F V10.” a J uh $1.. ‘ h... a ‘73.. K253 . I Alta)‘. 53'): y 3.1... i fis.‘....il.o€l.l Li. I. 5571:... 1.}, z. x.{r.ln!:tr :3. 3-! 52.....7 134456 .. . . tum.“ .15!!! .321...”- .. .. r . I9) t1! [Lichtsnln r . iii-nu): 335.31. . 4.2.0!szth tat.) 3.3.0.5 east: 31:1... . a.” 1! $1431.12! (2.5!!!! 1’ il‘:.£:(z A loinv .. I, . .. . . 1...... a . n: )2. (:22... .. ii“! \I.‘ 1 A a «w. 3 a}: ‘ .I x. hiatxx . 1’1. :x .1 1:13.}; LIBRARY Michigan State I Ink toroify VI II'VIUI$ This is to certify that the dissertation entitled VIBRATIONAL COHERENCE IN ZINC PORPHYRINS IN POLAR SOLUTIONS AND IN PROTEINS: VAN DER WAALS INTERACTIONS WITH SURROUNDING GROUPS IN THE FIRST SOLVATION SHELL presented by Kevin Lawrence Dillman has been accepted towards fulfillment of the requirements for the Ph.D. degree in Chemistry WWI. 73% Major Professor’s Signature / 51 Dcwném 2007 Date MSU is an Afiinnative Action/Equal Opportunity Employer no I—-—--a-o-I‘o-I—o-o-a-o-I-cnu-OOI-u- PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 5108 K:lProj/Aoc&Pres/ClRC/DateDm.indd VIBRATIONAL COHERENCE IN ZINC PORPHYRINS IN POLAR SOLUTIONS AND IN PROTEINS: VAN DER WAALS INTERACTIONS WITH SURROUNDING GROUPS IN THE FIRST SOLVATION SHELL By Kevin Lawrence Dillman A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Chemistry 2009 ABSTRACT VIBRATIONAL COHERENCE IN ZINC PORPHYRIN S IN POLAR SOLUTIONS AND IN PROTEINS: VAN DER WAALS INTERACTIONS WITH SURROUNDING GROUPS IN THE FIRST SOLVATION SHELL By Kevin Lawrence Dillman We have employed femtosecond dynamic-absorption spectroscopy to monitor the low-frequency vibrational coherence from ZnII tetralds(N-methylpyridyl)porphyrin (ZnTMPyP) in polar solutions and ZnH-substituted cytochrome c (ZnCytc). The ground-state vibrational coherence from ZnTMPyP is dominated by rapidly damped components whose intensities are perhaps an order of magnitude stronger than the more slowly damped components that arise from the skeletal normal modes of the porphyrin macrocycle. The mean frequencies of the rapidly damped features exhibit a solvent-dependent shift. The shift is consistent with a van der Waals potential dom- inated by London-dispersion interactions. In the excited-state vibrational coherence, these components exhibit an increase in frequency that suggests the potential is al- tered by the presence of charges. These components are assigned to intermolecular vibrational modes between the porphyrin and clustered, first-shell solvent molecules. The vibrational coherence from ZnCytc in the native state is exhibits similar behav- ior. Low-frequency metal-doming modes contribute ~10 and ~30 cm‘l oscillatory features, but the strongest component in the vibrational coherence is a very rapidly damped (damping time y<150 fs) contribution centered at 79 cm‘l. This feature is assigned to van der Waals interactions between the porphyrin chromophore and nearby nonpolar amino-acid residues in the surrounding protein medium. The vibra- tional coherence from the molten-globule state of ZnCytc does not exhibit the rapidly damped feature. Its resonance Raman activity is effectively quenched owing to ex- pansion of the hydrophobic core and randomization of the surrounding structure. The results strongly suggest that the van der Waals modes gain resonance Raman ac- tivity through a direct attack of the solvent or neighboring groups on the Tr-electron density of the porphyrin chromophore. These findings provide the first structural ac- counting for the rate-controlling, low-frequency vibrational motions that are coupled to the electron-transfer reactions in photosynthesis. ACKNOWLEDGMENTS The research presented in this dissertation could not have been performed with- out the support of many other individuals. First and foremost is that of my advisor, Professor Warren Beck. I thank him for his invaluable guidance and his mentorship. Without him and the effort he has put into teaching me, this work would have been impossible. I thank my committee members, Professor Marcos Dantus, Professor James McCusker and Professor Lynmarie Posey for their time and input. I thank Dr. Kathrine Shelly and Dr. Sanela Lampa-Pastirk for their patience and instruction. They taught me almost everything I know about how to perform fem- tosecond spectroscopy. I also thank my fellow graduate student Jagnyaseni Tripathi. Her curiosity and questions have led me to a better understanding of my own work. I have been fortunate to work with several talented undergraduate students, and I particularly wish to acknowledge Kenneth Barnes, Aaron Dame and Steven Miller. I must also thank my entire family, who have provided me with endless encour- agement. Specifically, I must thank my parents, Tom and Ellen Dillman, who have always been there to support me, and my brother, Peter, for his excellent friendship. I thank the Department of Chemistry at Michigan State University for a Teach- ing Assistantship (2005-2009). This research was supported by the National Sci- ence Foundation Biomolecular Systems Cluster/Molecular Biophysics program un- der grants MCB-009120 and MCB-OSZOOOZ. Additional support for instrumentation was provided by the Michigan Structural Biology Center at Michigan State University, which is supported by the Michigan Life Sciences Corridor. iv TABLE OF CONTENTS LIST OF TABLES vii LIST OF FIGURES viii LIST OF ABBREVIATIONS xii 1 Introduction 1 2 Vibrational Coherence and Photosynthetic Electron Transfer 4 2.1 Electron Transfer in Photosynthesis ....................... 5 2.2 Vibrational Coherence and Wavepackets .................... 9 2.3 Dynamic-Absorption Spectroscopy ........................ 1 1 2.4 Vibrational Coherence in the Purple-Bacterial Photosynthetic Reaction Center ......................................... 14 2.5 Vibrational Coherence from Myoglobin and Cytochrome c ......... 19 2.6 Vibrafl'onal Coherence in Bacteriochlorophyll Solutions and Proteins . . 25 2.6.1 Vibrational Coherence in Bacteriochlorophyll a in Polar solution . 25 2.6.2 Intermolecular Vibrational Coherence From B777 and B820 . . . . 30 2.7 Proposed Experiments ............................... 32 3 Ground-state vibrational coherence in polar solutions of Zn“ tetrakis(N- methylpyridyl)porphyrin with Soret-band excitation 38 3.1 Summary ........................................ 38 3.2 Introduction ...................................... 39 3.3 Experimental ..................................... 44 3.3.1 Sample Preparation ............................. 44 3.3.2 Continuous-Wave Absorption and Fluorescence Spectroscopy . . 44 3.3.3 Femtosecond Spectroscopy ........................ 45 3.4 Results ......................................... 49 3.5 Discussion ....................................... 68 3.5.1 Conclusions .................................. 81 4 Excited-state vibrational coherence in methanol solution of Zn“ tetrakis(N- methylpyridyl)porphyrin 83 4.1 Summary ........................................ 83 4.2 Introduction ...................................... 84 4.3 Experimental ..................................... 87 4.3.1 Sample Preparation ............................. 87 4.3.2 Continuous-Wave Absorption and Fluorescence Spectroscopy . . 87 4.3.3 Femtosecond spectroscopy ........................ 88 4.4 Results ......................................... 89 4.5 Discussion ....................................... 97 4.5.1 Conclusions .................................. 101 5 Vibrational coherence in the native and molten-globule states of Zn“- substituted cytochrome c 105 5.1 Summary ........................................ 105 5.2 Introduction ...................................... 106 5.3 Experimental ..................................... 108 5.3.1 Sample Preparation ............................. 108 5.3.2 Continuous-Wave Absorption and Fluorescence spectroscopy . . 1 1 1 5.3.3 Femtosecond Spectroscopy ........................ 1 1 1 5.4 Results ......................................... 1 12 5.5 Discussion ....................................... 1 1 7 5.5.1 Conclusions .................................. 129 BIBLIOGRAPHY 131 3.1 3.2 3.3 3.4 3.5 4.1 4.2 4.3 5.1 5.2 LIST OF TABLES Model parameters for the exponential decays in the pump—probe tran- sients in ZnTMPyP solutions at 22 °C. ..................... Solvent dependence of the asymmetric Gaussian lineshape parameters for the rapidly damped components observed in the vibrational coher- ence from ZnTMPyP. ................................ Gaussian lineshape parameters for the slowly damped modulation com- ponents observed in ZnTMPyP pump—probe transients. .......... Comparison of Gaussian linewidths and effective exponential damping times for the modulation components observed in the vibrational co- herence from ZnTMPyP in CH3 OH. ....................... Parameters for the model ZnTMPyP-CH3OH intermolecular potentials shown in Figure 3.12 ................................ Gaussian lineshape parameters for the slowly damped modulation components obtained from the 62 5-nm pump—probe transient from ZnTMPyP in methanol (see Figure 4.2). ..................... Gaussian lineshape parameters and effective exponential damping times for the rapidly damped modulation components obtained from the 625-nm pump—probe transient of ZnTMPyP in methanol (see Fig- ure 4.2). ........................................ Parameters for ZnTMPyP—CH3OH intermolecular potentials (see Fig- ure 4.4) ........................................ Gaussian lineshape parameters for the modulation components ob- tained from the pump-probe transient of ZnCytc in the native state (see Figure 5.3). ...................................... Gaussian lineshape parameters for the modulation components ob- tained from the pump—probe transient of ZnCytc in the molten-globule state (see Figure 5.4). ................................ 77 94 120 2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 2.10 2.11 2.12 2.13 2.14 LIST OF FIGURES Temperature dependence of the primary electron transfer rate constant in the reaction center of Blastochloris viridis. ................. 6 Driving force A130 and reorganization energy An for a model pair of non-adiabatic potential energy surfaces, Ea and E b .............. 8 Excited-state and ground-state coherent wavepacket motion in the dynamic-absorption experiment .......................... 10 Femtosecond dynamic-absorption transients of bacteriorhodopsin. . . . 12 Fourier-transform spectra and model calculations of the oscillatory part of the pump-probe transients obtained from bacteriorhodopsin ...... 13 Dynamic-absorption transient from the reaction centers of Rhodobacter sphaeroides R-26 at 100 K. ............................. 1 5 Dynamic-absorption transients at various temperatures of reaction cen— ters of Rhodobacter sphaeroides R-26 at 100 K, 10 K, and 293 K. ..... 16 Fourier-transform spectra of the oscillatory parts of the pump-probe signals of the dynamic absorption transients of Rhodobacter sphaeroides. 18 Open band femtosecond dynamic-absorption spectroscopy measure- ments on MbNO samples. ............................. 22 Sequence of dynamic-absorption transients using the open band detec- tion scheme on FeCytc ................................ 23 Femtosecond dynamic-absorption transients of deoxyMb. ......... 24 Dynamic-absorption transient from BChl a in pyridine ............ 26 Model of the rapidly damped vibrational coherence observed in BChl or solutions as a distribution of damped cosinusoids, £(w). ......... 27 Solvent dependence of the mean frequency (w) of the rapidly damped vibrational coherence observed in BChl a solutions in a range of polar solvents on the solvent dipole moment. .................... 28 f . 2.15 2.16 2.17 2.18 3.1 3.2 3.3 3.4 3.5 3.6 3.7 The Hu and Schulten model for the B820 subunit in LH1 from Rhodobac- ter sphaeroides, showing the pair of BChl macrocycles and the trans- membrane or helices in ribbon and surface renderings. ........... 31 Expanded view of the rapidly damped oscillation observed in the dynamic-absorption transient from BChl a in acetone. ........... 33 Expanded view of the rapidly damped oscillation observed in the dynamic-absorption transient from B777. ................... 34 Expanded View of the rapidly damped oscillation observed in the dynamic-absorption transient from B820. ................... 3 5 Optimized structure of ZnII meso—tetrakis(N-methylpyridyl)porphyrin (ZnTMPyP) complexed with a single methanol molecule as an axial lig- and to the Zn11 ion. ................................. 43 Femtosecond pump-probe spectrometer, consisting of a prism-pair pulse compressor (PC), second-harmonic generator (SHG), and a mod- ified Mach-Zehnder interferometer. ....................... 46 Soret (v > 20000 cm‘l) and Q-band region of the continuous-wave absorption (solid curve) and fluorescence (dotted curve) spectra from ZnTMPyP in CH3OH at room temperature (23 °C), plotted as the dipole strength, A/v and F/v3, respectively, and normalized to unit area with respect to the fluorescence spectrum and Q band ............... 50 Soret-band absorption absorption dipole-strength (A / v) spectra from ZnTMPyP in (a) CH30H, (b) DMF, (c) CH3 CN, and (d) DMSO. ........ 5 1 Femtosecond pump—probe dynamic-absorption transients detected at 422 nm (0.5-nm bandpass) with ZnTMPyP solutions in (a) CH3OH, (b) CD30D, (c) DMF, (d) CH3CN, and (e) DMSO. .................. 53 Oscillatory signals obtained from the pump—probe transients from ZnTMPyP solutions (see Figure 3.5) as the difference between the nor- malized signal and the fitted exponential decay functions: in (a) CH30H, (b) CD30D, (c) DMF, (d) CH3CN, and (e) DMSO. ................ 55 Fourier transform magnitude spectra from the oscillatory residuals (see Figure 3.6) of the pump-probe signals from ZnTMPyP solutions in (a) CH30H, (b) CD30D, (c) DMF, (d) CH3CN, and (e) DMSO. ........... 59 3.8 3.9 3.10 3.11 3.12 4.1 4.2 4.3 4.4 5.1 Comparison of the Gaussian waveforms (thick curves) obtained for the 3, 38, and 215-cm‘1 components in the oscillatory signals observed from ZnTMPyP in CH30H solution with fitted exponentially damped waveforms (thin curves, see Equation 3.4) .................... 64 Frequency domain representation of the time-domain models of the os- cillatory signals (see Figure 3.6) observed in ZnTMPyP solution: in (a) CH30H, (b) CD30D, (c) DMF, (d) CH3CN, and (e) DMSO. ........... 65 Time-domain representation of the rapidly damped oscillatory compo- nents (see Figures 3.6 and 3.9) in the pump-probe signals from ZnTMPyP solutions in (a) CH30H, (b) CD30D, (c) DMF, (d) CH3CN, and (e) DMSO. . 66 Dependence of the mean frequency of the rapidly damped oscillatory components (see Figures 3.6 and 3.9) observed in the pump-probe sig- nals from ZnTMPyP solutions in CH30H, DMF, CH3 CN, and DMSO. . . . . 69 Model potential-energy curves for the ZnTMPyP-CH3OH complex calcu- lated using Equation 3.1 1 and the experimentally observed mean inter- molecular mode frequency, 79 cm‘1 (see Table 3.2): (a) Q = 0; (b) Q = 1; and (c) the attractive, charge-dependent terms for Q = 1. ......... 76 Soret (v>20000 cm‘l) and Q-band region of the continuous-wave ab- sorption (solid curve) and fluorescence (dotted curve) spectra from ZnTMPyP in methanol at room temperature (23 °C), plotted as the dipole strength, A(v)/v and F(v)/v3, respectively, and normalized to unit area for the fluorescence spectrum and Q band. ............... 90 Femtosecond pump—probe dynamic absorption transient detected at 632 nm (4-nm bandpass) with ZnTMPyP in methanol. ............ 91 Comparison of the intensities and lineshapes observed in the excited- state and ground-state vibrational coherence from ZnTMPyP in methanol. 93 Model potential-energy curves for the ZnTMPyP-CH30H van der Waals complex as a function of the charge Q on the ZnTMPyP moiety: (a) Q = 0; (b) Q = 1; (c) the attractive, charge-dependent terms for Q = 1; ((1) Q = 2 and (e) the attractive, charge-dependent terms for Q = 2. . . . 102 Ribbon (left) and surface (right) renderings of the X-ray crystal structure of horse-heart ferricytochrome c (1hrc.pdb) ................... 110 5.2 5.3 5.4 5.5 5.6 Soret (v> 23000 cm’l) and Q-band regions of the continuous-wave ab- sorption spectrum of ZnCytc in the native (solid curves) and molten- globule (circles) states. ............................... 1 13 Soret-band (420-nm) pump—probe transient from the native state of Zn- Cytc. .......................................... 1 1 5 Soret-band (420-nm) pump—probe transient from the molten-globule state of ZnCytc. The signal points (circles) are superimposed on a model consisting of the sum of an exponential decay and two oscillatory com— ponents. ........................................ 1 16 Frequency-domain representations of oscillatory components of the fit- ted models of the vibrational coherence obtained from the (a) native state and (b) molten globule state of ZnCytc. ................. 118 Detail of the X-ray crystal structure of ferricytochrome c (1hrc.pdb) ex- amining possible interactions of the protein with the n-electron density of the porphyrin .................................... 126 LIST OF ABBREVIATIONS B777, monomeric subunit protein of B820 B820, dimeric subunit protein of the LH1 light-harvesting complex BChl, bacteriochlorophyll BPheL, primary electron acceptor in the photosynthetic reaction center CD3 OD, perdeuturated methanol CH3 OH, methanol CH3 CN, acetonitrile DMF, N,N-dimethylformamide DMSO, dimethylsulfoxide ESA, excited-state absorption FeCytc, FeIII cytochrome c fwhm, full width at half maximum K, lysene Mb, myoglobin N, asparagine P, primary electron donor in the photosynthetic reaction center PB, photobleaching Qa, primary quinone electron acceptor in the photosynthetic ter reaction cen- Qb, secondary quinone electron acceptor in the photosynthetic reaction cen- ter RISRS, resonant impulsive stimulated Raman scattering SE, stimulated emission T, threonine W, tryptophan Y, tyrosene ZnCytc, ZnII cytochrome c ZnTMPyP, ZnII mesatetrakis(N-methylpyridyl)porphyrin ZnT'PP, ZnII tetraphenylporphyrin CHAPTER 1 Introduction The electron-transfer reactions in the purple-bacterial photosynthetic reaction center are nearly barrierless.l Low-frequency vibrational modes that couple to the reaction account for the majority of the reorganization energy that matches the driving force in the Marcus theory for condensed-phase electron-transfer ratesz‘4 and contribute to the barrierless reaction kinetics. The structural origin of these modes has not been previously determined. This dissertation provides the first detailed structural accounting for the vibrational motions that are coupled to the electron-transfer re- actions in photosynthesis. The findings suggest that the relevant vibrational modes are intermolecular in nature. They arise from van der Waals interactions between the chromophore and nearby amino-acid residues in the surrounding protein medium. This dissertation is organized as follows: Chapter 2 introduces some of the previous contributions that have led to the prin- cipal work covered in this dissertation. We briefly cover the nature of the electron- transfer reactions in the photosynthetic reaction center, and we discuss the nature of the dynamic-absorption experiment. We examine the results of some of the early vibrational coherence measurements on the reaction center and the results of simi- lar experiments on myoglobin and cytochrome c. We finish by discussing previous work from this laboratory. Femtosecond dynamic-absorption spectroscopy on bacte- riochlorophyll (BChl) in polar solution, and on the light—harvesting subunit proteins B777 and 8820 reveal that the low-frequency vibrational coherence is dominated by rapidly damped components that are assigned to intermolecular vibrational modes. These modes exhibit a dependence on the dipole moment of the solvent. These re- sults inspire the experiments discussed in the subsequent chapters. In Chapter 3, we examine the origin of the rapidly damped vibrational coherence from ZnII tetrakis(N-methylpyridyl)porphyrin (ZnTMPyP) with Soret-band excitation. ZnTMPyP has a high degree of symmetry, so it has a much smaller dipole moment than BChl. We discuss the effect this should have on the rapidly damped mode. We discuss how the dependence of the mean frequency of the intermolecular mode on the solvent dipole moment is consistent with a van der Waals intermolecular poten- tial with the strongest terms coming from London-dispersion interactions. We show evidence of a isotope-dependent shift in the frequency of the intermolecular mode from methanol and perdeuturated methanol. We conclude that the rapidly damped vibrational modes arise from intermolecular modes with clustered, first—shell solvent molecules. Chapter 4 discusses the intermolecular vibrational coherence from ZnTMPyP with Q-band excitation. The vibrational coherence observed in these experiments arises from wavepacket motion on the $1 excited state. We observe that the intermolecular mode is shifted to substantially higher frequencies than previously obtained from ground-state vibrational coherence of ZnTMPyP. In the excited state of ZnTMPyP, the n-electron density extends over the N-methylpyridyl rings, so charge terms are added to the van der Waals intermolecular potential. We discuss how these terms dominate the potential and change its character relative to the neutral potential. We conclude with a brief discussion of how the formation of charged species in electron- transfer reactions in the photosynthetic reaction center can trap the charged species and contribute to their high quantum efficiency. Chapter 5 concludes this dissertation with a discussion of the van der Waals inter- action in ZnH-substituted cytochrome c (ZnCytc). We compare the vibrational coher- ence from the native and molten-globule states of ZnCytc In a protein environment, the van der Waals mode arises from interactions with nearby amino-acid residues. We examine how the potential is dominated by London-dispersion interactions, and how the absence of nearby polar or charged residues in the native state leads to inho- mogeneous linebroadening of the van der Waals mode. In the molten-globule state, the protein core expands and the interactions are randomized. These factors effec- tively quench the van der Waals mode in the molten globule. We discuss how this result suggests that the van der Waals modes gain resonance Raman activity through direct attack of the nearby groups on the Tr-electron density of the chromophore. We conclude that the addition of charged or polar groups could tune the frequency of the van der Waals mode, and that similar interactions are likely candidates for the rate-controlling interactions in the photosynthetic reaction center. CHAPTER 2 Vibrational Coherence and Photosynthetic Electron Transfer Over the last 25 years, there has been an effort to design light-harvesting and electron-transfer molecular donor-acceptor systems for use in solar energy conver- sion applications.5'17 A central design principle in this work has been to mimic various aspects of the structure and function of natural photosynthetic reaction cen- ters, such as that from the purple non-sulfur bacteria.8v10 The quantum efficiency of man-made systems falls far short of the essentially unity efficiency of the natural reaction center.18’ 19 It is quite likely that what has not been mimicked so far is the structure and dynamics of the surrounding protein medium; This dissertation describes an experimental study of how solvent structures around porphyrins in polar solutions and in a small redox protein are coupled to electronic transitions. The experiments employ a time-domain version of resonance Raman spectroscopy, dynamic-absorption spectroscopy. These studies provide the first detailed structural accounting for the vibrational motions that are coupled to the electron-transfer reactions in photosynthesis. In order to provide an introduc- tion to these questions, the structure of electron transfer in photosynthesis is re- Viewed. Vibrational coherence and wavepacket generation are then discussed, and the dynamic-absorption pump-probe experiment is introduced. Observations of vi- brational coherence first in the photosynthetic reaction center and then in myoglobin and cytochrome c will be reviewed. This chapter concludes with a review of the vibrational coherence observed from bacteriochlorophyll in polar solutions and in proteins. 2.1 Electron Transfer in Photosynthesis The electron-transfer rate for the primary and secondary charge-separation reactions in the purple-bacterial photosynthetic reaction center is almost completely indepen- dent of temperature},20 Figure 2.1 shows the rate constant as a function of tem- perature for the primary electron-transfer rate in the reaction center of Blastochlo- ris viridis. The figure indicates that as the temperature increases, the reaction rate decreases: at room temperature, the primary electron transfer event takes approx- imately 3 ps, while at 10 K, the reaction time decreases to ~700 fs. This behavior defies the typical temperature dependence expected from the Arrhenius equation, k = Ae‘Ea/RT, where the rate k decreases as the temperature T increases.21 This fact suggests that the electron-transfer reaction in the photosynthetic reaction center is nearly barrierless, so Ea is approximately 0. This both accounts for the rate behav- ior and for the long lifetimes of the non-neutral products: if the barrier is zero for the charge-separation reaction, then provided the reaction is enthalpically favored, the barrier for the charge-recombination reaction is high. Bixon and Jortner exam- ined the case where vibrational modes of the solvent coupled to the electron transfer reaction.1'22v23 In particular, for the case where the reaction is truly barrierless, or Ea = 0, the reaction rate k can be expressed as1 2111/2 (ehw/kBT— 1)1/2 = 2. h2w(2p)1/2 ehw/kBT+1 ( 1) 5 1(98) k (1011 sec") I J __ I o 100 200 300 T (K) Figure 2.1. Temperature dependence of the primary electron transfer rate constant in the reaction center of Blastochloris viridis. (Adapted from reference 1.) where V is the electronic coupling, 00 is the vibrational frequency, h is Planck’s con- stant divided by 211‘ and k3 is Boltzmann’s constant. From the Marcus theory for condensed-phase electron-transfer rates?“4 the ac- tivation energy for the electron-transfer reaction depends on both the driving force, the Gibbs free energy difference between the reactant and product AGO, and the re- organization energy A, the energy required by the reactant to assume the geometry of the product state. Under the Born-Oppenheimer approximation,21 the reactant must reorganize before the electron transfer can occur. Figure 2.2 shows the re- organization energy for a model pair of non-adiabatic potential energy surfaces. A spontaneous reaction (for which AGO is negative) has a positive driving force. The activation energy is given by Ed = hw(p — 5)2/45 (2.2) Here, p = AGO/ha), and S = A/hw. Since the rate constant is inversely proportional to the activation energy, it is clear from Equation 2.2 that for a given spontaneous reaction with a fixed AGO<0, the reaction rate will increase as the reorganization energy It increases between 0 and AGO. This is the so-called ‘normal regime’ in the Marcus theory.4 As the reorganization increases beyond the driving force, the ‘in- verted regime’, the reaction rate decreases. However, if the reorganization energy exactly matches the driving force, then the (p — S)2 term in Equation 2.2 will become zero, so the reaction barrier will vanish. Bixon and Jortner observed that the relevant vibrational modes coupled to the electron transfer reaction that provided the bulk of the reorganization energy needed to match the driving force were in the ~ 100 cm‘1 regimeLZO'22 The structural origin of these low-frequency modes, however, was indeterminate. The next section discusses how femtosecond spectroscopy can be used to moni- tor wavepacket motion on a potential energy surface. It also reviews the phenomenon Energy -.r--_ i. I'. I g | o l : I I l l I I I I I I I I I I :<—>: D '0 Qn,a Qn b Vibrational Displacement Figure 2.2. Driving force A150 and reorganization energy An for a model pair of non- adiabatic potential energy surfaces, Ea and Eb. The driving force corresponds to the Gibbs free energy difference between the minimum of the potential wells between the reactant (Qma) and product (@133) states. The reorganization energy is the difference between the minimum of the the potential well for the reactant and the energy of the reactant surface at the geometry of the product minimum. of vibrational coherence. These two concepts form the basis of the femtosecond dynamic-absorption experiment, which is discussed later. 2.2 Vibrational Coherence and Wavepackets Vibrational coherence arises from coherent wavepacket oscillation on a vibronic po- tential energy surface.24'26 When a coherent pulse of light that is resonant with a vibronic transition hits a sample, it launches a wavepacket provided that the pulse is impulsive (i.e., the duration of the pulse is shorter than the period of the vibrational mode).27 The first and second actions of the electric field of the pump with the mat- ter in the sample prepare a population in either the ground state or some resonant excited state.25'26 Figure 2.3 shows the formation and motion of a wavepacket in the dynamic—absorption experiment. The excited-state potential energy surface le) is displaced from the ground-state surface lg) in Figure 2.3. Because the two sur- faces are displaced along the vibronic reaction coordinate, the potential is sloped in the Franck-Condon region. The slope of the potential surface generates a force on the wavepacket. This imparts momentum to the wavepacket, which will begin to move away from the Franck-Condom regionZS-zgr29 on either the ground or excited state. For the case of a ground-state wavepacket, this mechanism is analogous to resonance-Raman scattering”, 25- ”‘33 Following its formation, the wavepacket oscillates in the potential well. For a ground-state wavepacket, the duration of the pump pulse controls the displacement and momentum of the wavepacket. For a particular vibrational mode, the optimum pulse duration is approximately one-third of the mode’s period. If the pulse is sig- nificantly shorter than the optimum duration, then the displacement away from the Franck-Condon region is small, and the experimentally detectable modulation by the wavepacket of a probe pulse is weak. If the pump pulse is too long, then it is no Energy 0 ya 0 O O I I rg re Vibrational Displacement Figure 2.3. Excited-state and ground-state coherent wavepacket motion in the dynamic-absorption experiment. The excited-state (Ie)) and ground-state (I g)) potential-energy surfaces are drawn as parabolas that are displaced with respect to a generalized multimode coordinate; 1'9 and re mark the equilibrium ground-state and excited-state geometries, respectively. Thick arrows represent the resonant pump- laser field; thin arrows show the direction that the wavepackets evolve during the first passage on the two surfaces. The numbers indicate event times, starting with the ground-state probability density (t = 0), creation of the excited-state wavepacket by the pump field (t = 1), creation of the ground-state wavepacket by the pump field (t = 2), and evolution during the first vibration (t = 3). The turning-point regions, where a wavepacket narrows and reverses its direction of motion, are marked on both surfaces with diamonds; white (unfilled) diamonds mark the turning points that conuibute to the interference pattern in the dynamic absorption spectrum. (From reference 29.) 10 longer impulsive, so this too decreases the modulation intensity.33 For an excited- state wavepacket, the intensity patterns are different. As long as the excited state is displaced from the ground state, the excited-state wavepacket will gain momentum if the pulse is impulsive. The strongest modulations will be those modes with the low- est frequency. The momentum of the wavepacket also trivially depends on the dis- placement of the excited state relative to the ground state. Not all modes are equally displaced; modes with a larger displacement will impart a greater momentum to the wavepacket than modes with a smaller displacement. The experimentally detectable modulation of a probe pulse will be greater for modes with larger displacements. The next section discusses the femtosecond dynamic-absorption pump—probe exper- iment, which can be used to monitor wavepacket motion and can generate informa- tion about the vibrational modes. 2.3 Dynamic-Absorption Spectroscopy Vibrational coherence can be detected using the dynamic-absorption tech- nique.?-6’32-35‘37 As the wavepacket oscillates back and forth in the potential well, changes in the transmitted intensity of a variably delayed probe pulse can be mea- sured. As the wavepacket oscillates, it is only resonant with the probe pulse at a particular geometry of the vibronic displacement along the reaction coordinate. As the wavepacket moves on- and off-resonance with the probe bandwidth, the inten- sity of the transmitted probe pulse is modulated with the probe delay. When the wavepacket is on-resonance, the absorption of the sample increases, and the intensity of the transmitted probe beam decreases. When the wavepacket is off-resonance, the transmitted intensity increases. The dynamic-absorption pump—probe experiment monitors the intensity of the transmitted probe beam as a function of the variable probe delay. The resonant transitions generate an oscillatory signal in the time do- 11 0.02 0.01 0 -0.01 '0'02-02 Tojode i i ' 1Z4 ' 1.8 i 2.2 2.6 Ielay(ps) - -_‘__. -._- _._“ ‘ ‘ A - w j'u v'v Figure 2.4. Femtosecond dynamic-absorption transients of bacteriorhodopsin. The laser spectrum was monitored at (a) 568 nm, (b) 620 nm, and (c) 656 nm probe wavelengths. (Adapted from reference 32.) 12 MEL. 800A 1000 1200 1400 1600 Figure 2.5. Fourier-transform spectra and model calculations of the oscillatory part of the pump-probe transients obtained from bacteriorhodopsin (see Figure 2.4). Panel a shows the spectrum obtained with a 568 nm probe wavelength and the 580 nm calculated spectrum. Panel b shows experimental and spectra at 620 nm. Panel c shows 656 nm experimental and 650 nm calculated spectra. The calculations used to generate the models are detailed in references 32, 33 and 34. (Adapted from reference 32.) 13 main that can be analyzed in the frequency domain to produce peaks corresponding to the coupled vibrational modes. Early use of this technique included characterization of torsional motions in retinal that are associated with formation of the photoisomerization prod- uct state in bacteriorhodopsin and rhodopsin by Shank, Mathies and cowork- ers.31'33' 36, 38“” Figure 2.4 shows the dynamic-absorption transients obtained from bacteriorhodopsin using 12 fs excitation pulses by Shank and co-workers.32 Follow- ing excitation by the pump-pulse, each transient exhibits a pattern of oscillations over at least the first picosecond. These oscillations correspond to the motions of the wavepacket described above. Fourier-transform spectra corresponding to the dynamic-absorption transients are shown in Figure 2.5. These spectra show peaks at 1010, 1160, 1200 and 1525 cm‘l. Below each experimental spectrum in the fig- ure is a calculated model spectrum. The models exhibit good agreement with the experimental spectra. The peaks correspond to vibrational modes at those frequen- cies. The results are significant because they reveal the particular vibrational modes that are coupled to vibronic transition. This methodology is equally applicable to the vibrational modes coupled to the electron transfer reaction in the photosynthetic reaction center. 2.4 Vibrational Coherence in the Purple-Bacterial Photosynthetic Reaction Center Vos, Martin and co-workers observed vibrational coherence in the photosynthetic re- action center of Rhodobacter sphaeroides R-26 following impulsive excitation of the primary electron donor, P. The vibrational coherence in the stimulated emission signal following excitation of the primary electron donor is primarily modulated pri- marily by low-frequency oscillations. The damping time of the vibrational coherence 14 oscillatory part (x3) 0.05 - 100 K probe 796 nm 0 1 , 2 3 Time (ps) Figure 2.6. Dynamic-absorption transient from the reaction centers of Rhodobacter sphaeroides R-26 at 100 K. The oscillatory part of the signal is shown magnified x3 above the transient. (From reference 43.) 15 probe 930 nm «3.232333%? 0.1 10 K W02. ............ ‘ 293 K l l I 1 l a I l l 0 2 4 6 8 Time (ps) Figure 2.7. Dynamic-absorption transients at various temperatures of reaction cen- ters of Rhodobacter sphaeroides R-26 at 100 K, 10 K, and 293 K. The oscillatory portions of the signals are shown magnified x3 (normalized to the maximum signal at 100 K). (From reference 43.) 16 (<1 ps at 10 K) was on the same time scale as the primary electron transfer reaction. Moreover, the frequencies of these modes were in the 30-100 cm‘1 regime.“‘51 A sample dynamic absorption transient from their early experiments43 is presented in Figure 2.6. At 100 K, the strongest contribution to the oscillatory portion of the signal is a low-frequency mode with a period greater than 500 fs. At lower tem- peratures, the low-frequency components provide greater contributions to the over- all vibrational coherence signal than at room temperature, as shown in Figure 2.7. The low-frequency components are perhaps an order of magnitude larger than any higher-frequency contributions. The low-frequency components are damped almost completely within the first two picoseconds. Vos, Martin and co-workers assigned these low-frequency vibrations to global modes of the protein medium“!52 These low-frequency modes are likely candidates for the modes that account for the major- ity of the reorganization energy in the charge-separation reactions in photosynthesis. Subsequent experiments revealed strong components between 80 and 1 50 cm’l, in addition to low frequency (15-80 cm’l) modes.47 The Fourier-transform spec- trum from these experiments is shown in Figure 2.8 The top panel shows two strong features at 84 and 145 cm“1, and a slightly weaker one at 192 cm'l. The bottom panel shows a strong 30 cm‘1 mode. The linewidths of these components are rela- tively broad, perhaps 50 cm"1 full-width at half—maximum (fwhm) for the 30-cm‘1 and 145-cm"1 peaks. The low-frequency features were once again assigned to mo- tions of the protein“,48 In particular, it was suggested that these low-frequency ~30 cm'1 modes were phonon-like in character, involving collective motions of the entire protein.48'SO However, the frequency of these modes are much lower than the ~100 cm‘1 modes that Bixon and Jortner suggested were coupled to the primary electron-transfer reaction},22 However, Figure 2.8A also suggests the possible pres- ence of a very broad (>100 cm"1 fwhm) component centered near 125 cm‘l. This feature was not discussed by the authors. The nature of the broad lineshape com- 17 145 1L ........ I 0 100 200 300 400 500 30 B 0 100 200 300 400 500 Frequency (cm'1) Figure 2.8. Fourier-transform spectra of the oscillatory parts of the pump—probe signals of the dynamic absorption transients of Rhodobacter sphaeroides. The pump— probe signals were obtained with pump pulses of 30 fs (A, solid curve) and 100 fs (B). For comparison, the same analysis for the data at 10 K (30-fs pump pulses) is shown (A, curve). (Adapted from reference 47.) 18 ponents in ZnII porphyrins in polar solutions and in proteins is discussed heavily in Chapters 3-5. If the modes are not global motions of the protein, as suggested by Vos, Martin and co-workers, then the structural origin of the coupled modes is still indeterminate. 2 . 5 Vibrational Coherence from Myoglobin and Cytochrome c Although they do not directly address the questions raised in the previous section, the pump-probe vibrational coherence experiments performed by Champion and co- workers are discussed here. Their experiments probe the vibrational coherence of the proteins cytochrome c (Cytc) and myoglobin (Mb). Both of these proteins con- tain a heme group,53 similarly to bacteriochlorophyll, although the proteins stud- ied by Champion typically contained iron as the metal ion bound to the porphyrin rather than magnesium. The ZnH-substituted form of cytochrome c (ZnCytc), how- ever shares some additional similarities to the magnesium-containing heme in bac- teriochlorophyll. We examine the the vibrational coherence from ZnCytc in Chap- ter 5, and it grants crucial insight into the nature of the modes that are relevant to the electron-transfer reactions. The instrumental setup used by Champion and co- workers implements some novel design elements. We take advantage of similar tech- niques in our experiments discussed in Chapters 3-5. The results found by Cham- pion and co-workers that are reviewed in this section revealed several low-frequency vibrational modes of the heme that provide a basis to which we later compare the results of the ZnCytc experiments in Chapter 5 These results are also significant be- cause they demonstrate that skeletal normal modes of the porphyrin can be observed in the vibrational coherence. 19 The experiments by Champion and co-workers were performed using a Tizsapphire laser to generate 45-120 fs pulses between 700 and 960 nm.53‘S7 These pulses were frequency-doubled by a BBQ crystal to generate blue pulses that could probe the Soret absorption band of the samples. A particularly novel approach used in these experiments was the use of spatially selective detection techniques that al- lowed a narrow region of the probe bandwidth to be analyzed.“56 This method could be used to selectively enhance the intensity of components in the vibrational coherence signals over certain ranges of frequencies compared to experiments where the entire probe bandwidth was integrated.“56 The narrow probe bandwidth en- hances the modulation depth of the transmitted probe, which increases the signal quality and also enhances the detected intensity of certain modes. Data analysis was performed using linear predictive singular value decomposition (LPSVD) meth- odssg‘60 to analyze the resulting pump—probe transients.S3 The LPSVD algorithm was used to generate frequency and damping time information for the various com- ponents from the overall vibrational coherence signal. The most prominent vibrational modes consistently observed in the dynamic ab- sorption signals, such as those shown in Figures 2.9 and 2.10, were near 40 and 80 cm'1.54'56' 57,51162 Figure 2.9 shows the dynamic-absorption transients and cor- responding LPSVD spectra from MbNO. The most intense components in the spectra are the low-frequency oscillations. They are up to an order of magnitude stronger than any of the high-frequency components. Figure 2.10 shows the transients and LPSVD spectra from ferrous cytochrome c (FeCytc). Similarly to MbNO, the most in- tense components are the from the low-frequency ~40-cm’1 and ~80-cm‘1 modes. The observed frequencies are somewhat higher in FeCytc than in MbNO. The 40-cm‘1 mode was assigned to doming motions of the central metal ion in the heme, while the 80-cm'l mode was assigned to an overtone of the doming mode.57 A relatively high-frequency mode near 220 cm'1 was also reported. This mode was assigned to a 20 metal-doming mode involving the metal ion and one of its axial ligands.57 In partic- ular, for the case of FeCytc, this mode was assigned to the stretch between the metal ion and its axial His-18 ligand.S4 While typically weaker than the doming modes, this mode was consistently observed in both myoglobin and cytochrome c.5415“;1 These results are important for comparison to our later experiments on ZnII meso- tetrakis(N-methylpyridyl)porphyrin (ZnTMPyP) in Chapters 3 and 4 and also in par- ticular to our experiments on ZnCytc in Chapter 5. While the results reported by Champion and co-workers give very good results for the slowly damped features in the vibrational coherence, they do not address any rapidly damped features. The focus of the experiments was on the origin of the low-frequency components near 40 cm‘l, but additional features in the vibrational coherence can be observed. In particular, a direct comparison of a Fourier-transform spectrum and LPSVD fit of a pump-probe transient obtained from deoxymyoglobin reveals a broad component near 300 cm‘1 in the Fourier-transform spectrum and LPSVD analysis,53 as shown in Figure 2.11 that is not directly addressed by the authors. Figure 2.11 shows the dynamic-absorption transient of deoxyMB; Panel b reveals a broad background centered near 300 cm“1 that underlies the narrower components. This lineshape is similar to that discussed earlier from the vibrational coherence from the reaction center of Rhodobacter sphaeroides (see Figure 2.8). Simi- lar features can be seen in Figures 2.9 and 2.10. In each case, the broad portion of the broad background has superimposed narrower components. The broad lineshape is characteristic of a rapidly damped feature in the time domain. The vibrational modes corresponding to these features likely play a critical role in the previously discussed electron transfer reactions in the photosynthetic reaction center. 21 AT (a.u.) -1 -1 0.0 Residual Fe-His 437 nm Residual Fe-His 439 nm Residual Fe-His 442 nm Residual F -H' e ls 445nm 1.5 2.0 2.5 3.0 Time (ps) 22 Intensity (a.u.) C? -L O —L Fe-His 167 0 0 100 200 300 400 Frequency (cm'1) Figure 2.9. Open band femtosecond dynamic-absorption spectroscopy measure- ments on MbNO samples. The left panels present the oscillatory parts of the the pump-probe signals (circles) with their LPSVD fits (solid line). The 40 cm‘1 com- ponent (dashed line) is superimposed on the oscillatory signal and the 220 cm—1 mode shown is shifted for clarity (thin solid line). The right panels show the LPSVD- generated power spectra. (Adapted from reference 54.) Ferrous cytochrome c 45 ’3 , \W 91 , . V_ 4 158 43 87 168 :i A 43 «ii =3 V G 9 s C C 3 -1. E 91 E _ 45 92 169 -1. 45 W o. / 1 45 . W 91 - o 1 2 -1. / _________ 91m? .......... 0.0 035 1.0 135 2:0 25 3:0 3.5 0 160 260 300 Time (ps) Frequency (and) Figure 2.10. Sequence of dynamic-absorption transients using the open band detec- tion scheme on FeCytc. The left panels present the oscillatory part of the signal, and the right panels show the corresponding power spectra. The ~45 cm‘1 oscillation (solid line and the ~91 nw oscillation (dashed line) used in the fit are shown displaced from the data. The inset in the lower-right panel displays the Morse potential used in the simulation. (Adapted from reference 57.) 23 (b) deoxyMb 12:1.76 ps lllll;l l4lllllLJlll Relative Intensity 1411'] l—Ll l l_LLl l 114. I l lllllllllllleJnlllLlilL 0.5 1.0 1.5 100 200 300 400 t (m) '17 (cm") Figure 2.11. Femtosecond dynamic-absorption transients of deoxyMb. The signal (a) is shown along with the LPSVD fit (solid line). The discrete Fourier transform ampli- tudes (dotted line) and the power spectrum derived from the LPSVD analysis (upper solid line) (b) are compared with the resonance Raman spectrum (lower solid line). The low frequency residual (c) is superimposed with the LPSVD fit that describes its oscillation and decay. (From reference 53.) 24 2.6 Vibrational Coherence in Bacteriochlorophyll Solutions and Proteins The experiments that will be discussed in Chapters 3-5 were inspired by work from this laboratory on bacteriochlorophyll a (BChl) that suggested a new interpretation for the structural origin of the ~100 cm“1 modes coupled to the electron trans- fer reactions in the photosynthetic reaction center. The results of the experiments suggested that these modes arise from intermolecular interactions between the chro- mophore and the solvent or surrounding protein medium. Experiments on BChl a in polar solutions revealed that the vibrational coherence was dominated by rapidly damped modes. These modes exhibited a shift in frequency with a greater than quadratic dependence on the solvent dipole moment. This behavior was consis- tent with a van der Waals intermolecular potential.63 Additional experiments on the light-harvesting subunit proteins B777 and B820 revealed that direct attack on the n-electron density of the chromophore by the solvent appeared to be necessary for the intermolecular mode to gain resonance Raman activity. Although the later experimental chapters cover this work In more detail, in the following the main ideas are reviewed. 2.6. l Vibrational Coherence in Bacteriochlorophyll a in Polar solution Pump-probe experiments with Q-band excitation on BChl a in pyridine solution revealed two distinct types of components in the low-frequency vibrational coher- ence.64 The first were a set of slowly damped components that persisted out to 8 ps. These components had frequencies over a range of 11—206 cm‘l; their damping times y ranged from 800 fs to 1.7 ps. In addition to these components, a very strong, rapidly damped feature was also observed. This component had a much shorter 25 x3 19‘ 3' 'E :3 OJ .2 2- E . (D J: L 1 ps I— 1- I— < o I l I l l I 0 2 4 6 8 10 Probe Delay (ps) Figure 2.12. Dynamic-absorption transient from BChl a in pyridine. The inset shows a magnified View of the first picosecond of the signal. (From reference 63.) 26 [1((0) I (t) \ A, II 2ps Figure 2.13. Model of the rapidly damped vibrational coherence observed in BChl a solutions as a distribution of damped cosinusoids, £(w). A log-normal disuibu- tion L (a), defined by its center frequency (00, width Aw and asymmetry (or skew) p, sampled over its full width. Each sample Lorentzian lineshape has its own center frequency 000,-, and corresponds in the time domain to an exponentially damped cos- inusoid in the time domain (b) with an intrinsic damping time y. Summing the set of Lorentzians obtained by sampling over the full width of the log-normal distribution £(w) (c) results in a superposition whose rapidly damped waveform (d) resembles the vibrational coherence observed in BChl solutions. (Adapted from reference 63.) 27 200 - acetone 175 . '5 Q 150 _ ’3‘ pyridine 125 .. 1-propanol 100 A ' I ' l I Dipole Moment, D Figure 2.14. Solvent dependence of the mean frequency (00) of the rapidly damped vibrational coherence observed in BChl (1 solutions in a range of polar solvents on the solvent dipole moment. The data points are superimposed on a model curve derived from an expression for the natural frequency for a 6-12 intermolecular potential for the BChl-solvent mode. The expressions for the intermolecular potential and natural frequency are detailed in reference 63 and in Chapters 3-5 (From reference 63.) 28 damping time (200 fs), and dominated the early-time portion of the transient absorp- tion signal. It was suggested that this 155 cm‘1 mode was due to an intermolecular interaction between BChl and a pyridine molecule in the first solvation shell. Subsequent experiments were performed by Katherine Shelly on BChl a in a se- ries of polar solvents."33 The observed vibrational coherence was again dominated by rapidly damped components (see Figure 2.12). The figure shows the dynamic- absorption transient from BChl a in pyridine solution. By far the strongest feature in the vibrational coherence is a rapidly damped component, which is shown in the inset of Figure 2.12. This feature is almost completely damped within the first 500 fs, but it is at least an order of magnitude larger than any other oscillatory component. The lineshapes corresponding to the rapidly damped components exhibited behavior consistent with the superposition of several Lorentzian lineshapes with a distribution of frequencies (see Figure 2.13). In such a picture, the overall lognormal distribution shown in Figure 2.13a arises from the sum of individual Lorentzian lineshapes. This behavior is consistent with a picture where the rapid damping times are due to inho- mogeneous line broadening. The individual Lorentzians correspond to the lineshape generated by a single molecule. The local variations between different molecules in the ensemble generate slightly different frequencies. The time-domain superposition of these individual waveforms is a rapidly damped oscillation. The rapidly damped components also exhibited a solvent-dependent frequency shift. The shift followed a trend consistent with a van der Waals intermolecular potential between the chro- mophore and a nearby solvent molecule (see Figure 2.14). The van der Waals potential is discussed in detail in the experimental chapters 3-5. The frequency of the rapidly damped mode increased with a greater than quadratic dependence on the dipole mo- ment of the solvent. The intermolecular potential was dominated by dipole-dipole interactions. In the nonpolar solvent limit, the mean frequency of the rapidly damped vibrational coherence was extrapolated to be ~100 cm‘l. The results further sug- 29 gested that interactions between the chromophore and clustered, first-shell solvent molecules dominated the low-frequency vibrational coherence. This conclusion pro- vides the movitvation for much of the work that follows in this dissertation. 2.6.2 Intermolecular Vibrational Coherence From B777 and B820 Additional experiments on the light-harvesting subunit proteins B777 and B820 yielded similar results.65 B820 is a dimer containing a pair of BChl macrocycles (see Figure 2.15). In the purple-bacterial LH1 light—harvesting complex, 15 B820 subunits combine to form a ring around the photosynthetic reaction center.66 The monomer that pairs to form B820 is known as B77767"71 In B777, a single BChl macrocy- cle binds to an or helix by coordinating the Mg11 ion to a histidine residue.“72 In the monomeric form, the BChl is directly exposed to the solvent. In B820, the BChl macrocycles form a van der Waals complex. Figure 2.15 shows that in the complex, the BChl macrocycles directly face each other. In this configuration, the macrocyles resemble the special pair in the primary electron donor P in the photosynthetic reac- tion center.65' 73' 74 In B777, the vibrational coherence strongly resembled that of BChl in acetone from the earlier experiments (see Figures 2.16 and 2.17). The rapidly damped por- tion of the signal was much more intense than the more slowly damped oscilla- tions. The rapidly damped feature could be modeled as the sum of two compo- nents. These components were assigned to hindered translational and rotational (librations) modes between the porphyrin chromophore and one of the surrounding solvent molecules.“65 In B820, the vibrational coherence, while still dominated by rapidly damped components, was significantly weaker (see Figure 2.18) than in B777. The frequencies of the rapidly damped components were also lowered. These results were consistent with an assignment of the rapidly damped vibrational coherence to intermolecular vibrational modes. In the monomeric B777 direct attack on the rt— 30 Figure 2.15. The Hu and Schulten75 model for the B820 subunit in LH1 from Rhodobacter sphaeroides, showing the pair of BChl macrocycles and the transmem- brane or helices in ribbon and surface renderings. (From reference 65.) 31 electron density of the chromophore is possible. The B820 dimer, however, shields the chromophore because the BChl macrocycles are paired. As a result, the strength of the intermolecular interaction is significantly attenuated. Additional interactions between the BChl macrocycle and a nearby tryptophan residue and BChl-BChl inter- actions were also observed in B820. The results further advanced the hypothesis that the intermolecular vibrational modes are the dominant ones in the low-frequency vi- brational coherence. They also suggested a mechanism for gaining resonance Raman activity: direct attack on the n-electron density of the chromophore. 2.7 Proposed Experiments The results from this work on BChl a, B777 and B820 suggest further experiments to clarify the nature of this intermolecular vibrational mode. These results suggest that the vibrational coherence from BChl is dominated by intermolecular modes with a clustered, first-shell solvent molecules that attack the Tr-electron density of the BChl macrocycle. This result should be generalizable to other tr-electron containing systems, such as a porphyrin. It is not clear, however, that the intermolecular mode from BChl in polar solution is fully understood; the vibrational coherence lacks nor- mal intramolecular (skeletal) vibrational modes of the porphyrin chromophore. The relative resonance Raman activities of the intramolecular and intermolecular modes cannot be directly compared. In order to compare the two types of modes, both must be present in the vibrational coherence. By varying the solvent environment, the in- termolecular van der Waals modes can be distinguished. The Soret-band excitation experiments performed by Champion and co-workers on myoglobin and cytochrome c revealed the presence of several skeletal modes in the vibrational coherence,53'S7 so they suggest that other porphyrins or porphyrin-containing proteins might make good candidates for further study. 32 103x AT/T I l I T I 200 300 400 500 600 Probe Delay (fs) CD '0 3 'E U) (U 2 (D .2 <15 (I) 0: 0 I I I I I 0 100 200 300 400 500 Frequency (cm") Figure 2.16. Expanded view of the rapidly damped oscillation observed in the dynamic—absorption transient from BChl a in acetone. The signal is superimposed with a model defined by the sum of two independent log-normal distributions £(w) of damped cosinusoids. The scaling of the ordinate is relative to the magnitude of the pump-probe ground-state depletion signal. Bottom: Plots of £(w) for the two com- ponents observed in acetone and their sum .’M(w) (thick curve). (From reference 63.) 33 80 — I'- \ l- 40 — G X m 0 '— o T' -40 _ I I I I I 200 300 400 500 600 Probe Delay (fs) 0 .o B 'E O) (O 2 0) ..>_ E G) D: 0 0 100 200 300 400 500 Frequency (cm'I) Figure 2.17. Expanded View of the rapidly damped oscillation observed in the dynamic-absorption transient from B777. The signal is superimposed with a model defined by the sum of two independent log-normal distributions £(w) of damped cosinusoids. The scaling of the ordinate is relative to the magnitude of the pump- probe ground-state depletion signal. Bottom: Plots of [(00) for the two components observed in B777 and their sum 5mm) (thick curve). (From reference 65.) 34 l- 4- \ l- _ <1 0 X m ’4'— o ‘- -8- I I I I I I I I I 400 600 800 1000 1200 1400 1600 1800 2000 Probe Delay (fs) (D '0 3 'E U) (U 2 (D .2 E SE 0 I I I I 0 50 100 150 200 Frequency (cm'I) Figure 2.18. Expanded view of the rapidly damped oscillation observed in the dynamic-absorption transient from B820. The signal is superimposed with a model defined by the sum of two independent log-normal distributions £(w) of damped cosinusoids. The scaling of the ordinate is relative to the magnitude of the pump- probe ground-state depletion signal. Bottom: Plots of £(w) for the two components observed in B820 and their sum .’M(00) (thick curve). (From reference 65.) We can examine the vibrational coherence from ZnII mesa-tetrakis(N- methylpyridyl)porphyrin (ZnTMPyP) in a series of polar solutions. These experiments use Soret-band excitation, so they monitor the ground-state vibrational coherence of ZnTMPyP in analogy to the experiments on BChl. By varying the dipole moment of the solvent, we can look for changes in the mean frequency of the rapidly damped mode. We can also compare the vibrational coherence from ZnTMPyP in methanol and perdeuturated methanol solutions to search for evidence of an isotpe effect. ZnTMPyP has a small permanent dipole moment compared to BChl, so the frequency of the rapidly damped mode should be smaller than that observed in the vibrational coherence from BChl solution. The rapidly damped vibrational coherence should still exhibit a solvent dependent shift, and it should be more intense than the slowly damped skeletal modes. The stimulated-emission experiments by Vos, Martin and co-workers on pho- tosynthetic reaction centers monitored vibrational coherence from the excited state.45’48' 50' 51' 76 By tuning the laser Q-band of the absorption spectrum, the vibra- tional coherence from the excited state ZnTMPyP can be similarly monitored. In the excited state, the Tr-electron density of ZnTMPyP extends over the N-methylpyridyl rings, so the van der Waals intermolecular potential senses charges on the porphyrin. The addition of charges to the intermolecular potential should have a significant ef- fect on the frequency of the intermolecular mode. We cannot fully characterize the behavior of the van der Waals mode in a protein environment from the B777 and B820 experiments. In B777, the BChl macrocycle is directly exposed to the solvent, while in B820 the BChl forms a dimer to provide shelter from the solvent. A monomeric porphyrin in the interior of a protein may behave differently. The issue of which amino-acid residues in the protein couple to the TI —>Tr* transition of the porphyrin macrocycle must be considered. In a protein such as ZnII cytochrome c (ZnCytc), we can compare the vibrational coherence from 36 the native and molten-globule states to test the response of the van der Waals mode to a change in the local environment of the porphyrin. The following chapters will address these issues. 37 CHAPTER 3 Ground-state vibrational coherence in ‘6‘“ m-«P-fi polar solutions of ZnII tetrakis(N-methylpyridyl)porphyrin with Soret-band excitation 3.1 Summary Ground-state coherent wavepacket motions arising from intermolecular modes with clustered, first-shell solvent molecules were observed using the femtosec- ond dynamic absorption technique in polar solutions of Zn11 mesa-tetrakis(N- methylpyridyl)porphyrin (ZnTMPyP) with excitation in the Soret absorption band. As was observed previously in bacteriochlorophyll a solution, the pump-probe tran- sients in ZnTMPyP solutions are weakly modulated by slowly damped (effective damping time y > 1 ps) features that are assigned to intramolecular modes, the skele- tal normal modes of vibration of the porphyrin. The 40-cm‘1 and 215-cm"1 modes from the metal-doming and metal—solvent-ligand modes, respectively, are members 38 of this set of modulation components. A slowly damped 2-4-cm‘1 component is assigned to the internal rotation of the N-methylpyridyl rings with respect to the porphyrin macrocycle; this mode obtains strong resonance Raman intensity enhance- ment from an extensive delocalization of rt-electron density from the porphyrin in the ground state onto the rings in the n* excited states. The dominant features observed in the pump-probe transients are a pair of rapidly damped (y < 250 fs) modulation components arising from intermolecular modes with solvent molecules. This structural assignment is supported by an isotope shift of the average mode frequencies in methanol and perdeuterated methanol. The solvent dependence of the mean intermolecular mode frequency is consistent with a van der Waals inter- molecular potential that has significant contributions only from the London disper- sion and induction interactions; ion-dipole or ion-induced-dipole terms do not make large contributions because the Tr-electron density is not extensively delocalized onto the N-methylpyridyl rings. The modulation depth associated with the intermolecu- lar modes exhibits a marked dependence on the electronic structure of the solvent that is probably related to the degree of covalency; the strongest modulations are observed in acetonitrile and dimethylsulfoxide. The results strongly support a struc- tural assignment of the low-frequency modes that are coupled to the primary and secondary electron-transfer reactions in photosynthetic reaction centers to inter- molecular modes between the redox-active chromophores and first-solvation shell groups from the surrounding protein, and an important additional function of the intermolecular modes in the stabilization of charged intermediates is suggested. 3.2 Introduction In the purple-bacterial photosynthetic reaction center, the primary and sec- ondary charge-separation reactions exhibit effectively activationless dynamics ow- 39 ing to coupling to low-frequency vibrational modes in the 100-cm'1 regirne.1'20’22 The main components detected in the vibrational coherence observed in pump- probe43“16"“3’51’76 and fluorescence upconversion77 transients following impulsive excitation of the primary electron donor, P, are modes with frequencies in the 30— 100-cm‘1 range. The possibility that the modes observed in the vibrational coher- ence are the ones that are coupled to the electron-transfer reaction coordinate has been discussed,4gv 78‘“ but their structural origin remains indeterminate. Bixon and Jortnerlrzoi 27- attributed the coupled modes to the protein medium, and Vos, Martin, and coworkers suggested that the modes that are active in the vibrational coherence are delocalized or phonon-like in character.“ 50 In recent work, the Beck laboratory has advanced an alternative hypothesis that the vibrational modes that control the electron-transfer dynamics in reaction cen- ters are intermolecular in origin, between the redox-active chromophores and groups in the first-solvation shell from the surrounding protein medium. This hypothesis is supported by our observations using femtosecond pump-probe spectroscopy of the rapidly damped low-frequency vibrational coherence from bacteriochlorophyll a (BChl) in polar solution“!64 and in the light-harvesting subunit proteins B777 and B820.65 In solution, BChl exhibits a slowly damped set of modulations over the 100— 8000-fs probe delay range; these components exhibit damping times in the >1-ps regime and mode frequencies ranging from 10-220-cm‘1. The magnitude spectrum obtained by Fourier transformation is comparable to the low-frequency region of the conventional resonance Raman spectrum from BChl in solution, in films, and in the reaction center,85'86 so these slowly damped features are assigned to the low- frequency skeletal modes of the BChl macrocycle.64 By far the strongest features in the vibrational coherence from BChl solutions, however, are a very rapidly damped set of modulation components in the sub-ps time scale. These features are well de- scribed by time-domain models that correspond in the frequency domain to inhomo- 40 geneously broadened lineshapes with very broad, asymmetric Gaussian profiles. The mean frequency of the rapidly damped vibrational coherence exhibits a greater than quadratic dependence on the gas-phase dipole moment of the solvent. This trend is consistent with a van der Waals intermolecular potential in which the London- dispersion and dipole—dipole interactions make large contributions; the extrapolated 100-cm‘1 frequency in the nonpolar limit arises predominantly from the London- dispersion term.63 The rapidly damped vibrational coherence observed in the B777 system is comparable to that observed with BChl in acetone solution because the BChl macrocycles are exposed to the polar head groups of the surrounding nonionic de- tergent. In the paired BChl system called B820 that forms when two B777 monomers associate, however, the supporting or helices sterically protect the BChl macrocycles from direct attack by the surrounding detergent. The vibrational coherence then re- veals weaker and more slowly damped components at 28 cm"1 and 46 cm‘1 that are assigned to BChl—tryptophan and BChl—BChl intermolecular modes.65 The intermolecular modes that contribute to the vibrational coherence in BChl-containing systems modulate the pump-probe ground-state depletion sig- nal owing to the resonant impulsive stimulated Raman scattering (RISRS) mecha- nism.24‘26'30'32'3"3 It was suggested that these modes are resonance Raman active because molecules in the first solvation shell make a direct attack on the Tr-electron density above and below the macrocycle and are displaced by the Tr—>1T* transi- tion.63v65 This hypothesis suggests that the spatial organization of the n-electron density in the ground state selects the first-shell solvent molecules that contribute to the vibrational coherence. As a test of these ideas, we discuss in this chapter the low-frequency ground-state vibrational coherence in polar solutions of ZnII meso- tetrakis(N-methylpyridyl)porphyrin (ZnTMPyP, see Figure 3.1). Owing to the nominal C4 symmetry axis normal to the macrocycle, ZnTMPyP has a small net dipole mo- ment; with a single methanol (CH3OH) axially ligated to the ZnII ion, the calculated 41 dipole moment is only 0.66 D (see Figure 3.1). ZnTMPyP is nevertheless soluble in a range of polar solvents because the peripheral N-methylpyridyl rings each carry a positive charge. Because the Tr-electron density in the ground state (see Figure 3.1a) is mostly confined to the region of the porphyrin, however, the resonance Raman active solvent molecules are effectively held at some distance from the peripheral charges, so charge-dependent terms will make only a small contribution to the intermolecular potential. In short, the ground-state intermolecular mode frequency in ZnTMPyP so- lution should be smaller than detected in the same solvents in BChl solution because the only terms in the intermolecular potential that will make large contributions are those that depend on the polarizability of the porphyrin: the London-dispersion and solvent-dipole-induced-dipole terms. The results show that the vibrational coherence from ZnTMPyP detected with Soret-band excitation contains dominant contributions from rapidly damped features arising from intermolecular modes with clustered solvent molecules. The compo- nents arising from intramolecular modes, from the porphyrin macrocycle proper or from the metal-axial-ligand interaction, can be distinguished from the intermolec— ular modes because they exhibit long damping times and narrow lineshapes. The intermolecular mode frequency follows the expected dependence on the solvent’s dipole moment; the 70-cm'1 frequency in the nonpolar limit and the observation of a isotope shift in perdeuterated methanol are consistent with the hypothesis dis- cussed above. In addition to supporting a structural assignment of the low-frequency modes that are coupled to the electron-transfer reactions in photosynthetic reaction centers, the results have additional significance because they suggest a role for in- termolecular modes in the stabilization of charged intermediates in redox catalysis in proteins. The results described in this chapter were previously published in the Journal of Physical Chemistry (see reference 88.) 42 Figure 3.1. Optimized structure of Zn11 meso—tetrakis(N-methylpyridyl)porphyrin (ZnTMPyP) complexed with a single methanol molecule as an axial ligand to the Zn11 ion. The structure was obtained from a B3LYP hybrid density functional electronic structure calculation with Gaussian 0387 at the 6-31G level of theory. The ball-and- stick structures are shown superimposed with density surfaces for (a) the highest occupied molecular orbital (HOMO) and for (b) the lowest unoccupied molecular or- bital (LUMO). 43 3.3 Experimental 3.3. 1 Sample Preparation ZnTMPyP (CAS 28850-44-4) was used as received from Sigma-Aldrich. CH3OH (spectrophotometric grade) and CD3OD (99.8 atom %D) were obtained from Sigma- Aldrich. CH3OD (99 atom %D) was obtained from Cambridge Isotope Labs. Ace- tonitrile (CH3CN, from Mallinkrodt, spectrophotometric grade), dimethylsulfoxide (DMSO, from EMD, ACS grade), and N,N-dimethylformamide (DMF, from Jade Sci- entific, reagent ACS grade) were used as received. For use in the femtosecond pump-probe experiments, solutions of ZnTMPyP were prepared by dissolving the dry ZnTMPyP powder in the indicated solvent to obtain an absorbance of 0.4 for a path length of 1.0 mm at the center of the laser spectrum at 420 nm, as detailed below; with the methanols and DMSO, the solution was passed through a 0.22-pm microfilter prior to checking the absorbance. The samples were held in the femtosecond pump-probe spectrometer at room temperature (23 °C) in a fused-silica flow cuvette (0.5-mm path length). A peristaltic pump was used to circulate a 10-mL reservoir of sample solution through the cuvette at 2.75 mL/min. The sample’s absorption spectrum was monitored during the experiment for changes arising from photochemistry or permanent photobleaching. The sample reservoir Was exchanged with fresh solution several times during each run. 3.3.2 Continuous-Wave Absorption and Fluorescence Spectroscopy Absorption spectra were obtained at 23 °C with a Hitachi U—2000 spectrophotometer (2-nm band pass). Fluorescence spectra were acquired at 23 °C with a Hitachi F-4500 spectrofluorimeter (5-nm band pass for the excitation and emission monochroma- tors). As presented as a function of wavenumber, the fluorescence intensities are 44 multiplied by the square of the wavelength in order to compensate for the fixed (in wavelength units) spectral bandpass of the emission spectrometer.89 3.3.3 Femtosecond Spectrosc0py Femtosecond pump-probe transients with impulsive excitation were recorded using the dynamic-absorption technique26v32'35'37 with a pump—probe spectrometer (see Figure 3.2) consisting of a frequency doubled, self-mode-locked Ti:sapphire oscil- lator (Coherent Mira-F oscillator and Verdi V5 (5 W) NszVO4 pump laser, Coher- ent/Inrad 5-050 second-harmonic generator), a SFlO Brewster prism-pair pulse com- pressor, and a rapid-scanning, modified Mach-Zehnder interferometer with confocal sample and autocorrelation-crystal positions. The present experiments employ 50-fs pulses with intensity spectra centered at 420 nm (4-nm FWHM, as measured with an Ocean Optics USE-2000 spectrometer/CCU detector with a 0.5-nm bandpass), and a fairly narrow bandpass (0.5 nm) of the transmitted probe beam. This approach cor- responds to that used by Champion and coworkers in their studies of low-frequency vibrational coherence in heme proteins”, 55162 The apparatus and methodology used in this work is largely that described in our recent study of rapidly damped vibrational coherence in BChl solutions with Q- band (780-nm) excitation,63 but a number of changes were required to permit work in the blue part of the spectrum. A 76-MHz train of 840-nm pulses is generated by a Coherent Mira 900E Ti:sapphire oscillator equipped with Coherent’s X-wave broad- tuning-range (690-1020-nm) cavity optics; the oscillator is pumped by a Coherent Verdi pump laser (S-W output). As shown in Figure 3.2, these red pulses first enter a pulse-compression system composed of a double-passed pair of SFlO Brewster- angled prisms and an adjustable interprism path length. The output pulses are di- rected to a Coherent/Inrad 5-050 second-harmonic generator, which is fitted with a 45 6506993 6n HoumBoEooaoE cmfidm noumumwmdouoa. .02 ”BB Houauafiouonn ESE ”Rambo 203235025 55.62 Enigma .U< ”5559353 08.308 .35 £552 305 5% 25 25c Bongo 5>flp-558o§>~ww :50 Eggnog 5:28 5mm_-:m_u .9 ”55:69: 35200533 .me 553553 5382-3,“ .3 55533 55:62: .8 5559953 “3:359:00 .6 ”5523832 55058855 .2? ”£385 550 55885533 Hovngfiomz “5558 m was .629 555 how oEoEHmnécouom .02 535358 339 5985.5 a Co 96228 5588553 unearnfisn “5:838:29.— .~.m wan-ME v 5mg aan 40>>-___uz Q m> 65> #55200 ' v 55580 oLEaammuF 2.08 9:2 “55:00 wmo wm< o ‘V o \ o< . >> >> LQmEoEoocos. a . . a . . :25 .. J 0 2mm 559qu0 523 H 4mm. 5.50 co>_5-55Eo5>_mo 53> 55Eo5t2£ 5occoNéoms. 46 type I B-barium borate (BBO) second—harmonic generation (SHG) crystal (l-mm path length). The co-propagating 420 and 840-nm output beams that radiate from the BBQ crystal are then directed to a modified Mach-Zehnder pump-probe interferometer. For this work, the interferometer was redesigned to support a dichroic mode of oper- ation that allows us to control and characterize pump-probe pairs of blue pulses in the 350—500-nm region of the spectrum. The interferometer employs a femtosecond autocorrelator beamsplitter (CV1 Laser, FABS-800-4SS) to split the input beam into pump and delayed probe pulses; as specified, the 800-nm beamsplitter divides the s-polarized 840-nm light evenly into two beams, but the p-polarized 420-nm light is divided by the beamsplitter into two beams with a 3:1 pump:probe intensity ratio. The pump-probe delay is scanned continuously by a galvanometer-driven retroreflec- tor (Clark-MXR, ODL-150) mounted in the probe beam’s arm of the interferometer. The pump beam’s intensity is modulated at 100 kHz by a A/2-retarding (at 420 nm) photoelastic modulator (Hinds Instruments) and a calcite polarizer (Karl Lambrecht), in series.90 Because the planes of polarization of the 420-nm and 840-nm compo- nents are orthogonal as they enter the photoelastic modulator owing to the use of the Type I SHG crystal, the 420-nm and 840-nm beams that emerge from the calcite polarizer exhibit amplitude-modulation envelopes that are 90° out-of-phase with re- spect to each other, but no their planes of polarization are parallel, as analyzed by the polarizer. The compensating calcite polarizer in the probe beam is oriented 45° with respect to the pump-beam’s polarizer, so the two orthogonally polarized colors are analyzed equivalently, and the output planes of polarization are parallel. Finally, a single dichroic beamsplitter removed from the Coherent/Inrad 5-050 SHG unit is mounted to separate the 420-nm and 840-nm light in the pump and probe beams, which propagate along parallel paths on the optical table as they exit the delay arms of the Mach-Zehnder interferometer. A fused-silica singlet lens (5- 47 cm focal length) is employed to focus the 420-nm pump and probe beams onto the sample; the 840-nm pump and probe beams are focused onto a thin (100nm thick- ness) BBO autocorrelation crystal by a matching lens. After it emerges from the sample, the probe beam is analyzed by a calcite polarizer oriented at 90° with re— spect to the pump-beam’s plane of polarization, and then the residual probe light enters an Acton-Research SP-150 monochromator. An amplified photodiode (Thor- labs PDA55) and a lock-in amplifier (SRS, SR8 50) referenced to the pump-modulation frequency are used to detect the pump—probe signal from a 0.5-nm bandpass of the probe beam. A photomultiplier (1P28, in a Clark-MXR housing) and a lock-in am- plifier (Femto, LIA-MV-200H) are used to detect the background-free autocorrelation signal that is generated by the 840-nm pump-probe beams incident upon the BBQ autocorrelation crystal. The pump-probe and autocorrelation signals are recorded and averaged simultaneously by a sample-and-hold amplifier and transient recording system, as described previously.29 In the present work, the probe delay is linearly scanned at 1.0 Hz over the -3 to +27 ps range. The pump-probe and autocorrelation signals are sampled at 17 kHz, and there are 13000 points in each scan; the effective dwell time during data acquisition is 2.3 fs/point. The autocorrelation signal obtained from the 840-nm pump and probe beams is used in the characterization of the group-delay dispersion of the entire apparatus. The effective instrument-response function for the 420-nm pump and probe pulses is minimized by adjusting the distance between the pair of Brewster prisms; the width of the 840-nm autocorrelation and the pump-probe photobleaching rise time obtained with the 420-nm pulses are monitored simultaneously. The autocorrelation signal is also used as a reference timing pulse for the transient recorder29 that is used to acquire simultaneously the autocorrelation signal and the pump—probe signal. The autocorrelation pulse is precisely correlated on the probe-delay axis with the zero of time, where the 420-nm pump and probe pulses overlap temporally in the sample, 48 so its timing is used to cancel the temperature-induced drift in the dimensions of the pump and probe arms of the interferometer and to compensate for jitter in the scanning of the probe-delay line. The use of the fundamental (red) pulses in the autocorrelation arm of the instrument permits us to operate this instrument with the blue pump and probe pulses tuned well into the ultraviolet. If the blue pulses were used directly to obtain the autocorrelation, the 200-nm transmission cutoff of the BBQ crystal would not permit tuning of the pump and probe pulses below 400 nm. Each dynamic-absorption transient that is presented here and in the subsequent chapters is the result of averaging together several scans. For each probe delay point, the mean and standard deviation from the set of individual scans were recorded. Statistical filtering was used to reject outlying points. The transients effectively rep- resent the result of several days worth of repeated experiments for each type of sample. 3.4 Results Figure 3.3 shows continuous-wave absorption and fluorescence spectra from ZnTMPyP in methanol (CH3OH) at 23 °C. The spectra are plotted as relative dipole strengthsggvgli92 as a function of wavenumber v, A(v) /v and F (v) /v3, respectively. The absorption spectrum features two bands, the Soret (or B) band and the Q band, in the blue and red parts of the spectrum, respectively. The fluorescence spectrum extends to the red of the Q band. Figure 3.4 shows the solvent dependence of the Soret-band region from solutions of ZnTMPyP in CH3 OH, dimethylformamide (DMF), acetonitrile (CH3CN), and dimethylsulfoxide (DMSO) at 23°C. Also shown in Figure 3 is the output spectrum from the frequency doubled Ti:sapphire oscillator as it was tuned for the dynamic-absorption experiments to 420 nm. The detected probe band- pass was obtained from the red side of the laser spectrum (422 nm). 49 Relative Dipole Strength Frequency (103 cm") Figure 3.3. Soret (v > 20000 cm‘l) and Q-band region of the continuous-wave absorption (solid curve) and fluorescence (dotted curve) spectra from ZnTMPyP in CH3OH at room temperature (23 °C), plotted as the dipole strength, A / v and F/v3, respectively, and normalized to unit area with respect to the fluorescence spectrum and Q band. 50 k 20 22 24 26 Frequency (103 cm“) Relative Dipole Strength C.’ ggg Figure 3.4. Soret-band absorption dipole-strength (A/v) spectra from ZnTMPyP in (a) CH3OH, (b) DMF, (c) CH3CN, and (d) DMSO. Superimposed with arbitrary scaling on (a) is the intensity spectrum of the 420-nm, 50-fs pulses used in all of the pump— probe experiments. 51 The dynamic-absorption transients obtained from the ZnTMPyP solutions under these experimental conditions are shown in Figure 3.5. Following an intense coher- ence spike that goes off the plotted scale,93'95 the transients exhibit a pattern of cosinusoidal modulations nearly all the way to the end of the 25-ps recording that is superimposed upon a single or double exponential decay function. The modu- lation pattern includes a strong very low-frequency component with positive-going recurrences at ~2 ps and ~18 ps; this oscillation carries a less intense set of higher frequency modulations that are even more slowly damped. Lastly, a very rapidly damped modulation contributes primarily to the 250-600-fs region of the transient. This portion of the signal is similar in character to the rapidly damped vibrational coherence observed in polar solutions of BChl, where the slowly damped portion of the signal is perhaps ten times weaker than observed in the ZnTMPyP transient.63 The dynamic-absorption signals from ZnTMPyP were fit in the time domain to a model consisting of a single- or double-exponential decay and an oscillatory portion containing slowly and rapidly damped parts. The models used previously with the signals observed in BChl solution and in the B820 or B777 systems were intended to handle only the rapidly damped components in the modulation pattern.63'65 After truncation of the signal prior to the 250-fs delay point in order to avoid contributions from the tail of the coherence spike and from nonresonant background signals,95 the oscillatory part was isolated by subtracting a fitted single- or double-exponential function: I(t) = A0(1 +A1e-t/T1+A2e-t/T2) (3.1) The fit parameters are listed in Table 3.1. The transients and fits were normalized by dividing by A0 in order to compare the amplitude of the modulation components to the intensity of the pump-induced ground-state depletion signal, which is effectively constant on the 0—25-ps time scale. 52 / x3 AT/T m A ca 00 O 0 Z, 01- f i 10 15 20 Time (ps) Figure 3.5. Femtosecond pump—probe dynamic-absorption transients detected at 422 nm (0.5-nm bandpass) with ZnTMPyP solutions in (a) CH3OH, (b) CD30D, (c) DMF, (d) CH3CN, and (e) DMSO. The >250—fs portion of the signal in CH3OH-is also shown in a x3 expanded View. The signal traces are shown superimposed on single- or double-exponential fit functions of the form A0(1 + A1 e't/ T1 + Aze‘t/TZ) for the 250-fs-25-ps range. The fit parameters are provided in Table 3.1. As plotted, the signals are normalized by dividing by the non-decaying fraction, A0. 53 Table 3.1. Model parametersa for the exponential decays in the pump-probe transientsb in ZnTMPyP solutions at 22 °C. Solvent A1 T1 (ps) A2 T2 (ps) CH3OH 0.194 1 4 — — CD3OD 0.260 2.5 — — DMF 2.32 2.4 — — CH3CN 0.160 1 2 0.293 1.4 DMSO 1.64 0 8 5.78 3.2 a I(t) = A0(1 +A1e‘t/Tl +A2e‘t/72); A0 is a normalization con- stant. b See Figure 3.5. 54 AT/T 1 l 1% > 1 I > > I I 99 93.055.09.09 00 N-L-‘OOO—t ON OO’IOU‘IOU‘IO l 4 MAKAA/lll ,‘Iin A II\ AAAAAAAAALI ‘ ‘ UV VWUWVVVWWVWV’VVVHW 1000 2000 3000 4000 5000 Time (fs) Figure 3.6. Oscillatory signals obtained from the pump—probe transients from ZnTMPyP solutions (see Figure 3.5) as the difference between the normalized sig- nal and the fitted exponential decay functions: in (a) CH30H, (b) CD30D, (c) DMF, (d) CH3CN, and (e) DMSO. The data points are shown superimposed on a time- domain model composed of slowly damped and rapidly damped components (see Equations 3.2-3.6). The model parameters are listed in Tables 3.2 and 3.3. 55 The oscillatory residuals, the difference between the normalized signal and the fitted decay function, were then fit by a multicomponent model that contains both slowly damped and rapidly damped oscillatory components (see Figure 3.6). The slowly damped part was modeled over the 250—7000-fs range as a sum of damped cosinusoids of the form _ 2 2 Ii“) = Aie t 0i /2 cos(w0it — (bi) /\/21T (3.2) with each modulation component i having a center frequency 000,- and phase (1),. These waveforms correspond to Gaussian lineshapes in the frequency domain, _ _ . 2 2 1,-(w) = A,- e (w “’01) “2‘5 ’ / (aidzn) (3.3) where the linewidth is controlled by the standard deviation, 0,- = Awi / 2m, with Awi representing the full width at half maximum. Equations 3.2 and 3.3 are nor- malized so that the amplitude Ai corresponds to the area of the lineshape in the frequency domain and to the intensity of the signal in the time domain; Equation 3.2 corresponds to the Fourier transform of Equation 3.3. As discussed below, an approximate description of the signal can be obtained with components described by exponentially damped cosinusoids, Ii(t) = Ai e‘t/yi cos(w0it — (1),) (3.4) which correspond to Lorentzians in the frequency domain, but a poorer fit to the signal was obtained especially at long delays t; the model walks out-of-phase with respect to the experimental signal, and the time dependence of the amplitude is poorly described. The rapidly damped part of the oscillatory residuals (Figure 3.6) was modeled in the time domain as the sum of two inhomogeneously broadened components with asymmetric Gaussian lineshapes. It was previously shown that such a model pro- vides a good description of the rapidly damped vibrational coherence from BChl in 56 polar solution;63'64 the reader is directed to that work for a detailed discussion and for tests of the model. The two rapidly damped components are assigned, in order of frequency, to the hindered translational and librational (hindered rotational) inter- molecular modes between the porphyrin or chlorophyll macrocycle and its clustered, first-solvation-shell solvent molecules.63v 96, 97 In the work on BChl solutions, each rapidly damped component was described in the time domain by an integral over a distribution of exponentially damped cos- inusoids with an intrinsic (or homogeneous) damping time y; the intensities of the cosinusoids are scaled by a lognormal lineshape,98 £(w): II”) = I0 dw£(w) cos(wt — (1),) e’t/l’ (3.5) £(w) is parameterized by its area, At, its center frequency, wot, its width , Awi, and an asymmetry (or skew) parameter, pi. In the present work, the £(w) distribution function was approximated with the piecewise sum of two half-Gaussian lineshapes, 6010,14,, 000,-, 0,, pi), and the integral is replaced by a sum over an evenly spaced set of sub-Gaussian components I j(t) of width 03/ 4 spaced by (Ti/2: 22 1,-(t) = Z 6(w0j = 000,- + (j — 8)oj,Ai,w0i,oi,pi)IJ-(t,woj,oj = oi/4) (3.6) j=0 The j sub-Gaussians are expressed in the time and frequency domain by expressions analogous to Equations 3.2 and 3.3; in the sum, they are scaled by the intensity of 6(a)) sampled at their center frequencies, woj. The right- and left-side widths ”ml of 6(a)) are determined by the overall width of the distribution 0,- and its asymmetry Pi as O'i= (O'l-l-0'r)/2=0'l(1+pi)/2 (3.7) The widths or and o; are related by p,- = 07/01; the left- (I) and right-hand (r) half- Gaussians are defined by Equation 3.3 over 00 < (001' and w 2. 000,, respectively. Note 57 that the choice of the width and spacing of the sub-Gaussians determines the num- ber required (here, 23) to obtain a smooth approximation of a lognormal distribution. The sum in Equation 3.6 was truncated at 0 cm’l, if necessary, in order to avoid the addition of negative frequencies to the waveform, and it was normalized so that the area of the lineshape in the frequency domain was set to the amplitude parame- ter At. The resulting rapidly damped waveform and lineshape are indistinguishable from those obtained in the previous work for a given set of parameters, but the nonlinear optimization program runs perhaps two orders of magnitude faster than before because the lognormal lineshape function itself98 is avoided and the number of sub-Gaussians in the sum is small. A further benefit is that no assumption of the intrinsic damping time y is required—a Raman echo experiment would be required to obtain a discrete measurement of the homogeneous linebroadening”, 100 The slowly and rapidly damped portions of the signal were modeled separately in order to simplify the optimization procedure. The slowly damped part of the residual oscillatory signal was modeled first using a set of components constrained to have narrow lineshapes (Awi < 10 cm‘l). Starting frequencies and widths for these components were obtained from a Hanning-windowed Fourier-transform spec— trum (see Figure 3.7); the window function attenuates the rapidly damped portion of the signal.92 The twelve most intense components in the Fourier-transform spec- trum were included in the slowly damped part of the model. The converged model produced a Hanning-windowed Fourier transform spectrum that matched the exper- imental spectrum; the frequencies, widths, and relative amplitudes are particularly well defined because the linewidths are narrow. The rapidly damped model discussed above was then constrained during the second phase of the optimization process to return only broad lineshapes (Awi > 10 cm‘l). In all the solvents studied, the sum of two rapidly damped asymmetric Gaussian components provided a good description of the rapidly damped residual function. 58 801 A 60- 40- 201 x8 40- B 30~ 20- x4 10- 1204 C 30i x14 0 r Relative Magnitude _. -b 01 o o I l J 0. 801 E 60‘ 40_ x1.75 20- 0 100 200 300 400 Frequency (cm“) Figure 3.7. Fourier transform magnitude spectra from the oscillatory residuals (see Figure 3.6) of the pump-probe signals from ZnTMPyP solutions in (a) CH30H, (b) CD3OD, (c) DMF, (d) CH3CN, and (e) DMSO. The Fourier transform was applied to the oscillatory input signals after a Hanning window was applied over the 250-7000-fs range. Table 3.2. Solvent dependence of the asymmetric Gaussian lineshape pa- rameters for the rapidly damped components observed in the vibrational coherence from ZnTMPyP. Solvent Component Parametera CH3 OH CD30D DMF CH3 CN DMSO 1 (00, cm-1 68 63 82 80 100 Aw, cm-1 49 55 59 50 48 Ab 9.81 10.1 7.67 69.8 62.2 p 1.18 1.17 1.45 1.21 1.17 2 000, cm-1 81 73 94 119 124 Aw, cm-1 59 45 50 59 56 Ab 15.1 17.2 8.10 82.3 125 p 1.18 1.18 1.29 1.15 1.55 Sum (w), cm-1 79 74 95 106 125 z,- A, 24.9 27.3 15.8 152 187 3‘ See Equations 3.2-3.6 and the text. b Normalized to the intensity of the ~215-cm'1 slowly damped compo- nent in each solvent. 60 .Am.m 588$ 88 82888 888.888 88:88.8: 8 $538 w 828888 .8“ 88:88 and n .w 888988 h8 88 .3 88:88.88 55 3.. 8882988 55. .88 58 85 o.mim.m 88888.”— 88 83.1786 83d £188 .93 a 588 8:8 888 Sod Sod 53.. 88 .88 .38 :8 .88 88 88 .8 3. .88 88 .88 .28 88 .488 8:. 88 28.88.88 88.24.58 28.88.88 83.88.88 88.88.88 2 88 .488. 88 Ed .88 .88 88 .88 888 88 .88 88 a. o .88 .88 8 88 .88 .88 88 .588 .28 :8 88 8 .28 88 888 .88 88 .88 .88 a 8 .8888 o. . 8888 3 88888 on. 8.38 on. 8. 88 w E .88 88 :8 83. 88 Sod 888 88 88 888. .2: 8o .88 .2: a E .88 8: :3 88 8: 88 .88 .4: 88 .88 8: 8o .88 8: m 88. 8883 88.8882 68.88.42 8888882 no.6. 882 m 88 82. .8 Rd .88 .8 8.8 8:. .8 88 888 .8 88 88 8 .8 u. 88 .88 .8 88 888 .5. 888 .28 .8 88 888 88 88 .88 88 8 33888.2 88.2.8 48888.8 888.888 88.8. .8 8 88 .38 8. 88 .88 8 K 8.8 8 88 .45 8 88 5.8 8 8 828 2688 8,8 888 888 88888 acmZOm 8888.88 58.81888 88ng 8 88588 8858388 8088888 8888 >387. 55 .8..— 8855888 5888:: 886880 .m.m 532. 61 The sum of the rapidly damped and slowly damped models is superimposed on the data points in Figure 3.6. Both the slowly damped and rapidly damped portions of the model were found to be robust with respect to the starting parameters used in both parts of the nonlinear optimization procedure. The final fitted amplitudes Ai have estimated confidence intervals that are no worse than 255%, and the center frequencies 000i are known with confidence intervals of i 2 cm’ 1. The parameters for the optimized time-domain models shown in Figure 3.6 are tabulated in Tables 3.2 and 3.3. The tabulated amplitudes A, are normalized with respect to the intensity of the ground-state depletion part of the signal (see the ordinate scaling for Figure 3.5) and to the amplitude for the ~2ZlS-cm‘1 component, the strongest slowly-damped, high-frequency (000i > 100 cm”1) component in all of the solvents but DMSO, where the 195-cm‘1 component is the strongest. Effective exponential damping times for the Gaussian or asymmetric Gaussian modulation components were estimated by fitting an exponentially damped cosinu- soid (see Equation 3.4) to the Gaussian waveform for each component in the model. This exercise corresponds in the frequency domain to optimizing a Lorentzian lineshape to the corresponding Gaussian lineshape. As an example, the Gaussian linewidths, A000, and the exponential damping times, y, are compared for ZnTMPyP in CH3 OH in Table 3.4; similar results (not shown) are obtained in the other solvents. Figure 3.8 compares the Gaussian and fitted Lorentzian waveforms for the 3, 38, and 215-cm‘1 components. The Lorentzian waveforms were optimized over the 250— 7000-fs range, as in Figure 3.6, but the results are comparable when the entire 150- ZSOOO-fs range shown in Figure 3.8 is fit. In general, the fitted Lorentzian waveform is more intense than Gaussian waveform at the beginning and end of the waveform; if the Lorentzian waveform is constrained to fit the Gaussian waveform at the end of the modulation pattern, the resulting damping time is longer. This issue is especially important for the slowly damped 3-cm‘1 asymmetric Gaussian component, which 62 Table 3.4. Comparison of Gaussian linewidths and effective exponential damping timesa for the modulation components observed in the vibra- tional coherence from ZnTMPyP in CH3OH. 000,cm'1 Aw, cm‘1 y, ps 3 3.47 2.941).C 21 8.88 1.36 38 2.03 6.71C 68d 49 0.223 81cl 59 0.182 89 2.39 5.64 138 2.78 4.85 177 2.96 4.57 191 2.78 4.87 215 1.67 8.14c 244 1.90 7.16 264 3.62 3.58 284 3.16 4.11 312 3.24 4.16 a See Equation 3.4 and Figure 3.8. b Lower limit; an upper limit of y = 21.6 ps is obtained by fitting to the t > 15 ps portion of the signal. C See Figure 3.8. d Rapidly damped component, see Equation 3.6 and the text. 63 Relative Magnitude O I 1 I l 1 5 20 25 OH 01 —L C Time (ps) Figure 3.8. Comparison of the Gaussian waveforms (thick curves) obtained for the 3, 38, and 215-cm‘1 components in the oscillatory signals observed from ZnTMPyP in CH30H solution with fitted exponentially damped waveforms (thin curves, see Equation 3.4). The Gaussian linewidths and damping times are listed in Table 3.4. 64 Relative Magnitude 1 0 100 200 300 400 Frequency (cm'l) Figure 3.9. Frequency domain representation of the time-domain models of the os- cillatory signals (see Figure 3.6) observed in ZnTMPyP solution: in (a) CH30H, (b) CD30D, (c) DMF, (d) CH3CN, and (e) DMSO. The broad features in the SO—ZOO-cm‘1 regime are shown superimposed with the spectra (dotted lines) of their underlying components; in (d) and (e) these features are attenuated in order to keep them on scale. The spectrum from the 2—4-cm‘1 slowly damped component in (a)-(d) is also attenuated. 65 AT/T 0 200 400 600 800 1000 1200 1400 Time (is) Figure 3.10. Time-domain representation of the rapidly damped oscillatory compo- nents (see Figures 3.6 and 3.9) in the pump-probe signals from ZnTMPyP solutions in (a) CH30H, (b) CD30D, (c) DMF, (d) CH3CN, and (e) DMSO. The dotted portion of the signal is extrapolated from the fitted region of time (see Figure 3.6). 66 exhibits a 2.94-ps damping time if the overall waveform is fitted, but if the fit is con- strained to the >10-ps region, a 21.6 ps damping time is obtained. As mentioned above, we find that Lorentzian waveforms result in a poor overall model for the os- cillatory residuals even if constrained over the 250-6000-fs region, so the damping times in Table 3.4 have to be regarded as crude lower-limit estimates of the damping time. Given that Gaussian lineshapes are observed, and that these lineshapes inher- ently report inhomogeneous broadening, the true homogeneous damping times may be significantly longer. Figure 3.9 shows the frequency domain representation of the overall time-domain model for the oscillatory residuals shown in Figure 3.6. The spectra are dominated by the broad lineshapes from the rapidly damped components over the SO--200-cm‘1 region. The slowly damped component whose lineshape is centered at 2—4 cm“1 is, however, comparable in intensity to the sum of the two rapidly damped components (see Tables 3.2 and 3.3). The other slowly damped components are an order of mag- nitude less intense. Figure 3.10 shows a time-domain representation of the summed rapidly damped components from the models shown in Figure 3.6. Because the effective damping time for both of the rapidly damped components is <250 fs in all of the solvents, a great deal of the amplitude of this component lies in the early-time truncated portion of the pump—probe transient (see Figures 3.5 and 3.6). As extrapolated in Figure 3.10, the waveforms in CH30H and CD3OD are very similar to the rapidly damped vibrational coherence observed previously in polar solutions of BChl, where the intramolecular, slowly damped vibrational coherence is comparatively weak.63' 64 The strong positive-going recurrence at ~200 fs for the waveform in CH30H is only partially observed in the analyzed time region of the residual; a weaker second re- currence at ~600 fs is followed by a broad negative-going trough at ~900 fs that recovers to the baseline over the 1000—1 SOO-fs range. As the solvent is varied, the 67 extent in time of the rapidly damped waveform contracts as the mean frequency, ((0), increases (see Figure 3.10). From the sum of the two underlining component lineshapes (see Figure 3.9), .’M(w) = 11(w) + 12(w), the mean frequency obtained using a normalized mean-value relation, __ 16°de(00)«) - J6°dwN(w) (w) (3.8) shifts over the 79—125-cm"1 range as the solvent is varied (see Table 3.2 and Fig- ure 3.11). As compared to those observed in CH30H, the spectra of the two rapidly damped components in CD30D (see Figure 3.9) are downshifted by S and 8 cm’1 and the mean frequency is downshifted by S cm‘1 (see Table 3.2). 3.5 Discussion The femtosecond pump-probe, dynamic-absorption transients obtained with Soret- band excitation of polar solutions of ZnTMPyP (see Figure 3.5) exhibit a complex modulated signal superimposed on an exponential or double-exponential decay func- tion with a ~2-ps time constant leading to a non-decaying offset over the O-ZS-ps time range. Similar decay components and time constants are attributed to the $2 —> 51 internal-conversion process in pump—probe or fluorescence-upconversion ex- periments with ZnII tetraphenylporphyrin (ZnTPP) in nonpolar solvents,101'102 but for ZnTMPyP in polar solvents the decays arise from vibrational relaxation and vibra- tional energy transfer to the solvent in the 81 state following an ultrafast (Ts lOO-fs) intemal-conversion process. In femtosecond transient-absorption experiments em- ploying continuum probe pulses and 1 SO-fs pump pulses in the Soret band at 404 nm, Fontaine-Aupart and co-workers103 failed to detect an Sg-state stimulated-emission band from ZnTMPyP in aqueous solution, so the Sz-state population relaxes to the Sl-state vibronic manifold almost as rapidly as it is prepared by the pump pulse. In contrast, a distinct stimulated-emission band was observed for ZnTPP under the 68 130+ DMSO 0 120 - 1104 Acetonitrile 100 - 90- Mean Frequency (cm") 80 - Methanol 70- 2 3 4 Dipole Moment (D) Cd _L Figure 3.11. Dependence of the mean frequency of the rapidly damped oscilla- tory components (see Figures 3.6 and 3.9) observed in the pump-probe signals from ZnTMPyP solutions in CH30H, DMF, CH3 CN, and DMSO. The data points are shown superimposed with a trend fitted to Equation 3.12. 69 same optical conditions; the intensity of the stimulated emission decays with a 2-ps time constant as the Sz-state population relaxes to the Sl-state. In ZnTMPyP, pop- ulation deposited in excited vibronic levels of the Sl-state manifold by the $2 —>Sl internal-conversion process is detected in terms of excited-state absorption (ESA) transitions to vibronic levels in the Soret band (states Sn). The time-resolved pump— probe spectra observed by Fontaine-Aupart and co-workers exhibit a broad-region of long-lived (> 5 ps) excited-state absorption at wavelengths above 450 nm; the rel- atively sharp, ground-state depletion band that lies to the blue has a maximum at 435 nm and extends to ~ 405 nm.103 At wavelengths below 435 nm, however, it is likely that the ground-state depletion and BSA bands overlap extensively. Thus, in our experiments, the net ground-state depletion signal decreases in intensity as the Sl-state population cools on the ps time scale because the ESA band shifts to the blue and overlaps more strongly with the static ground-state depletion band at the 42 2-nm probe bandpass. The intensity of the decays we observe in the ZnTMPyP pump-probe signals are solvent dependent, and the decay in two of the solvents (CH3 CN and DMSO) is biexpo- nential. The decays observed for ZnTMPyP in DMSO and CH3OH are not significantly altered (not shown) when the laser is tuned to 425 nm, closer to the Soret band’s max- imum but still on the blue side. These observations show that the observed dynamics are less dependent on the initial vibronic state prepared in the Sz-state manifold than they are on the nature of the solvent. Given that dynamic solvation does not con- tribute to the signal because the 82 state persists for such a short period following absorption of the pump pulse, the rate at which the excess vibrational energy of the S] state is dissipated is solvent dependent. Biexponentiality would arise in this case from the presence of two ensembles of ZnTMPyP with different first-solvation shell structures. A similar explanation was invoked by Zewafl and coworkers to account for the biexponential internal conversion for ZnTPP in methylene chloride.101 70 The conclusion that the 82 state of ZnTMPyP decays to the 81 state via an intemal- conversion process on the s 100-fs time scale is especially significant with respect to the present work because it restricts assignment of the modulation components observed in the pump—probe signals to coherent wavepacket motions on the ground- state potential-energy surface. Table 3.4 shows that all of the slowly damped modu— lation components have effective damping times that are > 1 ps; even the two rapidly damped components have effective damping times that are >150 fs. It is unlikely that the ESA transitions that overlap with the probed region of the spectrum con— tribute to the modulated signal; the Sg—>Sl intemal-conversion process would be expected to be incoherent with respect to the initial vibrational phase prepared by the pump pulses.104 Except for the 2-4-cm‘ 1 component, the modulation components with the largest amplitudes Ai are observed in the ZOO-cm‘1 region of the spectrum. This pattern of amplitudes is consistent with an assignment to coherent wavepacket motions on the ground-state potential energy surface because these experiments employed SO-fs pump pulses. If the ground- and excited-state potential-energy surfaces of the system are displaced with respect to the coordinate of a normal mode of vibration, moving wavepackets on the ground- and excited-state potential surfaces are generated by two successive actions of the the pump field.24'26'33 The ground-state wavepacket is created by a process that directly corresponds to resonance Raman scattering. The duration of the pump pulse controls the displacement and momentum of the ground-state wavepacket because it places a limit on the time interval between the first action of the pump field that prepares the ground—excited-state polarization and the second action that materializes the wavepacket on the ground-state surface. The pump—probe signal is modulated optimally by ground-state wavepacket motion when the pump-pulse duration is about one-third of the mode’s period; with pulses that are shorter than the optimum, the wavepacket is not displaced very far away 71 from the Franck-Condon geometry, so only a small modulation of the transmitted probe signal results. Of course, the depth of modulation of the pump—probe signal decreases as the pump-pulse duration increases beyond the optimum duration to- wards the mode’s period.33 With the SO-fs pulses used in the present experiments, the optimum mode frequency is 222 cm‘l; the relative strength of the ~215-cm’1 component observed in each solvent is probably due to the close matching of the pump—pulse duration with the mode frequency. Even though the measured ampli- tude of the 312—cm“1 mode is only about one-fifth that of the 215-cm‘1 mode (see Table 3.3), the relative resonance Raman activities of the two modes are compara- ble if one considers that the optimum pulse duration that would drive a 312-cm“1 wavepacket is ~32 fs. The unusually large amplitude of the very low frequency, 2—4-cm‘1 modula- tion component would seem to be in conflict with an assignment to a ground-state wavepacket motion. The modulation depth that would be expected with SO-fs pump pulses would be close to zero for such a low frequency mode unless it is charac- terized by an unusually large excited-state displacement. Most of the intramolec- ular vibrational modes of ZnTMPyP would be expected to exhibit comparable dis- placements, with the strongest modes associated with the n-electron density in the porphyrin macrocycle, so the rule-of-thumb about the optimum pulse duration dis- cussed above would reasonably apply.33 An exception might be anticipated for the internal rotation of the peripheral N-methylpyridyl rings (see Figure 3.1). In the ground state, the rings are rotated well away from the the plane of the porphyrin; in the excited state, owing to the flow of the Tr-electron density from the porphyrin re- gion, the minimum-energy geometry would extend the conjugated region by making the rings and porphyrin coplanar. Thus, an internal rotation of the rings is launched in the excited state by resonant excitation of the Soret band because the Sg-state potential-energy surface is sloped in the Franck-Condon region along that coordi- 72 nate; the internal rotation would obtain significant momentum prior to preparation of the ground-state wavepacket on the ground—state surface, where the rotation is bound only owing to the poor delocalization of the n-electron density and owing to steric hindrance. This assignment of the 2—4-cm‘1 mode is also prompted by its significant solvent dependence (see Table 3.3); the highest frequency (4 cm‘l) is ob- served in the most viscous solvent (DMSO), and perhaps the large amplitude in DMF relates to a specific clustering around the rings. The frequencies, linewidths, and relative intensities of the the other slowly damped modulation components are relatively insensitive to the choice of solvent (see Table 3.3). These components are well modeled (see Figure 3.6) by relatively narrow Gaussian components (Equation 3.2). The solvent independence of the mod- ulation frequencies supports a structural assignment of all of these features to the intramolecular, skeletal modes of the ZnH-porphyrin macrocycle. Normal-mode calculations using the B3LYP hybrid density functional for the ZnTMPyP-methanol complex (see Figure 3.1) are consistent with an assignment of the ~40-cm"1 and ~21 S—cm‘1 components to the metal-doming and ZnH—axial-ligand stretching modes, respectively; somewhat higher frequencies are assigned to these modes in Fe“— porphyrins.105 Of these, the former is evidently more solvent sensitive, but it should be kept in mind that the frequencies and intensities of the very lowest frequency components in the slowly damped set of features are known with less confidence than for the higher-frequency or more rapidly damped features. As compared to the amplitude of the rapidly damped components, the slowly damped components are much stronger in ZnTMPyP with Soret-band excitation than observed with Q-band excitation of BChl solution; the resonance Raman intensities of the intramolecular modes of porphyrins are strongly enhanced by excitation in the Soret band.106 The rapidly damped (y < 1 ps) part of the vibrational coherence observed in ZnTMPyP solution exhibits solvent-dependent time-domain waveforms that are very 73 similar to those observed in BChl solution. In the frequency domain, the two rapidly damped components present inhomogeneously broadened lineshapes that are anal- ogous to the ones assigned to the hindered translational and librational modes in instantaneous normal mode analyses of molecular dynamics simulations of polar liquids.96'97i107 A structural assignment of these components to intermolecular modes between the Zn-porphyrin and the clustered solvent molecules in the first sol- vation shell is directly supported by the observation of an isotope shift in methanol and perdeuterated methanol solution (see Table 3.2 and Figure 3.9) of approximately the right magnitude. The mean frequency observed in CH30H is 79 cm'l; in the harmonic-oscillator limit, if the mode involves an intermolecular mode between a sin— gle CH3 OH molecule and the ZnTMPyP molecule, the mode should exhibit a downth to 74.7 cm“1 using the to D = wH(uH/pp)0'5 relationship108 between the frequency w D of the deuterated species and the frequency w H of the protonated species and their reduced masses up and pH, respectively. The observation of a slightly lower mean frequency, 74 cm’l, might indicate the presence of some higher molecular weight species, perhaps from hydrogen-bonded chains,109‘1 12 but the confidence in- terval (:2 cm‘l) is larger than this discrepancy. Consistent with the intermolecular mode assignment, a smaller isotope shift was observed in the presence of CH30D than in CH30H (results not shown); the observed mean frequency of 76 cm"1 in CH3OD is within the confidence interval of the expected value of 77.9 cm-l. As noted in the Introduction, the solvent dependence of the rapidly damped com- ponents (see Table 3.2 and Figure 3.9) provides a test of the form of the van der Waals potential introduced in the previous work on BChl solutions,63 2 2 2 2 V = 12 _ Clull lu2| 0(2lfl1l + 0‘1IH2I 2 6 . 0(r) a/r (19011012 + (4ns0)2 + 47T80 D r (3 9) 74 A relation for the natural frequency of the intermolecular mode in the limit of small displacements from the equilibrium structure, 7/6 Clullzluzl2 0(2|l11|2+0¢1|112|2 7/3 1/2 _ _ + NB (ba1a2+ (4 so)? 4 £0 2 u (3.10) is obtained from a Taylor series expansion of Equation 3.9. Both expressions con— tain a series of terms arising from the physical components of the interaction: the constant at scales the Pauli exchange interaction, the London-dispersion interaction is scaled by the constant [9 = 31112/ 2(11 + I2), with [1,2 representing the ionization potentials for the two molecules, c = 2/3kBT scales the dipole-dipole interaction, and the 0(2IM1 I2 and 0:1 lugl2 terms arise from the solute-dipole and solvent-dipole induced-dipole interactions, respectively.113'116 In these equations, Hi and ori cor- respond to the dipole moment and polarizability for the solute (subscript 1) and solvent (subscript 2); r is the distance between the two molecules, [,1 is the reduced mass for the intermolecular oscillator, D represents the dielectric constant for the solvent, and £0 relates the permittivity of free space. If the solute chromophore is charged, the intermolecular potential gains two ad- ditional terms, for the ion-dipole and ion—induced dipole interactions: _ __ «211le2 erZ 2 4 VQ(r) — V0 ((4Tl'50)2 + 41r50)/2D r (3.11) The ion-dipole term depends linearly on the charge Q of the solute and on the square of the solvent’s dipole moment, pg; the ion—induced-dipole term contains the polar- izability 042 of the solvent and the square of the charge, Q2. Both terms depend on 7‘4, so they decay much less rapidly with increasing distance than the rest of the terms in the intermolecular potential. The mode frequency expression that follows from Equation 3.1 1 is too lengthy to be included here, but it predicts that the charge- dependent terms significantly increase the mode frequency over that predicted by Equation 3.10. 75 Potential Energy (cm'1) I r 2.5 3.0 ‘ I 1.5 2.0 Intermolecular Distance (A) Figure 3.12. Model potential-energy curves for the ZnTMPyP-CH30H complex cal- culated using Equation 3.11 and the experimentally observed mean intermolecular mode frequency, 79 cm'1 (see Table 3.2): (a) Q = O; (b) Q = 1; and (c) the attrac- tive, charge-dependent terms for Q = 1. Table 3.5 lists the other parameters for the potential curves and compares req and me for curves (a) and (b). 76 Table 3.5. Parameters for the model ZnTMPyP-CHgOH intermolecular potentials shown in Figure 3.12 Parametera Value exchange (Pauli) interaction coefficient, a 76.2 eV A12 London-dispersion interaction coefficient, 19 = 3I112 /2(I1 + 12) 9.35 eVb'C ZnTMPyP dipole moment, lull 0.663 Db CH3 OH dipole moment, Iflzl 1.70 DC ZnTMPyP polarizability, a1 109.8 A313 CH30H polarizability, a2 3.29 A3C ZnTMPyP-CH3OH reduced mass, (.1 30.7 Da CH3OH dielectric constant, D 33C req (Q = 0) 1.90 A me (Q = 0) 286 cm-1 Teq (Q = 1) 1.83 A Vmin(Q= 1) 510 cm‘1 ‘3 See Equations 3.9 and 3.11 and the text. b From B3LYP 6-31G electronic structure calculation. C From reference 1 l 7. 77 Figure 3.12 shows a pair of model potentials for the ZnTMPyP-CH3 OH intermolec- ular mode that illustrates the effect of turning on the charge-dependent terms in Equation 3.11 and of changing the intermolecular distance. The only parameter in the intermolecular potential that is not known from the literature, from a calculation, or from experiment is that for the exchange interaction, a, but it can be estimated given an experimental measurement of the mode frequency and an initial assump- tion of Q = 0 so that the charge-dependent terms are inactive. Combination of Equations 3.9 and 3.10 and the observed mean frequency of 79 cm'1 (see Table 3.2) returns a value for a of 76.2 eV A12; the rest of the parameters used in Figure 3.12 are listed in Table 3.5 along with estimates for the equilibrium bond distance, Teq, and the bond-dissociation energy, me, with Q = O and Q = 1. If a is held constant, the effect of changing Q from 0 to 1 is a sharpening of the potential near the mini- mum, which indicates an increased mode frequency, almost a doubling of me, and a modest 0.07 A reduction in req (compare curves (a) and (b) in Figure 3.12). Curve (c) in Figure 3.12 shows the magnitude of the charge-dependent terms in Equation 3.1 1, as calculated for the Q = 1 potential. This curve can be used to gauge the contribu- tion of the charged terms to the stabilization and mode frequency when the charges are located at some distance from the clustered solvent molecules. The ones that interact with the n-electron density in the ground state are those that contribute to the ground-state vibrational coherence, and on average these solvent molecules are constrained to be more than 2.5 A from the charges on the N-methylpyridyl rings. This distance estimate is based on the assumption that the charge is localized on the nearest carbon atom of the N -methylpyridyl ring and that the solvent molecule is fixed at Teq for the Q = 0 case normal to the porphyrin ring above the adjacent meso carbon atom, to which the N-methylpyridyl rings are attached. At this distance, the charge-dependent terms account for a 70-cm‘1 stabilization, about 25% of me for Q = 0. Given that the charge is actually delocalized over the N-methylpyridyl ring in 78 the ground state, the integrated interaction energy will be considerably smaller than this worst-case estimate. We conclude that it is a good approximation to neglect the charge-dependent terms in the ground-state intermolecular potential. In the excited state, because the n-electron density is extensively delocalized from the porphyrin region onto the N-methylpyridyl rings, the intermolecular mode frequency should be significantly increased from that in the ground state owing to a larger contribution of the charge-dependent terms in Equation 3.1 1. The expression for the intermolecular mode frequency in the absence of charge- dependent terms, Equation 3.10, can be simplified to give a relation for the depen- dence on the dipole moment of the solvent as v = (C1 + C2|u2|2)7/6 (3.12) where C1 and C2 are constants. This expression follows directly from Equation 3.10 with the assumption that all of the parameters other than the dipole moment of the solvent molecule are fixed; a simpler, approximate form was used in the previ- ous work on BChl solutions.63 Figure 3.11 plots the mean frequency of the rapidly damped components observed in the ZnTMPyP solutions as a function of the gas- phase dipole moment of the solvent. The mode frequency increases as the dipole moment increases in a manner that is consistent with the form of Equation 3.12. The fit of the mode frequencies to Equation 3.12 extrapolates to a mode frequency near 70 cm“1 in the nonpolar limit. As anticipated in the Introduction, this y-intercept value is significantly lower than the value of ~100 cm“1 observed in BChl solution. ZnTMPyP’s four-fold symmetry results in a dipole moment of zero in the plane of the porphyrin; the out-of-plane doming of the ZnII ion due to the axial coordination of a solvent ligand results in a small dipole-moment component orthogonal to the plane of the porphyrin (0.66 D, as obtained from the B3LYP 6-31G electronic struc- ture calculation for the methanol complex, see Figure 3.1). The dipole moment of BChl is significantly larger (3.6 D,118 so the 0(2IM1 I2 induction‘term in Equations 3.9 79 and 3.10 that contributes to the y-intercept constant C1 in Equation 3.12 is more than 25 times smaller for ZnTMPyP since this term scales as the square of the dipole moment. The remaining term in Equations 3.9 and 3.10 that contributes to the y- intercept term of Equation 3.12 is the London-dispersion term, boq 0:2. Thus, a great deal of the residual 70 cm‘1 of the y-intercept value obtained for ZnTMPyP is due to the London-dispersion term. This conclusion is consistent with the general ob- servation that the dipole—dipole and London-dispersion terms in the intermolecular potential are perhaps 10 times larger than either of the induction terms.113‘116 Some of the molecular details of the interaction between ZnTMPyP and the sur- rounding solvent are suggested by deviations from the fitted curve shown in Fig- ure 3.11 and by the dependence of the resonance Raman activity for the intermolec- ular mode on the solvent. The reduction of Equation 3.10 to Equation 3.12 requires that the polarizability 012 for the solvent molecule is constant as the dipole moment ug varies. The dispersion of the points for the solvents DMSO, CH3CN, and DMF from the fitted curve in Figure 3.11 suggests a possible contribution of the solvent polarizability in the strength of the intermolecular interaction with the ZnTMPyP so- lute. These solvents have similar dipole moments, yet the mean frequency of the intermolecular mode varies in them over the 95-125-cm‘1 range, a span of about one-fifth of the mean value. The increase of the mean frequency from that observed in methanol (a2 = 3.29 A3 117), 79 cm‘l, to that for acetonitrile (0:2 = 4.4 A3 117), 106 cm’l, depends mostly on the induction term 011 |fl2|2 since the dipole—dipole term C|M1 [2 lug |2 is relatively small owing to the small dipole moment of ZnTMPyP. The induction term is relatively large in porphyrin and BChl solutions because of the large polarizabilities (0(1) of these solutes. The difference between the mean fre- quency observed in acetonitrile and DMSO (0:2 = 8.87 A3 119), however, probably arises largely from the enhanced polarizability of the sulfonyl group over that of the carbonyl. Further, the decreased mean frequency in DMF (a2 = 7.81 A3 117) has 80 to be attributed to a specific solvent effect that decreases the strength of the inter- molecular interaction given that the polarizability of DMF is almost as large as that of DMSO. The solvent dependence of the amplitude of the rapidly damped modulation com- ponents (see Table 3.2) suggests that the resonance Raman intensity enhancement for the intermolecular mode partly depends on the covalency of the interaction, the degree to which the Tr-electron density of the solute chromophore is mixed with that of the clustered solvent molecule. The use of a van der Waals potential to follow trends in the mode frequency invokes a purely electrostatic picture, of course, but some involvement of covalency is suggested by the observation that the strongest modulation amplitudes occur in CH3CN and DMSO, which feature cyanide/nitrile and sulfonyl groups, respectively. This observation follows the same trend observed with Q-band excitation of polar solutions of BChl; the amplitude of the intermolecular vi- brational coherence was strongest by far in pyridine solution.63 As mentioned above, DMF seems to be an outlier; the intermolecular modes in DMF and CH3 OH have com- parable amplitudes even though the former contains delocalized n-electron density over the amide group. It follows that both the Albrecht A- and B-termslzov121 con— tribute to the resonance Raman intensity. The A-term would primarily include effects from the change of shape and extent of the n-electron density that yield a gradient in the excited-state potential-energy surface in the Franck-Condon region, whereas the B-term would include couplings of the excited state to nearby states. The part of the A-term enhancement that arises from the redistribution of n-electron den- sity in ZnTMPyP’s excited state (see Figure 3.1) is likely to be small given that the intermolecular modes are easily observed in BChl, which lacks a comparable ground- to-excited-state displacement of electron density. 81 3.5.1 Conclusions The results discussed in this chapter support the hypothesis that intermolecular modes between porphyrin or chlorophyll macrocycles and clustered molecules in the first shell of solvent contribute the most intense components to the low-frequency (0-300-cm‘1) vibrational coherence in polar media. The corresponding mode fre- quencies fall in the right range to account for the modes that dominate the Marcus reorganization energy for electron-transfer reactions in photosynthetic reaction cen- ters. In comparison to the intermolecular modes, the intramolecular modes from the porphyrin or chlorophyll macrocycle account for features in the vibrational co- herence that are at least an order of magnitude weaker. Further, an analysis of the electronic structure of ZnTMPyP suggests that there should be a significant ground- to-excited-state change in the mode frequency and stabilization of the intermolecular modes owing to the change in shape of the Tr-electron density. The contribution of ion-dipole and ion-induced-dipole terms to the intermolecular potential in the ex- cited state of ZnTMPyP should result in a significantly higher intermolecular mode frequency. These projections suggest an important role for intermolecular modes between the prosthetic groups and their first-solvation-shell surroundings in the sta- bilization of the charged intermediate and product species that occur in proteins during redox catalysis. 82 CHAPTER 4 Excited-state Vibrational coherence in methanol solution of ZnII tetrakis(N-methylpyridyl)porphyrin 4.1 Summary We have employed femtosecond dynamic-absorption spectroscopy to observe excited-state wavepacket motions in ZnII meso-tetrakis(N-methylpyridyl)porphyrin (ZnTMPyP) in methanol solution. The pump-probe transients observed in ZnTMPyP solution with Q-band excitation exhibit a net excited-state absorption character. The signals are weakly modulated by slowly damped (effective damping time y>1.5 ps) components that correspond to skeletal normal modes of the porphyrin macrocycle, but by far the dominant modulation features are a pair of rapidly damped (y<400 fs) components arising from intermolecular modes with first-shell solvent molecules. The 256 cm‘1 mean frequency of the rapidly damped components is significantly higher than observed previously in the ground state, 79 cm‘l. This increase in fre- quency arises from an extensive delocalization of the n-electron density from the 83 porphyrin region to the singly charged N-methylpyridyl rings. In the excited state, the van der Waals intermolecular potential accordingly obtains large ion-dipole and ion—induced-dipole terms that account for the increased mode frequency and result in a significant stabilization of the equilibrium structure. These results suggest an especially important functional role for the intermolecular modes between the re- dox chromophores and the surrounding protein medium in photosynthetic reaction centers that involves trapping of charged intermediates. 4.2 Introduction The observation of vibrational coherence in femtosecond pump—probe ex- periments with impulsive excitation of electronic chromophores in solu- tion” 26' 29. 33' 35' 122' 123 is usually associated with resonance Raman activity arising from an excited-state displacement of a small number of skeletal normal modes of the chromophore. In several classes of proteins where light-induced chemistry oc- curs, vibrational coherence in the skeletal modes of a prosthetic group is associated with the reaction coordinate. Shank, Mathies, and coworkers characterized coher- ent wavepacket motion in retinal’s torsional modes associated with formation of the photoisomerization product state in bacteriorhodopsin and rhodopsin.26' 32' 35' 3942 Champion and coworkers observed an analogous reaction-driven vibrational co- herence in the ground-state out-of—plane porphyrin modes in myoglobin and cy- tochrome c following photodissociation of axial ligands.53'55-61 Martin, Vos, and coworker8124 have observed vibrational coherence in cytochrome c oxidase aa3 upon photolysis and transfer of carbon monoxide from heme £13 to CuB. The excited-state vibrational coherence observed by Vos and Martin from the primary electron donor, P, in the purple-bacterial photosynthetic reaction center,4 3'51 however, is likely to in- volve modes that are derived from the surrounding protein mediumlv 22 rather than 84 just from the paired bacteriochlorophyll (BChl) chromophore itself.73'125'126 Be- cause of their low frequencies, the modes would be expected to be delocalized over the protein structure spanning the reaction center’s chromophores just as phonons are in molecular crystals.43v45’ 50 In the previous chapters, we discussed how the ground-state vibra- tional coherence observed in solutions of BChl and of Zn“ mesa-tetrakis(N- methylpyridyl)porphyrin (ZnTMPyP) suggest, in contrast, that by far the strongest contributions to the vibrational coherence in the 100-cm‘1 regime arise from local- ized intermolecular modes with clustered, first-shell solvent molecules.63‘65v 88 The pump—probe signals observed with Q-band excitation of BChl in polar solvents are modulated by rapidly damped (time constant y<200 fs) components that shift to higher frequencies as the dipole moment of the solvent increases. The mean fre- quency of these modulation components follows the natural frequency of a van der Waals intermolecular potential that contains dominant contributions from the London-dispersion and dipole-dipole interactions.63 The slowly damped (3122000 fs) modulation components that arise from the skeletal modes of the BChl macrocycle are perhaps ten times less intense than those from the solvent modes.54 Similar as- signments were made for the ground-state vibrational coherence observed in polar solutions of ZnTMPyP with Soret—band (420-nm) excitation. The normal modes of the porphyrin macrocycle and its single solvent-derived axial ligand contribute very slowly damped features to the vibrational coherence, some of which persist even to the 20-ps delay point. The rapidly damped, solvent-dependent modes again account for the dominant features. The dependence of the mean frequency of the rapidly damped components follows an intermolecular potential that exhibits a lower mode frequency in the nonpolar limit than observed in BChl (70 cm"1 and 120 cm‘l, respectively), and an isotope-dependent shift was observed when the signals from methanol and perdeuterated methanol were compared. The magnitude of the shift 85 indicates that the intermolecular modes arise in methanol solution predominantly from 1:1 complexes with the porphyrin solute.88 The intermolecular potential used in the discussion of the ground-state vibra- tional coherence in BChl and ZnTMPyP solutions includes only the terms that arise from the polarizability and dipole moment of the neutral solute and solvent. In ZnTMPyP, the Tr-electron density in the ground state is confined to the porphyrin macrocycle (see Figure 3.1a), so an attacking solvent molecule does not strongly sense the positive charges on the peripheral N-methylpyridyl rings. In the 11* excited states (see Figure 3.1b), however, the Tr-electron density is extensively delocalized from the porphyrin over two of the N-methylpyridyl rings, so the intermolecular potential between ZnTMPyP and clustered solvents should gain ion-dipole and ion—induced- dipole terms. These charge-dependent terms would be expected to contribute to a significant shift to higher frequency for the intermolecular oscillator. We have put these ideas to the test in this chapter. We show that pump-probe signals obtained with Q-band excitation of ZnTMPyP monitor excited-state wavepacket motions owing to detection of a net excited-state absorption signal. As expected, the rapidly damped portion of the vibrational coherence shifts to a much higher frequency than observed in the ground-state vibrational coherence observed with Soret-band excitation. These results advance the hypothesis that resonance Raman-active intermolecular modes with first-solvation-shell components contribute dominantly to the low-frequency vi- brational coherence in porphyrin and BChl systems. We conclude with a brief discus- sion of how the charge-dependent terms in the intermolecular potential should play an important role in the trapping of charge-separated intermediates in the photosyn- thetic reaction center. 86 4.3 Experimental 4.3. 1 Sample Preparation ZnTMPyP (CAS 28850-44-4) was used as received from Frontier Scientific. Methanol (CH3OH, spectrophotometric grade) was obtained from Sigma-Aldrich. For use in femtosecond pump-probe experiments, solutions of ZnTMPyP were prepared by dis- solving the dry ZnTMPyP powder in methanol to obtain an absorbance of 0.4 for a path length of 1.0 mm at the center of the laser spectrum at 625 nm, as detailed below. The solution was passed through a 0.22-m filter prior to checking the ab- sorbance. The samples were held in the femtosecond pump-probe spectrometer at room temperature (23 °C) in a fused-silica flow cuvette (0.5-mm path length). A peri- staltic pump was used to circulate a 10-mL reservoir of sample solution through the cuvette at 2.70 mL/min. The sample’s absorption spectrum was monitored during the experiment for changes arising from photochemistry or permanent photobleach- ing. The sample reservoir was exchanged with fresh solution several times during each run. 4.3.2 Continuous-Wave Absorption and Fluorescence Spectroscopy Absorption spectra were obtained at 23 °C with a Hitachi U-2000 spectrophotometer (2-nm bandpass). Fluorescence spectra were acquired at 23 °C with a Hitachi F-4SOO spectrofluorimeter (S-nm bandpass for the excitation and emission monochroma- tors). As presented as a function of wavenumber, the fluorescence intensities are multiplied by the square of the wavelength in order to compensate for the fixed (in wavelength units) spectral bandpass of the emission spectrometer.89 87 4.3.3 Femtosecond spectroscopy Femtosecond pump-probe transients with impulsive excitation were recorded using the dynamic—absorption technique, in which the probe beam is dispersed in a grating monochromator after passing through the sample.32v 33’ 35‘37 Most of the instrumen- tation and methodology is similar to that described in the previous work on ZnTMPyP solutions with Soret-band excitation.88 In this work, however, the pump and probe pulses were obtained from the signal-beam output of an optical parametric amplifier (Coherent CPA 9450), which was pumped by an amplified Ti:sapphire laser (Coher- ent Mira-seed oscillator and a modified Coherent RegA 9050 regenerative amplifier, with Coherent Verdi V5 and V10 pump lasers, respectively). The laser was operated at a repetition rate of 250 kHz. The experiments were conducted with 13-nJ, 45-fs pump and 7.2-nJ probe pulses centered at 625 nm (15 nm fwhm, as measured with an Ocean Optics USB-2000 spectrometer/CCD detector with a 0.5-nm bandpass). The pump and probe pulses were corrected for group-delay dispersion on the way to the sample by a SF 10 Brewster prism-pair pulse compressor. The pump-probe time delay was scanned using a rapid-scanning delay stage (Clark-MXR, ODL-150) in a modified Mach-Zehnder interferometer with confocal sample and autocorrelation- crystal positions. Calcite polarizers and A/2-retarding wave plates in the pump and probe beams set their planes of polarization at 90°; after passing through the sample, the probe beam was analyzed by another calcite polarizer oriented 90° relative to the pump-beam’s plane of polarization, and then it was passed through a monochroma- tor (Spex 270M, 4-nm bandpass) and detected by an amplified photodiode (Thorlabs PDA55). The monochromator’s slits were adjusted to obtain a fairly narrow band- pass (4 nm) compared to the width of the laser’s spectrum; the monochromator’s grating was detuned from the laser’s intensity maximum to the ~ 50% intensity point. This approach is similar to that used by Champion and co-workers in their studies of low-frequency vibrational coherence in heme proteins.53’55’61 The pump—probe 88 signal was obtained from the photodiode signal using a lock-in amplifier (Femto LIA- MV-200-H); the pump beam was modulated at 50 kHz by a A/4-retarding photoelastic modulator (Hinds Instrumentation) with A/4-retarding and A/2-retarding waveplates and a calcite polarizer in series. 4.4 Results Figure 4.1 shows continuous-wave absorption and fluorescence spectra from ZnTMPyP in methanol at 23 °C. The spectra are plotted as relative dipole strengths89'91'92 as a function of wavenumber v, A(v) /v and F(v) /v3 respectively. The fluorescence spectrum extends to the red from the 0-0 peak of the Q band, the absorption feature that corresponds to the first excited singlet (81) state. Also shown in Figure 4.1 is the output spectrum from the OPA as it was tuned to 625 nm for the dynamic-absorption experiments. Because we wanted to detect vibrational co- herence from coherent wavepacket motion on the 81 state’s potential-energy surface, the OPA was tuned as far to the red in the Q band as possible so that the probe bandpass would detect a large stimulated-emission signal. As estimated from the dipole-strength spectra, at the 632-nm center of the probe bandpass the stimulated- emission (SE) signal is five times larger than the photobleaching (PB) signal. The dynarnic-absorpn'on transient obtained from ZnTMPyP in methanol under these experimental conditions is shown in Figure 4.2. Following an intense bipolar spike that goes off the plotted scale near the pump-probe zero-delay point,93‘95 the transient exhibits a pattern of cosinusoidal modulations superimposed on a rising transient of net excited-state absorption (ESA) character. The oscillatory residual ob- tained by subtracting a fitted biexponential trend consists of a very rapidly damped portion over the 100-600-fs range followed by a more slowly damped part that per- sists at least to the 2000-f8 range. This damping character is similar to that observed 89 A/v Relative Magnitude l I I I I I 12 14 16 18 20 22 24 26 Frequency (103 cm") Figure 4.1. Soret (v>20000 cm‘l) and Q-band region of the continuous-wave absorp- tion (solid curve) and fluorescence (dotted curve) spectra from ZnTMPyP in methanol at room temperature (23 °C), plotted as the dipole strength, A(v)/v and F(v)/v3, respectively, and normalized to unit area for the fluorescence spectrum and Q band. Superimposed with arbitrary scaling is the intensity spectrum of the 62 5-nm, 45-fs pulses used in the pump-probe experiment. 90 -O.8 - —1.0- I r l l l -300 0 300 600 900 1200 1500 1800 Time (is) Figure 4.2. Femtosecond pump—probe dynamic absorption transient detected at 632 nm (4-nm bandpass) with ZnTMPyP in methanol. The pump-induced change in trans- mission signal (AT /T) is shown superimposed on a double-exponential fit function (dashed curve) over the > 100-fs delay range of the form A0(1 +A1e‘t/Tl +A2e“/T2 ), with A0 = 3.00 x 10—3, A1 = 0.204, T1 = 98 fs, A2 = 0.272, and 72 = 1.8 ps. As plotted, the signal is normalized by dividing by the nondecaying fraction, A0. The oscillatory residual, the difference between the signal and the fit function, is shown above the signal superimposed on a model composed of slowly and rapidly damped oscillatory components (see Figure 4.3 and Tables 4.1 and 4.2). 91 in polar solutions of BChl and in polar solutions of ZnTMPyP with Soret-band excita- tion.63'88 After truncation of the <100-fs portion to avoid consideration of the spike near time zero,95 the oscillatory residual was fit in the time domain to a model consisting of a sum of damped cosinusoids of the form _ 2 2 Ii(t) = Aie t 01' /2 cos(w0it — 49,-) /\/21‘r (4.1) with each component 1' having a center frequency (00,- and phase 43,-. These wave- forms correspond to Gaussian line shapes in the frequency domain, _ _ . 2 2 1,-(00) = Aie (w (”01) “201' ) Hap/217) (4.2) where the linewidth is controlled by the standard deviation, 0,- = Awi/ 2m, with Awi representing the full width at half maximum. Equations 4.1 and 4.2 are nor- malized so that the amplitude Ai corresponds to the area of the lineshape in the frequency domain and to the intensity of the signal in the time domain. Except as noted in the following, the fitting procedure was described-in detail in the chapter on the ground-state vibrational coherence from ZnTMPyP.88 The fit is shown super- imposed on the residual’s data points in Figure 4.2. ' Starting parameters for the slowly damped components (Aw<10 cm‘l) were ob- tained from a Hanning-windowed Fourier-transform spectrum (see Figure 4.3a). The window function suppresses the rapidly damped portion of the signal and applies linebroadening to the slowly damped components. The twelve most intense peaks in the Fourier-transform spectrum were included in the model. The fit parameters for the slowly damped components are listed in Table 4.1. The remaining, rapidly damped (Aw>10 cm‘l) part of the oscillatory residual was then modeled as the sum of two broad Gaussian components. The fit was not improved by inclusion in the lineshape of an asymmetry or skew parameter, 92 A o .. B 8 0 - B 'E C 8 2 a) .2 c_"6 0 6‘6 D E I ' ] I | ' I ' I ' 0 100 200 300 400 500 Frequency (cm'1) Figure 4.3. Comparison of the intensities and lineshapes observed in the excited- state and ground-state vibrational coherence from ZnTMPyP in methanol. Ex- cited state, from the 62 5-nm pump—probe signal, see Figure 4.2: (a) Hanning- windowed, Fourier-transform magnitude spectrum; spectra for (b) the slowly damped (Aw<10 cm‘l) and (c) the rapidly damped (Aw>10 cm‘l) components from the time-domain model. Ground state, from the 420-nm pump—probe signal, from ref- erence 88): spectra for (d) the slowly damped (Aw<10 cm'l) and (e) the rapidly damped (Aw> 10 cm‘l) components from the time-domain model. In (c) and (e), the sum of the two rapidly damped components (dotted lines) is shown as the solid trace. The scaling factor for (d) and (e) is 1.6 times smaller than that for (b) and (c). 93 Table 4.1. Gaussian lineshape parameters for the slowly damped modula- tion components obtained from the 62 S-nm pump-probe transient from ZnTMPyP in methanol (see Figure 4.2). Component wo (cm—1) Aw (cm‘l) Amplitudea 1 15 2.55 2.84 2 37 2.36 1.00 3 63 2.35 0.507 4 94 9.41 0.776 5 112 9.41 0.761 6 157 2.77 1.42 7 172 2.35 0.597 8 195 4.61 0.597 9 216 2.35 0.313 10 255 2.35 0.373 11 313 4.66 0.358 12 346 9.04 2.54 3‘ Normalized relative to component 2, which has an amplitude of 6.72 x 10'3 relative to the fitted nondecaying fraction, A0 (see Figure 4.2). 94 Table 4.2. Gaussian lineshape parameters and effective exponential damping times for the rapidly damped modulation components obtained from the 62 S-nm pump- probe transient of ZnTMPyP in methanol (see Figure 4.2). Component wo (cm‘ 1) Aw (cm‘ 1) y (fs)a AmplitudeID l 220 44.0 309 1 7.9 2 285 70.8 193 22.4 a Damping time of best-fit exponentially damped cosinusoid (see Equation 4.3). b Normalized relative to the 37-cm‘1 slowly damped component (see Table 4.1). 95 as was previously required in the modeling of the rapidly damped ground-state com- ponents.88 The two components are assigned, in order of frequency, to the hindered translational and librational (hindered rotational) intermolecular modes between the porphyrin macrocycle and its clustered, first-solvation-shell solvent molecules.63'88 The fit parameters for these components are listed in Table 4.2. Also included in Ta- ble 4.2 are effective damping times y for a best-fit exponentially damped cosinusoid of the form Ii(t) = Ai e‘t/yi cos(w0,-t — 4),) (4.3) which corresponds in the frequency domain to a Lorentzian line shape. The tabulated values of )7 were obtained as previously described by fitting each Gaussian compo- nent in the model individually to Equation 4.3. This procedure provides a rough estimate of the lower limit for the damping time.88 Figures 4.3b and c show frequency domain representations of the slowly and rapidly damped portions of the optimized model. These spectra should be com- pared to Figures 4.3d and e, which show spectra for the slowly and rapidly damped components in the ground-state vibrational coherence.88 The rapidly damped com- ponents in Figure 4.3c are clearly shifted to a much higher frequency from those in Figure 4.3c. From the sum of the two underlining component lineshapes (see Fig- ure 4.3c and e), Mu») = I1(w) + 12(w), the mean frequency was obtained using a normalized mean-value relation, _ 16° dw .’M(w)w (w) _ 16° dwjmw) (4.4) The mean frequency of the sum of the two rapidly damped components shown in Figure 4.3c is 256 cm‘l, whereas the mean frequency for those in Figure 4.3e is 79 cm-l. 96 4.5 Discussion The pump-probe signal from ZnTMPyP with Q—band excitation at 625 nm is modu- lated primarily by coherent wavepacket motions on the 31 -state potential-energy sur- face. This assignment is based on the tuning of the laser spectrum, the choice of the detected probe bandwidth (see Figure 4.1), and the observation of a rising net ESA sig- nal that persists to the end of the recording. This result is consistent with the obser- vations of Fontaine-Aupart and coworkers,103 who observed a long-lived ESA signal at probe wavelengths above 575 nm in their pump-continuum-probe time-resolved spectra. The laser was tuned so that the red side of the laser spectrum, where the probe bandpass was selected, overlaps favorably with the stimulated-emission spec- trum (see Figure 4.1). Owing to the 5:1 SEzPB ratio, as judged from the ratio of the continuous-wave fluorescence and absorption dipole strengths centered at the de- tected probe bandpass, respectively, and because the strength of the ESA part of the signal is larger than of the sum of the photobleaching and stimulated-emission parts, we estimate that ground-state wavepacket motions make less than a 10% contribu- tion to the pump—probe signal. The SE and BSA signals are synchronously modulated by the Sl-state wavepacket. The pattern of modulation intensities observed in Figure 4.3a and b is fully consis- tent with an assignment to excited-state coherent wavepacket motions. These slowly damped (y> 1.5 ps, Aw <10 cm’l) features are assigned to the skeletal normal modes of the ZnTMPyP molecule. In systems where the ground- and excited-state potential- energy surfaces are displaced with respect to the coordinate of the vibrational mode, moving wavepackets on the ground- and excited-state potential surfaces are launched by two successive actions of the pump field.24'26t 33 The ground-state wavepacket is created by a process that directly corresponds to resonance Raman scattering. The pump-probe signal is modulated optimally by ground-state wavepacket motion when the pump-pulse duration is about one-third of the mode’s period.26 As we explained 97 previously, the ~222-cm‘1 maximum observed in the intensity profile of the slowly damped features detected with Soret-band excitation (see Figure 4.3d) is consistent with an assignment to ground-state wavepacket motions because of our use of 50-fs pump pulses in that experiment.88 In contrast, for excited-state wavepacket motion, the amplitudes are expected to increase monotonically as the frequency decreases.26 The intensity profile indicated by Figure 4.3b is obviously consistent with this trend. The main exception to the expectation that the intensities for the excited-state wavepacket motions should decrease smoothly with increasing mode frequency is the relatively intense 346-cm"1 peak, which is assigned to an an out-of—plane defor- mation mode localized on the N-methylpyridyl rings based on assignments of peaks at similar frequencies in continuous-wave resonance Raman spectra of ZnTMPyP and other N-methylpyridyl porphyrins.127 The strength of the 346-cm‘1 peak is es- pecially notable considering our use of 45-fs pump and probe pulses, which lack enough impulsive bandwidth to make observation of modulation features above 300 cm"1 facile. Note also that the relatively intense 3-cm'l component observed in the ground state vibrational coherence (Figure 4.3d) but apparently absent from the excited-state vibrational coherence (Figure 4.3b) arises from internal rotation of the N-methylpyridyl rings with respect to the porphyrin. In the ground state, this motion is relatively unhindered because the n-electron density is confined to the porphyrin region (see Figure 3.1a); in the excited state, the n-electron density is extensively delocalized from the porphyrin to the rings (see Figure 3.1b), so the resulting partial double-bond character strongly clamps internal rotation. Except for the 3-cm‘1 and 346.cm'1 peaks discussed above, the frequencies of the components observed in the slowly damped excited-state vibrational coherence (Figure 4.3b) correlate well with those observed in the ground-state vibrational co- herence signal (Figure 4.3d). Most of these modes arise from out-of-plane defor- mations of the porphyrin macrocycle. The 37-cm‘1 and 216-cm‘1 components are 98 assigned to the metal-doming and Znn-axial-ligand stretching modes, respectively. The excited-state mode frequencies for these components differ by less than one cm‘1 from the frequencies observed in the ground state.88 The rapidly damped (y<400 fs, Aw>15 cm‘l) features shown in Figure 4.3d and e correspond to the intermolecular modes between the ZnTMPyP molecule and first-shell solvent molecules. In the excited state, the mean frequency of the summed intensity of these features is 256 cm‘l. This frequency is substantially higher than the ground-state mean frequency of 79 cm’l. This increase is consistent with the van der Waals intermolecular potential we introduced previously in our discussion of the trend of the ground-state frequency as a function of the dipole-moment of the solvent molecule.88 In the ground state of ZnTMPyP, because the 1t-electron density is confined to the porphyrin ring (see Figure 3.1), the intermolecular potential can be expressed as 2 2 2 2 = 12_ Clull |u2| 0(2IM1I +0t1|u2| 2 6 V0(r) a/r (bail 0:2 + (41TEo)2 + 4Tf€0 D r (4.5) As shown previously,63 a Taylor series expansion of Equation 4.5 around the equi- librium geometry gives an expression for the natural frequency of the intermolecular mode, 7/6 3 Clll1|2|uz|2 “2111112 +0(1|I12|2 7/3 1/2 = __ b 4.6 v (2a)2/3( 011012+ (411.50)2 + 41r£0 2WD l1 ( ) In the above equations, the constant a scales the Pauli exchange interaction, 19 scales the London-dispersion interaction, c = 2 / 3 kBT scales the dipole—dipole interactions, and the (x2 llll |2 and a1|u2|2 terms correspond to dipole-induced-dipole interac- tions (see Table 4.3).113‘116 In these equations, [1,- and 01,- correspond to the dipole moment and polarizability for the solute (subscript 1) and solvent (subscript 2); r is the distance between the two molecules, 11 is the reduced mass for the intermolecular oscillator, D represents the dielectric constant for the solvent, and £0 relates the per- mittivity of free space. As expressed above, the dipole-dependent terms apply to the 99 high-temperature limit, where k ET is much greater than the well depth, so the inter- action is averaged over all orientations.115 Since the well depth for ZnTMPyP—CH3 OH in the ground state (see Table 4.3 and reference 88) is 286 cm‘1 at T=298 K, the as- sumption that it is appropriate to average the interaction energy over all orientations may not be valid. In the excited state, the Tr-electron density is extensively delocalized from the porphyrin region to the N-methylpyridyl rings (see Figure 3.1), so the formal charges on the rings are also delocalized. As noted previously,88 the intermolecular potential should be expanded to include two additional solute-charge-dependent terms corre- sponding to the ion-dipole and ion-induced dipole interactions with the charge Q: _ CQIMZI2 0'szZ 2 4 VQm _ v0 —- (--———-( 41160)2 + ——4mo) / 21) r (4.7) Both of the new terms are attractive; the ion-dipole term depends linearly on the charge Q of the solute and on the square of the solvent’s dipole moment, [12; the ion— induced-dipole term contains the polarizability 012 of the solvent and the square of the charge, Q2. Both terms depend on r‘4, so the potential expands to much larger intermolecular distances than does the neutral potential (Equation 4.5). Thus, inclu- sion of the charge-dependent terms causes the potential to extend to much greater intermolecular distances than does the neutral potential. The relation for the nat- ural mode frequency that follows from Equation 4.7 is not shown here owing to its length, but it predicts that the mode frequency should significantly increase over that predicted by Equations 4.5 and 4.6. Equation 4.7 also predicts that the depth of the potential well will increase significantly over that predicted from the neutral potential (Equation 4.5), so the assumption of orientational averaging made for the uncharged case is no longer valid, and the result would be an even more stable structure for the porphyrin-solvent van der Waals complex. Figure 4.4 shows three model curves for the ZnT'MPyP—CH3OH intermolecular mode to illustrate the effect of adding charges to the intermolecular potential. Most 100 of the parameters in Figure 4.5 and 4.7 are obtained from the literature (see Ta- ble 4.3). The coefficient of the exchange interaction, a, can be estimated given the experimental measurement of the ground-state mode frequency (79 cm'l) and the assumption that Q = 0 in the ground state so that the charge-dependent terms do not contribute to the intermolecular potential. In the excited state, the Tr-electron density extends over two of the N-methylpyridyl rings, so it is reasonable to sup- pose that two charges are introduced. As Q increases from 0 to 2, the potential-well depth me increases by almost a factor of five but the equilibrium bond length req experiences only a ten percent contraction (0.18 A). The potential well sharpens sig- nificantly as Q increases, so the intermolecular mode will accordingly increase in frequency. The predicted mode frequency for the Q = 2 case obtained from Equa- tion 4.7 is 168 cm"1, which is about two-thirds of the mean frequency determined from the rapidly damped components shown in Figure 4.3c, 256 cm‘l. A better prediction of the excited-state mode frequency would be obtained from a improved potential that includes explicitly the excited-state charge distribution and employs a molecular dynamics simulation to gauge the distribution of solvent-solute interac- tion distances and orientations. 4.5.1 Conclusions The ground-to-excited-state increase in the frequency of the intermolecular mode between ZnTMPyP and its first-shell solvent molecules allows firm conclusions about the nature of the intermolecular potential and the nature of the coupling of the mode to the 17 —> 17* transition. First, the magnitude of the increase in the mode frequency shows the importance of charge-dependent terms in the intermolecular potential rel- ative to the nonpolar dipole-dependent terms. The effective absence of the charge- dependent terms in the intermolecular potential detected in the ground state further suggests that the solvent molecules that contribute to the detected intermolecular 101 250- A 0 E, —250 4 3 oh) "500 -' C ’0'... c .4" E —750 4 g, —1000 J c? —1250 .- —1500 — , 1.5 2.0 2.5 3.0 Intermolecular Distance (A) Figure 4.4. Model potential-energy curves for the ZnTMPyP-CH3OH van der Waals complex as a function of the charge Q on the ZnTMPyP moiety: (a) Q = 0; (b) Q = 1; (c) the attractive, charge-dependent terms for Q = 1; (d) Q = 2 and (e) the attrac- tive, charge-dependent terms for Q = 2. The curves were calculated using Equa- tion 3.10 and the experimentally observed ground-state intermolecular mode fre- quency, 79 cm"1 (see reference 88). Table 4.3 lists the parameters and compares req and me for curves (a), (b), and (d). 102 Table 4.3. Parameters for ZnTMPyP-CH30H intermolecular potentials (see Fig- ure 4.4) Parametersa Symbol Value exchange (Pauli) interaction coefficient a 76.2 eV A12 London-dispersion interaction coefficient 17 9.35 eVb'C ZnTMPyP dipole moment in 0.663 Db CH3OH dipole moment 112 1.70ID ZnTMPyP polarizability 0(1 109.8 A3 b CH3OH polarizability ‘ 012 3.29 A3 C CH3OH dielectric constant D 33C Equilibrium geometry with Q = 0 Teq 1.90 A me 286 cm‘1 Equilibrium geometry with Q = 1 req 1.83 A me 510 cm-1 Equilibrium geometry with Q = 2 req 1.72 A vmm 1336 cm-1 a See Equations 4.5-4.7 and the text. b From B3LYP/6-31G calculation (see reference 88). C From reference 1 1 7. 103 mode are those that directly contact the n-electron density. In the ground state, the Tr-electron density is essentially confined to the porphyrin region, so the charges on the N-methylpyridyl rings are not sensed by the solvent molecules that interact with the n-electron density. The extensive excited-state delocalization of the Tr-electron density over the N—methylpyridyl rings then turns on the charge-dependent terms that control the intermolecular mode frequency. These observations allow us to con- clude further that the displacement of the intermolecular potential in the excited state that underlies the ground-state and excited-state wavepacket motions arises at least in part from the excited-state change of extent of the Tr-electron density; the expansion of the Tl' electrons pushes the solvent molecules away from the frame of the porphyrin and launches the wavepacket motions. These conclusions are significant because they imply that substantially increased activation and reorganization energies will be encountered for electron-transfer re- actions in polar solution owing to the trapping of net charges by the intermolecular modes with the surrounding medium. A particular impact would be expected on the well depths of the intermolecular modes with first-shell groups that we suggest make dominant contributions to the reorganization energies for the electron-transfer reac- tions in photosynthetic reaction centers. The rate of the secondary electron-transfer reaction that shuttles the negative charge from BPhef to QA in the purple-bacterial photosynthetic reaction center would be maximized by locating the BPheL acceptor in a nonpolar region, as far as possible from polar or charged side chains. In contrast, the barriers for charge-recombination reactions are raised by placing the primary electron-donor P and the quinones QA and QB in polar regions of the reaction cen- ter. These structural features73'125-126 clearly enhance the quantum efficiency of the two-step, transmembrane charge-transfer/energy-storage mechanism in the reaction CEI'IIEI'. 104 CHAPTER 5 Vibrational coherence in the native and molten-globule states of ZnH-substituted cytochrome c 5.1 Summary A crucial feature of the long-distance electron-transfer reactions in the purple- bacterial photosynthetic reaction center is that they exhibit activationless dynamics. In the previous chapters, we have advanced the hypothesis that van der Waals inter- actions between the redox-active chromophores and adjacent groups in the protein- derived medium account for the majority of the reorganization energy in these reac- tions and accordingly control the activation energy. In this chapter, we test several aspects of this hypothesis by characterizing the ground-state low-frequency vibra- tional coherence from ZnH-substituted cytochrome c (ZnCytc) in the native and acid- induced molten-globule states using femtosecond dynamic-absorption spectroscopy. The vibrational coherence observed from the native state exhibits two types of com- ponents. The first type is a pair of slowly damped (damping time y>1.5 ps) com- 105 ponents with frequencies of ~10 and ~30 cm‘l that are assigned to metal-doming modes with the M80 and H18 axial ligands. The second type is a strong, very rapidly damped (y<150 fs) 79-cm‘l component that is assigned to van der Waals inter- actions with nonpolar groups from the surrounding protein. The resonance Raman activity of this set of modes is effectively quenched in the molten-globule state owing to the expansion of the hydrophobic core and to the randomization of the surround- ing structure. These findings show that the van der Waals modes obtain resonance Raman activity through a direct attack of the neighboring groups upon the Tr-electron density of the ZnII porphyrin. Similar interactions are likely candidates for the rate- controlling interactions in the photosynthetic reaction center. 5.2 Introduction Perhaps the key feature that allows the purple bacterial photosynthetic reaction cen- ter to perform efficient conversion of solar photons into transmembrane chemical potential is that the electron-transfer reactions are optimized by the protein struc- ture so that they occur with activationless dynamics. The rate of electron transfer is effectively independent of temperature in the theory of Bixon and Jortner because the reactions are coupled to displacements along vibrational modes of the interven- ing protein medium that supports the redox-active chromophores}, 22 The tempera- ture dependence of the rate constant for the primary electron transfer reaction, from the paired bacteriochlorophyll primary electron donor, P, to the bacteriopheophytin acceptor, BPheL, is modeled adequately when the mean frequency of the protein- derived modes lies in the 80-100 cm‘1 regime.20 The observation of vibrational coherence in the stimulated emission from P* in this frequency range by Vos, Mar- tin, and co-workers43‘46'48‘51- 76 is probably consistent with the expectations of the 106 Bixon and Jortner theory but the structural assignment of the active modes has re- mained indeterminate. Vos and Martin suggest that the modes that appear in the vibrational coherence in the reaction center are like the phonons of molecular crystals in being delocal- ized over the protein structure spanning the redox chromophores.43'45v50 In the previous chapters, we raise an alternative hypothesis that the modes arise from van der Waals interactions: from localized, formally nonbonding interactions between the chromophores and groups in the first-solvation shell.63‘65v88 We have explic- itly compared the interactions that occur in polar liquids to those of the solvent molecules that are coordinated in gas-phase clusters.128 The vibrational modes aris- ing from these intermolecular interactions contribute rapidly damped features to the ground-state and excited-state vibrational coherence observed in polar solutions of bacteriochlorophyll (BChl) and of Zn11 porphyrins. In comparison to the compo- nents arising from the slowly damped intramolecular (or skeletal) normal modes of vibration of the chromophores in the < BOO-cm“1 regime, the intermolecular modes exhibit perhaps ten times the resonance Raman activity.63‘65’88 The broad lineshapes observed in solution for the van der Waals modes arise pre- dominantly from the inhomogeneously broaded ensemble of chromophore—solvent interactions in the first shell.63'88 In a folded protein structure, the interactions between a chromophore and the neighboring protein-derived groups are formally analogous to the those observed in polar solution but are likely to be more highly ordered. Thus, the lineshape of a particular chromophore-protein interaction should exhibit a narrower lineshape. In this chapter, in order to test these hypotheses, we have characterized the ground-state vibrational coherence from ZnH-substituted cy- tochrome c (ZnCytc) in the native state and in the molten-globule state. The structure of cytochrome c surrounds the intrinsic metalloporphyrin with an essentially nonpo- lar environment. The closest polar amino-acid residue lies at a distance of about 3 A. 107 f This arrangement suggests that the low-frequency vibrational coherence from the na- tive state should be dominated by London-dispersion interactions with the nonpolar residues that are at closer range. In the molten-globule state, however, the loss of or- der and the expansion of the hydrophobic core should turn off the resonance Raman activity of the modes that contribute to the vibrational coherence in the native state. The results allow a comparison of the spectrum of the vibrational coherence in the native and molten-globule states of ZnCytc with that observed from ZnII por- phyrins in polar solution. The main finding is that the rapidly damped component that dominates the vibrational coherence in the native state is completely absent in the vibrational coherence from the molten-globule state. We conclude that the protein-dependent changes in lineshape and intensity are consistent with the form of the van der Waals potential and the mechanism of resonance Raman activity we introduced previously.63v65'88 The present results strongly support a structural as- signment of the low-frequency modes that are active in photosynthetic reaction cen- ters to van der Waals interactions, and we discuss briefly how the mean frequency of these interactions can be tuned by the introduction of polar residues to the environ- ment of the redox-active chromophores. 5.3 Experimental 5.3.1 Sample Preparation ZnCytc was prepared from horse-heart ferricytochrome FeCytc (see Figure 5.1) us- ing the procedure developed in the Vanderkooi laboratory.129 liquid anhydrous hy- drogen fluoride (Iinde) was employed as the demetalating agent; the reaction was run on a home-built gas-handling system in Teflon reaction vessels. Metal recon- stitution of the free-base Cytc product with ZnII was performed in the presence of a lO-fold molar excess of zinc acetate (Sigma 379786-5G, 99.999%). The extent of 108 demetalation and metal-reconstitution reactions was monitored spectrophotometri- cally; the starting and product species are easily distinguished from each other in terms of the number and position of bands in the Q-band region of the absorption spectrum. ZnCytc product solutions were subsequently worked up using methods derived from those of Winkler and co-workers130 and Kostic and co-workers.131 Af- ter desalting, the protein was isolated by cation-exchange chromatography, first on a Whatrnan CM-S2 column and optionally then on a Mono-S 4.6/100 PE FPLC colunm (GE Healthcare Life Sciences). Fractions corresponding to ZnCytc were equflibrated with a 25-mM sodium phosphate buffer solution at pH 6.9 by repeated concentra- tion using an Amicon ultrafilter and dflution with the buffer solution. After a final ultrafiltration step, the product was analyzed on a Superdex 75 FPLC gel-filtration column (GE Healthcare Life Sciences). The eluent was monitored simultaneously at 280 and 420 nm. The gel-filtration column was calibrated using a solution of low- molecular-weight protein standards (GE Healthcare Life Sciences, LMW standards kit). The chromatograms (not shown) indicate that the porphyrin binding fractions corre- spond exclusively to monomeric cytochrome c species and that the samples are free of low-molecular-weight peptide fragments. Solutions of the acid-induced molten-globule state of ZnCytc were prepared ac- cording to the methods developed by Goto and co-workers132 and described by Kostic and co-workers.133 For use in the femtosecond pump-probe experiments, solutions of ZnCytc in the native or molten-globule states were prepared by dilut- ing the stock protein solution with pH 7.0 SO-mM sodium phosphate buffer solution to obtain an absorbance of 0.8 for a path length of 1.0 mm at the center of the laser spectrum at 420 nm for the native-state samples or at 418 nm for the molten-globule- state samples. The samples were held in the femtosecond pump-probe spectrometer at room temperature (23 °C) in a fused-silica flow cuvette (0.5-mm path length). A peristaltic 109 Figure 5.1. Ribbon (left) and surface-(right) renderings of the X-ray crystal structure of horse-heart ferricytochrome c (1hrc.pdb). The porphyrin and the axial ligands to the FenI ion, M80 and H18, are shown as stick representations in the ribbon picture, and space-filling representations in the surface picture. 110 pump was used to circulate a IO-mL reservoir of sample solution through the cuvette at 2.70 mL/rnin. The sample’s absorption spectrum was monitored during the exper- iment for changes arising from photochemistry or permanent photobleaching. The sample reservoir was exchanged with fresh solution several times during each run. 5.3.2 Continuous-Wave Absorption and Fluorescence spectroscopy Absorption spectra were obtained at 23 °C with a Hitachi U-2000 spectrophotometer (2-nm band pass). Fluorescence spectra were acquired at 23 °C with a Hitachi F-4500 spectrofluorimeter (S-nm band pass for the excitation and emission monochroma- tors). As presented as a function of wavenumber, the fluorescence intensities are multiplied by the square of the wavelength in order to compensate for the fixed (in wavelength units) spectral bandpass of the emission spectrometer.89 5.3.3 Femtosecond Spectroscopy Femtosecond pump—probe transients with impulsive excitation were recorded using the dynamic-absorption technique in which the probe beam is dispersed in a grating monochromator after passing through the sample.26i32i35'37 The present experi- ments were conducted with 50-fs pulses with intensity spectra centered at 420 nm for the native state and 418 nm for the molten-globule state (both 4 nm FWHM, as measured with an Ocean Optics USE-2000 spectrometer/CCD detector with a 0.5-nm bandpass) and a fairly narrow bandpass (1 nm) of the transmitted probe beam. This approach is similar to that used by Champion and co-workers in their studies of low-frequency vibrational coherence of heme proteins.53i 55,62 The pump—probe spectrometer used in this work consists of a frequency dou- bled, self-mode-locked Ti:sapphire oscillator (Coherent Mira-F oscillator and Verdi V5 (5 W) NszVO4 pump laser, Coherent/Inrad 5-050 second-harmonic generator), a 111 SFIO Brewster prism-pair pulse compressor, and a rapid-scanning, modified Mach— Zehnder interferometer with confocal sample and autocorrelation-crystal positions. The planes of polarization of the pump and probe beams were set to be 45° apart using calcite polarizers and A/2-retarding wave plates. After passing through the sample, the probe beam was analyzed by another calcite polarizer oriented 90° rel- ative to the pump-beam’s plane of polarization, and then it was passed through a monochromator (Acton Research SP-150) and was detected by an amplified photo- diode (Thorlabs PDAS). The pump-probe signal was obtained from the photodiode signal using a lock-in amplifier (Femto LIA-MV-200-H); the pump beam was modu- lated at 100 kHz by a A/ 2-retarding photoelastic modulator (Hinds Instrumentation) and a calcite polarizer. The instrumentation and methodology is identical to that described in Chapter 3. 5.4 Results Figure 5.2 shows continuous-wave absorption and fluorescence spectra from the native state of ZnCytc at 23 °C. The spectra are plotted as relative dipole strength889'91’92 as a function of wavenumber v, A(v) /v and 1-"(v)/v3 respectively. The absorption spectrum features two bands, the Soret (or B) band and the Q band, in the blue and red parts of the spectrum, respectively. The fluorescence spectrum extends to the red from the 0-0 peak of the Q band. Also shown in Figure 5.2 is the absorption spectrum from the molten-globule state of ZnCytc. The absorption spectrum of the molten-globule sample is blue shifted relative to that of the native state sample. The dynamic-absorption transients obtained for the native and molten-globule states are shown in Figures 5.3 and 5.4, respectively. Each transient exhibits an in- tense biphasic spike near the zero of time that goes off the plotted scale.93‘95 The 112 +12 F/v3 Relative Dipole Strength 0 O O O O 0 O 0 O O O O O O O o 0 O O O O O J . I I I 15 17 19 21 23 25 Frequency (103 cm") O J Figure 5.2. Soret (v> 23000 cm’l) and Q-band regions of the continuous-wave ab- sorption spectrum of ZnCytc in the native (solid curves) and molten-globule (circles) states. The fluorescence spectrum of the native state of ZnCytc at room temperature (23 °C) is shown overlapping and extending to the red of the absorption spectrum. The spectra are normalized to unit area with respect to the fluorescence spectrum and Q band. 113 native-state transient (see Figure 5.3) then exhibits a very rapidly damped oscillation over the early-time portion of the signal. The signal drops almost to the baseline and is then followed by a positive-going recurrence at the 170-fs delay point. The oscillation appears to be largely damped within the next 600 fs. The transient from the molten-globule state (see Figure 5.4) does not exhibit this oscillation. The signal exhibits a smooth transition from the positive side of the spike to a positive offset at the ~150 fs point. Both transients then exhibit a pattern of slowly damped cosinu- soidal oscillations superimposed on an exponential decay that extends to the ~4 ps delay point. The most intense oscillations have periods longer than 1 ps. Following truncation of the < IOO-fs portion,95 the transients were fit in the time domain to the sum of an exponential decay and multiple oscillatory components of the form 2 2 Ii(t) = Aie-t 0i /2 cos(w0it — (bi) /x/2F (5.1) with each component i having a center frequency (00; and phase (bi. These wave- forms correspond to Gaussian lineshapes in the frequency domain, _ _ . 2 2 1,-(w) = A,- e (w “’01) “201' ’ / (0,7211) (5.2) where the line width is controlled by the standard deviation, ai=Awi mm, with Awi standing for the full width at half maximum. Equations 5.1 and 5.2 are nor- malized so that the amplitude A,- corresponds to the area of the lineshape in the frequency domain and to the intensity of the signal in the time domain. Except as noted in the following, the modeling procedure was as described in detafl in previ— ous work on the ground-state vibrational coherence from ZnTMPyP.88 The models are shown superimposed on the data points of the transient in Figures 5.3 and 5.4. The two strongest slowly damped components observed below 100 cm“1 in a Hanning- windowed Fourier-transform spectrum (not shown) were included in the model. The Fourier-transform spectra from both samples exhibit higher-frequency peaks near 114 1.5 1.0 4 ': ’21 0.5 .- 0.0 .. I I I I I 0 1000 2000 3000 4000 Time (is) Figure 5.3. Soret-band (420-nm) pump-probe transient from the native state of Zn- Cytc. The signal points (circles) are superimposed on a model consisting of the sum of an exponential decay and two oscillatory components. The exponential has a time constant of 1.8 ps; the plotted signal is normalized to the exponential’s intensity. The parameters for the oscillatory components are listed in Table 5.1. 115 1.5 1.0 - ‘f ‘ ': '2 0 5 7 0.0 I I l l I O 1000 2000 3000 4000 Time (is) Figure 5.4. Soret-band (420-nm) pump-probe transient from the molten-globule state of ZnCytc. The signal points (circles) are superimposed on a model consisting of the sum of an exponential decay and two oscillatory components. The exponential has a time constant of 1.8 ps; the plotted signal is normalized to the exponential’s intensity. The parameters for the oscillatory components are listed in Table 5.2. 116 120 cm‘1 and 220 cm‘l, but these features are comparable in intensity to the noise level. Similar features were observed by Champion and co-workers in ferri- cytochrome c and myoglobin.“57 The relative weakness of the higher-frequency modulation components is consistent with the tuning of the pump spectrum to the center of the absorption spectrumssi 56 Frequency-domain representations of the models obtained for the oscillatory part of the native and molten-globule signals are shown in Figure 5.5a and b, respectively. The parameters for these models are listed in Tables 5.1 and 5.2. The model for the native-state signal includes two slowly damped features, at 10 and 31 cm‘l, and a single rapidly damped feature centered at 79 cm’l. The slowly damped components probably arise from metal-doming modes in sites having axial-ligand interactions to the ZnII ion with the M80 and H18 side chains, respectively. Comparable features appear at 1 1 and 26 cm"1 in the molten-globule spectrum, but they are less intense and narrower in lineshape than those observed in the native state. The broad feature (A00= 106 cm‘l) at 79 cm‘1 that appears only in the native-state model is assigned to van der Waals modes with the surrounding protein. It is several times more intense than the slowly clamped, metal-doming features. 5.5 Discussion The main finding of this chapter is that the vibrational coherence observed in the molten-globule state of ZnCytc lacks the strong, rapidly damped modulation compo- nent from van der Waals interactions with nearby protein residues that is observed in the native state. This observation suggests that the expanded hydrophobic core and the randomization of the local structural order that accompanies the formation of the molten-globule state destroys the coupling of the ZnII porphyrin’s 1r—>Tr* transition to the surrounding protein. In the following, we account for this conclu- 117 ii A an .0 3 'E or <0 '2' J m ......... .2 0 - E SE B o _ I I I I 0 50 100 150 200 250 Frequency (cm'1) Figure 5.5. Frequency-domain representations of oscillatory components of the fitted models of the vibrational coherence obtained from the (a) native state and (b) molten globule state of ZnCytc. The model parameters are listed in Tables 5.1 and 5.2. 118 Table 5.1. Gaussian lineshape parameters for the modulation components obtained from the pump—probe transient of ZnCytc in the native state (see Figure 5.3). Component 000 (cm'l) Aw (cm'l) Amplitudea 1 10 8.02 0.809 2 3 l 3.25 0.1 77 3 79 106 2.84 3 Normalized relative to the amplitude of the exponential decay of the transient (see Figure 5.3). 119 Table 5.2. Gaussian lineshape parameters for the modulation components obtained from the pump-probe transient of ZnCytc in the molten-globule state (see Figure 5.4). Component wo (cm'l) Aw (cm‘l) Amplitudea 1 1 1 7.76 0.1 57 2 26 0.575 7.08 x 10-2 3‘ Normalized relative to the amplitude of the exponential decay of the transient (see Figure 5.4). 120 sion first by considering the spectroscopic origin of the vibrational coherence that we observe in both samples and then by considering the nature of the change in struc- ture that modulates the resonance Raman activity and possibly also the lineshape of the van der Waals modes. We conclude with a brief discussion of the significance of these findings with respect to the function of the van der Waals interactions in purple-bacterial reaction centers and how the frequency of these interactions can be tuned by changing the structure of the protein around the chromophore. The pump-probe signals obtained with Soret-band excitation of ZnCytc in the na— tive and molten-globule states are modulated primarily by ground-state vibrational coherence. Both signals exhibit a net ground-state depletion character over the 4-ps experimental window, but there is a ~2-ps exponential decay that returns the sig- nal to the baseline level at the end of the recording (see Figures 5.3 and 5.4). This behavior is just like that observed previously with ZnTMPyP.88 The exponential de- cay, then, most likely arises from vibrational cooling in the Sl-state manifold follow- ing an ultrafast (t<100 fs) internal conversion of the Sz-state population prepared by the pump pulse.103 Given that the impulsive character of the excitation is not retained following the Sg—>Sl nonradiative process, the modulations observed in the pump-probe signals have to be assigned to coherent wavepacket motions on the ground-state potential energy surface. These motions are launched along the coor- dinates of the normal modes that are displaced by the 1t—>1T* transition, assuming the Albrecht A-term mechanism for resonance Raman activity;120'121 the amplitudes of the associated modulation components that appear in the pump-probe signal are analogous to relative resonance Raman activities.24‘26' 30' 32’ 33' 121 The intramolecu- lar or skeletal modes of the ZnII porphyrin obtain resonance Raman activity owing to the changes in the electronic structure that accompany the Tr—>Tr* transition. The van der Waals modes between the Zn11 porphyrin and groups that pack around its rt-electron density obtain resonance Raman activity because they are reoriented or 121 displaced by the changes in shape or extent of the n-electron density that accompany the Tr->Tr* transition. The slowly damped modulation components (y> 1.5 ps, Aw < 10 cm‘l) correspond to intramolecular modes, from the porphyrin macrocycle and the axial ligands to the ZnII ion.65'88 The two modulation features that were modeled in the native and molten-globule signals are very likely to correspond to metal-doming motions. In their studies of myoglobin and ferricytochrome 6, Champion and co-workers ob- served low-frequency modes near 40 cm‘1 that were assigned to the metal-doming mode.57 In ZnII porphyrins, however, the central metal ion is not as strongly co- ordinated by the porphyrin, so a decrease in the metal-doming frequency might be expected. Previous studies have shown that the ZnII ion in porphyrin complexes in solution prefers a five-coordinate crystal field, so only a single axial ligand is likely to be bound in ZnCytc.134‘137 Six-coordinate ZnII porphyrins are observed in crys- tals, but the porphyrin is distorted by crystal-packing forces.133‘140 In ZnCytc, the resonance Raman spectra obtained by Kostic and co-workers suggest that an ortho- dox six-coordinate configuration for the ZnII ion is not present.131 The picosecond time-resolved fluorescence results obtained previously in this laboratory, however, show that there are two phases of axial-ligand photodissociation with Q-band exci- tation.141 We suggested in that work that the folded native structure might impose a six-coordinate structure on the Zn11 ion owing to packing forces from the folded protein in analogy to those in crystals. The present work shows that two slowly damped modulation features, at 10 and 30 cm‘l, are present in the signals from the native and molten-globule states of ZnCytc. We suggest that both of these features arise from the metal-doming mode of the ZnII porphyrin in two different five-coordinate configurations. The lower fre- quency feature, at 10 and 11 cm‘1 in the native and molten-globule states, respec- tively, probably arises from the sub-ensemble of molecules in which the Zn11 ion is 122 domed so that an axial-ligand interaction can be made to the side chain of the me- thionine residue M80. The higher-frequency feature, at 31 and 26 cm"1 in the native and molten-globule states, would then arise from the sub-ensemble with the ZnII ion domed so that the axial-ligand interaction is made to the side chain of the histidine residue H18. The two structures would be distinct because of the significant activa- tion required for the ZnII ion to move across the plane of the porphyrin. The lower frequency and broader width of the 10-cm‘1 feature probably arises from the weaker ligand-metal bond formed between the ZnII ion and the thioether side chain of the methionine residue. The observation that the both the 10-cm‘1 and 30-cm‘1 fea- tures are significantly weaker in modulation amplitude in the molten-globule state than observed in the native state (see Figure 5.5 and Tables 5.1 and 5.2) suggests that a smaller fraction of the overall ensemble of molecules coordinates either of the two amino acids to the ZnII ion at a given instant in time. Several aspects of the change in protein structure that accompanies the forma- tion of the molten-globule state account for the disappearance of the rapidly damped modulation feature that apparently arises from van der Waals interactions in the native state. The molten-globule state of cytochrome c has been extensively stud- ied.132-133’142'148 It is characterized by the loss of the ordered interactions be- tween neighboring groups that characterizes the native structure, but most of the secondary structural content of the native state is retained.143 The protein structure adopts an expanded configuration that rapidly samples a range of internal organi- zations. X-ray light-scattering experiments show that the Stokes radius of the cavity that surrounds the porphyrin in the molten-globule state is between four and fif- teen percent larger than that of the native state.142i143 The blue—shifted absorption spectrum of the molten-globule state (see Figure 5.2) probably results from the low- ered average polarizability of the protein-derived surroundings that results from the expanded cavity. Such an expansion would be expected to attenuate strongly the res- 123 onance Raman activities of the van der Waals modes because on average the distance between the adjacent groups and the rt-electron density of the Zn“ porphyrin would be increased and the associated sensing of the ground-to-excited-state changes in the n-electron density would be concomitantly decreased. The rapid sampling of different internal configurations would be expected to broaden the range of configu- rations that contribute to the inhomogeneously broadened ensemble, with a broad- ened lineshape the likely result. In the time domain, in a pump-probe experiment, the especially rapid damping of such a component would mitigate further against its detection. The relatively low mean frequency of the van der Waals modes observed in the native state of ZnCytc is consistent with the structure of the FeCytc because there are no nearby polar residues that are close enough to the n-electron density of the porphyrin to obtain strong coupling to the 1r—-n* transition and high mode fre- quencies. In the following, we analyze the range of possible interactions in terms of the X-ray crystal structure and in terms of the known behavior of the van der Waals potential. The frequency of the van der Waals modes can be understood in terms of the potential that was discussed in the previous chapters. When there are no nearby charges present, the potential can be expressed as the sum of a repulsive, exchange term scaled by 7‘12 and a series of attractive terms scaled by 7’6: 2 2 2 2 V 1. =61 1.12_ b +Clull |u2| +0t2|u1| +0t1|fl2| 2 6 53 0( ) / ( Otlotz (41Tfo)2 41T£0 D 1’ ( ) As shown previously,63 a Taylor series expansion of Equation 5.3 around the equi- librium geometry gives an expression for the natural frequency of the van der Waals mode, 7/6 3 clullzluzl2 Otzlml2 + 0<1|M2|2 7/3 1/2 v = —— bar or + - (2a)2/3 ( 1 2 (41750)2 + 4TT£0 2WD [1 (5 4) 124 In both equations, the constant a scales the Pauli (exchange) interaction, 17 scales the London-dispersion interaction, c = 2 / 3k3T scales the dipole—dipole interactions, and the (xiu 1,2 terms correspond to dipole-induced-dipole interactions; pi and 0:,- correspond to the dipole moment and polarizability of the choromphore solute and solvent (subscripts 1 and 2, respectively). Lastly, r is the distance between the two molecules, u is the reduced mass of the oscillator, D is the dielectric constant of the medium, and so is the permittivity of free space.113‘116 As expressed above, the dipole-dependent terms apply to the high-temperature limit, where kBT is much greater than the well depth, so the interaction is averaged over all orientations.115 If the potential well depth is greater than k3 T, the assumption that it is appropriate to average the interaction energy over all orientations may not be valid. When the solute or solvent has a charge Q12, the potential gains attractive terms corresponding to the ion-dipole and ion—induced-dipole interactions: 2 2 2 + 2 VQ(1’) = V0 - (CQIIHZI + CQZIMII + 0‘qu 0(1Q2) /2D21’4 (5.5) (41T€0)2 41T£0 This form of the potential is similar to that introduced in Chapters 3 and 4, ex- cept that we now explicitly account for any charges that may be located either on the solute or solvent. The ion—dipole term depends linearly on the charge and the square of the dipole moment pi, and the ion—induced-dipole term depends on the polarizability (x,- and exhibits quadratic charge dependence. Additionally, the charge- containing terms have an r'4 dependence on the solute-solvent distance. Thus, the charge-dependent terms cause the potential to extend to much greater distances than does the neutral potential. Equation 5.5 also predicts that the depth of the potential well will increase significantly over that predicted from the neutral potential (Equa- tion 5.3), so the assumption of orientational averaging made for the uncharged case is no longer valid, and the result would be an even more stable structure for the van der Waals complex. Overall, the largest contributions to the van der Waals potential 125 Figure 5.6. Detail of the X—ray crystal structure of ferricytochrome c (1hrc.pdb) ex- amining possible interactions of the protein with the n-electron density of the por- phyrin. Polar and charged residues are rendered as light and dark gray stick figures, respectively. Nonpolar residues are omitted from the figure so that the resulting non-polar cavity is made clear. Nearest-neighbor distances are shown between the porphyrin and several residues: the hydroxyl group on Y67 (3.27 A), the nitrogen atom on W59 (5.29 A), and the nitrogen on K79 (6.71 A). The Y67, W59 and K79 residues are shown in black. The interior water molecule (Wat-112, black sphere) is 3.77 A from the porphyrin. 126 are those arising from the London-dispersion interactions and those involving nearby polar or charged residues. An assessment of the protein-derived structures that are near enough to the por- phyrin in ZnCytc to form a coupled van der Waals mode can be made on the basis of the X-ray crystal structure of FeCytc because the two proteins assume the same or similar structures in solution. As noted above, the conformation of the porphyrin and the configuration of the axial ligands to the metal ion are likely to differ in FeCytc and ZnCytc because of the specifics of the coordination chemistry of the two metals. The structures of horse-heart ZnCytc and ferrocytochrome c in solution have been compared using 2D-NMR spectroscopy.149'150 The configurations of the polypep- tide backbone and amino-acid side chains of the Fe11 and ZnII proteins are apparently quite comparable judging from the proton chemical shifts. Anni et a1.149 conclude that the side chains of H18 and M80 are suitably arranged so that they can interact with the ZnII ion as axial ligands just as indicated by the FeCytc X-ray crystal struc- ture. The X-ray crystal structure of ZnCytc from horse heart is not available; however, the structures of Fe:ZnCytc cocrystals from tuna have been solved by Winkler, Gray, and coworkers.151 The Fem-containing sites in the asymmetric unit exhibit struc- tures that are qualitatively comparable to those of the ZnH-containing sites. Thus, although some differences from the crystal structure should be expected in solution, it is reasonable to use the crystal structure of FeCytc from horse heart as a guide for that of ZnCytc in the following analysis. Figure 5.6 depicts the polar and charged residues in FeCytc that contribute to the environment of the porphyrin. None of these groups are close enough to the n-electron density of the porphyrin to make a coupled van der Waals mode. The distances shown in Figure 5.6 are those for the shortest nearest-neighbor vectors between the porphyrin and a labeled polar or charged residue. The Tr-electrons are delocalized over the entire porphyrin ring, however, so the magnitude of the van der 127 Waals interactions would be smaller than indicated by the distance measurements. The nearest charged residue is a lysine residue (K79), but it lies greater than 6 A at closest approach from the nearest atom in the porphyrin macrocycle. The lysine residue (K13) that crosses in front of the heme in Figure 5.6 is approximately 10 A from the porphyrin. The nearest polar groups in the surrounding protein medium that are not directly coordinated to the metal ion are the tyrosine and tryptophan residues, Y67 and W59. As shown in Figure 5.6, these residues are at least 3 and 5 A from the porphyrin macrocycle, respectively, so their interactions with the N'- electron density are weak. Polar interactions with nearby bulk water molecules can be neglected because the porphyrin is exposed along the equatorial plane (see Fig- ure 5.1). The conserved interior water molecule Wat-112 (Wat-166 in yeast iso-I- cytochrome c152) is hydrogen bonded to the nearby N52, Y67 and T78 residues.153 It is ~3.8 A from the porphyrin (see Figure 5.6), so its contribution to the signal will also be negligible. The main contributions to the coupled van der Waals interactions between the protein and the porphyrin’s n-electron density in ZnCytc arise from London- dispersion interactions with non-polar residues. These residues fill the majority of the cavity adjacent to the porphyrin in the native state (see Figure 5.6). The line- shape resulting from these interactions wfll be very broad because of the wide range of distances and orientations with respect to the porphyrin. The modulation compo- nents from polar residues, if they were present, would exhibit higher frequencies and narrower lineshapes. The trend of the mode frequency as a function of the solvent dipole moment obtained in the previous work on ZnTMPyP suggests a frequency in the nonpolar limit near 70 cm‘l.88 The mean frequency (79 cm‘l) and broad line- shape (or rapid damping) of the van der Waals interaction observed in the native state of ZnCytc is consistent with these expectations. 128 These observafions allow a suggestion about the nature of the vibrational coher- ence in the purple-bacterial photosynthetic reaction center43‘45’48'51v 76 that arises from van der Waals modes. The overall spectrum should contain an underlying broad lineshape arising from the interactions of non-polar residues in the surroundings of the primary electron donor P. Narrower lineshapes, from interactions of dipolar or charged residues, will be superimposed on the broad background. This description is consistent with the observations made by Vos, Martin and co-workers in their stud- ies of the effects of point mutations on the vibrational coherence in the Rhodobacter sphaeroides reaction centers.50 The mutations swapped nonpolar for polar, polar for nonpolar, or polar for charged residues. The spectrum from each mutant was unique in having one or more relatively narrow lines at different frequencies but these vary- ing features were superimposed on a broad background lineshape that was indepen- dent of the mutation. Thus, we suggest that the sharper components arise from dipole-dipole interactions involving an amino acid near the porphyrin, and the in- variant broad background lineshape arises from the London-dispersion interactions with adjacent groups. 5.5.1 Conclusions The principal result of this chapter is that rapidly damped contributions to the vi- brational coherence in ZnCytc arise from van der Waals interactions between the porphyrin and the surrounding protein medium. The expansion of the core and ran- domization of the surrounding structure effectively quenches these interactions in the molten-globule state. The van der Waals potential that accounts for the modu- lation components in the vibrational coherence is dominated by London-dispersion interactions; charged and polar residues are too distant in the native structure from the porphyrin to contribute significantly to the signal. The specific interactions that gain resonance Raman intensity via coupling to the Tr—>Tr* transition of the por- 129 phyrin are those that are capable of directly attacking the Tr-electron density. In the photosynthetic reaction center, the low-frequency vibrational modes that are coupled to the primary and secondary charge-separation reactions probably arise from van der Waals interactions between the redox-active chromophores and the surrounding protein medium. The local structure around a chromophore can tune the mean fre- quency of the interaction through the addition of polar or charged groups. These findings provide a possible mechanism that accounts for the ability of the surround- ing protein structure to optimize the efficiency of the energy-storing electron-transfer reactions in photosynthesis. 130 BIBLIOGRAPHY (1) Bixon, M.; Jortner, J. Activationless and pseudoactivationless primary electron transfer in photosynthetic bacterial reaction centers. Chem. Phys. Lett. 1989, 1 5 9, 17—20. (2) Marcus, R. A. On the theory of oxidation-reduction reactions involving electron transfer. I. J. Chem. Phys. 1956, 24, 966—989. (3) Jortner, J. Temperature dependent activation energy for electron transfer be- tween biological molecules. J. Chem Phys. 1976, 64, 4860-4867. (4) Marcus, R. A.; Sutin, N. Electron transfers in chemistry and biology. Biochim. Biophys. Acta 1985, 811, 265—322. (5) Wasielewski, M. R. Photoinduced electron transfer in supramolecular systems for artificial photosynthesis. Chem. Rev. 1992, 92, 436—461. (6) Wasielewski, M. R.; Wiederrecht, G. P.; Svec, W. A.; Niemczyk, M. P. Chlorin- based supramolecular assemblies for artificial photosynthesis. Solar Energy Mat. Solar Cells 1995, 38, 127-134. (7) Gust, D.; Moore, T. A.; Makings, L. R.; Liddell, P. A.; Nemeth, G. A.; Moore, A. L. Photodriven electron transfer in triad molecules: a two-step charge recombina- tion reaction. J. Am. Chem. Soc. 1986, 108, 8028-8031. (8) Gust, D.; Moore, T. A. Mimicking photosynthesis. Science 1989, 244, 35-41. (9) Gust, D.; Moore, T. A.; Moore, A. L.; Lee, S. -J.; Bittersmann, E.; Luttrull, D. K.; Rehms, A. A.; DeGaraziano, J. M.; Ma, X. C.; Gao, F.; Belford, R. E.; Trier, T. T. Efficient multistep photoinitiated electron transfer in a molecular pentad. Science 1990, 248, 199-201. (10) Gust, D.; Moore, T. A.; Moore, A. L. Mimicking photosynthetic solar energy transduction. Acc. Chem. Res. 2001, 34, 40-48. (11) Prathapan, 5.; Johnson, T. E.; Lindsey, J. S. Building-block synthesis of por- phyrin light-harvesting arrays. J. Am. Chem. Soc. 1993, 115, 7519-7520. (12) Hsiao, J. -S.; Krueger, B. P.; Wagner, R. W.; Johnson, T. E.; Delayen, J. K.; Mauzer- all, D. C.; Fleming, G. R.; Lindsey, J. 5.; Bocian, D. F.; Donohoe, R. J. Soluble Syn- thetic Multiporphyrin Arrays. 2. Photodynarnics of energy-transfer processes. J. Am. Chem. Soc. 1996, 118, 11181—11193. 131 (13) Stenberg-Yfrach, G.; Rigaud, J. -L.; Durantini, E. N.; Moore, A. L.; Gust, D.; Moore, T. A. Light-driven production of ATP catalysed by FOFl-ATP synthase in an artificial photosynthetic membrane. Nature 1998, 392, 479-482. (14) McCusker, J. K Fuel from photons. Science 2001, 293, 1599—1601. (15) McCusker, J. K Femtosecond absorption spectroscopy of transition metal charge-transfer complexes. Acc. Chem. Res. 2003, 36, 876—887. (16) Redmore, N. P.; Rubtsov, I. V.; Therien, M. J. Synthesis, electronic structure, and electron transfer dynamics of (aryl)ethynyl—bridged donor-acceptor systems. J. Am. Chem. Soc. 2003, 125, 8769—8778. (17) Splan, K. E.; Stem, C. L.; Hupp, J. T. Two coordinatively linked supramolecu- lar assemblies constructed from highly electron deficient porphyrins. Inorg. Chim. Acta 2004, 357, 4005—4014. (18) Kalyanasundaram, K.; Vlazhopoulos, N.; Krishnan, V.; Monnier, A.; Gratzel, M. Sensitization of TiOz in the visible light region using zinc porphyrins. J. Phys. Chem 1987, 91, 2342. (19) Kay, A.; Gratzel, M. Artificial photosynthesis. 1. Photosensitization of TiOZ so- lar cells with chlorophyll derivatives and related natural porphyrins. J. Phys. Chem. 1993, 97, 6272-6277. (20) Fleming, G. R.; Martin, J. -L.; Breton, J. Rates of primary electron transfer in photosynthetic reaction centers and their mechanistic implications. Nature 1988, 333, 190-192. (21) Atkins, P.; Jones, L. Chemical Principles: The Quest for Insight; W. H. Freeman: New York, 2002. (22) Bixon, M.; Jortner, J. Coupling of protein modes to electron transfer in bacterial photosynthesis. J. Phys. Chem. 1986, 90, 3795-3800. (23) Jortner, J.; Bixon, M. lntramolecular vibrational excitations accompanying solvent-controlled electron transfer reactions. J. Chem. Phys. 1988, 88, 167- 170. (24) Pollard, W. T.; Fragnito, H. L.; Bigot, J. -Y.; Shank, C. V.; Mathies, R. A. Quantum- mechanical theory for 6 fs dynamic absorption spectroscopy and its applica- tion to nfle blue. Chem. Phys Lett. 1990, 168, 239—245. (25) Pollard, W. T.; Lee, S. -Y.; Mathies, R. A. Wave packet theory of dynamic absorp- tion spectra in femtosecond pump-probe experiments. J. Chem Phys. 1990, 92, 4012-4029. 132 (26) Pollard, W. T.; Mathies, R. A. Analysis of femtosecond dynamic absorption spec- tra of nonstationary states. Annu. Rev. Phys. Chem. 1992, 43, 497—523. (27) Jonas, D. M.; Bradforth, S. E.; Passino, S. A.; Fleming, G. R. Femtosecond wavepacket spectroscopy: influence of temperature, wavelength, and pulse du- ration. J. Phys. Chem. 1995, 99, 2594-2608. (28) Voth, G. A.; Hochstrasser, R. M. Transition state dynamics and relaxation pro- cesses in solutions: a frontier of physical chemistry. J. Phys. Chem. 1996, 100, 13034-13049. (29) Carson, E. A.; Diffey, W. M.; Shelly, K. R.; Lampa-Pastirk, 5.; Dillman, K. L.; Schleicher, J. M.; Beck, W. F. Dynamic-absorption spectral contours: vibrational phase-dependent resolution of low-frequency coherent wave-packet motion of IR144 on the ground and excited 11 —> 11* surfaces. J. Phys. Chem. A 2004, 108, 1489-1500. DOI: 10.1021/jp035176c. (30) Walsh, A. M.; Loring, R. F. Theory of resonant and nonresonant impulsive stim- ulated Raman scattering. Chem. Phys. Lett. 1989, 160, 299-304. (3 1) Pollard, W. T.; Brito Cruz, C. H.; Shank, C. V.; Mathies, R. A. Direct observation of the excited-state cis-trans photoisomerization of bacteriorhodopsin: multilevel line shape theory for femtosecond dynamic hole burning and its application. J. Chem. Phys. 1989, 90, 199-208. (32) Dexheimer, S. L.; Wang, Q.; Peteanu, L. A.; Pollard, W. T.; Mathies, R. A.; Shank, C. V. Femtosecond impulsive excitation of nonstationary vibrational states in bacteriorhodopsin. Chem Phys. Lett. 1992, 188, 61-66. (33) Pollard, W. T.; Dexheimer, S. L.; Wang, Q.; Peteanu, L. A.; Shank, C. V.; Math- ies, R. A. Theory of dynamic absorption spectroscopy of nonstationary states. 4. Application to 12-fs resonant impulsive Raman spectroscopy of bacteri- orhodopsin. J. Phys. Chem. 1992, 96, 6147-6158. (34) Myers, A. B.; Mathies, R. A. Resonance Raman intensities: a probe of excited- state structure and dynamics. In Biological Applications of Raman Spectroscopy; Spiro, T. G., Ed.; Wiley-Interscience: New York, 1987; Vol. 2, Resonance Raman Spectra of Polyenes and Aromatics, pp 1-58. (35) Fragnito, H. L.; Bigot, J. -Y.; Becker, P. C.; Shank, C. V. Evolution of the vibronic absorption spectrum in a molecule following impulsive excitation with a 6-fs optical pulse. Chem. Phys. Lett. 1989, 160, 101-104. (36) Schoenlein, R. W.; Peteanu, L. A.; Mathies, R. A.; Shank, C. V. The first step in vision: femtosecond isomerization of rhodopsin. Science 1991, 254, 412—415. 133 (37) Wang, Q.; Schoenlein, R. W.; Peteanu, L. A.; Mathies, R. A.; Shank, C. V. Vibra- tionally coherent photochemistry in the femtosecond primary event of vision. Science 1994, 266, 422-424. (38) Mathies, R. A.; Brito Cruz, C. H.; Pollard, W. T.; Shank, C. V. Direct observation of the femtosecond excited-state cis-trans isomerization in bacteriorhodopsin. Science 1988, 240, 777-779. (39) Schoenlein, R. W.; Peteanu, L. A.; Wang, Q.; Mathies, R. A.; Shank, C. V. Femtosecond dynamics of cis-trans isomerization in a visual pigment analog: isorhodopsin. J. Phys. Chem. 1993, 97, 12087-12092. (40) Peteanu, L. A.; Schoenlein, R. W.; Wang, Q.; Mathies, R. A.; Shank, C. V. The first step in vision occurs in femtoseconds: complete blue and red spectral studies. Proc. Natl. Acad. Sci. U.S.A. 1993, 90, 1 1762-11766. (41) Wang, (1.; Kochendoerfer, G. G.; Schoenlein, R. W.; Verdegem, P. J. E.; Lugten- burg, J.; Mathies, R. A.; Shank, C. V. Femtosecond spectroscopy of a 13— demethylrhodopsin visual pigment analogue: the role of nonbonded interac- tions in the isomerization process. J. Phys. Chem. 1996, 100, 17388-17394. (42) Bardeen, C. J.; Yakovlev, V. V.; Wilson, K. R. Quantum control of population transfer in green fluorescent protein by using chirped femtosecond pulses. J. Am. Chem Soc. 1998, 120, 13023. (43) V08, M. H.; Lambry, J. -C.; Robles, S. J.; Youvan, D. C.; Breton, J.; Martin, J. -L. Direct observation of vibrational coherence in bacterial reaction centers using femtosecond absorption spectroscopy. Proc. Natl. Acad. Sci. U.S.A. 1991, 88, 8885-8889. (44) V08, M. H.; Lambry, J. -C.; Robles, S. J.; Youvan, D. C.; Breton, J.; Martin, J. -L Femtosecond spectral evolution of the excited state of bacterial reaction centers at 10 K. Proc. Natl. Acad. Sci. U.S.A. 1992, 89, 613-617. (45) Vos, M. H.; Rappaport, F.; Lambry, J. -C.; Breton, J.; Martin, J. -L. Visualization of the coherent nuclear motion in a membrane protein by femtosecond spec— troscopy. Nature 1993, 363, 320-325. (46) Vos, M. H.; Jones, M. R.; Hunter, C. N.; Breton, J.; Lambry, J. -C.; Martin, J. —L. Co- herent dynamics during the primary electron-transfer reaction in membrane- bound reaction centers of Rhodobacter sphaeroides. Biochemistry 1994, 33, 6750-6757. (47) Vos, M. H.; Jones, M. R.; McGlynn, P.; Hunter, C. N.; Breton, J.; Martin, J. -L. Influence of the membrane environment on vibrational motions in reaction 134 (48) (49) (50) (51) (52) (53) (54) (55) (56) (57) centers of Rhodobacter sphaeroides. Biochim. Biophys. Acta 1994, 1186, 117- 122. Vos, M. H.; Jones, M. R; Breton, J.; Lambry, J. -C.; Martin, J. -L. Vibrational dephasing of long- and short-lived primary donor states in mutant reaction centers of Rhodobacter sphaeroides. Biochemistry 1996, 35, 2687-2692. Vos, M. H.; Jones, M. R.; Martin, J. -L. Vibrational coherence in bacterial reac- tion centers: spectroscopic characterisation of motions active during primary electron transfer. Chem. Phys. 1998, 233, 179—190. Rischel, C.; Spiedel, D.; Ridge, J. P.; Jones, M. R.; Breton, J.; Lambry, J. -C.; Martin, J. -L.; Vos, M. H. Low frequency vibrational modes in proteins: changes induced by point-mutations in the protein-cofactor matrix of bacterial reaction centers. Proc. Natl. Acad. Sci. U.S.A. 1998, 95, 12306-12311. Vos, M. H.; Martin, J. -L. Femtosecond processes in proteins. Biochim. Biophys. Acta 1999, 1411, 1-20. Go, N.; Noguti, T.; Nishikawa, T. Dynamics of a small globular protein in terms of low-frequency vibrational modes. Proc. Nat. Acad. Sci. USA 1983, 80, 3696- 3700. Zhu, L.; Li, R; Huang, M.; Sage, J. T.; Champion, P. M. Real time observation of low frequency heme protein vibrations using femtosecond coherence spec- troscopy. Phys. Rev. Lett. 1994, 72, 301-304. Rosca, F.; Kumar, A. T. N.; Ye, x; Sjodin, T.; Demidov, A.; Champion, P. M. Investigations of coherent vibrational oscillations in myoglobin. J. Phys. Chem. A 2000, 104, 4280—4290. Rosca, F.; Ionascu, D.; Kumar, A. T. N.; Demidov, A. A.; Champion, P. M. Fem- tosecond coherence spectroscopy using spectrally selective differential pho- todetection. Chem. Phys. Lett. 2001, 337, 107-116. Rosca, F.; Kumar, A. T. N.; Ionascu, D.; Sjodin, T.; Demidov, A. A.; Champion, P. M. Wavelength selective modulation in femtosecond pump-probe spectroscopy and its application to heme proteins. J. Chem. Phys. 2001, 114, 10884-10898. Rosca, F.; Kumar, A. T. N.; Ionascu, D.; Ye, X.; Demidov, A. A.; Sjodin, T.; Whar- ton, D.; Barrick, D.; Sligar, S. G.; Yonetani, T.; Champion, P. M. Investigations of anharmonic low-frequency oscillations in heme proteins. J. Phys. Chem. A 2002, 106, 3540—3552. 135 (58) Barkhuijsen, H.; de Beer, R.; Bovée, W. M. M. J.; van Ormondt, D. Retrieval of frequencies, amplitudes, damping factors, and phases from time-domain signals using a linear least-squares procedure. J. Magn. Reson. 1985, 61, 465- 481. (59) Barkhuijsen, H.; de Beer, R.; van Ormondt, D. Aspects of the computational efficiency of LPSVD. J. Magn. Reson. 1985, 64, 343—346. (60) Barkhuijsen, H.; de Beer, R.; van Ormondt, D. Improved algorithm for noniter- ative time-domain model fitting to exponentially damped magnetic resonance signals. J. Magn. Reson. 1987, 73, 553-557. (61) Wang, W.; Ye, X.; Demidov, A. A.; Rosca, F.; Sjodin, T.; Cao, W.; Sheeran, M.; Champion, P. M. Femtosecond multicolor pump-probe spectroscopy of ferrous cytochrome c. J. Phys. Chem. B 2000, 104, 10789-10801. (62) Wang, W.; Demidov, A.; Ye, X.; Christian, J. F.; Sjodin, T.; Champion, P. M. Appli- cation of femtosecond coherence spectroscopy to the observation of nuclear motions in heme proteins and transparent solutions. J. Raman Spectr. 2000, 31, 99-105. (63) Shelly, K. R.; Golovich, E. C.; Beck, W. F. Intermolecular vibrational coherence in bacteriochlorophyll a with clustered polar solvent molecules. J. Phys. Chem. B 2006, 1 1 0, 20586-20595. DOI: 10.102 l/jp062909v. (64) Shelly, K. R.; Carson, E. A.; Beck, W. F. Vibrational coherence from the dipyridine complex of bacteriochlorophyll a: intramolecular modes in the 10-2 20 cm“1 regime, intermolecular solvent modes, and relevance to photosynthesis. J. Am. Chem. Soc. 2003, 1 25, 1 1810-1 181 1. DOI: 10.102 l/ja0366890. (65) Shelly, K. R.; Golovich, E. C.; Dillman, K. L.; Beck, W. F. Intermolecular vibra- tional coherence in the purple-bacterial light-harvesting proteins B777 and B820 from Rhodospirillum rubrum. J. Phys. Chem. B 2008, 112, 1299—1307. DOI: 10.1021/jp077103p. (66) Roszak, A. W.; Howard, T. D.; Southall, J.; Gardiner, A. T.; Law, C. J.; Isaacs, N. W.; Cogdell, R. J. Crystal structure of the RC-LHl core complex from Rhodopseudomonas palustris. Science 2003, 302, 1969-1972. (67) Sturgis, J. N.; Robert, B. Thermodynamics of membrane polypeptide oligomer- ization in light-harvesting complexes and associated structural changes. J. Mol. Biol. 1994, 238, 445-454. (68) Mfller, J. F.; Hinchigeri, S. B.; Parkes-Loach, P. S.; Callahan, P. M.; Sprinkle, J. R.; Riccobono, J. R.; Loach, P. A. Isolation and characterization of a subunit form 136 (69) (70) (71) (72) (73) (74) (75) (76) (77) (78) of the light-harvesting complex of Rhodospirillum rubrum. Biochemistry 1987, 26, 5055-5062. Parkes-Loach, P. S.; Sprinkle, J. R; Loach, P. A. Reconstitution of the B873 light- harvesting complex of Rhodospirillum rubrum from the separately isolated or- and B-polypeptides and bacteriochlorophyll a. Biochemistry 1988, 27, 2718- 2727. Pandit, A.; Visschers, R. W.; van Stokkum, I. H. M.; Kraayenhof, R.; van Gron- delle, R. Oligomerization of light-harvesting I antenna peptides of Rhodospiril- lum rubrum. Biochemistry 2001, 40, 12913-12924. Arluison, V.; Seguin, J.; Robert, B. The reaction order of the dissociation reac- tion of the B820 subunit of Rhodospirillum rubrum light-harvesting I complex. FEBS Lett. 2002, 516, 40-42. Wang, Z. Y.; Gokan, K.; Kobayashi, M.; Nozawa, T. Solution structures of the core light-harvesting alpha and beta polypeptides from Rhodospirillum rubrum: Implications for the pigment-protein and protein-protein interactions. J. Mol. Biol. 2005, 347, 465—477. Deisenhofer, J.; Epp, O.; Mild, K.; Huber, R.; Michel, H. X-ray structure analy- sis of a membrane protein complex: electron density map at 3 A resolution and a model of the chromophores of the photosynthetic reaction center from Rhodopseudomonas viridis. J. Mol. Biol. 1984, 180, 385—398. Diffey, W. M.; Homoelle, B. J.; Edington, M. D.; Beck, W. F. Excited-state vi- brational coherence and anisotropy decay in the bacteriochlorophyll a dimer protein B820. J. Phys. Chem. B 1998, 102, 2776-2786. Hu, X.; Schulten, K. Model for the light-harvesting complex I (B875) of Rhodobacter sphaeroides. Biophys. J. 1998, 75, 683-694. Vos, M. H.; Jones, M. R.; Hunter, C. N.; Breton, J.; Martin, J. -L. Coherent nuclear dynamics at room temperature in bacterial reaction centers. Proc. Nat. Acad. Sci. USA 1994, 91, 12701-12705. Stanley, R. J.; Boxer, S. G. Oscillations in the spontaneous fluorescence from photosynthetic reaction centers. J. Phys. Chem. 1995, 99, 859-863. Vos, M. H.; Rischel, C.; Jones, M. R.; Martin, J. -L. Electrochromic detection of a coherent component in the formation of the charge pair P+HL in bacterial reaction centers. Biochemistry 2000, 39, 8353-8361. 137 (79) Yakovlev, A. G.; Shkuropatov, A. Y.; Shuvalov, V. A. Nuclear wavepacket motion producing a reversible charge separation in bacterial reaction centers. FEBS Lett. 2000, 466, 209-212. (80) Yakovlev, A. G.; Shkuropatov, A. Y.; Shuvalov, V. A. Nuclear wave packet motion between P* and P B A' potential surfaces with a subsequent electron transfer to HA in bacterial reaction centers at 90 K. Electron transfer pathway. Biochem- istry 2002, 41, 14019—14027. (81) Yakovlev, A. G.; Shkuropatov, A. Y.; Shuvalov, V. A. Nuclear wavepacket motion between P* and P+BA‘ potential surfaces with a subsequent electron transfer to HA in bacterial reaction centers. 1. Room temperature. Biochemistry 2002, 41, 2667-2674. (82) Shuvalov, V. A.; Yakovlev, A. G. Coupling of nuclear wavepacket motion and charge separation in bacterial reaction centers. FEBS Lett. 2003, 540, 26—34. (83) Sporlein, S.; Zinth, W.; Wachtveitl, J. Vibrational coherence in photosynthetic re- action centers observed in the bacteriochlorophyll anion band. J. Phys. Chem. B 1998, 102, 7492-7496. (84) Huppmann, P.; Spérlein, S.; Bibikova, M.; Oesterhelt, D.; Wachtveitl, J.; Zinth, W. Electron transfer in reaction centers of Blastochloris viridis: photosynthetic reactions approximating the adiabatic regime. J. Phys. Chem. A 2003, 107, 8302-8309. (85) Palaniappan, V.; Aldema, M. A.; Frank, H. A.; Bocian, D. F. Qy-excitation reso- nance Raman scattering from the special pair in extitRhodobacter sphaeroides reaction centers. Implications for primary charge separation. Biochemistry 1992, 31, 11050—11058. (86) Czamecki, K.; Diers, J. R.; Chynwat, V.; Erickson, J. R; Frank, H. A.; Bocian, D. F. Characterization of the strongly coupled, low-frequency vibrational modes of the special pair of photosynthetic reaction centers via isotopic labeling of the cofactors. J. Am. Chem. Soc. 1997, 119, 415-426. (87) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheese- man, J. R.; Montgomery, J. A. J.; Vreven, T.; Kudin, K. N.; Burant, J. _C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajirna, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, 0.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.; Salvador, 138 P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, 1.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Iaham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez, C.; Pople, J. A. Gaussian 03, Revision D.01; Gaussian, Inc.: Wallingford, CT, 2004. (88) Dillman, K L.; Shelly, K. R.; Beck, W. F. Vibrational coherence in polar solutions of Zn11 tetrakis(N-methylpyridyl)porphyrin with Soret-band excitation: rapidly damped intermolecular modes with clustered solvent molecules and slowly damped intramolecular modes from the porphyrin macrocycle. J. Phys. Chem. B 2009, 113, 6127-6139. DOI: 10.1021/jp807795x. (89) Lakowicz, J. R. Principles of Fluorescence Spectroscopy; Kluwer Aca- demic/Plenum Publishers: New York, 1999. (90) Diffey, W. M.; Beck, W. F. Rapid—scanning interferometer for ultrafast pump- probe spectroscopy with phase-sensitive detection. Rev. Sci. Instrum. 1997, 68, 3296-3300. (91) Cantor, C. R.; Schimmel, P. R. Biophysical Chemistry. Part [1: Techniques for the Study of Biological Structure and Function; W. H. Freeman and Company: San Francisco, 1980. (92) McHale, J. L. Molecular Spectroscopy; Prentice Hall: Upper Saddle River, New Jersey, 1999. (93) Cong, P.; Deuhl, H. R; Simon, J. D. Using optical coherence to measure the ultrafast electronic dephasing of large molecules in room-temperature liquids. Chem. Phys. Lett. 1993, 211, 367-373. (94) J00, T.; Jia, Y.; Yu, J. -Y.; Lang, M. J.; Fleming, G. R. Third-order nonlinear time domain probes of solvation dynamics. J. Chem. Phys. 1996, 104, 6089-6108. (95) Xu, Q.; Ma, Y.; Stiopkin, I. V.; Fleming, G. R. Wavelength-dependent resonant homodyne and heterodyne transient grating spectroscopy with a diffractive optics method: Solvent effect on the third-order signal. J. Chem. Phys. 2002, 116, 9333-9340. (96) Stratt, R. M. The instantaneous normal modes of liquids. Acc. Chem. Res. 1995, 28, 201-207. 139 (97) (98) (99) (100) (101) (102) (103) (104) (105) (106) (107) Stratt, R. M.; Maroncelli, M. Nonreactive dynamics in solution: the emerg- ing molecular view of solvation dynamics and vibrational relaxation. J. Phys. Chem. 1996, 100, 12981-12996. Siano, D. B.; Metzler, D. E. Band shapes of the electronic spectra of complex molecules. J. Chem. Phys. 1969, 51, 1856-1861. Vanden Bout, D.; Berg, M. Ultrafast Raman echo experiments in liquids. J. Raman Spectr. 1995, 26, 503-511. Berg, M.; Vanden Bout, D. A. Ultrafast Raman echo measurements of vibrational dephasing and the nature of solvent-solute interactions. Acc. Chem. Res. 1997, 30, 65—71. Yu, H. -Z.; Baskin, J. S.; Zewail, A. H. Ultrafast dynamics of porphyrins in the condensed phase: 11. Zinc tetraphenylporphyrin. J. Phys. Chem. A. 2002, 106, 9845-9854. DOI: 10.1021/jp0203999. Yoon, M. -C.; Jeong, D. H.; Cho, 8.; Kim, D.; Rhee, H.; Joo, T. Ultrafast tran- sient dynamics of Zn(II) porphyrins: observation of vibrational coherence by controlling chirp of femtosecond pulses. J. Chem. Phys. 2003, 118, 164-171. Enescu, M.; Steenkeste, K.; Tfibel, F.; Fontaine-Aupart, M. -P. Femtosec- ond relaxation processes from upper excited states of tetrakis(N-methyl-4- pyridyl)porphyrins studied by transient absorption spectroscopy. Phys. Chem. Chem. Phys. 2002, 4, 6092—6099. Jean, J. M.; Fleming, G. R. Competition between energy and phase relaxation in electronic curve crossing processes. J. Chem. Phys. 1995, 103, 2092-2101. Franzen, 8.; Fritsch, K.; Brewer, S. H. Experimental observation of anharmonic coupling of the heme-doming and iron-ligand out-of-plane vibrational modes confirmed by density functional theory. J. Phys. Chem. B 2002, 106, 11641- 11646. Morikis, D.; Li, R; Bangcharoenpaurpong, 0.; Sage, J. T.; Champion, P. M. Reso- nance Raman scattering as a probe of electron-nuclear coupling: applications to heme proteins. J. Phys. Chem. 1991, 95, 3391—3398. Ladanyi, B. M.; Stratt, R. M. Short-time dynamics of solvation: linear solvation theory for polar solvents. J. Phys. Chem. 1995, 99, 2502-2511. (108) Wilson, Jr., E. B.; Decius, J. C.; Cross, P. C. Molecular Vibrations: The Theory of Infrared and Raman Vibrational Spectra; McGraw-Hill: New York, 1955. 140 (109) Garg, S. K.; Smyth, C. P. Microwave absorption and molecular structures in liquids. LXII. The three dielectric dispersion regions of the normal primary al- cohols. J. Phys. Chem. 1965, 69, 1294-1301. (110) Jorgensen, W. L. Structure and properties of liquid methanol. J. Am. Chem. Soc. 1980, 102, 543—549. (111) Bertolini, D.; Cassettari, M.; Salvetti, G. The dielectric properties of alcohols- water solutions. I. The alcohol rich region. J. Chem. Phys. 1983, 78, 365—372. (112) Sumi, H.; Marcus, R. A. Dielectric relaxation and intramolecular electron trans- fers. J. Chem. Phys. 1986, 84, 4272—4276. (113) London, F. The general theory of molecular forces. Trans. Faraday Soc. 1937, 33, 8-26. (114) Margenau, H. Van der Waals forces. Rev. Mod. Phys. 1939, 11, 1—35. (115) Kauzmann, W. Quantum Chemistry: An Introduction; Academic Press: New York, 1957. (116) Berry, R. 8.; Rice, S. A.; Ross, J. Physical Chemistry; Oxford University Press: New York, 2000. (117) Iide D. R., Ed. CRC Handbook of Chemistry and Physics; CRC Press, Inc.: Boca Raton, Florida, 2006. (118) He, 2.; Sundstrom, V.; Pullerits, T. Excited states of carotenoid in LH2: an ab initio study. Chem. Phys. Lett. 2001, 334, 159-167. (119) Jin, R. -H.; Aoki, S.; Shima, K. Phosphoniumyl cationic porphyrins: self- aggregation origin from 11-11 and cation-11 interactions. J. Chem. Soc. Faraday Trans. 1997, 93, 3945-3953. (120) Albrecht, A. C. “Forbidden” character in allowed electronic transitions. J. Chem. Phys. 1960, 33, 156-169. (121) Tang, J.; Albrecht, A. C. Developments in the theories of vibrational Raman intensities. In Raman Spectroscopy: Theory and Practice; Szymanski, H. A., Ed.; Plenum: New York, 1970; Vol. 2, pp 33-68. (122) Arnett, D. C.; Vohringer, P.; Scherer, N. F. Excitation dephasing, product forma- tion, and vibrational coherence in an intervalence charge-transfer reaction. J. Am. Chem. Soc. 1995, 117, 12262-12272. 141 (123) Yang, T. -S.; Chang, M. -S.; Chang, R.; Hayashi, M.; Lin, S. H.; Vohringer, P.; Dietz, W.; Scherer, N. F. Femtosecond pump-probe study of molecular vibronic structures and dynamics of a cyanine dye in solution. J. Chem. Phys. 1999, 110,12070-12081. (124) Liebl, U.; Lipowski, G.; Negrerie, M.; Lambry, J. -C.; Martin, J. -L.; Vos, M. H. Coherent reaction dynamics in a bacterial cytochrome c oxidase. Nature 1999, 401,181-184. (125) Deisenhofer, J.; Epp, O.; Mild, K; Huber, R.; Nfichel, H. Structure of the protein subunits in the photosynthetic reaction center of Rhodopseudomonas viridis at 3.1 A resolution. Nature 1985, 318, 618-624. (126) Deisenhofer, J.; Michel, H. The photosynthetic reaction center from the purple bacterium Rhodopseudomonas viridis. Science 1989, 245, 1463-1473. (127) Blom, N.; Odo, J.; Nakamoto, K; Strommen, D. P. Resonance Raman studies of metal tetrakis(4-N-methylpyridyl)porphine: band assignments, structure- sensitive bands, and species equilibria. J. Phys. Chem. 1986, 90, 2847-2852. DOI: 10.1021/j100404a015. (128) Spence, T. G.; Trotter, B. T.; Burns, T. D.; Posey, L. A. Metal-to-ligand charge transfer in the gas-phase cluster limit. J. Phys. Chem. A 1998, 102, 6101—6106. (129) Vanderkooi, J. M.; Adar, F.; Ericir'lska, M. Metallocytochromes c: characteri- zation of electronic absorption and emission spectra of Sn‘“ and Zn4+ cy- tochromes c. Eur. J. Biochem. 1976, 64, 381—387. (130) Elias, H.; Chou, M. H.; Winkler, J. R. Electron-transfer kinetics of Zn-substituted cytochrome c and its Ru(NH3)5(I-Iistidine-33) derivative. J. Am. Chem. Soc. 1988, 110, 429-434. (1 31) Ye, 5.; Shen, C.; Cotton, T. M.; Kostié, N. M. Characterization of zinc-substituted cytochrome c by circular dichroism and resonance Raman spectroscopic meth- ods. J. Inorg. Biochem. 1997, 65, 219—226. (132) Goto, Y.; Calciano, L. J.; Fink, A. L Acid-induced folding of proteins. Proc. Natl. Acad. Sci. U.S.A. 1990, 87, 573—577. (133) Tremain, S. M.; Kostic, N. M. Molten-globule and other conformational forms of zinc cytochrome c. Effect of partial and complete unfolding of the protein on its electron-transfer reactivity. Inorg. Chem. 2002, 41, 3291-3301. (134) Nappa, M.; Valentine, J. S. The influence of axial ligands on metalloporphyrin visible absorption spectra. Complexes of tetraphenylporphynatozinc. J. Am. Chem. Soc. 1978, 100, 5075-5080. 142 (135) Humphry-Baker, R.; Kalyanasundaram, K Influence of axial ligation on the flu- orescence of tetrakisphenylporphyrins. J. Photochem. 1985, 31, 105-112. (136) Vogel, G. C.; Beckmann, B. A. Binding of pyridine to phenyl-substituted deriva- tives of zinc tetraphenylporphine. Inorg. Chem. 1976, 15, 483—484. (137) Scheidt, R. W. Porphyrin stereochernistry. In The Porphyrins“, Dolphin, D., Ed.; Academic Press: New York, 1978; Vol. 3, pp 463-511. (138) Fleischer, E. B.; Miller, C. K.; Webb, L. E. Crystal and molecular structures of some metal tetraphenylporphines. J. Am. Chem. Soc. 1964, 86, 2342-2347. (139) Schauer, C. K; Anderson, 0. P.; Eaton, 5. 5.; Eaton, G. R. Crystal and molecular structure of a six-coordinate zinc porphyrin: bis(tetrahydrofuran)(5,10,15,20- tetraphenylporphinato)zinc(ll). Inorg. Chem. 1985, 24, 4082—4086. (140) Scheidt, R. W.; Eigenbrot, C. W.; Ogiso, M.; Hatano, K. Stereochemistry of a por- phyrin atropisomer. The molecular and crystal structure of six-coordinate [501, 1OB-bis(o-nicotinamidophenyl)-l5,20-diphenylporphinato] zinc(II). Bull. Chem. Soc. Jpn. 1987, 60, 3529-3533. (141) Lampa-Pastirk, 5.; Iafuente, R. C.; Beck, W. F. Excited-state axial-ligand photodissociation and nonpolar protein—matrix reorganization in Zn(II)- substituted cytochrome c. J. Phys. Chem. B 2004, 108, 12602-12607. DOI: 10.1021/jp049587k. (142) Damaschun, G.; Gernat, C.; Damaschun, H.; Bychkova, V. E.; Ptitsyn, O. B. Com- parison of intramolecular packing of a protein in native and ‘molten globule’ states. Int. J. Biol. Macromol. 1986, 8, 226-230. (143) Ptitsyn, O. B. The molten globule state. In Protein Folding; Creighton, T. E., Ed.; W. H. Freeman and Company: New York, 1993; pp 243-300. (144) Kataoka, M.; Hagihara, Y.; Mihara, K.; Goto, Y. Molten globule of cytochrome c studied by small angle x-ray scattering. J. Mol. Biol. 1993, 229, 591—596. (145) Hamada, D.; Kuroda, Y.; Kataoka, M.; Aimoto, 5.; Yoshimura, T.; Goto, Y. Role of heme axial ligands in the conformational stability of the native and molten globule states of horse cytochrome c. J. Mol. Biol. 1996, 256, 172-186. (146) Kamiyama, T.; Sadahide, Y.; Nogusa, Y.; Gekko, K. Polyol-induced molten glob- ule of cytochrome c: an evidence for stabilization by hydrophobic interaction. Biochim. Biophys. Acta 1999, 1434, 44-57. 143 (147) Lyubovitsky, J. G.; Gray, H. B.; Winkler, J. R. Structural features of the cy- tochrome c molten globule revealed by fluorescence energy transfer kinetics. J. Am. Chem. Soc. 2002, 124, 14840-14841. (148) Bongiovanni, C.; Sinibaldi, F.; Ferri, T.; Santucci, R. Glycerol-induced formation of the molten globule from acid-denatured cytochrome c: Implication for hier- archical folding. J. Protein Chem. 2002, 21, 35-41. (149) Anni, H.; Vanderkooi, J. M.; Mayne, L. Structure of zinc-substituted cy- tochrome c: nuclear magnetic resonance and optical spectroscopic studies. Biochemistry 1995, 34, 5744—5753. (150) Qian, C.; Yao, Y.; Tong, Y.; Wang, J.; Tang, W. Structural analysis of zinc- substituted cytochrome c. J. Biol. Inorg. Chem. 2003, 8, 394—400. (151) Teczan, F. A.; Crane, B. R.; Winkler, J. R.; Gray, H. B. Electron tunneling in protein crystals. Proc. Natl. Acad. Sci. U.S.A. 2001, 98, 5002—5006. (152) Berghuis, A. M.; Brayer, G. D. Oxidation state-dependent conformational changes in cytochrome c. J. Mol. Biol. 1992, 223, 959-976. (153) Pielak, G. J.; Auld, D. 5.; Betz, S. F.; Hilgen-Willis, 5. E.; Garcia, L. L. Nuclear magnetic resonance studies of class I cytochromes c. In Cytochrome c: a Multi- disciplinary Approach; Scott, R. A., Mauk, A. G., Eds.; University Science Books: Sausalito, California, 1996; pp 203-284. 144 "11111111111111fililllsllliflliill“