university of zurich - uzh · 2010. 11. 29. · transient 2d-ir spectroscopy of non-equilibrium...

16
University of Zurich Zurich Open Repository and Archive Winterthurerstr. 190 CH-8057 Zurich http://www.zora.uzh.ch Year: 2008 Two-dimensional infrared spectroscopy of photoswitchable peptides Hamm, P; Helbing , J; Bredenbeck, J Hamm, P; Helbing , J; Bredenbeck, J (2008). Two-dimensional infrared spectroscopy of photoswitchable peptides. Annual Review of Physical Chemistry, 59:291-317. Postprint available at: http://www.zora.uzh.ch Posted at the Zurich Open Repository and Archive, University of Zurich. http://www.zora.uzh.ch Originally published at: Annual Review of Physical Chemistry 2008, 59:291-317.

Upload: others

Post on 30-Jan-2021

7 views

Category:

Documents


0 download

TRANSCRIPT

  • University of ZurichZurich Open Repository and Archive

    Winterthurerstr. 190

    CH-8057 Zurich

    http://www.zora.uzh.ch

    Year: 2008

    Two-dimensional infrared spectroscopy of photoswitchablepeptides

    Hamm, P; Helbing , J; Bredenbeck, J

    Hamm, P; Helbing , J; Bredenbeck, J (2008). Two-dimensional infrared spectroscopy of photoswitchable peptides.Annual Review of Physical Chemistry, 59:291-317.Postprint available at:http://www.zora.uzh.ch

    Posted at the Zurich Open Repository and Archive, University of Zurich.http://www.zora.uzh.ch

    Originally published at:Annual Review of Physical Chemistry 2008, 59:291-317.

    Hamm, P; Helbing , J; Bredenbeck, J (2008). Two-dimensional infrared spectroscopy of photoswitchable peptides.Annual Review of Physical Chemistry, 59:291-317.Postprint available at:http://www.zora.uzh.ch

    Posted at the Zurich Open Repository and Archive, University of Zurich.http://www.zora.uzh.ch

    Originally published at:Annual Review of Physical Chemistry 2008, 59:291-317.

  • Two-dimensional infrared spectroscopy of photoswitchablepeptides

    Abstract

    We present a detailed discussion of the complimentary fields of the application of two-dimensionalinfrared (2D-IR) spectroscopy in comparison with two-dimensional nuclear magnetic resonance(2D-NMR) spectroscopy. Transient 2D-IR (T2D-IR) spectroscopy of nonequilibrium ensembles isprobably one of the most promising strengths of 2D-IR spectroscopy, as the possibilities of 2D-NMRspectroscopy are limited in this regime. T2D-IR spectroscopy uniquely combines ultrafast timeresolution with microscopic structural resolution. In this article we summarize our recent efforts toinvestigate the ultrafast structural dynamics of small peptides, such as the unfolding of peptidesecondary structure motifs. The work requires two ingredients: 2D-IR spectroscopy and the possibilityof triggering a structural transition of a peptide on an ultrafast timescale using embedded or intrinsicphotoswitches. Several photoswitches have been tested, and we discuss our progress in merging thesetwo pathways of research.

  • 2D-IR Spectroscopy of Photoswitchable Peptides

    Peter Hamm, Jan Helbing, Jens BredenbeckPhysikalisch–Chemisches Institut, Universität Zürich,Winterthurerstrasse 190, 8057 Zürich, Switzerland

    We present a detailed discussion of the complimentary fields of application of 2D-IR spec-troscopy in comparison with 2D-NMR spectroscopy. Transient 2D-IR spectroscopy of non-equilibrium ensembles is probably one of the most promising strengths of 2D-IR spectroscopy,as the as the possibilities of 2D-NMR spectroscopy are very limited in this regime. Transient2D-IR spectroscopy uniquely combines ultrafast time-resolution with microscopic structural res-olution. Our recent efforts to investigate ultrafast structural dynamics of small peptides, suchas the unfolding of peptide secondary structure motifs, are summarized. The work requires twoingredients: 2D-IR spectroscopy and the possibility to trigger a structural transition of a peptideon an ultrafast timescale using embedded or intrinsic photo-switches. Several photo-switcheshave been tested, and our progress in merging these two pathways of research is discussed.

    I. INTRODUCTION

    The importance of structural dynamics of biomoleculesfor their function is widely appreciated. 2D-IR spec-troscopy is a novel spectroscopic tool that uniquely com-bines ultrafast time resolution with appreciable structureresolution. The transferability of the concepts of multi-dimensional NMR spectroscopy [1] to IR spectroscopyhas been postulated already in the earliest publicationson two-dimensional NMR [2], however, the first two-dimensional IR spectrum was measured only relativelyrecently [3]. In analogy to 2D-NMR spectroscopy, theconnectivity of various vibrational states can be relatedto local contacts of the corresponding molecular groups,which is the basic principle of structure determination inboth cases. Equivalents of all fundamental NMR experi-ments (COSY [4, 5], NOESY [6], and EXSY [7–9]) havebeen demonstrated in the IR range in the meanwhile [10].

    By nature of the close analogy, it might seem that thebiggest competitor of 2D-IR spectroscopy will always be2D-NMR spectroscopy. That would put the bar high!Both 2D-NMR and 2D-IR spectroscopy are capable ofresolving certain functional groups in a molecule, work(mostly) in the solution phase and aim to elucidate struc-ture as well as dynamics of solution phase systems. Overthe course of the last 20-30 years NMR spectroscopy hasproven to be so much more versatile that it has largelysuperseded conventional (1D) IR spectroscopy for stan-dard analytic purposes. This has essentially two reasons:(i) NMR spectra are more structured and allow one toresolve individual resonances from much larger moleculesthan IR spectroscopy. (ii) The interpretation and assign-ment of NMR spectra works on the basis of simple em-

    pirical rules, whereas the assignment of infrared spectrais currently still a challenge to even the most powerfulquantum-chemistry codes for only a mid-size moleculewith, say, 30 atoms.

    When we started out to develop 2D-IR spectroscopyas a tool to determine molecular structure, we concen-trated on a small peptide, trialanine [6, 11–13]. Much toour surprise, the conformation of that molecule, or bet-ter the distribution of conformations, was not known atthe time although it has important model character: The’backbone’ of trialanine contains only one central (φ, ψ)-pair of dihedral angles and hence reflects the intrinsicstructure propensity of the peptide bond itself withoutthe additional complication of intermolecular hydrogenbonds in larger peptides that form secondary or tertiarystructure motifs. In a series of papers [6, 11–13], we de-termined the preferred conformation, the relative popula-tions of conformations, as well as the flexibility of triala-nine. We found that trialanine populates predominantlythe so-called poly-prolin II structure (PII) with about80%. Other structure elements like α-helical and β-sheetmotifs seemed to contribute only little. This result wasunexpected since the PII structure is not common inlarger proteins, and since standard molecular dynamics(MD) force fields predicted different structure distribu-tions [14]. Our results (among others) lead to suggestionsfor modifications of the AMBER force field [15].

    Very recently, our structure prediction of trialanine hasessentially been confirmed in a combined NMR and MDsimulation study, going just a little bit beyond the mostelementary methods of the rich toolbox of NMR spec-troscopy [16]. This is good and bad news: On the onehand, it validates that 2D-IR spectroscopy can indeed be

  • 2

    EXSYspectral diffusion

    T1, T2, HetNOE

    ps ns µs ms s

    T1ρ, T2 EXSY

    pump-probeRDC

    a

    b

    fs

    “NOE”

    ps ns µs ms s

    pump-probe

    fs

    ‘real time’

    ‘real time’

    FIG. 1: Time arrow from femtoseconds to seconds togetherwith time regimes that can be covered by various techniquesof (a) 2D-NMR and (b) 2D-IR spectroscopy. Panel (a) isadapted from Ref. [17].

    used to determine the unknown structure of small pep-tides, even when the experimenter cannot put any biasedexpectation into the structure analysis. On the otherhand, it says that if 2D-NMR spectroscopy tries hardenough it almost always will beat 2D-IR spectroscopy interms of pure structure resolution capabilities.

    It has often been argued that the big potential of 2D-IR spectroscopy, as compared to 2D-NMR spectroscopy,is the by many orders of magnitudes higher intrinsic timeresolution of the former. However, one must very care-fully specify what exactly is meant by ’time resolution’.Fig. 1a shows the time regimes that can be covered byvarious variants of NMR spectroscopy, which do in factinclude the ultrafast picosecond and even sub-picosecondrange [17]. For example the orientational diffusion timeof bulk water has been deduced already in the mid-70’sby NMR spectroscopy [18], revealing a value of 2.3 psfor D2O. Only about 20 years later [19], IR spectroscopycame into the position to repeat the measurement, re-vealing that the process is more complex. The gener-ally accepted picture now is that the anisotropy decayis biphasic with a sub-100 fs librational component priorto a 2.5-3 ps diffusive component [20, 21]. Hence, whileNMR spectroscopy is essentially right for to the diffusivepart, it completely misses the fast component and cannotdifferentiate the contributions in a heterogeneous process– a subtle but important difference.

    The way how the ultrafast time regime is accessed bythe intrinsically very slow NMR spectroscopy is indirectthrough relaxation methods (T1, T2, HetNOE, T1ρ andresidual dipolar couplings (RDC) in Fig. 1a). For ex-ample, in the case of orientational diffusion of water dis-cussed above, a simple theory exist which relates the T1relaxation time of the proton spin transition, which iswhat is actually measured in the experiment and whichis quite slow (0.45 s), to the fast orientational correlation

    time τc of the molecule [18]. Both quantities are inverselyproportional, hence the faster the molecular structuralprocess, the slower the spin relaxation. The importantpoint is that this theory relies on a model assumption –free orientational diffusion – which is not extremely welljustified in the case of water [22]. NMR relaxation tech-niques cannot validate the model assumption made tointerpret the data, and hence tend to oversimplify com-plex systems such as liquid water. For a truly diffusiveprocess, for example the reorientation of a larger, closeto spherical molecule in solution, the results from NMRspectroscopy are usually perfectly correct.

    The important difference between IR and NMR spec-troscopy is the time range that can be accessed in ’realtime’. That is, the experimenter varies a time variablesomewhere in the measuring instrument – e.g. the tim-ing of one part of a NMR pulse sequence with respect tothe rest, or the sledge of an optical delay line – and thattime is related directly to the time a particular struc-tural process takes in a one-to-one manner, i.e. the timefor interconversion of two molecular species. (For thepurpose of this review, we distinguish between structuraland energy relaxation processes, the latter of which canbe investigated as well by 2D-IR spectroscopy.) In thecontext of this definition, this is exchange spectroscopy(EXSY) and pump-probe spectroscopy in Fig. 1:

    • In an EXSY experiment, an initial part of a pulsesequence (preparation phase) is used to label asub-ensemble of molecules in a certain conforma-tion by exciting a spectrally distinct spin or vibra-tional state. The sub-ensemble is then given timeto change its conformation during a waiting period,after which the conformation is finally read out. Byvarying the waiting period one observes the con-formational transition in real time. EXSY has anupper limit of the accessible time window which isdictated by the excitation lifetime of the spin or vi-brational state used as a label. In the case of 2D-IRspectroscopy, the time-window ranges from about100 fs for e.g. the frequency fluctuations of theOH vibration in water (i.e. spectral diffusion) [23–26] to typically a few 10’s of picoseconds for e.g.hydrogen bond on- and off-times in solute-solvent-systems [7–9].

    • In a pump-probe experiment, a molecular systemis perturbed in some way (e.g. by triggering a pho-toreaction, or by a temperature jump) and therebybrought into a non-equilibrium state. As a re-sult, all molecules in the sample volume will un-dergo a conformational transition in a concertedway, and 2D-NMR or 2D-IR spectroscopy is thenused to elucidate a snap-shot structure of that en-semble on the fly. The lower time limit of pump-probe spectroscopy is dictated by the spectroscopicmethod (about 1 ps in the case 2D-IR spectroscopy,and 10’s of milliseconds for 2D-NMR spectroscopy),whereas no intrinsic limitation is imposed towards

  • 3

    longer timescales.

    From all the methods indicated in Fig. 1, pump-probespectroscopy is the only one which allows one to accessthe non-equilibrium regime. Comparing Fig. 1a withFig. 1b, clearly the biggest potential of IR spectroscopy,as compared to NMR spectroscopy, lies in the almost un-limited time window accessible to such non-equilibriumexperiments. This is why we made a strategic decisionsome time ago to explore and develop exactly this capa-bility of 2D-IR spectroscopy: Measuring snap-shot struc-tures of fast evolving molecular systems by means oftransient 2D-IR (T2D-IR) spectroscopy. This is wherethe dynamical aspect of NMR spectroscopy almost com-pletely breaks down.

    T2D-IR spectroscopy is a versatile technique that canbe applied to all sorts of problems in chemistry and bio-physics. For example, we have studied in detail the po-larization dependence of 2D-IR spectra of a metal-to-ligand charge transfer complex in its electronically ex-cited state [27], have established a technique that al-lows one to correlate vibrations in the electronic groundto those in a photo-product state (’labeling vibrationsby light’) [28], applied that technique to elucidate theconnectivity of binding sites after photo-dissociation ofCO in myoglobin [29], and studied solvation phenomenabeyond the linear response regime [30]. Very recently,T2D-IR experiments have also been reported by Tok-makoff and coworkers [31, 32] as well as Kubarych andcoworkers [33]. In this review we focus on the foldingdynamics of photoswitchable peptides [34–36]. Our ap-proach requires two ingredients that we currently developfurther in parallel: (i) 2D-IR and T2D-IR spectroscopy(Sec. IIA) and (ii) concepts of ultrafast photo-switchingof peptides (Sec. II B). In Sec. III, we report on the stagewe currently reached in merging these two pathways ofresearch.

    II. METHODOLOGY

    A. 2D-IR and Transient 2D-IR Spectroscopy

    While the first 2D-IR experiment was a pulsed fre-quency domain experiment [3], a technique that has beenadapted and refined by various groups [4, 6, 7, 11–13, 37–41], time domain pulsed Fourier transform tech-niques have been devised as well [8, 9, 23–26, 42–57]. Variants of 2D vibrational spectroscopy are 2D-Raman spectroscopy [58–60] or DOVE spectroscopy (i.e.doubly vibrationally enhanced four wave mixing, a hy-brid frequency-time domain spectroscopy) [61, 62]. 2Dspectroscopy has also been applied to electronic transi-tions [63–68], has been addressed extensively by theoreti-cal groups [10, 69–82] and has been reviewed from variousperspectives [5, 83? –91].

    The information content of the time domain and fre-quency domain implementations of 2D-IR spectroscopy is

    |0〉〉〉〉

    |1〉〉〉〉

    |2〉〉〉〉

    |0‘〉〉〉〉

    |1‘〉〉〉〉

    |2‘〉〉〉〉

    a

    |10’〉〉〉〉

    |20’〉〉〉〉|11‘〉〉〉〉

    |00‘〉〉〉〉

    |01‘〉〉〉〉

    |02‘〉〉〉〉

    b

    FIG. 2: 2D-IR spectrum of two vibrators that are (a) notcoupled or (b) coupled.

    largely identical [55, 92]: The pulsed-frequency domainexperiment does a pre-selection of the pump frequencywith the help of a tunable Fabry-Perot filter, whereasin the time-domain Fourier transform techniques thepump process is achieved by two phase-locked, spectrallybroad pulses that interfere in the sample. In the lat-ter case, the spectral information is regained by Fouriertransforming the data after the measurement. With re-spect to the probe process, both techniques are identical.The time-domain Fourier transform technique has con-ceptual advantages (highest possible spectral resolution,higher sensitivity, more control over pulse parameters ofall three field interactions) which, however, are boughton the expense of a more difficult experiment (interfer-ometic stability, accurate determination of the absolutephase) [92]. Zanni and coworkers have recently proposeda hybrid method based on an acousto-optical modulatorthat seems to combine the advantages of the two ap-proaches [54, 55].

    It is more intuitive to discuss the outcome of a 2D-IRexperiment in a frequency domain picture. First considertwo vibrators that are uncoupled, in which case a char-acteristic double-peak structure is obtained for each vi-brator on the diagonal of the 2D-IR spectrum (Fig. 2a).Whenever the pump-frequency is resonant with one ofthe vibrators, the latter is promoted into the first ex-cited state (|1〉 or |1′〉, respectively), and the probe beamthen probes the fundamental 0-1 transition (bleach andstimulated emission) as well as the first 1-2 overtone tran-sition (excited state absorption). Since chemical bondsare always weakly anharmonic with slightly different fre-

  • 4

    ω

    pu

    ωpr

    ωpr

    ω

    pr

    b ca

    FIG. 3: Schematic illustration of transient 1D (top) and 2D-IR (bottom) spectroscopy. The absorption bands shift fromthe steady state a to b, each yielding a characteristic 2D-IRspectrum. In c, difference 1D and 2D-IR spectra are shown.Negative response is depicted in blue, positive response in red.Adapted from Ref. [34].

    quencies of the 0-1 and the 1-2 transitions, the two peakswill separate along the probe frequency axis. Both peakshave opposite signs, and we usually color-code them asblue (negative, bleach and stimulated emission) and red(positive, excited state absorption).

    When the coupling between the two vibrators isswitched on (Fig. 2b) by bringing them into close spa-cial proximity, additional cross peaks appear in the off-diagonal region of the 2D-IR spectrum. This is since(for example) the |00′〉 → |10′〉 transition –exciting thefirst vibrator while the second is in its ground state– isdifferent in frequency from the |01′〉 → |11′〉 transition,where the second vibrator is already in its first excitedstate. This so-called intermode anharmonicity is a di-rect manifestation of the coupling between the two vi-brators [86]. The coupling can be of static nature (in thesense of a level scheme as shown in Fig. 2b, right), orof dynamic nature by the flow of energy between twovibrators [6]. In any case, the intensity of the crosspeak carries information on the distance, whereas theanisotropy of the cross peak reveals the relative orien-tation of the corresponding molecular groups. Both areimportant structure indicators which can be used to de-termine the unknown structure of small peptides from2D-IR spectroscopy [4, 11, 12, 43, 49].

    For reasons discussed in the introduction, we concen-trate on an extension of 2D-IR spectroscopy, transient2D-IR (T2D-IR) spectroscopy [27–30, 34–36]. For thatpurpose, we introduce an additional optical or UV pulsepreceding the 2D-IR pulse sequence, which triggers aphotochemical reaction, or a conformational transition,of the molecular system under study. The evolving non-equilibrium ensemble is then probed by the 2D-IR partof the experiment, and we may call the outcome a UV-pump-2D-IR-probe experiment.

    Since the UV pump pulse will in general not con-vert 100% of the reactant species into the photo-productspecies, the 2D-IR spectrum in the presence of the UV-switch pulse always contains contributions from both. In

    order to eliminate contributions of molecules that havenot absorbed any UV photon, two sets of 2D-IR spectraare recorded simultaneously - one with the UV-switchpulse on and one with the UV-switch pulse off - and sub-tracted from each other [34]. Hence, just like it is com-mon practice in conventional pump-probe spectroscopy,our T2D-IR spectra are in fact T2D-IR difference spec-tra.

    To facilitate the interpretation of T2D-IR spec-troscopy, we introduce the schematic response of an ideal-ized molecule with one single homogeneously broadenedoscillator, the frequency of which shifts upon triggeringthe photo-reaction. In analogy to Fig. 2, the vibratorwould give rise to a 2D-IR spectrum with two peaks ineach of the initial and product states: a negative (blue)bleach and stimulated emission signal at the vibrator’soriginal frequency and a positive (red) excited state ab-sorption signal, which is red-shifted owing to the intrin-sic anharmonicity of the oscillator. The T2D-IR spec-trum (Fig. 3c), however, is the difference of the 2D-IRspectra of the initial (Fig. 3a) and the transient species(Fig. 3b). In a transient 1D difference spectrum, nega-tive contributions arise from the depleted initial popula-tion, while positive contributions stem from the transientphoto-product (Fig. 3c, top). Accordingly, the signs ofthe various 2D-IR peaks (color-coded in Fig. 3) of theinitial and the product state are interchanged also in theT2D-IR difference spectrum (Fig. 3c, bottom).

    If the frequency shift upon the photo-reaction, as wellas the anharmonicities of reactant and product states, arelarge compared to the linewidth of a band, in total fourpeaks are expected along the diagonal for each individualresonance. Principally speaking, the same is true for eachcross peak. However, we will see that this is in generalnot the case, and the various contributions overlap andtend to strongly cancel due to their opposite signs. Inparticular for transient cross peaks, where we anticipatea change of coupling and hence a change of intermodeanharmonicity, only the two peaks related to the confor-mation with the stronger coupling tends to survive thiscancelation effect [35]. However, T2D-IR line shapes arenot yet thoroughly studied from the theoretical side.

    B. Photoswitchable Peptides

    The second important ingredient for studying non-equilibrium structural dynamics is the possibility to initi-ate a conformational transition of a peptide or a proteinon an ultrafast picosecond timescale. The best estab-lished approach to achieve this is a laser induced temper-ature jump to initiate, for example, folding or unfoldingof a peptide [93–101]. Temperature jumps can, in prin-ciple, be very fast and are only limited by the thermalequilibration time of the bulk solvent that is in the rangeof a few tens of picoseconds, depending on the concentra-tion of the molecule used to absorb light and to depositenergy into the solvent [102]. Nevertheless, most tem-

  • 5

    perature jumps studies so far were limited to the 10 nsregime for technical reasons: Q-switched Raman-shiftedNd-YAG lasers were used to generate the intense IR pulsewith an energy of typically 1 mJ needed to heat the sol-vent water directly by exciting an overtone of either itsOH or OD stretch vibration.

    However, the intrinsic time resolution of 2D-IR spec-troscopy is significantly faster (1 ps) than the pulse du-ration of suitable Q-switched lasers (10 ns). We felt thatlimiting the overall time resolution of the experiment to10 ns would potentially miss important advantages of 2D-IR spectroscopy (although this time resolution might besufficient for most of the problems of peptide dynamicsstudied by temperature jump experiments).

    Temperature jumps introduce another problem whencombining it with a 2D-IR experiment as a probe pro-cess: A significant amount of energy is deposited in arelatively small volume which causes a sudden jump ofpressure in a small sample volume. Strategies to reduceartifacts from the resulting shock waves and the result-ing index of refraction gradients have been worked outfor conventional pump-probe experiments [103], but 2D-IR spectroscopy, consisting of a sequence of IR pulses, ismore critical in this regard. The fact that the pulsed-frequency domain experiment is not a phase sensitivetechnique has proven to be a crucial advantage. Only re-cently, Tokmakoff and coworkers succeeded to cope withthe problem of phase distortions using the time-domainFourier transform technique [31, 32].

    We have chosen another approach to initiate confor-mational transitions in small peptides. By incorporat-ing a photoswitch into a peptide which undergoes a pho-tochemical reaction after electronic excitation, one mayalter the conformation of the peptide in a very con-trolled way. The switching time is ultrafast in mostcases (≈1 ps). Furthermore, since the photoswitch is di-rectly incorporated into the molecular system, its actionis much more local than that of a temperature jump.As a consequence, the overall energy pumped into thesample volume is much smaller (at least a factor of 10),and hence problems with shock waves are considerablyreduced (albeit still present in some cases). We distin-guish between photoswitches, when they are reversible,or phototriggers, when it is a one-time switch. Variousphotoswitches have been proposed in the literature, ofwhich we tested a few for our purpose to combine it withT2D-IR spectroscopy: azobenzene (Fig. 4a,b) as wellas disulfide-bridges and thiopeptides as intrinsic photo-switches (Fig. 4c,d). The use of photoswitches for variousapplications is reviewed in Refs. [104–106].

    a

    d

    c

    430 nm

    366 nm

    260 nm

    260 nm

    280 nm

    b430 nm

    366 nm

    FIG. 4: Photoswitchable peptides studied in our lab: Azoben-zene (a) embedded into peptide backbone (b) or cross-linkingamino acid side chains. Intrinsic photoswitches: (c) Photo-cleavable disulfide-bridge and (d) Thiopeptides.

    III. TRANSIENT 2D-IR SPECTROSCOPY OFPHOTOSWITCHABLE PEPTIDES

    A. Azobenzene

    The cis-trans isomerization of azobenzene is widelyused for photo-switching [106]. It is characterized by ahigh quantum yield (50%), high photostability, ultrafastswitching speed (about 1 ps) and the fact that cis andtrans isomers have absorption bands which are well sep-arated, and which therefore allow for a selective switch-ing in both directions [107]. Azobenzene has been incor-porated directly into the backbone of peptides by vari-ous groups [108–121], switching the conformation of thelatter by photo-exciting the former. In a first example,we concentrated on a small cyclized peptide designed byMoroder and coworkers. The peptide resembles the ac-tive site of thioredoxin reductase (Fig. 4a), and was cy-clized by an azobenzene-based ω-amino acid [112]. UVspectroscopy has been applied, which reports on the re-sponse of the chromophore [122–124], as well as IR spec-troscopy, which addresses directly the dynamics of thepeptide backbone [125].

    Isomerization of the photoswitch couples very directlyto a conformational transition of the peptide back-bone. In fact we find that the peptide backbone canbe stretched on an ultrafast 20 ps timescale upon cis-trans isomerization of an azo-moiety [125]. After this

  • 6

    1640

    1660

    1680

    1700

    3

    4

    γt = 12.2 cm-1

    γt = 12.8 cm-1

    γt = 13.5 cm-1

    1640

    1660

    1680

    1700

    1620 1640 1660 1680 1700 1720

    1640

    1660

    1680

    1700

    probe frequency [cm-1]

    pum

    p fr

    eque

    ncy

    [cm

    -1]

    pum

    p fr

    eque

    ncy

    [cm

    -1]

    probe frequency [cm-1]

    pum

    p fr

    eque

    ncy

    [cm

    -1]

    1620 1640 1660 1680 1700 1720 1740

    1 h

    l

    200 ps

    1.7 ns

    20 ps

    b

    c

    f

    e

    g

    j

    i

    k

    m

    d

    2

    a

    d

    b

    c

    e

    f

    g

    i

    h

    j

    l

    k

    FIG. 5: (a)–(l): Time resolved pump-probe (top of eachpanel) and T2D-IR spectra (bottom of each panel) atT=20 ps, 200 ps and 1.7 ns, respectively. The left columnshows experimental results, the right column the simulation.The labelled arrows refer to features described in the text.Adapted from Ref. [34].

    initial driven phase, the molecule keeps on equilibratingon a slower nanosecond timescale, but the major part ofthe conformational transition clearly occurs within thefirst 20 ps. This is concluded from the transient pump-probe signals which has built up already 20 ps after theUV pump pulse (Fig. 5a, Pos. 1) and changes onlymarginally between 20 ps and 1.7 ns (Fig. 5a,c,e). The20 ps timescale might seem surprisingly fast for a con-formational transition of a peptide. However, given thefact that a major fraction of the energy of the excitationphoton (290 kJ/mol) is used to drive the conformationaltransition, we concluded that the peptide backbone isstretched with an artificially high force

    Nevertheless, such cyclized azo-peptides appear to beideal model systems to test the capability of 2D-IR spec-troscopy in combining ultrafast time-resolution with di-rect structural resolution. This is why we used exactlythis system for our first demonstration of transient 2D-IR spectroscopy [34]. Despite the fact that hardly any

    time-dependence is observed in the pump-probe spectrabetween 20 ps and 1.7 ns (Fig. 5a,c,e), we find large dif-ferences in the T2D-IR spectra (Fig. 5b,d,f), illustratingthe information gain of the latter as compared to theformer. In particular, we find only a small signal fromthe photoswitched peptides in the T2D-IR spectrum af-ter 20 ps (expected at Pos. 2 in Fig. 5b) although it givesrise to a band in the 1D spectrum (Fig. 5a, Pos. 1). Onlythe T2D-IR signal of the initial cis state is observed. Af-ter 200 ps, however the band of the trans-ensemble hasgrown in (Fig. 5d, Pos. 3) and within 1.7 ns we observe atilt of this band towards the diagonal of the 2D spectrum(Fig. 5f, Pos. 4).

    Interestingly, the T2D-IR spectra can be modeled al-most quantitatively by just varying the homogeneous linewidth γt of the amide I band of the photo product by 10-15% between 20 ps and 1.7 ns (Fig. 5h, j, l). In particularthe low intensity of the transient product band at 20 ps aswell as its rise at later times are perfectly matched. Tran-sient 2D-IR spectroscopy is sensitive to relatively smallchanges in the lineshape function due to strong cancela-tion effects in the double-difference spectrum.

    The homogeneous linewidth γt in 2D-IR experimentsis a measure of fast (< 1 ps) fluctuations of the pep-tide backbone. The time scale relevant for homogeneousbroadening (

  • 7

    1625 1650 1675 1700 1725

    1625

    1650

    1675

    1700

    ωpu

    [cm

    -1]

    ωpr

    [cm -1]

    1625 1650 1675 1700 1725

    ωpr

    [cm -1]

    Equilibrium 2D Transient 2D

    100 ps

    a

    db

    c

    FIG. 6: (a+b) Equilibrium and (c+d) transient non-equilibrium spectra 100 ps after photo-cleavage of cyclo-(Boc-Cys-Pro-Aib-Cys-OMe). The top panels (a+c) show 1D-IR spectra and the bottom panels (b+d) the corresponding2D-IR spectra. The equilibrium 2D-IR spectrum (b) is theweighted difference between parallel and perpendicular po-larization measurement (to enhance the cross-peak contribu-tion), whereas the transient 2D-IR spectrum (d) is with allthree laser beams polarized parallel. The crosspeak markedby circles in either spectrum relates to the coupling across thehydrogen bond in the β-turn structure. The example struc-tures were obtained form an accompanying MD simulation.Adapted from Ref. [35]

    B. Intrinsic Photoswitches

    The disulfide bridge between two cysteines, that occursin natural peptides and proteins, may be used as pre-determined breaking point that photo-dissociates uponelectronic excitation at ≈260 nm, and hence as photo-trigger. This concept was first reported in pioneeringwork of Hochstrasser and coworkers [129, 130]. A 20residue peptide was designed to fold into an α-helix af-ter photo-cleavage, while no distinct secondary struc-ture was possible in the initial bridged state. Unfor-tunately, geminate recombination of the liberated thiylradicals occurred before helix formation could be de-tected. We have recently used this approach for a muchsmaller peptide (Fig. 4c) [131] that forms a β-turn struc-ture in the disulfide-bridged configuration. Upon photo-cleavage of the disulfide bridge, the hydrogen bond inthe β-turn is destabilized and the ring structure opens.In contrast to the molecule studied by Hochstrasser andcoworkers [129, 130], the molecule is much more straineddue to the shortness of the backbone and stays open onlong timescales. In fact, only about half of the radicalsare quenched after 1 ms either intra- or intermolecular-ily [132].

    The peptide features a completely resolved and easy

    to assign IR spectrum of the five C=O groups withouthaving to invoke isotope labeling. This fortunate sit-uation made it the ideal test case for transient 2D-IRspectroscopy that allowed us to directly extract micro-scopic structural information on an ultrafast timescalefor the conformational transition of the molecule [35].Fig. 6b shows an equilibrium 2D-IR spectrum of themolecule which is rather congested since essentially allC=O groups are coupled to each other. Fig. 6d, in con-trast, shows a transient 2D-IR spectrum 100 ps afterphoto-cleavage of the disulfide bridge. The diagonal isdominated by signals similar to those in Fig. 5. Moreimportant for the structural resolution aspect is the off-diagonal region, where only one dominant peak remains,related to the coupling between Cys(1)-Aib(3) across theintramolecular hydrogen bond. The corresponding crosspeak is weak in the equilibrium case (Fig. 6b) and ishidden underneath the wings of the much stronger crosspeaks of Boc-Cys(1) and Cys(1)-Pro(2), both related tonearest neighbors in the sequence. In contrast, Cys(1)and Aib(3) are ’far in sequence’, but close in space in thefolded structure. Therefore, upon opening of the ringstructure, changes in coupling are expected only for thelatter in the transient 2D-IR spectrum, which is a differ-ence 2D-IR spectrum between photo-product state andreactant state. A time series showed that the transientcross-peak rises on a 100 ps timescale, in good agreementwith accompanying results from MD simulations [35].The transient cross-peak constitutes a direct and easyto interpret structural indicator.

    Another potential intrinsic photoswitch is the -CO-NH- peptide unit itself, which is thermodynamically morestable in the trans state, but which can be switched to thecis conformation by electronic ππ∗ excitation [133], forc-ing the peptide to change its secondary structure [134,135]. However, the required excitation wavelength liesdeep in the UV (190 nm) which is hard too reachtechnically and which can destroy the molecule. Ex-citation at 190 nm is also non-selective and can ran-domly switch any peptide unit in a polypeptide. Bothproblems are avoided by replacing one peptide unit bya thiopeptide (i.e. -CSNH-) unit (Fig. 4d) [136–140].This red-shifts the wavelength of the ππ∗ transition ofthe trans isomer to ≈260 nm [141, 142] which is eas-ily accessible by frequency-tripling a Ti:S laser. Fur-thermore, incorporation of a thioamide linkage resultsin only minor changes of the secondary structure of apeptide [143, 144]. A combination of theoretical andtime-resolved experimental studies on the photoswitchitself (N -Methylthioacetamide) have shown that the iso-merization quantum yield is high (30-40% in the trans-cis and 60-70% in the cis-trans direction) although themolecule remains trapped in a relatively long-lived ex-cited state for 200-300 ps, where it is almost free torotate about the thioamide bond [145, 146]. Neverthe-less, thiopeptides are particularly useful as model sys-tems for T2D-IR spectroscopy, because the isomeriza-tion of the thioamide bond changes the relative orienta-

  • 8

    tion of the adjacent peptide units, which are sufficientlyclose to each other for amide I mode coupling. In anal-ogy to trialanine in 2D-IR spectroscopy the smallest testmolecule for T2D-IR is thus a tetrathiopeptide, with twoamide I oscillators separated by the thio-photoswitch.The first thiopeptide we have investigated by T2D-IRspectroscopy (Boc-Ala-Gly(=S)-Ala-Aib-OMe) [36, 147],in addition contains two protecting groups (Boc andAib/OMe, providing two more C=O oscillators), whichstabilize the trans-molecules in a looped conformation viaan intramolecular hydrogen bond. In the T2D-IR spec-tra, we observe resolution enhancement along the diago-nal as well as a transient cross peak due to the breakingof that hydrogen bond [36].

    IV. DISCUSSION

    Comparing the results of Ref. [34] (Fig. 5) with thoseof Refs. [35, 36] (Fig. 6), one may suggest concrete de-sign criteria to render T2D-IR spectroscopy an expressivemethod to time-resolve transient structures. In the caseof the azobenzene switched cyclic octapeptide of Ref. [34],the first example we investigated, the amide I′ band wasnot spectrally resolved, hence, we could not interpret theT2D-IR spectra in terms of structural motifs. The struc-tural information did not go any deeper than that of thecorresponding 1D spectroscopy; nevertheless, additionaldynamic information could be extracted. In contrast, incase of the disulfid-bridged β-turn [35], we deal with asomewhat smaller peptide whose amide I′ spectrum isfully resolved. As a result, specific local contacts can beidentified whose time-dependence can be interpreted interms of the breaking of certain loop structures.

    In the examples of Refs. [35, 36], bands are resolvedsince the different C=O groups are sitting in differentchemical environments. For larger peptides, this will ingeneral not be the case. However, spectral resolutioncan always be regained when employing isotope label-ing [12, 46, 49, 147]. The synthesis of site-selectivelyisotope labelled peptides is not particularly demanding.For example, in our studies of the folding of a photo-switchable α-helix (Fig. 4b), we employ double 13C16Oand 13C18O isotope labeling in order to single out twospecific sites from the unstructured main amide I′ bandof unlabeled groups. 13C16O-labeling causes a frequencydownshift of the amide I′ band of about 30 cm−1, whereas13C18O shifts it by 50 cm−1. In that way, the local con-tact between a specific pair of amino acids can be iden-tified (e.g. the i → i + 4 configuration of a α-helicalloop [49]), and its time-dependence can potentially bemeasured. These experiments promise to give unprece-dented structural insights into which helix-loop is form-ing at what time to what extend during the folding pro-cess.

    An important, but not yet thoroughly explored issue isthe lineshapes of T2D-IR spectra [148]. If the frequencyshifts induced by the photo-reaction are larger than the

    linewidth of the band, in total four spectrally resolvedbands are expected for each diagonal and cross peak inthe T2D-IR spectrum (Fig. 3). This is the case for e.g.the metal-to-ligand charge transfer of a metal-carbonylcompound we studied extensively ([Re(CO)3(dmbpy)Cl],dmbpy=4,4’-dimethyl-2,2’bipyridine) [27]. However, inthe case of photoswitchable peptides, the transient fre-quency shifts that are induced by a conformational tran-sition are typically very small – much smaller than theline width of the amide I band. In this case the two con-tributions Fig. 3a and Fig. 3b will overlap and tend tocancel each other in Fig. 3c. If the lineshape functionsof reactant and product species are identical (which willnot be the case in general), then the T2D-IR is expectedto be symmetric, as in Fig. 3c. Due to the strong cance-lation between reactant and product contributions, how-ever, small changes in the line shape functions of eithercontribution may cause strong effects for the resultingdifference T2D-IR spectrum (see Fig. 5b, Position 2).

    The theoretical modeling of these cancelation effects isquite demanding and has not yet been achieved. First,it requires very high accuracy to obtain still reasonableresults after subtracting the two contributions from re-actant and product state, and second, so far almost alllineshape calculations are based on an equilibrium as-sumption [70]. However, it is not clear after which timeone can think about an evolving structural ensemble asbeing in a quasi-equilibrium. We have shown that this as-sumption is definitely not valid on a 10 ps timescale [30],but we do not know whether it might become reasonableon a 100 ps to 1 ns timescale [34, 35]. This problem callsfor significant effort from the theoretical side.

    V. CONCLUSION

    NMR is a high-resolution spectroscopy, capable tospectrally isolate thousands of individual resonances.However, the narrow line width implies that direct timeresolution of NMR spectroscopy is intrinsically slow. Thetime resolution is ultimately limited by the dephasing(T2) time of the spectroscopic transition used as molecu-lar probe (typically 1 ms for spin transitions), rather thanby the measurement instrument. In contrary, typical de-phasing times of vibrational transitions in the solutionphase are in the range of 1 ps, hence IR bands are broad,and the number of resolved bands in a given spectralwindow is restricted. Nevertheless, the fast dephasingtime is responsible for the high direct time resolution of2D-IR spectroscopy. Spectral congestion can be circum-vented to a significant extent by isotope labeling. In thatsense, 2D-IR and 2D-NMR spectroscopy should not beviewed as competing methods, rather 2D-IR spectroscopyshould be further developed as a complementary method.We have focussed here on transient non-equilibrium 2D-IR spectroscopy as one of the most promising strengthsof 2D-IR spectroscopy (it certainly is not the only one).Learning to use the structure resolution capabilities of

  • 9

    2D-IR spectroscopy, also in equilibrium, is of course aninevitable prerequisite before extending it to the non-equilibrium regime, and work in this direction is still re-quired.

    Since the typical length and time scales of T2D-IRspectroscopy coincides exactly with what is accessibleto MD simulations, the latter can directly be testedagainst experiment. MD simulations currently are theonly approach to understand and visualize the compli-cate structure and dynamics - and hence the function -of biomolecules on an atomic level. Yet, it is clear thatthe MD approach is based on, in part very crude, approx-imations, while experimental techniques which can pro-vide pictures of similar clarity are scarce. With Ref. [35],we made a significant step towards a ’molecular movie’of ultrafast structural dynamics extracted directly fromexperimental data.

    Acknowledgement: We are thankful to numerouscontributions from our coauthors, in particular ChristophKolano, Mariusz Kozinski, Luis Moroder, Christian Ren-ner, Wolfram Sander, Gerhard Stock, Josef Wachtveitl,Sander Woutersen and Wolfgang Zinth. The work hasbeen financially supported by the University of Zurich,the Swiss Science Foundation (SNF) and the GermanScience Foundation (DFG).

    REFERENCES

    [1] Ernst RR, Bodenhausen G, Wokaun A. Principlesof Nuclear Magnetic Resonance in One and TwoDimensions. Clarendon Press, Oxford, 1987.

    [2] Aue WP, Bartholdi E, Ernst RR. 1976. Two-dimensional spectroscopy. application to nuclearmagnetic resonance. J. Chem. Phys. 64:2229.

    [3] Hamm P, Lim M, Hochstrasser RM. 1998. Struc-ture of the amide I band of peptides measuredby femtosecond nonlinear-infrared spectroscopy. J.Phys. Chem. B 102:6123–6138.

    [4] Hamm P, Lim M, DeGrado WF, Hochstrasser RM.1999. The two-dimensional IR nonlinear spec-troscopy of a cyclic penta-peptide in relation to itsthree-dimensional structure. Proc. Natl. Acad. Sci.USA 96:2036–2041.

    [5] Ge NH, Hochstrasser RM. 2002. Femtosecond two-dimensional infrared spectroscopy: IR-COSY andTHIRSTY. Phys. Chem. Comm. 5:17–26.

    [6] Woutersen S, Mu Y, Stock G, Hamm P. 2001. Sub-picosecond conformational dynamics of small pep-tides probed by two-dimensional vibrational spec-troscopy. Proc. Natl. Acad. Sci. USA 98:11254–11258.

    [7] Woutersen S, Hamm P. 2001. Hydrogen-bondlifetime measured by time-resolved 2D-IR spec-troscopy: N-methylacetamide in methanol. Chem.Phys. 266:137–147.

    [8] Zheng J, Kwak K, Asbury J, Chen X, Piletic IR,Fayer MD. 2005. Ultrafast dynamics of solute-

    solvent complexation observed at thermal equilib-rium in real time. Science 309:1338–1343.

    [9] Kim YS, Hochstrasser RM. 2005. Chemical ex-change 2d ir of hydrogen-bond making and break-ing. Proc. Natl. Acad. Sci. USA 102:11185–11190.

    [10] Scheurer C, Mukamel S. 2001. Design strategies ofpulse sequences in multidimensional optical spec-troscopy. J. Chem. Phys. 115:4989–5004.

    [11] Woutersen S, Hamm P. 2000. Structure determi-nation of trialanine in water using polarization sen-sitive two-dimensional vibrational spectroscopy. J.Phys. Chem. B 104:11316–11320.

    [12] Woutersen S, Hamm P. 2001. Isotope-edited two-dimensional vibrational spectroscopy of trialaninein aqueous solution. J. Chem. Phys. 114:2727–2737.

    [13] Woutersen S, Pfister R, Hamm P, Mu Y, KosovDS, Stock G. 2002. Peptide conformational hetero-geneity revealed from nonlinear vibrational spec-troscopy and molecular dynamics simulations. J.Chem. Phys. 117:6833–6840.

    [14] Mu YG, Kosov DS, Stock G. 2003. Conformationaldynamics of trialanine in water. 2. comparison ofAMBER, CHARMM, GROMOS, and OPLS forcefields to NMR and infrared experiments. J. Phys.Chem. B 107:5064–5073.

    [15] Gnanakaran S, Garcia AE. 2003. Validation ofan all-atom protein force field: From dipeptides tolarger peptides. J. Phys. Chem B 107:12555–12557.

    [16] Graf J, Nguyen PH, Stock G, Schwalbe H. 2007.Structure and dynamics of the homologous seriesof alanine peptides: A joint molecular dynam-ics/NMR study. J. Am. Chem. Soc. 129:1179–1189.

    [17] Lakomek NA, Carlomagno T, Becker S, GriesingerC, Meiler J. 2006. A thorough dynamic interpre-tation of residual dipolar couplings in ubiquitin. J.Biomol. NMR 34:101–115.

    [18] Fung BM, McGaughy TW. 1976. Molecular mo-tions in liquid. I. rotation of water and small al-cohols studied by deuteron relaxation. J. Chem.Phys. 65:2970–2976.

    [19] Woutersen S, Emmerichs U, Bakker HJ. 1997. Fem-tosecond mid-infrared pump-probe spectroscopy ofliquid water: evidence for a two-component struc-ture. Science 278:658–660.

    [20] Fecko CJ, Loparo J, Roberts ST, Tokmakoff A.2005. Local hydrogen bonding dynamics andcollective reorganization in water: Ultrafast in-frared spectroscopy of HOD/D2O. J. Chem. Phys.122:054506.

    [21] Rezus YLA Bakker HJ. 2005. On the orientationalrelaxation of HDO in liquid water. J. Chem. Phys.123:114502.

    [22] Laage D, Hynes JT. 2006. A molecular jump mech-anism of water reorientation. Science 311:832–835.

    [23] Asbury JB, Steinel T, Stromberg C, Gaffney KJ,Piletic IR, Goun A, Fayer MD. 2003. Hydro-gen bond dynamics probed with ultrafast infraredheterodyne-detected multidimensional vibrational

  • 10

    stimulated echoes. Phys. Rev. Lett. 91:237402.[24] Yeremenko S, Pschenichnikov MS, Wiersma DA.

    2003. Hydrogen-bond dynamics in water exploredby heterodyne-detected photon echo. Chem. Phys.Lett. 369:107–113.

    [25] Cowan ML, Bruner BD, Huse N, Dwyer JR, ChughB, Nibbering ETJ, Elsaesser T, Miller RJD. 2005.Ultrafast memory loss and energy redistribution inthe hydrogen bond network of liquid H2O. Nature434:199–202.

    [26] Eaves JD, Loparo JJ, Fecko CJ, Roberts ST, Tok-makoff A, Geissler PL. 2005. Hydrogen bonds inliquid water are broken only fleetingly. Proc. Natl.Acad. Sci. USA 102:13019–13022.

    [27] Bredenbeck J, Helbing J, Hamm P. 2004. Transienttwo-dimensional IR spectroscopy - exploring thepolarization dependence. J. Chem. Phys 121:5943.

    [28] Bredenbeck J, Helbing J, Hamm P. 2004. Labelingvibrations by light: Ultrafast transient 2D-IR spec-troscopy tracks vibrational modes during a photore-action. J. Am. Chem. Soc. 126:990–991.

    [29] Bredenbeck J, Helbing J, Nienhaus K, NienhausGU, Hamm P. 2007. Protein ligand migrationmapped by nonequilibrium 2D-IR exchange spec-troscopy. Proc. Natl. Acad. Sci. USA in press.

    [30] Bredenbeck J, Helbing J, Hamm P. 2005. Solvationbeyond the linear response regime. Phys. Rev. Lett.95:083201.

    [31] Chung HS, Ganim Z, Jones KC, Tokmakoff A.2007. Transient 2D IR spectroscopy of ubiquitinunfolding dynamics. Proc. Natl. Acad. Sci. USA inpress.

    [32] Chung HS, Khalil M, Smith AW, Tokmakoff A.2007. Transient two-dimensional IR spectrometerfor probing nanosecond temperature-jump kinetics.Rev. Sci. Instrum. 78:063101.

    [33] Kubarych KJ and coworkers. unpublished work pre-sented at the ACS Meeting, Chicago, March 2007.

    [34] Bredenbeck J, Helbing J, Renner C, Behrendt R,Moroder L, Wachtveitl J, Hamm P. 2003. Transient2D-IR spectroscopy: Snapshots of the nonequilib-rium ensemble during the picosecond conforma-tional transition of a small peptide. J. Phys. Chem.B 107:8654–8660.

    [35] Kolano C, Helbing J, Kozinski M, Sander W,Hamm P. 2006. Watching hydrogen bond dynam-ics in a β-turn by tranient two-dimensional infraredspectroscopy. Nature 444:469–472.

    [36] Cervetto V, Pfister R, Hamm P, Helbing J. 2007.Transient 2D-IR spectroscopy of thiopeptide iso-merization and hydrogern bond breaking. J. Phys.Chem. B to be submitted.

    [37] Rubtsov IV, Wang J, Hochstrasser RM. 2003. Dualfrequency 2D-IR of peptide amide-A and amide-Imodes. J. Chem. Phys. 118:7733–7736.

    [38] Larsen OFA, Bodis P, Buma WJ, Hannam JS,Leigh DA, Woutersen S. 2005. Probing thestructure of a rotaxane with two-dimensional in-

    frared spectroscopy. Proc. Natl. Acad. Sci. USA102:13378–13382.

    [39] Volkov VV, Chelli R, Righini R. 2006. Domain for-mation in lipid bilayers probed by two-dimensionalinfrared spectroscopy. Opt. Lett. 110:1499–1501.

    [40] Barbour LW, Hegadorn M, Asbury JB. 2006. Mi-croscopic inhomogeneity and ultrafast orientationalmotion in an organic photovoltaic bulk heterojunc-tion thin film studied with 2D IR vibrational spec-troscopy. J. Phys. Chem. B 110:24281–24286.

    [41] DeCamp MF, DeFlores LP, Jones KC, TokmakoffA. 2007. Single-shot two-dimensional infrared spec-troscopy. Opt. Express 15:233–241.

    [42] Asplund MC, Zanni MT, Hochstrasser RM. 2000.Two-dimensional infrared spectroscopy of peptidesby phase-controlled femtosecond vibrational pho-ton echoes. Proc. Natl. Acad. Sci. USA 97:8219–8224.

    [43] Zanni MT, Gnanakaran S, Stenger J, HochstrasserRM. 2001. Heterodyned two-dimensional infraredspectroscopy of solvent-dependent conformations ofacetylproline-NH2. J. Phys. Chem. B 105:6520–6535.

    [44] Khalil M, Demirdöven N, Tokmakoff A. 2003. Ob-taining absorptive line shapes in two-dimensionalinfrared vibrational correlation spectra. Phys. Rev.Lett. 90:047401.

    [45] Krummel AT, Mukherjee P, Zanni MT. 2003. Interand intrastrand vibrational coupling in DNA stud-ied with heterodyned 2D-IR spectroscopy. J. Phys.Chem. B 107:9165.

    [46] Mukherjee P, Kass I, Arkin IT, Zanni MT. 2006. Pi-cosecond dynamics of a membrane protein revealedby 2D IR. Proc. Natl. Acad. Sci. USA 103:3528–3533.

    [47] Demirdöven N, Cheatum CM, Chung HS, Khalil M,Knoester J, Tokmakoff A. 2004. Two-dimensionalinfrared spectroscopy of antiparallel beta-sheet sec-ondary structure. J. Am. Chem. Soc. 126:7981–7990.

    [48] Cowan ML, Ogilvie JP, Miller RJD. 2004. Two-dimensional spectroscopy using diffractive opticsbased phase-locked echoes. Chem. Phys. Lett.386:184–189.

    [49] Fang C, Hochstrasser RM. 2005. Two-dimensionalinfrared spectra of the 13C=18O isotopomers of ala-nine residues in an alpha-helix. J. Phys. Chem. B109:18652–18663.

    [50] Kurochkin DV, Naraharisetty S. RG, Rubtsov IV.2005. Dual-frequency 2D IR on interaction of weakand strong IR modes. J. Phys. Chem. A 109:10799–10802.

    [51] Volkov V, Schanz R, Hamm P. 2005. Active phasestabilization in Fourier-transform two-dimensionalinfrared spectroscopy. Opt. Lett. 30:2010–2012.

    [52] Maekawa H, Toniolo C, Moretto A, BroxtermanQB, Ge NH. 2006. Different spectral signaturesof octapeptide 3(10) and alpha-helices revealed by

  • 11

    two-dimensional infrared spectroscopy. J. Phys.Chem. B 110:5834–5837.

    [53] Ding F, Mukherjee P, Zanni MT. 2006. Passivelycorrecting phase drift in two-dimensional infraredspectroscopy. Opt. Lett. 31:2918–2920.

    [54] Shim SH, Strasfeld DB, Fulmer EC, Zanni MT.2006. Femtosecond pulse shaping directly in themid-IR using acousto-optic modulation. Opt. Lett.31:838–840.

    [55] Shim SH, Strasfeld DB, Ling YL, Zanni MT.2007. Using pulse shaping to automate 2D-IR spec-troscopy for more precise characterization of struc-ture and dynamcis. Proc. Natl. Acad. Sci. USApage in press.

    [56] Kozinski M, Garrett-Roe S, Hamm P. 2007. Vibra-tional spectral diffusion of CN− in water. Chem.Phys. page in press.

    [57] Nee MJ, McCanne R, Kubarych KJ, Joffre M. 2007.Two-dimensional infrared spectroscopy detected bychirped pulse upconversion. Opt. Lett. 32:713–715.

    [58] Golonzka O, Demirdöven N, Khalili M, TokmakoffA. 2000. Separation of cascaded and direct fifth-order Raman signals using phase-sensitive intrinsicheterodyne detection. J. Chem. Phys. 113:9893–9896.

    [59] Milne CJ, Li YL, Jansen T. LC, Huang L, MillerRJD. 2006. Fifth-order Raman spectroscopy ofliquid benzene: Experiment and theory. J. Phys.Chem. B 110:19867–19876.

    [60] Kaufman LJ, Heo J, Ziegler LD, Fleming GR. 2002.Heterodyne-detected fifth-order nonresonant Ra-man scattering from room temperature CS2. Phys.Rev. Lett. 88:207402.

    [61] Zhao W, Wright JC. 2000. Doubly vibrationallyenhanced four wave mixing: The optical analog to2D-NMR. Phys. Rev. Lett. 84:1411–1414.

    [62] Donaldson PM, Guo R, Fournier F, Gardner EM,Barter LM, Palmer DJ, Barnett CJ, Willison KR,Gould IR, Klug DR. 2007. Direct identificationand decongestion of fermi resonances by control ofpulse time-ordering in 2d-ir spectroscopy. J. Chem.Phys. in press.

    [63] Gallagher Faeder SM, Albrecht AW, Hybl JD,Landin BL, Rajaram B, Jonas DM. 1998. Hetero-dyn detection of the comlete electric field of fem-tosecond four-wave mixing signals. J. Opt. Soc.Am. B 15:2338–2345.

    [64] Tian PF, Keusters D, Suzaki Y, Warren WS. 2003.Femtosecond phase-coherent two-dimensional spec-troscopy. Science 300:1553–1555.

    [65] Brixner T, Stenger J, Vaswani HM, Cho M,Blankenship RE, Fleming GR. 2005. Two-dimensional spectroscopy of electronic couplings inphotosynthesis. Nature 434:625–628.

    [66] Li X, Zhang T, Borca CN, Cundiff ST. 2006.Many-body interactions in semiconductors probedby optical two-dimensional fourier transform spec-troscopy. Phys. Rev. Lett. 96:057406.

    [67] Engel GS, Calhoun TR, Read EL, Ahn TK, Man-cal T, Cheng YC, Blankenship RE, Fleming GR.2007. Evidence for wavelike energy transfer throughquantum coherence in photosynthetic systems. Na-ture 446:782 – 786.

    [68] Szocs V, Palszegi T, Lukes V, Sperling J, MilotaF, Jakubetz W, Kauffmann HF. 2006. Two-dimensional electronic spectra of symmetric dimers:Intermolecular coupling and conformational states.J. Chem. Phys. 124:124511.

    [69] Tanimura Y, Mukamel S. 1993. 2-Dimensionalfemtosecond vibrational spectroscopy of liquids. J.Chem. Phys. 99:9496–9511.

    [70] Mukamel S. Principles of Nonlinear Optical Spec-troscopy. Oxford University Press, Oxford, 1995.

    [71] Hamm P, Woutersen S. 2002. Coupling of the amideI modes of the glycine di-peptide. Bull. Chem. Soc.Jpn. 75:985–988.

    [72] Kwac K, Cho M. 2003. Molecular dynamics sim-ulation study of N-methylacetamide in water. II.Two-dimensional infrared pump-probe spectra. J.Chem. Phys. 119:2256–2263.

    [73] Kwac K, Lee H, Cho M. 2004. Non-Gaussianstatistics of amide I mode frequency fluctuation ofN-methylacetamide in methanol solution: Linearand nonlinear vibrational spectra. J. Chem. Phys.120:1477–1490.

    [74] Schmidt JR, Corcelli SA, Skinner JL. 2004. Ultra-fast vibrational spectroscopy of water and aqueousn-methylacetamide: Comparison of different elec-tronic structure/molecular dynamics approaches.J. Chem. Phys. 121:8887.

    [75] Dijkstra AG, Knoester J. 2005. Collective oscilla-tions and the linear and two-dimensional infraredspectra of inhomogeneous beta-sheets. J. Phys.Chem. B 109:9787–9798.

    [76] Gorbunov RD, Kosov DS, Stock G. 2005. Ab initio-based exciton model of amide i vibrations in pep-tides: Definition, conformational dependence, andtransferability. J. Chem. Phys. 122:224904.

    [77] Zhuang W, Abramavicius D, Hayashi T, MukamelS. 2006. Simulation protocols for coherent fem-tosecond vibrational spectra of peptides. J. Phys.Chem. B 110:3362–3374.

    [78] Jansen TL, Knoester J. 2006. Nonadiabatic effectsin the two-dimensional infrared spectra of peptides:Application to alanine dipeptide. J. Phys. Chem.B 110:22910–22916.

    [79] Bounouar M, Scheurer C. 2006. Reducing the vi-brational coupling network in N-methylacetamideas a model for ab initio infrared spectra computa-tions of peptides. Chem. Phys. 323:87–101.

    [80] Torii H. 2006. Effects of intermolecular vibra-tional coupling and liquid dynamics on the po-larized Raman and two-dimensional infrared spec-tral profiles of liquid N,N-dimethylformamide ana-lyzed with a time-domain computational method.J. Phys. Chem. A 110:4822–4832.

  • 12

    [81] Gorbunov RD, Nguyen PH, Kobus M, Stock G.2007. Quantum-classical description of the amide Ivibrational spectrum of trialanine. J. Chem. Phys.126:054509.

    [82] Kjellberg P, Brüggemann B, Pullerits T. 2007.Two-dimensional electronic spectroscopy of an ex-citonically coupled dimer. Phys. Rev. B 74:024303.

    [83] Mukamel S. 2000. Multidimensional femtosecondcorrelation spectroscopies of electronic and vibra-tional excitations. Annu. Rev. Phys. Chem. 51:691–729.

    [84] Hamm P, Hochstrasser RM. Structure and dy-namics of proteins and peptides: Femtosecond two-dimensional infrared spectroscopy. In Fayer MD,editor, Ultrafast Infrared and Raman Spectroscopypages 273–347, New York, 2001. Marcel Dekker.

    [85] Zanni MT, Hochstrasser RM. 2001. Two-dimensional infrared spectroscopy: a promisingnew method for the time resolution of structures.Curr. Opin. Struct. Biol. 11:516–522.

    [86] Woutersen S, Hamm P. 2002. Nonlinear 2D vibra-tional spectroscopy of peptides. J. Phys.: Condens.Matter 14:R1035–R1062.

    [87] Jonas DM. 2003. Two-dimensional femtosecondspectroscopy. Ann. Rev. Phys. Chem. 54:425–463.

    [88] Cho M. 2006. Coherent two-dimensional opticalspectroscopy. Bull. Korean Chem. Soc. 27:1940–1960.

    [89] Zheng J, Kwak K, Fayer MD. 2007. Ultrafast 2D IRvibrational echo spectroscopy. Acc. Chem. Research40:75–83.

    [90] Bredenbeck J, Helbing J, Kolano C, Hamm P. 2007.Ultrafast 2D-IR spectroscopy of transient species.ChemPhysChem. in press.

    [91] Scheurer C, Steinel T. 2007. 2D infrared chemicalexchange spectroscopy of ultrafast isomerizations.ChemPhysChem 8:503–505.

    [92] Cervetto V, Helbing J, Bredenbeck J, Hamm P.2004. Double-resonance versus pulsed fourier trans-form 2D-IR spectroscopy: An experimental andtheoretical comparison. J. Chem. Phys. 121:5935–5942.

    [93] Williams S, Causgrove TP, Gilmanshin R, Fang KS,Callender RH, Woodruff WH, Dyer RB. 1996. Fastevents in protein folding: Helix melting and forma-tion in a small peptide. Biochemistry 35:691–697.

    [94] Thompson PA, Eaton WA, Hofrichter J. 1997.Laser temperature jump study of the helix-coil ki-netics of an alanine peptide interpreted with a ’ki-netic zipper’ model. Biochemistry 36:9200–9210.

    [95] Lednev IK, Karnoup AS, Sparrow MC, Asher SA.1999. α-Helix peptide folding and unfolding activa-tion barriers: a nanosecond UV resonance Ramanstudy. J. Am. Chem. Soc. 121:8074–8086.

    [96] Sabelko J, Ervin J, Gruebele M. 1999. Observationsof strange kinetics in protein folding. Proc. Natl.Acad. Sci. USA 96:6031–6036.

    [97] Thompson PA, Muñoz V, Jas GS, Henry ER, Eaton

    WA, Hofrichter J. 2000. The helix-coil kinetics ofa heteropeptide. J. Phys. Chem. B 104:378–389.

    [98] Huang CY, Klemke JW, Getahun Z, DeGrado WF,Gai F. 2001. Temperature-dependent helix-coiltransition of an alanine based peptide. J. Am.Chem. Soc. 123:9235–9238.

    [99] Werner JH, Dyer RB, Fesinmeyer RM, AndersenNH. 2002. Dynamics of the primary processes ofprotein folding: Helix nucleation. J. Phys. Chem.B 106:487–494.

    [100] Huang CY, Getahun Z, Zhu Y, Klemke JW, De-Grado WF, Gai F. 2002. Helix formation via con-formation diffusion search. Proc. Natl. Acad. Sci.USA 99(5):2788–2793.

    [101] Petty SA Volk M. 2004. Fast folding dynamics ofan α-helical peptide with bulky side chains. Phys.Chem. Chem. Phys. 6:1022–1030.

    [102] Phillips CM, Mizutani Y, Hochstrasser RM. 1995.Ultrafast thermally induced unfolding of RNase A.Proc. Natl. Acad. Sci. USA 92:7292–7296.

    [103] Wray WO, Aida T, Dyer RB. 2002. Photoacous-tic cavitation and heat transfer effects in the laser-induced temperature jump in water. Appl. Phys.B-Lasers Opt. 74:57–66.

    [104] Mayer G, Heckel A. 2006. Biologically activemolecules with a ’light switch’. Angew. Chem. Int.Ed. 45:4900–4921.

    [105] Pieroni O, Fissi A, Angelini N, Lenci F. 2001. Pho-toresponsive polypeptides. Accounts Chem. Res.34:9–17.

    [106] In Feringa BL, editor, Molecular Switches Wein-heim, 2001. Wiley.

    [107] Nägele T, Hoche R, Zinth W, Wachtveitl J.1997. Femtosecond photoisomerization of cis-azobenzene. Chem. Phys. Lett. 272:489–495.

    [108] Ulysse L, Cubillos J, Chmielewski J. 1995. Pho-toregulation of cyclic peptide conformation. J. Am.Chem. Soc. 117:8466–8467.

    [109] Behrendt R, Renner C, Schenk M, Wang FQ,Wachtveitl J, Oesterhelt D, Moroder L. 1999. Pho-tomodulation of the conformation of cyclic peptideswith azobenzene moieties in the peptide backbone.Angew. Chem. Int. Ed. 38:2771–2774.

    [110] Kumita JR, Smart OS, Woolley GA. 2000. Photo-control of helix content in a short peptide. Proc.Natl. Acd. Sci. USA 97:3803–3808.

    [111] Renner C, Behrendt R, Spörlein S, Wachtveitl J,Moroder L. 2000. Photomodulation of conforma-tional states. I. mono- and bicyclic peptides with(4-amino)phenylazobenzoic acid as backbone con-stituent. Biopolymers 54:489–500.

    [112] Renner C, Cramer J, Behrendt R, MoroderL. 2000. Photomodulation of conformationalstates. II. mono- and bicyclic peptides with (4-aminomethyl)phenylazobenzoic acid as backboneconstituent. Biopolymers 54:501–514.

    [113] Flint DG, Kumita JR, Smart OS, Woolley GA.2002. Using an azobenzene cross-linker to either in-

  • 13

    crease or decrease peptide helix content upon trans-to-cis photoisomerization. Chem. Biol. 9:391–397.

    [114] Woolley GA. 2005. Photocontrolling peptide alphahelices. Acc. Chem. Res. 38:486–493.

    [115] Borisenko V, Woolley GA. 2005. Reversibil-ity of conformational switching in light-sensitivepeptides. J. Photochem. Photobiol. A: Chemistry173:21–28.

    [116] Aemissegger A, Krautler V, van Gunsteren WF,Hilvert D. 2005. A photoinducible beta-hairpin. .J. Am. Chem. Soc. 127:2929–2936.

    [117] Renner C, Moroder L. 2006. Azobenzene as con-formational switch in model peptides. Chem. Bio.Chem. 7:869–878.

    [118] Jurt S, Aemissegger A, Gntert P, Zerbe O, HilvertD. 2006. A photoswitchable miniprotein basedon the sequence of avian pancreatic polypeptide.Angew. Chem. Int. Ed. 45:6297–6300.

    [119] Dong SL, Loweneck M, Schrader TE, Schreier WJ,Zinth W, Moroder L, Renner C. 2006. A pho-tocontrolled beta-hairpin peptide. Chemistry - AEuropean Journal 12:1114–1120.

    [120] Rehm S, Lenz MO, Mensch S, Schwalbe H,Wachtveitl J. 2006. Ultrafast spectroscopy of aphotoswitchable 30-amino acid de novo synthesizedpeptide. Chem. Phys. 323:28–35.

    [121] Kusebauch U, Cadamuro SA, Musiol HJ, Lenz MO,Wachtveitl J, Moroder L, Renner C. 2006. Photo-controlled folding and unfolding of a collagen triplehelix. Angew. Chem. Int. Ed. 45:7015–7018.

    [122] Spörlein S, Carstens H, Satzger H, Renner C,Behrendt R, Moroder L, Tavan P, Zinth W,Wachtveitl J. 2002. Ultrafast spectroscopy revealssubnanosecond peptide conformational dynamicsand validates molecular dynamics simulation. Proc.Natl. Acad. Sci. USA 99:7998–8002.

    [123] Satzger H, Root C, Renner C, Behrendt R, MoroderL, Wachtveitl J, Zinth W. 2004. Picosecond dynam-ics in water-soluble azobenzene-peptides. Chem.Phys. Lett. 396:191–197.

    [124] Löweneck M, Milbradt AG, Root C, Satzger H,Zinth W, Moroder L, Renner C. 2006. A confor-mational two-state peptide model system contain-ing an ultrafast but soft light switch. Biophys. J.90:2099–2108.

    [125] Bredenbeck J, Helbing J, Sieg A, Schrader T, ZinthW, Renner C, Behrendt R, Moroder L, WachtveitlJ, Hamm P. 2003. Picosecond conformational tran-sition and equilibration of a cyclic peptide. Proc.Natl. Acad. Sci. USA 100:6452–6457.

    [126] Bredenbeck J, Helbing J, Kumita JR, Woolley GA,Hamm P. 2005. α-helix formation in a photoswitch-able peptide tracked from picoseconds to microsec-onds by time resolved IR spectroscopy. Proc. Natl.Acad. Sci. USA 102:2379–2384.

    [127] Hamm P, Helbing J, Bredenbeck J. 2006. Stretchedversus compressed exponential kinetics in α-helixfolding. Chem. Phys. 323:54–65.

    [128] Ihalainen JA, Bredenbeck J, Pfister R, Helbing J,Chi L, van Stokkum IH, Woolley GA, Hamm P.2007. Folding and unfolding of a photoswitchablepeptide from picoseconds to microseconds. Proc.Natl. Acad. Sci. USA page in press.

    [129] Volk M, Kholodenko Y, Lu H. SM, Gooding EA,DeGrado WF, Hochstrasser RM. 1997. Peptideconformational dynamics and vibrational Stark ef-fects following photoinitiated disulfide cleavage. J.Phys. Chem. B 101:8607–8616.

    [130] Lu H. SM, Volk M, Kholodenko Y, Gooding E,Hochstrasser RM, DeGrado WF. 1997. Aminothio-tyrosine disulfide, an optical trigger for initiation ofprotein folding. J. Am. Chem. Soc. 119:7173–7180.

    [131] Kolano C, Gomann K, Sander W. 2004. Smallcyclic disulfide peptides: Synthesis in preparativeamounts and characterization by means of NMRand FT-IR spectroscopy. Eur. J. Org. Chem.20:4167–4176.

    [132] Kolano C, Helbing J, Bucher G, Sander W, HammP. 2007. Intramolecular disulfide bridges as a pho-totrigger to monitor the dynamics of small cyclicpeptides. J. Chem. Phys. B submitted.

    [133] Li P, Chen XG, Shulin E, Asher SA. 1997. UVresonance Raman ground and excited state studiesof amide and peptide isomerization dynamics. J.Am. Chem. Soc. 119:1116–1120.

    [134] Sifferlen T, Rueping M, Gademann K, Jaun B, See-bach D. 1999. β-Thiopeptides: Synthesis, NMR so-lution structure, CD spectra, and photochemistry.Helv. Chim. Acta 82:2067–2093.

    [135] Frank R, Jakob M, Thunecke F, Fischer G,Schutkowski M. 2000. Thioxylation as one-atom-substitution generates a photoswitchable elementwithin the peptide backbone. Angew. Chem. Int.Ed. 39:1120–1122.

    [136] Jensen OE, Lawesson SO, Bardi R, Piazzesi AM,Toniolo C. 1985. Studies on amino-acids andpeptides. 8. synthesis and crystal-structure of twomonothiated analogs of Boc-Gly-S-Ala-Aib-OMe.Tetrahedron 41:5595–5606.

    [137] Lehmann J, Linden A, Heimgartner H. 1999. Site-selective incorporation of thioamide-linkages into agrowing peptide. Tetrahedron 55:5359–5376.

    [138] Schutkowski M, Frank R, Jakob M, Thunecke F,Fischer G, Schutkowski M. 2000. Thioxylationas one-atom-substitution generates a photoswitch-able element within the peptide backbone. Angew.Chem. Int. Ed. 39:1120.

    [139] Zhao J, Wildemann D, Jakob M, Vargas C, Schiene-Fischer C. 2003. Direct photomodulation of peptidebackbone conformations. Chem. Commun. pages2810–2811.

    [140] Satzger H, Root C, Gilch P, Zinth W, WildemannD, Fischer G. 2005. Photoswitchable elementswithin a peptide backbone-ultrafast spectroscopyof thioxylated amides. J. Phys. Chem. B 109:4770–4775.

  • 14

    [141] Harada I, Tasumi M. 1980. Formation of transientcis N-methylthioacetamide under ultraviolet-laserirradiation. Chem. Phys. Lett. 70:279–282.

    [142] Kato C, Hamaguchi H, Tasumi M. 1985. Tran-sient resonance Raman-study on the trans cis pho-toisomerization of N-methylthioacetamide. J. Phys.Chem. 99:407–410.

    [143] Miwa JH, Patel AK, Vivatrat N, Popek SM, MeyerAM. 2001. Compatibility of the thioamide func-tional group with beta-sheet secondary structure:Incorporation of a thioamide linkage into a beta-hairpin peptide. Organic Letters 3:3373–3375.

    [144] Miwa JH, Pallivathucal L, Gowda S, Lee KE.2002. Conformational stability of helical peptidescontaining a thioamide linkage. Organic Letters4:4655–4657.

    [145] Helbing J, Bregy H, Bredenbeck J, Pfister R,

    Hamm P, Huber R, Wachtveitl J, Vico LD, OlivucciM. 2004. A fast photoswitch for minimally per-turbed peptides: Investigation of the trans–cis pho-toisomerization of N-methylthioacetamide. J. Am.Chem. Soc. 126:8823–8834.

    [146] Cervetto V, Bregy H, Hamm P, Helbing J.2006. Time-resolved IR spectroscopy of N-methylthioacetamide: Trans-cis isomerizationupon nπ∗ and ππ∗ excitation and cis-trans pho-toreaction. J. Phys. Chem. A 110:11473–11478.

    [147] Cervetto V, Pfister R, Kolano C, Bregy H, Heim-gartner H, Helbing J. 2007. Coexistence ofcyclic h-bonded and open tetrapeptide conforma-tion. Chem. Eur. J. in press.

    [148] Gorbunov RD Stock G. to be published.