spectra, the IRpump-IRprobe delay is 500 fs. The corresponding 1D Vispump-IRprobe spectra are shown as a black line on top of each 2D map. The diagonalpeaks at 1630 and 1660 cm−1 are marked with an asterisk in the left panel.

D. Transient 2D infrared measurements

The spectral evolution observed in chloroform when theexcitation is set on the blue side of the peridinin absorptionband could indicate that multiple peaks are hidden under thebroad envelope observed in the 1630-1660 cm−1 region. Todisentangle the contribution of different bands, we measuredthe 2D-IR spectra of the excited state (T2D-IR) in chloroform,exciting the sample at both 400 nm and at 530 nm. In thecase of all-trans-β-apo-8′-carotenal, we showed that the anal-ysis of non-equilibrium 2D-IR spectra can reveal interestinginformation on the molecular excited state evolution.26 In fact,the narrow band infrared pump pulses used to measure two-dimensional pump-probe spectra, unveil multiple bands, oftenhidden under broad envelopes in the linear spectra. Figure 8shows the T2D-IR maps recorded for peridinin in chloroform,upon 400 nm visible excitation, at four different Vispump/IRprobedelays and at the fixed IRpump/IRprobe delay of 500 fs.

Our transient 2D measurements clearly show that twobands underlie the broad absorption in the 1620-1670 cm−1

region: two diagonal negative peaks at ca. 1630 and 1660 cm−1

are resolved in this spectral region. The kinetic traces reportedin Figure 9 show that the two negative peaks grow in a few pico-

FIG. 9. Kinetic traces of the diagonal negative bands peaking at 1630 and1660 cm−1 registered with selective IRpump excitation.

second time scale, and that the rise time of the low frequencycomponent at 1630 cm−1 is markedly slower.

The experimental evidence provided by the transient 2Dspectra provides substantial confirmation that the time evolu-tion observed in the 1D infrared transient spectra for the broadband in the 1630-1660 cm−1 region (see previous paragraph)arises from the presence of two underlying components show-ing a different rise time. Those two transitions can be separatedin the 2D experiment, since they are now excited selectivelyand independently.

When the measurement is repeated by exciting the sampleat 530 nm, a single broad peak is observed in the 1630 cm−1

region on the diagonal, similar to that observed, in the sameconditions, in the linear transient spectra, where an appre-ciable but minor spectral evolution was observed. The 2D mapsin Figure 10 also show that the dynamic evolution of this1630 cm−1 peak is limited. In Figure 10, the linear T1D-IRspectra are shown for comparison on top of the 2D maps forcorresponding Vispump/IRprobe delays.

IV. DISCUSSION

The results reported in this work show that non-linearvibrational spectroscopy can be particularly informative aboutthe complex excited state evolution in peridinin. In particular,the application of T2D-IR resolves the presence of additionalbands that are hidden under broad envelopes in the T1D-IRspectra. Our measurements show that the excited state evolu-tion of peridinin changes remarkably when dissolved in non-polar and in polar media and as a function of the excitationwavelength.

To discuss the solvent and wavelength effects on theexcited state relaxation pathway and on the role and natureof the ICT state in peridinin, we provide an assignment of thevibrational excited state spectrum, based on our measurementsand on previous results.

Infrared band shifts upon excitation depend on the elect-ron density redistribution in the excited states. In the T1D-IR spectra reported in our work, several vibrational bands ofperidinin in the excited state were observed in the spectral

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 137.111.13.200 On: Tue, 31

May 2016 00:48:46

Page 9: Femtosecond transient infrared and stimulated Raman … · THE JOURNAL OF CHEMICAL PHYSICS 142, 212409 (2015) Femtosecond transient infrared and stimulated Raman spectroscopy shed

212409-8 Di Donato et al. J. Chem. Phys. 142, 212409 (2015)

FIG. 10. Transient 2D-IR spectra mea-sured for peridinin at 530 nm ex-citation in chloroform. Spectra regis-tered at different Vispump-IRprobe de-lays are reported. For all spectra, theIRpump-IRprobe delay is 500 fs. The cor-responding 1D Vispump-IRprobe spectraare shown as a black line on top ofeach 2D map. The diagonal peak at1630 cm−1 is marked with an asterisk.

window 1450-1800 cm−1, the most informative region beingthat between 1600 and 1800 cm−1. In cyclohexane, two intensetransitions appear, with peaks at 1705 and 1743 cm−1. Similarbands are observed in chloroform, although these are muchbroader and shifted to lower frequencies (1630 and 1715 cm−1).The assignment of the excited state infrared transitions ofperidinin is not completely clear. Femtosecond time resolvedinfrared spectra were previously recorded for peridinin40 andseveral band assignments have been proposed based on quan-tum chemical calculations.35 Recently, we measured the T1D-IR and T2D-IR spectra of all-trans-β-apo-8′-carotenal26,27 andproposed an assignment for the excited state bands in the 1600-1800 cm−1 region based on quantum chemical calculationsand on transient Raman spectra.36,39,41 It is expected that inperidinin, as well as in other polar carotenoids, the vibrationalmodes mostly contributing to the excited state absorption inthe 1600-1800 cm−1 region arise from carbonyl groups andpolyene chain vibrations. It is known that the symmetric C==Cstretching (the ν1 vibration) of linear polyenes and carotenoidsundergoes significant upshift in the excited state by vibroniccoupling of S0, S1, and S2.36,42 In all-trans-β-apo-8′-carotenal,this mode has been observed to shift from 1530 cm−1 in S0 to∼1700 cm−1 in S1 in non-polar solvents.26,27,41 In peridinin, dueto its reduced molecular symmetry, the shift is expected to besmaller.36 The extent of this shift mostly depends on the sym-metry of the carotenoid structure, which regulates the vibroniccoupling between S1(2Ag

−) and S2(1Bu+) states through the ag

C==C stretching mode.36 Since the symmetric C==C stretchingis a Raman active vibration, the corresponding band in theexcited state is observable in time-resolved FSRS spectra, asshown for all-trans-β-apo-8′-carotenal41 and β-carotene.43 Aswe noticed in Sec. III C, in the FSRS spectra of peridininin cyclohexane, we distinguish two excited state vibrationalbands, peaking at 1700 and 1750 cm−1, whose position almostcoincides with the two bands observed in the excited state IRspectrum in the same solvent (see Figures 2 and 7).

The T1D-IR spectra in chloroform show, in case of 400 nmexcitation, a dynamic evolution of the excited state absorptionband peaking at 1630 cm−1 with a pronounced shoulder at1660 cm−1, which closely resembles the behavior previouslyreported for the ∼1680 cm−1 band in the T1D-IR spectra of all-trans-β-apo-8′-carotenal in the same solvent. In that case, theband was assigned to the excited state C==C stretching mode.26

Based on this analogy, we tentatively assign the 1630 cm−1

band of peridinin to the C==C ν1 mode and the 1720 cm−1 bandto a carbonyl mode; the corresponding frequencies in cyclo-hexane are 1715 cm−1 for the C==C stretching and 1740 cm−1

for the C==O stretching. This assignment is supported by theappearance, in the FSRS spectrum of peridinin dissolved incyclohexane, of the two excited state bands (see Figure 7) inthe same spectral region.

Having assigned the most important excited state modes,we can now interpret the spectral evolution in the IR region andextract information on the nature of the charge transfer state inperidinin.

The T1D-IR spectra recorded in cyclohexane showedlimited evolution, except at very short time delays, when thespectral dynamics signals the rise of the S1 population from theinitially excited S2 state. Any spectral change registered at latertimes (t > 1 ps) can be safely interpreted in terms of vibrationalcooling on the S1 potential energy surface. In principle, theexcited state decay path of peridinin in cyclohexane could beexplained by considering only three excited states, S0, S1, andS2, without involving the formation of an ICT state. How-ever, the wavelength dependence of the excited state lifetimeobserved when moving the excitation wavelength towards thered suggests that an ICT state could be formed even in non-polar media, although with low yields.

In chloroform, we observe a significantly different evolu-tion, particularly evident when the sample is excited at 400 nm.Here, the excited state C==C stretching band initially upshiftson a few hundreds femtosecond timescale (evolution from theblack to red EADS in Figure 5). Then, on a few ps timescale,the red side of the band grows in intensity (see the increasedintensity at 1630 cm−1 in the blue EADS in Figure 5 and thekinetics reported in Figures 4 and 9). A very similar spec-tral evolution was recently observed in the T1D-IR spectrameasured for the all-trans-β-apo-8′-carotenal in chloroformand several other solvents.26,27 Furthermore, in chloroform,spectral evolution depends on the excitation wavelength: byexciting the sample on the red-tail of its absorption, bands inthe T1D-IR spectrum become narrower and the evolution ofthe C==C band observed upon 400 nm excitation is almostabsent. This observation alone would suggest that two differentexcited states are involved in the photodynamics, one selec-tively excited with a red-most pump.

Further confirmation comes from the measure of 2D-IRspectra of peridinin in the excited state. Here, we are able to

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 137.111.13.200 On: Tue, 31

May 2016 00:48:46

Page 10: Femtosecond transient infrared and stimulated Raman … · THE JOURNAL OF CHEMICAL PHYSICS 142, 212409 (2015) Femtosecond transient infrared and stimulated Raman spectroscopy shed

212409-9 Di Donato et al. J. Chem. Phys. 142, 212409 (2015)

clearly distinguish two diagonal bands, respectively, peakingat 1630 and 1660 cm−1 which are partially hidden under thebroad envelope recorded in the T1D-IR spectra (see Fig. 8).The lower frequency band rises slower than the higher fre-quency one, as shown by the kinetic traces reported in Figure 9.Even more interestingly, no trace of cross peaks coupling thetwo diagonal peaks appears in the 2D spectra. This is a clearindication that the two components at 1630 and 1660 cm−1

originate from different molecular states. By analogy with theinterpretation proposed for all-trans-β-apo-8′-carotenal, whichwas also supported by an extensive theoretical analysis,26 webelieve that the two sub-bands resolved in the 2D maps corre-spond to the C==C stretching in the ICT (1630 cm−1) and S1(1660 cm−1) states, respectively. The time evolution of the twocomponents is also in line with that observed in the case ofall-trans-β-apo-8′-carotenal, where the lower frequency peak,measured in chloroform at 1680 cm−1 and assigned to the ICTstate, shows a slower rise component than the higher frequencyband at 1710 cm−1, assigned to S1.

It remains to clarify the nature of the ICT state and themechanism determining its formation and population follow-ing light absorption in polar solvents. In case of all-trans-β-apo-8′-carotenal, we proposed that the major contributor to theICT state is the bright 1Bu

+ state, which is repopulated on a fewpicosecond timescale with a mechanism involving a moleculardistortion. This interpretation was based both on the exper-imental results and on a theoretical MS-CASPT2//CASSCFanalysis of the ground and excited state energy minima andprofiles of the energy surfaces.

Recently, it has been shown that in case of peridinin, subtlegeometrical variations can significantly alter the nature, ener-gies, and electronic properties of the excited states, modifyingthe 2Ag

−/1Bu+ degree of mixing and, consequently, the ratio

between covalent and ionic contributions to the excited statewavefunctions.44 The structural parameter mainly influenc-ing the properties of the low-lying excited states is the bondlength alternation (BLA). The theoretical analysis reported inRef. 44 showed that the variations in the BLA significantlyinfluenced the transition oscillator strength, the ratio of singleand double excitation character, and the dipole moment ofthe S1 and S2 excited states. Small BLA values correspondto an overall delocalized charge on all the conjugated chain.Geometrical distortions determine an increase of the BLA,which corresponds to a less conjugated molecular form, withsingle and double bond characters becoming more localized.Furthermore, Knecht et al.44 demonstrated that S1 gains moreionic character as BLA increases, while, on the contrary, S2gains more doubly excited (i.e., covalent) character when theelectronic energy gap between the first two low-lying excitedstates gets smaller as a consequence of their increasingly mixednature. Slightly distorted peridinin molecules, with the electro-negative C==O group more isolated from the conjugated chain,are likely to be stabilized in polar solvents. Thus, polar environ-ments can stabilize conformations with increased BLA, whichwould suggest that the formation of ICT in peridinin followsa very similar mechanism to the one proposed for all-trans-β-apo-8′-carotenal.26

The idea that the electronic relaxation in peridinin is asso-ciated with structural deformations has already been proposed

in the literature, based on the results of visible pump-probe andpump-dump-probe spectroscopy.9,10 In their studies, Papagian-nakis et al.9 isolated a ground state intermediate upon dumpingthe population of the excited state of peridinin, which mightrepresent a distorted molecular structure generated during therelaxation of the photoexcited molecule in polar solvents. Ourinfrared studies support and further strengthen this interpreta-tion: the observation of different frequencies associated withthe same molecular vibration, namely, the C==C stretchingin the excited state, is a direct evidence of a conformationalvariation of the molecule. Also, the occurrence of conforma-tional changes allows to rationalize the effect of the excita-tion wavelength on the peridinin’s photodynamics reported inour work, which has also been previously documented.10 Ourmeasurements and previous visible pump-probe experimentsagree in suggesting that excitation of peridinin on the red-tailof its absorption populates a higher amount of the ICT state.This observation can be likely explained assuming that, whenthe excitation is at wavelengths larger than 500 nm, a sub-population of distorted red-absorbing molecules is excited.These distorted molecules would make a major contribution ofdirectly excited forms to the ICT population and would explainthe minor spectral evolution observed in this case.

V. CONCLUSIONS

Our time-resolved experimental results on peridinin pointout characteristic aspects of its behavior in the excited state,which appear to be common to other carbonyl carotenoidsand reveal that other chain substituents, such as the allenegroup, have a minor influence on its photodynamics. All ourmeasurements confirm the main points of the scenario that werecently proposed for trans-β-apo-8′-carotenal. This commonbehavior could not be expected a priori, given the relevantstructural difference between these two molecules, which isparticularly evident in the highly substituted polyene chainof peridinin. From our study, it appears that the excited stateevolution of peridinin and the formation of the ICT stateare driven by structural relaxation processes occurring uponphoto-excitation. Soon after the direct population of the S2excited state, induced by green light absorption, molecularrelaxation and solvent dynamics drive the excited state pop-ulation far from the Frank Condon region, determining amixing of covalent and ionic characters in the excited statewavefunction. Even slight molecular distortions produce anincrease of the bond length alternation and consequently ofthe ionic character of the low lying excited state. Polar solventrelaxation stabilizes these distorted structures, thus determin-ing the appearance of marker bands characteristic of the ICTstate in the transient spectra. Also the observed dependenceof the excited state evolution on the excitation wavelength isconsistent with a dynamics on the excited state potential energysurface that mostly takes place far from the Franck Condonregion.

ACKNOWLEDGMENTS

The authors gratefully acknowledge support from the Ital-ian MIUR: Program No. PRIN-2010ERFKXL-004; programFIRB “Futuro in Ricerca 2010” Grant No. RBFR10Y5VW to

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 137.111.13.200 On: Tue, 31

May 2016 00:48:46

Page 11: Femtosecond transient infrared and stimulated Raman … · THE JOURNAL OF CHEMICAL PHYSICS 142, 212409 (2015) Femtosecond transient infrared and stimulated Raman spectroscopy shed

212409-10 Di Donato et al. J. Chem. Phys. 142, 212409 (2015)

M.D.D. and RBFR109ZHQ supporting A.L. The contributionfrom Project No. EFOR-CABIR L. 191/2009 art. 2 comma 44is also gratefully acknowledged. This research has receivedfunding also from LASERLAB-EUROPE (Grant AgreementNo. 284464, EC’s 7th Framework Programme).

1H. A. Frank and R. J. Cogdell, Photochem. Photobiol. 63(3), 257 (1996); J.Jiang, H. Zhang, Y. Kang, D. Bina, C. S. Lo, and R. E. Blankenship, Biochim.Biophys. Acta, Bioenerg. 1817(7), 983 (2012).

2E. Hofmann, P. M. Wrench, F. P. Sharples, R. G. Hiller, W. Welte, and K.Diederichs, Science 272(5269), 1788 (1996).

3M. F. Hohmann-Marriott, T. Polivka, and E. Hofmann, The Structural Basisof Biological Energy Generation (Springer, Netherlands, 2014), Vol. 39,p. 39.

4T. Å. Polivka, R. G. Hiller, and H. A. Frank, Arch. Biochem. Biophys.458(2), 111 (2007).

5M. Di Valentin, C. E. Tait, E. Salvadori, L. Orian, A. Polimeno, and D.Carbonera, Biochim. Biophys. Acta, Bioenerg. 1837, 85 (2014).

6C. Bonetti, M. T. A. Alexandre, R. G. Hiller, J. T. M. Kennis, and R. v.Grondelle, Chem. Phys. 357(1–3), 63 (2009).

7J. A. Bautista, R. E. Connors, B. B. Raju, R. G. Hiller, F. P. Sharples, D.Gosztola, M. R. Wasielewski, and H. A. Frank, J. Phys. Chem. B 103, 8751(1999).

8E. Papagiannakis, D. S. Larsen, I. H. M. van Stokkum, M. Vengris, R. G.Hiller, and R. van Grondelle, Biochemistry 43(49), 15303 (2004).

9E. Papagiannakis, M. Vengris, D. S. Larsen, I. H. M. van Stokkum, R. G.Hiller, and R. van Grondelle, J. Phys. Chem. B 110(1), 512 (2005).

10D. Zigmantas, R. G. Hiller, A. Yartsev, V. Sundstrom, and T. Å. Polivka, J.Phys. Chem. B 107(22), 5339 (2003).

11D. Zigmantas, T. Å. Polivka, R. G. Hiller, A. Yartsev, and V. Sundstrom, J.Phys. Chem. A 105(45), 10296 (2001).

12S. Pillai, J. Ravensbergen, A. Antoniuk-Pablant, B. D. Sherman, R. vanGrondelle, R. N. Frese, T. A. Moore, D. Gust, A. L. Moore, and J. T. M.Kennis, Phys. Chem. Chem. Phys. 15(13), 4775 (2013).

13M. Di Valentin, S. Ceola, G. Agostini, G. M. Giacometti, A. Angerhofer, O.Crescenzi, V. Barone, and D. Carbonera, Biochim. Biophys. Acta, Bioenerg.1777(3), 295 (2008); M. Di Valentin, S. Ceola, E. Salvadori, G. Agostini, andD. Carbonera, ibid. 1777(2), 186 (2008); D. Carbonera, G. M. Giacometti,and U. Segre, J. Chem. Soc., Faraday Trans. 92(6), 989 (1996); M. T. A.Alexandre, D. C. Lührs, I. H. M. van Stokkum, R. Hiller, M.-L. Groot, J.T. M. Kennis, and R. van Grondelle, Biophys. J. 93(6), 2118 (2007); C. C.Gradinaru, J. T. M. Kennis, E. Papagiannakis, I. H. M. van Stokkum, R. J.Cogdell, G. R. Fleming, R. A. Niederman, and R. van Grondelle, Proc. Natl.Acad. Sci. U. S. A. 98(5), 2364 (2001).

14E. Pinto, L. H. Catalani, N. P. Lopes, P. Di Mascio, and P. Colepicolo,Biochem. Biophys. Res. Commun. 268(2), 496 (2000); R. Edge, D. J. Mc-Garvey, and T. G. Truscott, J. Photochem. Photobiol., B 41(3), 189 (1997).

15M. M. Enriquez, M. Fuciman, A. M. LaFountain, N. L. Wagner, R. R. Birge,and H. A. Frank, J. Phys. Chem. B 114(38), 12416 (2010); N. L. Wagner, J.A. Greco, M. M. Enriquez, H. A. Frank, and R. R. Birge, Biophys. J. 104(6),1314 (2013).

16H. A. Frank, J. A. Bautista, J. Josue, Z. Pendon, R. G. Hiller, F. P. Sharples,D. Gosztola, and M. R. Wasielewski, J. Phys. Chem. B 104, 4569 (2000).

17N. M. Magdaong, D. M. Niedzwiedzki, J. A. Greco, H. Liu, K. Yano, T.Kajikawa, K. Sakaguchi, S. Katsumura, R. R. Birge, and H. A. Frank, Chem.Phys. Lett. 593, 132 (2014); T. Polivka, S. Kaligotla, P. Chabera, and H.A. Frank, Phys. Chem. Chem. Phys. 13(22), 10787 (2011); D. Zigmantas,

R. G. Hiller, V. Sundström, and T. Å. Polivka, Proc. Natl. Acad. Sci. U. S.A. 99(26), 16760 (2002).

18D. M. Niedzwiedzki, N. Chatterjee, M. M. Enriquez, T. Kajikawa, S.Hasegawa, S. Katsumura, and H. A. Frank, J. Phys. Chem. B 113(41), 13604(2009).

19T. Polivka and V. Sundstrom, Chem. Rev. 104(4), 2021 (2004).20S. Wormke, S. Mackowski, T. H. P. Brotosudarmo, C. Jung, A. Zumbusch,

M. Ehrl, H. Scheer, E. Hofmann, R. G. Hiller, and C. Brauchle, Biochim.Biophys. Acta, Bioenerg. 1767(7), 956 (2007).

21D. Zigmantas, R. G. Hiller, F. P. Sharples, H. A. Frank, V. Sundström, andT. Polivka, Phys. Chem. Chem. Phys. 6(11), 3009 (2004).

22E. Papagiannakis, I. H. M. van Stokkum, M. Vengris, R. J. Cogdell, R. vanGrondelle, and D. S. Larsen, J. Phys. Chem. B 110(11), 5727 (2006).

23H. M. Vaswani, C.-P. Hsu, M. Head-Gordon, and G. R. Fleming, J. Phys.Chem. B 107(31), 7940 (2003).

24V. S. Pavlovich, Photochem. Photobiol. Sci. 13(10), 1444 (2014).25D. M. Niedzwiedzki, T. Kajikawa, K. Aoki, S. Katsumura, and H. A. Frank,

J. Phys. Chem. B 117(23), 6874 (2013).26M. Di Donato, M. Segado Centellas, A. Lapini, M. Lima, F. J. Avila Ferrer,

F. Santoro, C. Cappelli, and R. Righini, J. Phys. Chem. B 118, 9613 (2014).27E. Ragnoni, M. Di Donato, A. Iagatti, A. Lapini, and R. Righini, J. Phys.

Chem. B 119, 420 (2015).28J. Bredenbeck, J. Helbing, C. Kolano, and P. Hamm, ChemPhysChem 8(12),

1747 (2007).29P. Kukura, D. W. McCamant, and R. A. Mathies, J. Phys. Chem. A 108(28),

5921 (2004).30P. T. Touceda, S. M. Vazquez, M. Lima, A. Lapini, P. Foggi, A. Dei, and R.

Righini, Phys. Chem. Chem. Phys. 14(2), 1038 (2012).31M. T. Zanni and P. Hamm, Concepts and Methods of 2D Infrared Spectros-

copy (Cambridge University Press, 2011).32I. H. M. van Stokkum, D. S. Larsen, and R. van Grondelle, Biochim.

Biophys. Acta, Bioenerg. 1657(2–3), 82 (2004).33J. J. Snellenburg, S. P. Laptenok, R. Seger, K. M. Mullen, and I. H. M. van

Stokkum, J. Stat. Software 49(3), 1 (2012).34D. Bovi, A. Mezzetti, R. Vuilleumier, M.-P. Gaigeot, B. Chazallon, R.

Spezia, and L. Guidoni, Phys. Chem. Chem. Phys. 13(47), 20954 (2011).35E. Kish, M. M. M. Pinto, D. Bovi, M. Basire, L. Guidoni, R. Vuilleumier, B.

Robert, R. Spezia, and A. Mezzetti, J. Phys. Chem. B 118(22), 5873 (2014).36T. Noguchi, H. Hayashi, M. Tasumi, and G. H. Atkinson, J. Phys. Chem.

95(8), 3167 (1991).37S. Saito and M. Tasumi, J. Raman Spectrosc. 14(5), 310 (1983); L. Rimai,

M. E. Heyde, and D. Gill, J. Am. Chem. Soc. 95(14), 4493 (1973).38R. W. Hendler and R. I. Shrager, J. Biochem. Biophys. Methods 28(1), 1

(1994).39H. Hashimoto and Y. Koyama, Chem. Phys. Lett. 162(6), 523 (1989).40C. Bonetti, M. T. A. Alexandre, I. H. M. van Stokkum, R. G. Hiller, M. L.

Groot, R. van Grondelle, and J. T. M. Kennis, Phys. Chem. Chem. Phys.12(32), 9256 (2010); A. J. Van Tassle, M. A. Prantil, R. G. Hiller, and G. R.Fleming, Isr. J. Chem. 47(1), 17 (2007).

41T. M. Kardas, B. Ratajska-Gadomska, A. Lapini, E. Ragnoni, R. Righini,M. Di Donato, P. Foggi, and W. Gadomski, J. Chem. Phys. 140(20), 204312(2014).

42G. Orlandi and F. Zerbetto, Chem. Phys. 108(2), 187 (1986); G. Orlandi, F.Zerbetto, and M. Z. Zgierski, Chem. Rev. 91(5), 867 (1991); F. Zerbetto, M.Z. Zgierski, G. Orlandi, and G. Marconi, J. Chem. Phys. 87(5), 2505 (1987).

43D. W. McCamant, J. E. Kim, and R. A. Mathies, J. Phys. Chem. A 106(25),6030 (2002).

44S. Knecht, C. M. Marian, J. Kongsted, and B. Mennucci, J. Phys. Chem. B117(44), 13808 (2013).

Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 137.111.13.200 On: Tue, 31

May 2016 00:48:46

Page 12: Femtosecond transient infrared and stimulated Raman … · THE JOURNAL OF CHEMICAL PHYSICS 142, 212409 (2015) Femtosecond transient infrared and stimulated Raman spectroscopy shed

0021-9606 Advanced Search

Search My Library's Catalog: ISSN Search | Title SearchSearch Results

Search Workspace Ulrich's Update Admin

Enter a Title, ISSN, or search term to find journals or other periodicals:

The Journal of Chemical Physics

Log in to My Ulrich's

Macquarie University Library

Related Titles

Alternative MediaEdition (4)

Lists

Marked Titles (0)

Search History

0021-9606 - (1)0004-9522 - (1)0031-949X - (1)

Save to List Email Download Print Corrections Expand All Collapse All

Title The Journal of Chemical Physics

ISSN 0021-9606

Publisher American Institute of Physics

Country United States

Status Active

Start Year 1933

Frequency Weekly 48/yr.

Language of Text Text in: English

Refereed Yes

Abstracted / Indexed Yes

Serial Type Journal

Content Type Academic / Scholarly

Format Print

Website http://jcp.aip.org/

Email [email protected]

Description Publishes original research at the interface between physics and chemistry.

Save to List Email Download Print Corrections Expand All Collapse All

Title Details Table of Contents

Contact Us | Privacy Policy | Terms and Conditions | Accessibility

Ulrichsweb.com™, Copyright © 2013 ProQuest LLC. All Rights Reserved

Basic Description

Subject Classifications

Additional Title Details

Publisher & Ordering Details

Price Data

Online Availability

Abstracting & Indexing

Other Availability

Demographics

Reviews

ulrichsweb.com(TM) -- The Global Source for Periodicals http://ulrichsweb.serialssolutions.com/title/1358810459994/47912

1 of 1 22/01/2013 10:56 AM