u ‘4‘ 3‘? '(Z ‘v . .. .7“. . - . :95”: II¢>.‘_' '. A.‘ . “9'" ‘1’? :var :. .1 I «h ‘ .I-n, .u‘o'i‘e‘a‘l“ u _ » has. ‘ “I?" ‘3 N ' y. ‘ a .I 4 \v 0‘ . ‘ ‘10.". ..‘ 121’ Ah - it to." r m . . u p .. . .‘m “ 5' (k '44 .v» a 5‘: . r. . ‘ ~»-. ' «1": 1.;ti'a‘1' u . n. . $13, . 1‘4 51' ‘1 ‘ 44"}. "0' “5"" w- , ‘0 ,. . x \l . av; 3a 4’5) Ilr E ‘95. “PS 1’. f3; ,. . u «‘5 fl ., n r fi‘rL‘ \J ' w ‘ 59* M (‘l' “‘5'.“ ‘ ._ :‘ .‘ffL. ' . .35. 3., .~ r131???» will? ? C. :.<‘ 'v“"‘% . :fifvflkfl'nrw “5"}. y'v'lr'l'fi. 'v . 1&an ‘tlf'q'k "h" "“' _ ‘ ’5. 0 .‘¢ k ‘\ . 13.?“ *~ ‘ ' .. 4. 3‘0?" \ , .. Mk *1 , ‘. ’ ‘I em “ ‘ £11 563?: V i .3} .‘“S' "‘59. l “' Y ' ' J 3'1? ".‘5‘. . A y a; :10 ,J‘. ~.> '5 aria m’%‘~-& kl” Jr “5“." '" ~ )- ‘g, :x‘goggrminza‘fi 'g tag). :93?» 3,! Wrrfik‘! -" e 1‘,‘ 2’. 1.. ~ ' sérv‘m ‘ ’ ‘ “3391439,, 7 {1‘ ‘ er.” ‘5’ ’53“). u. 53:" 52.2 - . ‘~_ fin“ ”‘3’“ ‘3 , ‘x‘.’"=:e"";€7§21 “ r3 6" man-53‘: A 4. 1 n ‘1\ ~ «VS-A'U-it.‘ ‘4; v w ‘ Aw g. ".4512, tam; .r» 3W k 4-. afigggé , . .’ at . L3 in!" ... “ms-3&5" r.£&§%fl4}£fi¢ ‘ "5031»? }‘*:$\ . '¢;:.:.A\‘1§1 ‘ AM”: a 3% werksfi'w‘i'vfi' ‘zzaaw‘u ' hr? ‘4" J V", 0 I \ V), V I 1}" _, _‘ :- ‘Wx’i‘: “3' ‘5 'v' ‘ .1 . llllllllllllllllllllllllllll 3 1293 01399 555 This is to certify that the dissertation entitled Ultrafast Stimulated£Spectroscoby Studies of Vibrational Relaxation and Short Range Solvent Organization in Organic Solutions presented by Ying Jiang has been accepted towards fulfillment of the requirements for Ph . D degree in Chemis t ry fiMZ/w‘g Date %/ / d: 7/145, MS U is an Affirmative Action/Equal Opportunity Institution 042771 LEERARY Michigan State University PLACE It RETURN BOXto mnovothb checkout om your record. TO AVOID FINES rotum on or baton date due. DATE DUE DATE DUE DATE out ULTRAFAST STMULATED SPECTROSCOPY STUDIES OF VIBRATIONAL RELAXATION AND SHORT RANGE SOLVENT ORGANIZATION IN ORGANIC ' SOLUTIONS by YING JIANG A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirement for the degree of DOCTOR OF PHILOSOPHY DEPARTMENT OF CHEMISTRY 1995 ABSTRACT ULTRAFAST STIMULATED SPECTROSCOPY STUDIES OF VIBRATIONAL RELAXATION AND SHORT RANGE SOLVENT ORGANIZATION IN ORGANIC SOLUTIONS ' By YING JIANG Understanding solute-solvent molecular interactions and local solvent organization are of great importance for chemical reactions because the solution phase is the most widely used medium for chemical reactions. In solution, however, molecules are not isolated from one another like in the gas phase and are not spatially fixed like in the solid phase. Chemically important events take place on the femtosecond and picosecond time scale, and these processes depend sensitively on intermolecular interactions. Organization and intermolecular interactions are, at present, not well understood in the liquids. This thesis focuses on understanding solute-solvent interactions and local solvent organization through studies of the vibrational energy relaxation and rotational diffusion dynamics of fluorescent probe molecules. Subsequent to excitation and any optical emission, virtually all of the excess of energy in the system is dissipated as heat or vibrational energy. The 7‘! redistribution of this excess vibrational energy into surrounding molecules depends crucially on the environment surrounding the probe molecule. A novel laser technique has been developed to study vibrational energy relaxation dynamics in dilute solutions. A pump—probe strategy is used to monitor the stimulated emission response of probe molecule in dilute solutions. The measurement scheme can be modeled as a coupled three level system, where the levels are the vibrationless electronic ground state, the vibrational state of interest in electronic ground state and the vibrationless electronic excited state. The pump laser is operated at the frequency of 0-0 transition, and the probe laser is operated at a frequency corresponding to the difference between the pump laser and the vibrational state of interest. The stimulated response is S(t) = - a exp(-t/T1) + b exp(- t/Tclcc). The vibrational population relaxation times, T1, of four perylene vibrational modes were measured in both polar and nonpolar solvents using this technique. We found that T; times range from <10 ps to a few hundred picoseconds and are strongly mode- and solvent-dependent. Measuring perylene T1 times as a fimction of aliphatic chain length in a series of normal alkanes revealed the presence of solvent organization on a few A length scale. Comparison of the vibrational energy relaxation and rotational diffusion dynamics of perylene and l-methylperylene provides information on the persistent length of the local solvent organization. For perylene in the n-alkanes, the exchange of vibrational energy proceeds through in quadrupole-quadrupole interactions. For perylene, T1 relaxation and rotational diffusion measurements do not correlate in an obvious way. For 1- methylperylene, the solute-solvent vibrational energy exchange is through dipole- quadrupole interactions, which operate over a longer range (cc r“) than quadrupole- quadrupole interactions (at r’7). For l-methylperylene, there is a direct correlation between T1 and rotational diffusion dynamics. To China, where I was born and grew up, where my parents, sister, brother and their families are living. iv ACKNOWLEDGMENTS I am immensely grateful to my advisor, Dr. Gary Blanchard, for providing me with an intellectual and pleasant environment, for the financial support and enlightening advice and inestimable training. Perhaps the most valuable thing I learned from him is how enjoyable science can be. I also want to thank my committee members, Dr. Bob Cukier, Dr. Greg Baker and Dr. Jeff Ledford for their stimulating and helpful discussion through the course of years. My thanks also go to Dr. Tom Carter, MSU laser lab manager, who helped me with the lasers and the spectroscopy without reservation. Without the cooperative help of Ron. Tom and Scott in the Electronic shop this thesis may not have been possible. It is impossible to separate my thesis work from my colleagues in the group. Selezion Hambir taught me how to take care of my worries and gave me a lot of help and suggestions for my research work. It will be a precious memory for me to think of Lee Dewitt, Pat McCarthy, Dave Karpovich, Jeff Rasimas, Jen Horn and their willingness to share their knowledge, expertise and life experience (family stories, traveling, wild life in the woods and hunting...) with me. I am specially indebted to Geurt for his continuous encouragement and support during the composition of this thesis. Finally I am grateful for the grant from National Science Foundation that made thesis possible. TABLE OF CONTENTS ABSTRACT LIST OF TABLES LIST OF FIGURES Page CHAPTER 1. INTRODUCTION ............................ 1 CHAPTER 2. EXPERIMENTAL ........................... 11 CHAPTER 3. ULTRAFAST STIMULATED EMISSION SPECTROSCOPY 0F PERYLENE IN DILUTE SOLUTION - MEASUREMENT OF GROUND STATE VIBRATIONAL POPULATION RELAXATION ............................. 18 Summary 3.1. Introduction 3.2. Experimental 3.3. Results and Discussion 3.4. Conclusion 3.5. Literature cited CHAPTER 4. VIBRATIONAL POPULATION RELAXATION OF PERYLENE IN ITS GROUND AND EXCITED ELECTRONIC STATES ....... 49 Summary 4.1. Introduction 4.2. Theory 4.3. Experimental 4.4. Results and Discussion 4.5. Conclusion 4.6. Literature cited CHAPTER 5. VIBRATIONAL POPULATION RELAXATION OF PERYLENE IN n- ALKANES - THE ROLE OF THE LOCAL SOLVENT ORGANIZATION IN LONG RANGE VIBRATIONAL ENERGY TRANSFER ...... 68 Summary 5.1. Introduction 5.2. Experimental 5.3. Results and Discussion 5.4. Conclusion 5.5. Literature cited CHAPTER 6. ROTATIONAL DIFFUSION DYNAMICS OF PERYLENE IN n - ALKANES - OBSERVATION OF SOLVENT LENGTH DEPENDENT CHANGE OF BOUNDARY CONDITION ............... 92 Summary 6.1. Introduction 6.2. Background 6.3. Experimental 6.4. Results and Discussion 6.5. Conclusion 6.6. Literature cited CHAPTER 7. VIBRATIONAL POPULATION AND ORIENTATIONAL RELAXATION DYNAMICS OF l-METHYLPERYLENE IN n- ALKANES - THE EFFECTIVE RANGE OF DIPOLAR ENERGY RELAXATION IN SOLUTION .................... 114 Summary 7.1. Introduction 7.2. Experimental 7.3. Results and Discussion 7.4. Conclusion 7.5. Literature cited CHAPTER 8. SUMMARY AND FUTURE WORK .................. 147 vii 2.1. 3.1. 3.2. 4.1. 4.2. 5.1. 6.1. 7.1. 7.2. 7.3. 7.4. LIST OF TABLES Information for each dye laser: including making of the dye solutions, operating wavelength range and the mirror sets. ......... ‘ ........ 13 q The pump, probe wavelengths and probed vibronic region for each solvent ..... 2.: Build-up and decay times determined from the experimental data. All times are given in ps. The vibrational frequencies refer to the final state in the stimulated transition. The spectral origin for each solvent was estimated from the static spectroscopic data ......................... 39 The pump, probe wavelengths and dyes for each solvent .............. 55 Vibrational population relaxation times for the ground state and excited state v7 mode of perylene in n-alkane solvents ................... 61 T1 times for the V7 and (v7 + V”) modes of perylene in n-alkanes .......... 78 Experimental zero time anisotropies and reorientation times for perylene in several n-alkanes. The asterisksindicate an excited electronic state measurement. ..... q ................. . ............ 99 Spectral origin of l-methylperylene in the solvents used in this work, determined from the linear optical response .................... 122 T1 relaxation times for the l-methylperylene 1370 cm'1 mode in the n-alkane solvents .................................. 126 Experimental rotational diffusion time constants and zero-time anisotropies. All uncertainties are reported as standard deviations (i 10). . ‘. . . 138 Cartesian components of the rotational diffusion constant extracted from experimental data using Equations 5 ........................ 141 viii 2.1. 3.1. 3.2. 3.4. 3.6. 3.7. LIST OF FIGURES Schematic of the stimulated emission pump-probe spectrometer ......... Absorption and emission spectra for perylene in n-octane. The boxed emission feature indicates the Spectral region over which time resolved .12 stimulated emission spectra were recorded ..................... 25 Time resolved stimulated emission spectra of perylene in n-octane over the spectral region indicated in Figure l. (o) = 10 ps delay, (0) = 100 ps, (V) = 200 ps, (inverted filled triangle) = 300 ps, (Cl)= 400 ps, (I) = 500 ps, (A) = 600 ps, (filled triangle) = 700 ps, (0) = 800 ps, (9) = 900 ps ....... 26 . Time resolved stimulated emission spectra of perylene in l-butanol. (o) '= 10 ps delay, (0) = 100 ps, (V) = 200 ps, (inverted filled triangle) = 300 ps, (El)= 400 ps, (I) = 500 ps, (A) = 600 ps, (filled triangle) = 700 ps, (0) = 800 ps, (9) = 900 ps ............................ Time resolved stimulated emission spectra of perylene in l-octanol. (o) = 10 ps delay, (0) = 100 ps, (V) = 200 ps, (inverted filled triangle) = 300 ps, (El)= 400 ps, (I) = 500 ps, (A) = 600 ps, (filled triangle) = 700 ps, (0) = .27 800 ps, (0) = 900 ps ................................. 28 . Time resolved stimulated emission Spectra of perylene in DMSO. (0) = 10 ps delay, (0) = 100 ps, (V) = 200 ps, (inverted filled triangle) = 300 ps, (Cl)= 400 ps, (I) = 500 ps, (A) = 600 ps, (filled triangle) = 700 ps, (0) = 800 ps, (9) = 900 ps ................................ 29 Time resolved stimulated emission spectra of perylene in toluene. (o) = 10 ps delay, (0) = 100 ps, (V) = 200 ps, (inverted filled triangle) = 300 ps, (D): 400 ps, (I) = 500 ps, (A) = 600 ps, (filled triangle) = 700 ps, (0) = 800 ps, (9) = 900 ps ................................ 3O Schematic of coupled three level system used to model the experimental data. The terms k are described in the text .................... ix 33 3.8. 3.9. 4.1. 4.2. 4.3. 4.4. 4.5. 5.1. 5.2. LIST OF FIGURES (Continued) Time scans for perylene in n-octane at probe wavelengths corresponding to distinct vibronic resonances, where the ground state vibrational level is indicated Build-up and decay times are given for these data in Table 3.2. . . . Time scans for different IMW ratio ...................... (a) Schematic of the coupled three-level system used for interpretation of 0-0 excitation experiments. (b) Schematic of the coupled four-level system used for .38 41 interpretation of excited state T1 measurements .................. 52 Linear optical response of perylene in n-hexane. The absorption and emission - spectra have been normalized. Arrow “a” and “b” indicate the excitation wavelengths used for the 0-0 and v7* experiments, respectively .......... 57 Schematic of the exaggerated atomic displacements for the perylene v7 mode. The directions of the displacements were estimated from semiempirical calculation results ................................. Experimental stimulated response and laser cross-correlation for measurement of the ground state v7 mode of perylene in n-hexane. For this experiment, AP.” = 432 nm and AW” = 462 nm. ........................ (3) Experimental stimulated response for the 0-0 and v7* excitation of perylene. For these scans the ground state v5 mode of perylene in n-hexane is probed; A,” = 432 nm for 0-0 excitation and 409 nm for v7* excitation and Ami” = 466 nm for both excitation conditions. See Table 3.1 for the best fit results. (b) Difference signal, S(t), for the two scans shown in (a), with the best fit 58 6O fiInction shown as a solid line through the data ................. 62 Absorption and emission Spectra for perylene in n-octane .............. 73 Exaggerated atomic displacements associated with the v7 and v1, normal modes of perylene ..................................... 74 5.3. 5.4. 5.5. 6.1. 6.2. 6.3. 6.4. LIST OF FIGURES (Continued) Stimulated response of the (v-, + V”) mode of perylene in n-octane, presented with the cross correlation response function .................... 76 T1 relaxation times for v7 (0) and “(v7 + v15) (Cl) modes in perylene as a function of solvent chain length ............................... 77 Calculated probability, <

>, for long-range energy transfer for exact donor- acceptor resonance, 0) = 0 (Equation 5) and co = 300 cm'l, as a function of donor-acceptor separation, d ................ ' .......... y . . 85 Normalized absorption and spontaneous emission spectra of perylene in n-octane ..................................... 98 (a) Time scans for ground state recovery response of perylene in n-octane. The laser cross-correlation is presented with the time scans for pump and probe electric-fields polarized parallel (larger AT/T at early time) and perpendicular to one another. (b) R(t) signal produced from the experimental data shown in (a) using Equation 3. These data fit best to a single exponential decay functionality, as indicated by the dashed line .......................... 100 (a) Time scans for ground state recovery response of perylene in n-hexadecane. The laser cross-correlation is presented with the time scans for pump and probe electric fields polarized parallel (larger AT/T at early time) and perpendicular to one another. (b) R(t) signal produced from the experimental data shown in (a) using Equation 3. These data fit best to a single exponential decay 'fiinctionality, as indicated by the dashed line .................... 101 Ground state and excited state reorientation times for perylene as a function of n-alkane solvent viscosity. For all measurements the ground state and excited state reorientation times are the same to within the experimental uncertainty. See text for a discussion of these data ........................ 103 6.5. 6.6. 7.1. 7.2. 7.3. 7.4. 7.5. 7.6. 7.7. 7.8. LIST OF FIGURES (Continued) Illustration of prolate and oblate ellipsoid rotor shape and the axial ratio, p. . . . 105 Dimensions and Cartesian axis assignments for perylene .............. 108 Steady-state absorption and emission spectra of l-methylperylene in n-hexadecane. The arrow indicates the 0-0 transition energy, and the box the range over which the stimulated response, shown in Figure 7.3, was taken. . .. . .. . . . . . . . . .121 (a) Infrared and (b) Raman spectra of l-methylperylene. The asterisks indicate the vibrational resonance for which we determined T1 times ............ 123 The time-resolved stimulated spectra of 1-methylperylene in n-hexadecane, from 21978 cm’1 to 21231 cm“. This range corresponds to vibrational frequencies between 1170 cm" and 1920 cm"1 ......................... 125 The stimulated response of 1-methylperylene (1370 cm‘1 mode) in eight alkanes. The instrumental response function indicates zero delay time ........... 127 Solvent chain length dependence of the l-methylperylene 1370 cm'1 ’1‘, relaxation time ......................................... 129 (a) Tail-matched parallel and perpendicular intensity data and (b) anisotropy decay, R(t), of 1-methylperylene in n-pentane ....................... 136 (a) Tail-matched parallel and perpendicular intensity data and (b) anisotropy decay, R(t), of l-methylperylene in n-hexadecane ..................... 137 Orientational relaxation time, 10,, as a function of solvent viscosity. For the long chain alkanes (n-nonane to n-hexadecane), two exponential decays are found in R(t) ......................................... 139 CHAPTER 1. INTRODUCTION Dissimilar molecules interact with one another in the liquid phase according to their chemical compositions and conformations. Gaining a detailed understanding of these interactions has attracted significant research attention. Liquids are by far the most commonly used medium for performing chemical syntheses and analyses, and gaining predictive control over molecular interactions in liquids would be of great value to a large portion of the chemical community. Unfortunately, the moderately strong intermolecular interactions that give rise to the existence of the liquid phase are difficult to probe experimentally because of the lack of either long range or long time organization in this medium. For a given molecule in solution, there exist many energetically similar solvent “cage” configurations, and exchange between these configurations occurs at a rate that is fast compared to almost all experimental measurement schemes. Despite this inherent structural complexity, there is a great deal of steady state spectroscopies and solubility data that demonstrates the existence of highly Specific interactions between dissimilar molecules. Gaining an accurate picture of local organization in liquids and relating this information to macroscopic properties, such as solubility or reactivity, continues to be an area open to investigation and debate. Themost common experimental approach to the detection and characrerization of local organization in solution is to interrogate, in some manner, the spectroscopic response of a probe molecule dissolved in the liquid of interest. Because the probe molecule is different from the liquid, its presence necessarily disrupts any local organization of the solvent, and it is not possible to interrogate intrinsic solvent intermolecular interactions directly with probe molecules. The systems of interest to us are solutions of either chemically reactive or spectroscopically active molecules, where the interactions between dissimilar molecules define the important properties of the system. There are a variety of measurement schemes that have used spectroscopically active molecules as probes of solvent organization, and each of these approaches senses a different component of the probe molecule local environment. One can divide these measurements into two broad categories; molecular motion and energy relaxation. Measurements of probe molecule rotational diffusion in solution have shown that, when the probe molecule is large compared to individual solvent molecules, there is little need to account for specific intermolecular interactions, and the interaction between solvent and solute can be treated as largely friCIional. This limit. described by the Debye-Stokes— ”"91 and with Einstein equation,” has been shown to be valid for many polar systems. appropriate treatment of the friction coefficient.'2"'22' for nonpolar systems as well.'23'25] For solvent-solute systems where the molecular volume of the solvent and solute moieties are similar, the molecular nature of solvation processes must be taken into account. In this thesis the work on the orientational relaxation dynamics of perylene in n-alkanes shows that, as the hydrodynamic volumes of the solvent and the solute become Similar, the b) solvent cage formed around the solute can alter the ability of the solute to rotate about specific axes. Such data do not necessarily point to any specific structure within the solvent cage, but do demonstrate that the motional freedom of the solute can be predicted by relatively intuitive models. Under selected circumstances, such as those included in this thesis, one can gain insight into the shape of the volume swept out by the reorienting molecule. For this comparatively special case, we can gain some insight into the average organization of solvent molecules around the solute. - - - 6-30 - Energy relaxation measurements, such as excrtation transport,‘2 ' transrent spectral 4.50 - - - H l provrde information on energy shiftm'm and vibrational population relaxation, dissipation both within the probe molecule and between the probe molecule and the surrounding medium. Excitation transport measurements have been used to determine whether or not diffusive behavior dominates at short times or low concentrations in solution and, as such, have placed limits on the ability to treat energy relaxation events [27.29] statistically. One of the more popular techniques for measuring “solvation” times in liquids has been detection of the dynamic spectral shifts exhibited by modified 35.36.314.40 . . . . l ' In these measurements, the evolution Oftlie coumarin emlSSlOl'l band coumarins. is monitored after excitation with a short light pulse. The timescale of this spectral relaxation has been correlated with the bulk dielectric relaxation time(s) of the solvents examined. Recent experimental and computational work has shown that the spectral relaxation behavior of the coumarins is dominated by intramolecular relaxation between 53' . . 51.- . . . several overlapped electronic manifolds' The experimental Signature of this intramolecular relaxation is a pronounced excitation energy dependence of the coumarin emission band dynamics. Despite the experimental difficulties associated with using coumarins as probes of solvation, much valuable information on solvent relaxation has been gained from these experiments. The physical “picture” of solvent dynamics developed to explain these data is appealing and will likely be proven correct. if a probe molecule with a sufiiciently simple spectroscopic response can be found. In addition to using probe molecule electronic relaxation dynamics to interrogate local solvent organization, vibrational population relaxation has found its use for this purpose as welll‘w'w' Using vibrational relaxation dynamics to interrogate local solvent organization is explored in this thesis. The motivation for using vibrational states instead of electronic states stems from the comparatively short length scale over which vibrational energy transfer processes operate and intrinsic directionality of molecular vibrational motions. Numerous studies have shown that there is indeed Short range order in liquids.'5‘”’”' For instance, in liquids composed Of molecules with an anisotropic shape. such as the long- chain n-.alkanes, there are thermodynamic effects associated with the presence of a short- range molecular order in which the more extended conformations are stabilized by a cooperative effect. There is some spectroscopic evidence that gauche conformations are [61 relatively scarce ' and depolarized Rayleigh scattering in liquids shows that orientational (12 ' ' The pressure dependence Of order is higher in normal alkanes than in branched alkanes, the excess enthalpy, dHE/dP has been used to illustrate order destruction and order creation in liquids by E. Aicart eta/.156] The discrepancy between measured dHE/dP values and calculated values is due to the presence of short-range orientational order in the higher n-alkane liquids which makes dH/dP more negative and which, upon mixing, is destroyed, producing a positive contribution to dH/dP not accounted for by theory. Snyder‘s” found that, in the case of n-alkanes, the observed C-H stretching frequencies tend to fall in clusters that are regularly spaced with an average separation of about 145:1 cm". The clustering occurs because the isolated C-H stretching frequencies are determined by the '58] and structure of the n-alkanes in the immediate vicinity of the C-H bond. Ohtaki Marcusm’ made an attempt to paramerize the “structuredness” of a solvent from the viewpoint of intermolecular interactions using the structuredness parameter Sp. Stengle er aim” used the NMR chemical shift of Xe(l) to probe liquid structure. The Xe nucleus has a spin 1 = 3/2', it has an electric quadrupole moment which causes short relaxation times and leads to broad NMR lines. The relaxation rate is sensitive to the environment in a way that differs from the chemical shift. In this thesis, a novel pump-probe measurement scheme to detect the vibrational population relaxation dynamics of dilute fiuorophores in solution is developed. We have chosen the chemical system carefully so that the vibrational energy relaxation rate reflects the local solvent organization, and is not dominated by intramolecular processes. Chapter 2 describes the pump-probe laser experimental set-up we use to measure vibrational relaxation of probe molecules in dilute solution, In Chapter 3, the stimulated emission measurement scheme is discussed extensively and the vibrational relaxation of four modes of perylene in various solvents is presented. Chapter 4 demonstrates the capability of measuring T. in both electronic ground and excited states using this technique. In Chapters 5, 6 and 7 the focus is on the studies of vibrational energy relaxation and rotational diffusion dynamics of perylene and l-methylperylene in a series of normal alkanes. Information on the solvent local organization and the nature of intermolecular vibrational energy exchange is extracted from these data. Literature Cited 1. P. Debye; Polar Solvents, Chemical Catalog CO., New York, p. 84 (1929). 2. M. J. Sanders; M. J.Wirth; Chem. Phys. Lett, 101 361 (1983). b) . E. F. Gudgin-Templeton; E. L. Quitevis; G. A. Kenney-Wallace; J. Phys. Chem, _8_9_ 3238 (1985). 4. A. Von Jena; H. E. Lessing; Chem. Phys, 40 245 (1979). 1J\ . AVon Jena; H. E. Lessing; Ber. Bunsen-Gm: Phys. Chem, Q 181 (1979) 6. A. Von Jena; H. E. Lessing; Chem. Phys. Let/., E 187 (1981). 7. K. B. Eisentha1;Acc. Chem. Res., 8 118 (1975). 8. G. R. Fleming; J. M. Morris; G. W. Robinson; Chem. Phys., 1_7 91 (1976). 9, C. V. Shank; E. P. lppen; Appl. Phys. Lett, E 62 (1975) 10. D. P. Millar; R. Shah; A. H. Zewail, ('hem. Plots. l.etl., 99 435 (1979). 11. E. F. Gudgin-Templeton; G. A. Kenney-Wallace. .1. Phys. Chem, fl) 2896 (1986). 12. G. J. Blanchard; M. J. Wirth, ./. Phil's. Chem, 9_ 2521 (1986). 13. G. J. Blanchard, J. Chem. Phil-5., §_7 6802 (1987). 14. G. J. Blanchard; C. A. Cihal; .l. Phys. ('hem., 22 5950 (1988). 15. G. J. Blanchard; J. Phys. Chem, 9;, 6303 (1988). 16. G. J. Blanchard; J. Phys. Chem, fl 4315 (1989). 17. G. J. Blanchard;; Anal. Chem, _6_1_ 2394 (1989). 18. D. S. Alavi; R. S. Hartman; D. H. Waldeck; J. Phys. Chem, fi 6770 (1991). 19. R. S. Hartman; D. S. Alavi; D. H. Aldeck; J. Phys. Chem, 5 7872 (1991). 20. C. M. Hu; R. Zwanzig;°J. Chem. Phys, @ 4354 (1974). 21. G. K. Youngren; A. Acrivos; J. Chem. Phys, 6_3 3846 (1975). 22. R. Zwanzig; A. K. Harrison; J. Chem. Phys, 88 5861 (1985). 23. D. Ben-Amotz; T. W. Scott, J. Chem. Phys, 8_7 3739 (1987). 24. D. Ben-Amotz; J. M. Drake; J. Chem. Phys, E 1019 ( I988). 25. Y. Jiang; G. J. Blanchard; J. Phys. Chem, 98 6436 (1994). 26. Th. Forster; Ann. Phys. Lie/121g, 2_55 (1948). 27. C. H. Gochanour; H. C. Andersen; M. D. Fayer. .l. Chem. thx, m 4254 (1979). 28. S. G. Fedorenko; A. I. Burshtein; Chem. Phi/.81, % 341 (1985). 29. P. A. Anfinrud; w. s. Struve; J. Chem. Phys, E 4256 (1987). ’ 30 T. P. Causgrove; S. Yang; W. S. Struve; .1. Phys. Chem, 98 6844 (1989). 31. A. Mokhtari; J. Chesnoy; A Laubereau, Chem. Phys. l.eu.. 155 593 (1989). 32 A. Wagener; R. Richert. Chem. l’lll‘x. l.e/I.. 1_76 329 (1991) 33. A. Declemy; C. Rulliere; Ph. Kottis; Chem. Phys. l.eII., 133 448 ( I987). 34. E. W. Castner; M. Maroncelli; G. R. Fleming, J. Chem Plum, & 1090 (1987) 35. V. Nagarajan; A. M. Brearly; T. J. Kang; P. F. Barbara; .1. Chem. Phys, _8_6 3183 (1987) 9 36. M. Maroncelli; G. R. Fleming; J. Chem. Phys, & 6221 (1987). 37. J. D. Simon; Acc. Chem. Res, _2_1_ 128 (1988). 38. P. F. Barbara; W. Jarzeba; Adv. Photochem, 1_5 l (1990). 39. M. Maroncelli; R. S. Fee; C. F. Chapman; G. R. Fleming; J. Phys. Chem, fi 1012 (1991) 40. W. Jarzeba; G. C. Walker; A. E. Johnson; P. F. Barbara; Chem. Phys, 1-2 57 (1991). 41. T. Elsaesser; W. Kaiser; Annu. Rev. Phys. Chem, g 83 (1991). 42. R. Lingle, Jr; X. Xu; S. C. Yu; H. Zhu; J. B. Hopkins; J. Chem. Phys, 23_ 5667 (1990) 43. P. A. Anfinrud; C. Han; T. Lian, R. M. Hochstrasser; J. Phys. Chem, 9_4 1 180 (1990). 44. E. J. Heilweil; M. P. Casassa, R. R. Cavanagh; J. C. Stephenson, Ann. Rev. Phys. Chem, Q 143 (1989). 45. E. J. Heilweil; R. R. Cavanagh; J. C. Stephenson; Chem. ths. Lett, 134 181 (1987). 46. E. J. Heilweil; R. R. Cavanagh; J. C. Stephenson; J. Chem. Phys, 8_ 230 (1989). 47 E. J. Heilweil; M. P. Casassa. R. R Cavanagh; J. C. Stephenson, .l. Chem. Phys, E 5004(1986) 48. S. A. Hambir, Y. Jiang; G. l. Blanchard; .1. Chem. Phys, 28 6075 ( 1993) 49. Y. Jiang; G. J. Blanchard; J. Phys. Chem, 28 941 1 (1994). 50. Y. Jiang; G. J. Blanchard; J. Phys. Chem, 28 9417 (1994). 51 N. Agmon; J. Phys. Chem, 24_ 2959 (1990). 52. P. K. McCarthy, G. J. Blanchard; .l. Phys. Chem., 9_7 12505 (1993) 53. Y. Jiang;; McCarthy, P. K; G. J. Blanchard;; Chem. Phys, 183, 249 (1994). 54. J. J. Moura Ramos; J. Sol. Chem; _18 957, (1989) it) 55. P. Padila; S. Toxvaerd; J. Chem. Phys, 28 509, (1991). 56. E. Aicart; G. Tardajos; M. Costas; J. Sol. Chem, Q 369, (1989). 57. R. G. Snyder; A. L. Aljibury; H. L. Strauss; J. Chem. Phys, 81 53 52, (1984). 58. H. Ohtaki; J. Sol. Chem; 2_1 39, (1992). 59. Y. Marcus; J. 501. Chem, 2_1 1217, (1992). 60. T. R. Stengle; S. M. Hosseini; K. L. Williamson; J. Sol. Chem, 1_5 777, (1986). 61. (a).R. G. Snyder; J. Chem. Phys, 47 1316, (1967). (b). R. G. Snyder; J. 'H. Schachtschneider; Spectroch/micaAela1_9 85, ( I963 ). 62. S. N. Battacharyya; D. Patterson; J. Sol. Chem, 2 753, (1980). 11 CHAPTER 2. EXPERIMENTAL A mode-locked CW Nd:YAG laser (Coherent Antares 76-8) is used to produce 30 W average power at 1.064 um (IR) with 100 ps pulses at 76 MHz repetition rate (See schematic of the spectrometer in Figure 2.1). The output of this laser is frequency doubled using a Type I temperature tuned LBO SHG crystal (7 mm) to produce 3 W of average power at 532 nm (green), with the same pulse characteristics as for the fundamental. Both the collinear green and residual IR light are combined in an angle tuned Type I BBO SHG crystal to produce 1 W of average power at 355 nm (UV), again with the same pulse width and repetition rate as the fundamental. The 355 nm light is divided using a 56/44 (reflectance/transmittance) beam splitter and used to pump synchronously two cavity dumped dye lasers (Coherent 701-3). Both dye lasers are Operated with three plate birefriengent filters as the wavelength tuning element and no saturable absorber is used. The dye circulating in each laser is cooled to ~ 2 °C to increase the viscosity of the dye solutions and to reduce the rate of thermal degradation of the dyes. Stilbene 1, Stilbene 3 (Stilbene 420, Exciton) and Coumarin 1 (Coumarin 460, Exciton) dyes, as well as different sets of laser cavity mirrors were used, depending on the wavelength requirements of the experiments. 12 ' .—> AT r— mvi #1 —>T /L_l probeDL ._ lg ’ I ‘ delay stage S(t) = AT/T 1064nm(IR) V\“_ me : SHG CWImde-lockede'YAG 355nm(UV) 532nm(Green) 1064m0R) P - Polarization rotator DL - Dye laser PD - Photod/oale detector Figure 2.1. Schematic of the stimulated emission pump-probe Spectrometer. The dye solutions and the corresponding wavelength range as well as the optics set used are summarized in Table 2.1. Table 2.1. Information for each dye laser: including making of the dye solution, operating wavelength range and the mirror sets. Dye Preparation Operating Mirror set wavelength range (coherent coating rage) dissolve 1 g Stilbene 1 directly 405 - 450 nm 03 Stilbenel in 1.2 L warm ( ~ 100 °C) ethylene glycol. dissolve 2 g Stilbene 3 in 150 Stilbene 3 ml benzyl alcohol, then dilute 425 - 470 nm 03 (Stilbene 420) into 1.2 L warm ethylene glycol. dissolve 2 g coumarin 460 in Coumarin l 150 ml benzyl alcohol, then 460 - 490 nm 04 (Coumarin 460) dilute into 1.2 L warm ethylene glycol. Both dye lasers were cavity dumped at 7.6 MHz. This repetition rate was found to be optimum for the comparatively low gain blue laser dyes. The laser pulses are characterized by background-free non-collinear second order autocorrelation (Spectra-Physics model 409). The average autocorrelation trace was found to be ~ 7 - 10 ps depending on wavelength and dye. The pulses can be modeled with the noise-burst model,'” and the '2' The instrumental response frequency resolution was found to be ~ 4 cm'l for each laser function, determined by cross correlation ofthe pump and the probe laser pulse trains, is typically 10 ps FWHM and the cross correlation is taken to establish the zero time ofthe experiment. 14 The pump laser is used for instantaneous excitation and the probe laser is used to initiate the stimulate emission. The vibration of interest is excited at the difference frequency between pump and probe lasers. The pump and probe beams are focused on to the sample and the transmitted probe beam is directed to a monochromator and photodiode detector while the pump beam is stOpped before the monochromator. The signal we detect is the transient gain or loss of the probe laser intensity, which is usually very small. For our experimental condition, changes in probe laser intensity are on the order of 10" to 10'5 of the probe laser intensity, and the lifetime of the signal is significantly less than the inverse of the laser repetition rate, so no cumulative gain or loss can be used to advantage. Because the low frequency fluctuations of the probe laser can be as high as several percent of the average output intensity, a detection scheme is required that can separate the transient response from background noise. Based on the noise power spectrum ( N at f' l), the shot noise limit can be achieved by shifting the detection frequency to a few MHz. The magnitude of the shot noise is between ~ 10"’ to 10") of the laser beam intensity. We use a radio and audio frequency triple modulation shot noise limited detection scheme '3'“ to encode the signal. Both dye lasers are modulated using electro-optic modulators. Each modulator contains KTP crystals ( four of ~ 2.5 cm each) and a Glan-Thompson polarizer. A Sinusoidal electric field applied across the crystal causes a voltage dependent birefi'iengce in the crystal, leading to a rotation of the polarization of light passing through. The polarizer lets the vertically polarized component through to the mechanical delay line and the sample. The Sinusoidal electric fields used to drive electro-optic modulators have a maximum amplitude of~ 300 V The pump laser is amplitude modulated at 3.01 1 MHz 15 (con), and the probe laser at 2.110 MHz ((99). The pump laser is further modulated by a mechanical chopper at ~ 100 Hz ((0,). The signal detected is of the form AT/T, where AT is the difi‘erence in the probe laser transmitted intensity for the pump beam on and off. The form of the signal implies that the signal of interest interacts with both incident lasers, and in so doing acts as a molecular mixture for the modulations applied to the two lasers. In effect the sample multiplies the two modulation frequencies, where the efficiency of this modulation is a measure of the Signal magnitude. 1 1 coswa o coswfl = Ecos(a)a +wfl)+§cos(a)a — (05) [1] and the signal of interest is detected at a modulation frequency ((1),, + (Op), which is preset ‘ against a shot noise limited background. In time-resolved experiments most often the radiation applied is linearly polarized, so there is selectivity for the excitation and the collection of the response relating to the orientation of the molecules that are investigated. It is frequently of interest to measure experimental signals uncontaminated by orientational relaxation information. This can be done by collecting response at so-called Magic angle, 6..., which is the angle between polarization of the pump and probe in this thesis work. The induced anisotropy associated with the transition dipole y is defined by Equation 1. N t, —N t, r (1,7): //( 7) .L( 7’) [2] N//(t,7)+2Ni(l.r) 16 NI/(t) and N1(t) are the population with their orientation parallel and perpendicular-to the polarization of the pump laser respectively. In the case where only one initial state is prepared, r(t) is given by the expression involving the second Legendre polynomial (P;) of the correlation of the transition moment direction at time zero with that at time t. 2 . rm) = ng (cos6)3 1.0 - "‘ ’0 f0” /. A ”0". =1 3 L .3 £9. .E 0f5'— 00 i 1 I I i 1 I 20800 21200 21600 22000 frequency (cm") Figure 3.4. Time resolved stimulated emission spectra of perylene in l-octanol. (o) = 10 ps delay, (0) = 100 ps, (V) = 200 ps, (inverted filled triangle) = 300 ps, (D)= 400 ps, (I) = 500 ps, (A) = 600 ps, (filled triangle) = 700 ps, 0) = 800 ps, (0) = 900 ps. 29 1934 1734 1578 I375 1.5 — e C - /:\./s :1 V v I / /2 . A" \A 10 - /‘\543 ’5 ‘ ° =1 / 0“. 3 . / 3 ‘O :E) ’ 0 \v .V/ :1 If \ >‘A I . a . \ <>/ 0 0.5 — ’ J l J l 1 1 20400 20800 21200 21600 frequency (cm") Figure 3.5. Time resolved stimulated emission spectra of perylene in DMSO. (0) = 10 ps delay, (0) = 100 ps, (V) = 200 ps, (inverted filled triangle) = 300 ps, (CI)= 400 ps. (I) = 500 ps, (A) = 600 ps, (filled triangle) = 700 ps, (0) = 800 ps, (0) = 900 ps. 30 1934 1734 1578 1375 1.5 F 1.0 '- 3 3 .3“ § 0.5 — . 1 1 1 1 1' 1 20400 20800 2 1200 2 1600 frequency (cm") Figure 3.6 Time resolved stimulated emission spectra of perylene in toluene. (O) = 10 ps delay, (0) = 100 ps, (V) = 200 ps, (inverted filled triangle) = 300 ps, (Cl)= 400 ps, (I) = 500 ps, (A) = 600 ps, (filled triangle) = 700 ps. (0) = 800 ps. (0) = 900 ps. 31 The most obvious feature in these stimulated emission profiles is the presence of solvent-_ dependent sub-structure within the band, and the gradual evolution of the stimulated emission profile to correspond with the spontaneous emission spectrum. The physical origin of these features and the chemical information that is contained within them will be addressed below. The individual features seen in the time resolved spectra are red-shified from the perylene spectroscopic origin by vibrational resonance frequencies. These features correspond to individual vibronic transitions within the spontaneous emission envelope that was studied. The spectral width of the individual bands is dominated by the thermal width of the vibrationless excited state. Over the wavelength range that was examined, four Raman active vibrational modes at 1375 cm'l, 1578 cm'l, 1734 cm'1 and 1934 cm‘1 dominate the spectrum. The modes at 1375 cm'1 and 1578 cm'1 have been assigned previously as the v7 and v5 ag modesmm The 1734 cm'1 and 1934 cm'1 modes are combination modes; v1734 z v7 + 358 cm'l, and v1934 2: v7 + 550 cm'1 or v5 + 358 cm'lml Both the v15 mode at 358 cm’1 and the v14 mode at 550 cm'1 are prominent Raman active ag . 16 . . . . . . . . . modes in perylene.l ] It 15 not possrble to distinguish between these possrbilities for v1934 based on the frequency domain spectra. The temporal responses of these modes give an indication of their composition. 32 The single most unusual characteristic of these data is the persistence time of the individual vibronic features. The instrumental response time of the laser system is 10 ps, and thus a signal that persists for hundreds of ps can not be attributed to a coherent Raman response. These vibronic features do, however, dwindle to a level that is difficult to detect before the population of S] is exhausted. If this were a simple excited state absorption response the features would persist for the duration of the 51 population, and such a process would likely produce a spectral Signature very different from the one that was Observed. Coherent hyper-Raman scattering from excited state vibrational .modes with a resonant intermediate state could be invoked to explain the observed energy dependence in the data, but such a signal would persist only for the duration of the instrument response function. Modeling these data within the framework of pump pulse absorption followed by simple Stimulated emission from the excited state of perylene does not provide satisfactory agreement with the experimental data either, because such a process should nOt expose vibronic sub-structure. To understand these results we must consider a more complete model of the spectroscopic processes possible for our experimental conditions. For simplicity we model these data as a linear superposition of coupled three level systems. This simplification ultimately leads to small differences between the experimental data and the model, but the qualitative features and chemical information Contained in the data are represented well by this model. The three states in each coupled three level system are the electronic ground State (So). the electronic excited state (S 1) and a vibrational level in the 50 manifold, as depicted in 33 Figure 2.7, where the S] state is assigned the label A, the ground vibrational state B and the vibrationless ground state C. It is assumed that the excitation laser pulsewidth is Short compared to the processes we detect. Experimentally the pump wavelength is chosen to be v0.0, and to probe the vibrational state vvib the probe wavelength is chosen to be v0.0 - v..,.,. This is how two laser radiations forces a molecule to vibrate at the difference frequency. ‘ A 1 1., k ”probe :( (”O-0’ (0v) (11 0-0 = (”pump 2 I B ( m. ) Figure 3.7. Schematic of coupled three level system used to model the experimental data. The terms k are described in the text. In 1917 Einstein proposed a model for transitions between two states that relates absorption, spontaneous emission and stimulated emission'm For a two level (AB = hv) an incident electric field at frequency v will serve to diminish the population difference between two states. This happens because the population in the higher energy state will be transfered to the lower energy state and the population in the lower energy State absorbs the radiation to make a transition to the higher energy state (absorption). This 34 pOpulation exchange will not proceed indefinitely due to the finite lifetime of the.upper state (A) and vibrational state B. In this model k1 is the rate constant for emission from level A to level B. k1 = klstimulated + klspontaneous' The rate constant k2 is for absorption from level B to level A. The rate constant k3 represents the vibrational population relaxation from state B to state C. Because this is a spontaneous process, the mechanism of depopulation from B to C are not well determined. It is assumed that the B —> C transition is irreversible since the vibrational levels are at energies high enough that Boltzmann population of state B is negligible. The complete model is shown in Figure 3.7. The Einstein coefficients 8,: and 82) for a two level system are equal save for differences in the degeneracy of the two states involved (k) z k2). By virtue of the way the transient response is detected, the observed Al/I signal is the sum of the population changes for states A and B: S(t) = A(t) + B(t). This assertion appears outwardly to be counter-intuitive because stimulated emission produces a gain in the number of photons in the probe laser electric field while absorption involves the loss of photons from the same electric field. The response we sense is the net gain on the probe laser electric field, Le. (Stimulated emission - absorption). ”The reason that the experimental signal we detect is sensitive to the sum, and not the difference, in populations lies in the manner in which the experimental Signal is encoded and decoded. Although the details of the experimental laser beam modulation scheme are more complex, as discussed in Chapter 2, one can consider conceptually that the pump electric field imposes a sinusoidal amplitude modulation, (Dmod (“'5 MHz), on the excited state (A) population. For 35 a DC probe laser electric field applied at a time t after excitation, the stimulated emission signal will appear as a sinusoidal gain at comod, and the absorption component of the signal will appear as a sinusoidal loss, also at mm), but because of the population exchange between the emitting and absorbing states, the phase of the gain modulation is shifted by It from the phase of the loss modulation. In other words, When stimulated emission from A to B is at a maximum because the population of A is at its maximum level, absorption from B to A will be at a minimum because the population of B will be at its lowest level during the modulation cycle, and vice versa. Gam(a)mod.t) = A(t)sin(wm0d ~t) Loss(rumod.t) = B(t)srn(a)mod -t + 12') S(wnde) = (Jam((um0d.t) — Loss(wm0d .t) [l] S((umodJ) = A(t)sin(wm0d 0!) — B(t)sin(a)mod -t + II) = (A(t) + B(t))s1n(w,,,0d -t) By synchronous demodulation detection. we obtain S(t) = A(t) + B(t). The'coupled differential equations that describe the time-dependent population changes for the three level system are [2] "1 "3 A<:>B—>C "2 11114414192181 dt 337?] = I‘ll/11— (*2 + @1113} {[5] = "3l’31 dt 36 . . [18.19] . The solution to these equations has been reported by several authors, and the time dependence of each species can be found by making a series of substitutions. Following [191 the treatment and nomenclature of Szabo, a = [Al/[Ale fl = [Bl/[Ale 7 = [Cl/[Ala T=k11 K1=k2/kl K2=k3/k1 d ——=-a+K dr 1'8 gig: —K1fl—K7,B+a dr " d7 —=K dr 2'6 The solutions for or and B are given by K2-l3 K‘z-Az at = exp —}mkt — exp -/1 kt ) 41302-13) ( J 1) 1202-43) ( 2 1) 1 12-13 [exp(—/l3k1t)- exp(-/12k1t)] A(t) = where I 1 2.2 = %{1+K‘] +K‘2 +[(1+K1+K2)2 -4K2] '} l l. 7 ./ 13 = Elz—{l-i-tq +K‘2 —[(1+KI+K2)" —4K2]' 2} S(t) = or(t) + B(t). S(t) has the form of two exponential decays, 37 -__’_‘_2__ L. _ __1_. ._ S(t)— (’12 -,l3){,l3 exp( A3k1t) ’12 exp( 121(10} The exponential decay terms in S(t) containing 71.2 and A3 are related in a straightforward manner to the transition rate constants k} and k3 when the vibrational population relaxation rate constant, k3, is substantially larger than the overall rate constant k1 for the A -—> B transition. l/klspontaneous z 6.4 ns for perylenem] and, while there is no T1 data available for perylene in solution, Zinth et al. have reported a T1 for the v2 mode of anthracene to be ~240 psm The condition k3 >> k1 thus appears to be satisfied. For large values of k3/k1, 3.3 approaches unity. The exponential decay term in S(t) containing A3 will therefore be determined by k1 when vibrational relaxation from B is fast. Comparison Of A2 and k3/k1 reveals that, except near k3 ~ k), the term Azkl ~ k3. The decay term in S(t) containing 712 will therefore be related directly to k3. A build-up with a time constant T) = l/k3 and a decay with a time constant Telec = l/k] (te hereafter) are expected in our experimental time scans. In Figure 3.8 the time scans for perylene in n-octane at four probe wavelengths corresponding to the vibronic resonances are shown, where the relevant ground vibrational state is indicated. Data for times < 20 ps delay have been truncated so that there is no contribution to these decays from the instrumental response function. For all cases a fast build-up followed by a slow decay was observed. Both of these time constants vary as a fiinction of perylene vibrational mode and solvent, and are 38 1.5 l" r. ‘L AITT(att) U. \l m l.) 2110 delay finie (ps) Figure 3.8. Time scanes for perylene in n-octane at probe wavelengths corresponding to distinct vibronic resonances, where the ground state vibrational level is indicated Build-up and decay times are given for these data in Table 3.1. reported in Table 3.2. Analysis of these data show that the approximations of A3 ~ 1 and Azkl ~ k3 are good to within <5°/o, i.e. better than the uncertainty in the values of T. and Te. Table 3.2. Build-up and decay times determined from the experimental data. All times are given in ps. The vibrational frequencies refer to the final state in the stimulated transition. The spectral origin for each solvent was estimated from the static spectroscopic data. vibrational mode 500W" [3 75 cm'1 1578 cm'1 1734 cm'1 1934 cm'1 n-octane T1=25i3 T1=51i6 T1=55i4 T1=362i87 te=2163:10 te=2858i27 te=1828i13 te=872i15 l-OCIanOl T1=40i3 T1=100i1 l T1=54i5 T1:75i9 te=2521i20 te=21 1 1:42 te=2654i42 te=2165i41 l-butanol T1=357i86 T1—79i6 T1=3 541-40 Tl=188i25 te=1928i285 Te 1833:32 te=1575i58 te=2236il 17 DMSO T.=66:9 T.=32i12 T1=80i7 T1=26i4 te=3682i 147 te=256 1:47 te=2006i3 5 18:1526i13 toluene T.=364:133 T1=- --- T1: ---- Tl=55i5 te=l962i288 te=2805i20 te=3874i78 te=1697:15 There are several points regarding these data that require further attention. First, it is not immediately apparent that the B —> A transition needs to be considered to explain these data. The probe pulse stimulates emission from state A and, we assert, also stimulates absorption from state B. If only stimulated emission occurs then the effective rate constant for this process should depend on the probe laser intensity, and we should 40 measure a probe laser intensity dependent decay constant. If absorption from state B also contributes to the observed response one should observe no probe laser intensity dependence because the effective absorption rate constant depends on probe laser intensity in exactly the same manner as stimulated emission. Two time scans for perylene in octane where Iprobe E Ipump/ 1000 and Iprobe _—“: Ipump were measured (Figure 3.9). The slow decay constant is the same within experimental uncertainty for the two scans, establishing the contribution of absorption from B —> A during the probe pulse. Over the duration of the probe laser pulse, a population equilibrium is established between states A and B with the equilibrium constant given by klstimulated/kz. An intensity dependence in the signal was observed only when the laser pulses are overlapped in time, and this effect is attributed to stimulated Raman gain, a response that is expected to scale with the intensity Of both the pump and probe lasers and persist only for the duration of the instrument . 121] response function. Another noteworthy feature of these data is that the measured rate constant for the A —> B transition varies with probe laser wavelength. For a given transition, the stimulated . . . . . . 1221 em1s510n rate constant 15 related directly to the spontaneous emnssron rate constant, and both of these rate constants depend on the oscillator strength of the particular transition . 1231 . . . . . . be1ng probed. Because of the comparatively narrow detection bandwrdth 1ntnnsrc to . . . . [241 . . our stimulated emissron experiments, different total rate constants for each specrfic transition accessed were observed. 41 stimulated Raman gain f IPump ~ Iprobe ‘ 'l ‘1 . til H 1'. 1 l 6- Ii i Ill )1" ’K. i,“ 3. 4 '— r 1 Hi 1 i g .1 .' 1 l l i . ). 1 1 1 ~ . E ‘ i .’ I10.1I ‘ Q 2 1- 1 Iprobe ~ 0.001 ipump 0 _2 n . 1 1 l 1 I 1 I A 1 l I A L A 1 l I 1 _1 0 1 00 200 300 400 500 600 700 800 900 1 000 scan time Figure 3.9. Time scans for two different pump/probe intensity ratios. 42 The detail chemical meaning of the T, values reported in Table 3.2 is discussed as followed. The vs (1578 cm-l), V7 (1375 em-l), v14 (550 em-l) and v15 (358 cm-1) modes in perylene are all ag in-plane ring distortion modes. It is not surprising that the different modes relax at. different rates. The solvent "cage" surrounding perylene is almost certainly anisotropic, and each solute vibrational mode couples difierently to its immediate surroundings for both geometric and energetic reasons. Comparison of the relaxation times T1 for v7 and v1734 (V1734 = v7 + v15) reveals that for all solvents except toluene, the v7 mode couples more efficiently to the available solvent bath modes than the v15 mode. For toluene, however, T11375 = 364 i 133 ps, and T11734 < 10 ps, indicating efficient relaxation mediated by v15 at 358 cm'l. The v15 mode in perylene is essentially a stretching mode between the two naphthalene moieties and represents a significant concerted motion of different portions of the perylene It systemml The data indicate Strong coupling of this mode to the surrounding toluene solvent, suggesting at least partial alignment of the solvent and solute 1: systems. The first step inthis process appears to be intramolecular relaxation into v15 followed by energy transfer between the perylene modes and 'toluene. The infrared spectrum of toluene Shows a band at ~350 cm‘hml implicating resonant energy transfer as the dominant intermolecular relaxation mechanism. The perylene v5 mode, a symmetric breathing mode of the individual naphthalene moieties, also appears to be coupled strongly to the toluene solvent bath modes, and there are strong Raman and infrared active toluene vibrations between 1550 cm‘1 and 1650 cm” 43 1261 . . . .. 1. This spectral correlation also suggests the dominance of non-collis10nal energy transfer between perylene and toluene. The operative coupling mechanism for the alcohols appears to be fundamentally different. There are-substantial differences in T1 for perylene in l-butanol and l-octanol. The Raman and infrared spectra of l-butanol and l-octanol are similar, suggesting that if non- collisional coupling of perylene to the solvent bath is operative then we should observe a correlation in the measured T) times in the two alcohols. The non-correspondence we observe suggests that inelastic collisions may be an important relaxation mechanism in the alcohols. It is also possible that non-collisional relaxation efficiency depends critically on local organization in the solvents. This possibility is suggested by the data present in Chapter 5. The generally more efficient relaxation of perylene in the longer chain alcohol suggests that solute inelastic collisions with the solvent aliphatic moiety are more efficient than interactions with the polar hydroxyl group. The v1934 mode can be described as either v7 + v14 or v5 + v15. If v1934 = v5 + v15 then its T) values should correspond with either those of v5 or v1734. No such correlation was noted in the data. Conversely, if V1934 = v7 + v14 then one expects a correlation between T. values for this mode and v7. Again no such correlation was observed, implying, by process of elimination the relaxation of this mode is dominated by v14 at 550 cm'l. It is likely that the lack of correspondence with other modes that have been examined is due to the participation of all of these modes in this spectral region, where v14 provides the dominant relaxation pathway. 44 If the stimulated emission response of perylene is modeled as a linear superposition of coupled three level systems, the vibrational features in the data Should not disappear completely until population of the 51 state is exhausted. Experimentally, however, these vibrational features do not persist at a detectable level for the lifetime of the 51 state (Figures 3.2-6). We believe that the failure of our model to account for this observation lies in two factors, both of which must be present. First, the vibrational modes must exhibit some anharmonicity for us to observe combination modes. This anharmonicity can also allow coupling to other (anharmonic) perylene modes, and this possibility was not considered. Second, it is implicitly assumed that once the vibrational energy is dissipated from state B. it is lost irreversibly. For high frequency modes this is a good approximation. but for low frequency modes it is not. The v15 mode at 358 cm'1 is < 2 kT at 300 K. Boltzmann population will be significant for low frequency components of combination modes. It is believed that the slight discrepancies between the experimental data and the model that was used to understand it arise from the shortcomings of the simplistic model. Finally, in comparison with other methods like 1R pump/ IR probe and IR pump/ UV,VlS probe, this technique uses visible lasers and has high spectroscopic selectivity. A wide range of vibrational resonances can be accessed with our pump-probe scheme. A significant advantage of this technique, which are touch upon in later chapters, is the ability to excite probe molecule vibrational resonances that are degenerated with solvent 45 vibrations. Using IR lasers to excite such a vibration is not possible owing to the size of the solvent background absorption. 46 3.5. Conclusion A new technique for the measurement of T1 ground state vibrational population relaxation times using time resolved stimulated emission spectroscopy was developed. The measured population relaxation times were found to vary dramatically with the identity of both the solute vibrational mode and the solvent. Both intermolecular and intramolecular relaxation pathways depend on the identity of the solute vibrational mode. The T] variations from solvent to solvent for a particular mode serve to underscore the importance of intermolecular relaxation pathways in the vibrational relaxation process. The experimental data is modeled qualitatively by assuming that the spectrosc0pic response of perylene can be treated as a linear superposition of coupled three level systems. Differences between the experimental data and the model are likely due to anharmonic coupling between vibrational modes in perylene that is not accounted for. It is believed that other pOchyclic aromatic hydrocarbons will exhibit these effects due to the comparatively modest anharmonicities they exhibit. Polar dye molecules may not yield an analogous stimulated emission response because of the greater extent of anharmonic coupling in these systems. Nonetheless. the limitations of the model do not detract from the physical significance of the build-up and decay times seen in our experimental data. This means of measuring T1 is anticipated to prove useful in elucidating the molecular nature of solvent-solute interactions. 47 3 .6. Literature Cited 1. W. Zinth; C. Kolmeder; B. Benna; A. Irgens-Defregger; S. F. Fischer; W. Kaiser; J. Chem. Phys, 18 3916 (1983). N .N H. Gottfried; W. Kaiser, Chem. Phys. Lett, __1331 (1983) . A. Seilmeier; P. O. J. Scherer; W. Kaiser; Chem. Phys. Lett, 121140 (1984). b) 4. E. J. Heilweil; M. P. Casassa; R. R. Cavanagh and J. C. Stephenson; J. Chem. Phys, Q 5004(1986) LII ..E J. Heilweil; R R Cavanagh andJ. C. Stephenson; Chem. Phys. Lett, ___4 181 (1987) 6. E. J. Heilweil; R R. Cavanagh andJ C. Stephenson, J. Chem. Phys, _982 30(1988). \I . P. Anfinrud; C. Han; P. A. Hansen; J. N. Moore and R. M. Hochstrasser; Ultrafast Phenomena VI, T. Yajima; K. Yoshihara; G. B. Harns; and S. Shionoya; Eds. Springer, Berlin, p. 442-446, (1988). 8. P. A. Anfinrud; C. Han; T. Lian and R. M. Hochstrasser; J. Phys. Chem, % 1180 (1990) 9. R. Lingle Jr.; X. Xu; S.-C. Yu; H. Zhu and J. B. Hopkins; J. Chem. Phys, 9_3 5667 (19.90). 10 R Lingle Jr. X Xu; S -.C Yu; Y J Chang and] B. HOpkins; J. Chem. Phys, 92 4628(1990) 11.S-.C Yu; X Xu; R Lingle Jr and]. B. Hopkins; J Am. Chem. Soc, 12 3668 (1990). 12. X. Xu; R. Lingle, Jr.; S.-C. Yu; Y. I. Chang and J. B. Hopkins; J. Chem. Phys, 22 2106(1990) 13. J. B. Hopkins andP M Rentzepis; Chem. Phys. Lett, __2503 ( 1987) 14. N. H. Gottfried; W. Kaiser; Chem. Phys Lett, 1(1983) 48 15. S. J. Cyvin; B. N. Cyvin and P. Klaeboe; Spec. Lett, _l_§ 239 (1983). 16. S. Matsunuma; N. Akamatsu; T. Kamisuki; Y. Adachi; S. Maeda and C. Hirose; J. Chem. Phys, 88 2956 (1988). 17. A. Einstein; Zur Quantentheort’e der Strahlung, Phys. 2., 18 121 (1917). 18. D. H. McDaniel and C. R. Smoot; J. Chem. Phys, 60 966 (1956). 19. Z. G. Szabo in Comprehensive Chemical Kinetics, Vol. 2, ed by C. H. Bamford and C. F. H. Tipper; Elsevier, 24-26 (1969). ' 20. I. B. Berlman; Handbook of Fluorescence Spectra of Aromatic Molecules, Academic Press, p. 399, (1971). 21. B. F. Levine; C. V. Shank and J. P. Heritage; IEEE J. Quant. Electron, QE-lS 1418 (1979) 22. J. T. Verdeyen; Laser Electronics, Prentice Hall, p. 139, (1981). 23. G. Herzberg; Molecular Spectra and Molecular Structure, Vol. 111, Van Nostrand, p. 418,(1966) 24. G. J. Blanchard;J. Chem. Phys, 95 6317 (1991). 25. We have calculated the normal mode atomic displacement vectors for these perylene modes using the Spartan v.2.0 software package. 26. H. A. Szymanski; Correlation of Infrared and Raman Spectra of Organic Compounds, Hertillon Press, Cambridbge Springs, PA, (1969). CHAPTER 4. VIBRATIONAL POPULATION RELAXATION OF PERYLENE IN ITS GROUND AND EXCITED ELECTRONIC STATES Summary A novel scheme for the measurement of molecular vibrational population relaxation in both ground and excited electronic states using ultrafast Stimulated spectroscopy is proposed. This technique was demonstrated using 10'5 M perylene in n-pentane, n-hexane and n-heptane. The vibrational population relaxation times (T.) of the perylene V7 mode are 304:44 ps in the ground state and 140:1 1 ps in the first excited singlet state, and both T. times are solvent independent to within the experimental uncertainty. In contrast, the T. relaxation time of the perylene ground state v5 mode exhibits a measurable solvent dependence, ranging from 160 i 37 ps in n-pentane to 308 i 41 ps in n-heptane. 49 50 4.1. Introduction Despite the high density of states intrinsic to comparatively large molecules, certain chemical systems exhibit slow vibrational population relaxation. Perylene, for example. exhibits T. times that can vary from less than 10 ps to hundreds of ps, depending on the specific vibrational mode examined and the chemical identity of the surrounding medium.“I Perylene is non-polar, and the nature of its interactions with its surroundings are, for the most part, shorter range than for polar systems, where electrostatic and dipolar processes dominate. Perylene is a USCfUl probe molecule for vibrational population relaxation experiments because it exhibits only modest anharmonicity and Strongly mode- dependent relaxation rates in liquids. Perylene has been chosen as a well-characterized probe molecule for the experiments reported in this thesis. It is found that the vibrational population relaxation time, T., is mode sensitive and solvent sensitive. The purpose of this work is two-fold. First. the ability to measure vibrational population relaxation rates in excited electronic states of perylene in dilute solution will be demonstrated, and second, the state-dependent vibrational population relaxation times for the perylene v7 mode will be presented. T. for both 5.. and S. perylene for the v7 and v). modes in three n-alkanes were measured, and it was found that relaxation in the S. state proceeds a factor of two more rapidly than in the Sn State. These findings will be discussed in the context of inter- and intramolecular relaxation pathways that are expected to exhibit an electronic state- dependence. 51 4.2. Theory As discussed in the previous chpater, the measurement of ground state vibrational population relaxation times using transient stimulated spectroscopy can be understood in the context of a coupled three level system (Figure 4.1a). To recap briefly, the experimental signal is of the form, B S(t) = B(t) + ((1): HOT. {(12 + k3)exp(—klt) - (k2 + k1)exp(-k3t)} [1] For the measurement of vibrational population relaxation within an excited electronic state, one needs to consider the population dynamics for four states (Figure 4.1b). In this four level system. the states B, C and D are the same as those Shown for the three level system. State A is the vibrational mode ofinterest in the excited electronic state. The rate constants k., k; and k. are the same for the four level system as for the three level system. The primary differences between the observed relaxation dynamics for these two systems lie in the initial populations of states A and B. For the excited state vibrational population relaxation measurement, the initial population in state B is zero, where for the three level system. state B is populated directly by the exciting laser pulse The transient population relaxation dynamics for the four level system are described by the four coupled differential equations shown in Equation 2 B a 1 A k1 k C”probe (0 2 pump V\ C k 3 D b A 0’ pump Figure 4.1. (a) Schematic of the coupled three-level system used for interpretation of 0-0 excitation experiments. (b) Schematic of the coupled four-level system used for interpretation of excited state T. measurements. 53 A—->B<:>C—>D dA —=-k A d1 4 Eg=k4A—le+k2C ‘” In dC ——=kB— k~+k C cl] 1 (.1 2) ink3C dt For the three level system, it was implicitly assumed that k; >> k. E k2 we apply this same assumption plus one other, k. ~ k., for the solution of the four level system. Integration of these equations yields (k2 + A3) ("3 ‘k1)("4 "'11 + (k1 +k2 +k3 —k4) l Mt-kuw4-kn ("3 ‘k1)(’*'3 ‘ "41 1 exp(—klt) + exp(—k3t) S(t) = B(l)+ ("(1) = AOL-4 cxp(—k4t) The Signal from our experiments contains three exponential terms. In principle, one can fit the experimental decay curve using Equation 3, but with five variable parameters and a finite signal-to-noise ratio, it is difficult to obtain fits to the data where the fitted parameters possess the requisite mutual independence. This is, however, not an insurmountable problem. Using the information obtained from experiments where excitation was at the origin, one can determine k. with acceptable certainty. With k. and k; obtained from the 0-0 excitation experiments, either the raw data from the blue excitation experiments can be fitted directly, or the difference between blue and red excitation data can be taken and fitted with the requisite certainty. It is possible for a 54 ground state vibrational mode and an excited State vibrational mode to decay with the same time constant, i.e. k3. = k... If this condition were to occur, it would lead to undefined pre-exponential terms in Equation 3. Such a condition does not pose a problem experimentally because the ground state and excited state vibrational modes used in the measurements can be different, as long as the same ground state vibrational mode is used for both excitation frequencies in the determination of the excited state T. time. In fact, this is a strategy which was employed in the acquisition of T. times in the first excited singlet electronic state of perylene, as detailed below. 4.3. Experimental Spectrometer: The spectrometer used for these measurements has been described in detail in Chapter 2. Stilbene l and Stilbene 420 dyes (Exciton) were used for the pump dye laser and Stilbene 420 dye was used for the probe dye laser. The pump laser is operated at 434 nm for 0-0 excitation of perylene in n-pentane and n-hexane, and at 437 nm for tt—heptane. For excitation of the v7 mode in the first excited Singlet electronic state, the pump dye laser iS operated at 409 nm for n-pentane and n-hexane. and at 412 nm for n-heptane. For red (0-0) excitation experiments measuring v. T. times, the probe dye laser is operated at 462 nm for n-pentane and n—hexane and at 466 nm for n-heptane (See Table 4.1). For the measurement of v7‘ we used probe laser wavelengths of 466 nm for perylene in n-pentane and n-hexane, and 470 nm for perylene in n-heptane. Table 4.1 . The pump, probe wavelengths and dyes for each solvent. 55 0-0 excitation 1_J-* excitation solvent pump probe dyes pump .probe dyes (nm um) (nm 1 nm) n-pentane 434 /462 Stilbene 420/ Stilbene 420 409 / 466 Stilbene l / Stilbene 420 n-hexane 434 /462 Stilbene 420/ Stilbene 420 409 / 466 Stilbene 1 / Stilbene 420 n-heptane 437 I466 Stilbene 420/ Stilbene 420 412 / 470 Stilbene 1 / Stilbene 420 Steady state .spectroscopt‘es: The steady absorption spectra of the perylene solutions were measured using a Beckman DU-64 spectrophotometer, with ~ 1 nm resolution. Fluorescence spectra were recorded using a Perkin-Elmer model LS-S fluorescence spectrophotometer, with ~ 1 nm resolution. These data were used to estimate the spectral origin of perylene in each solvent. Chemicals and sample handling: Perylene and the n-alkanes were purchased from Aldrich Chemical Co. and were used as received. The solutions used for the time resolved Stimulated measurements were ~ 10 ul_vl_ and were flowed through a 1 mm pathlength flow cell to minimize thermal lensing contributions to the signal. The sample was temperature controlled at 300:0.1 K using a thermostatted bath 4.4. Results and Discussion 56 The static absorption and emission spectra of perylene in n-hexane are presented in Figure 4.2. The corresponding spectra of perylene in n—pentane and n-heptane are virtually identical. For excitation of the perylene 0-0 transition, the pump laser is set to the wavelength indicated by the arrow labeled “a” in Figure 4.2, and for excitation of the v7. mode, the pump laser wavelength is at ~1400 cm'1 above the origin, indicated by arrow “b” in Figure 4.2. While the ground state v7 mode has a characteristic frequency of 1372 cm", in the first excited singlet state this mode shifts to ~ 1393 cm".m The v7 mode in perylene is an in-plane Raman-active ring distortion mode of (1,. symmetry?“ and an exaggerated schematic of its atomic displacements is shown in Figure 4.3. The difference in resonance frequency between the ground and excited electronic states suggests that the potential well for this mode is slightly different in the electronic states, and it is therefore not surprising that the vibrational population relaxation dynamics for this mode will be unique to each state. Before the results for v7 and v7. population relaxation times are discussed, the method of acquisition and analysis of the raw data is described briefly. As noted in the experimental section, two groups of experiments were performed; the first aimed at measurement of the V7 model and the second group designed for measurement of the V7. mode. In the first group of experiments, the V7 mode (state C in Figure 4.1a) was accessed directly. For the 57 1.0 - 08 .— t. . . Absorp ton Emrssmn 0.4 r- linear response (a.u.) 350 400 450 500 wavelengh (nm) Figure 4.2. Linear optical response of perylene in n-hexane. The absorption and emission spectra have been normalized. Arrow “a” and “b” indicate the excitation wavelengths used for the 0-0 and v7* experiments, respectively. 58 V7 Figure 4.3. Schematic of the exaggerated atomic displacements for the perylene v7 mode. The directions of the displacements were estimated from semi- empirical calculation results. 59 second group of experiments, the ground state v5 mode (state C in Figure 4. lb) instead of the v7 mode was used. As discussed in the Theory section, it is possible that, for a given, vibrational mode, T. = T.‘ (k3=k1), and such an experimental condition could lead to an ill- defined condition in the interpretation of our data. The use of different ground and excited state modes serves to minimize, but not eliminate, the possibility that k3=k... In fact. for perylene in n-pentane, k;(v5) ~ k..(v7'°), but because the agreement is not exact, we are able to extract information on k. for v7. from the experimental data. A second reason for using v5 instead of vv in the determination of v7. is that the highest possible signal to noise ratio in the data is desired, and the stimulated transition cross section for perylene is larger for a probe laser energy of(v0-.. - 1578 cm'l) than for (V...) - 1372 cm'l). For the 0—0 excitation experiments on both the v7 and v5 modes, the time resolved stimulated signal S(t) was fitted with a double exponential function, and this process has been detailed previously. A time resolved scan used for the determination of T. for v7 is shown in Figure 4.4. For the determination of V). the k. and k;(v5) information obtained from the 0-0 excitation experiments are used to fit the data from the v7. excitation experiments In principle. one can either fit the data directly, or take the difference between the two experimental signals, normalizing for intensities at long delay times, where spontaneous and stimulated emission from state B are the dominant relaxation processes in both data sets. Both of these methods yield identical results, save for arbitrary pre-exponential factors, and the difference signal approach was chosen because it presents a clearer view of the difference in responses for the two excitation conditions. 60 l 1.0 *- 411.] > ' - ’ "-41“ ‘ ,‘ 08 1— ' ." ""r'"'1".-.. . 'Vli. l'hl‘." _ T111 ‘ .i‘.',"L_ .‘ t E—' 04 1- G 11.2 — l . ‘1 0.0 I L l I l M 1 I l I J I 0 200 400 600 800 delay time (ps) Figure 4.4. Experimental stimulated response and laser cross-correlation for measurement of the ground state v7 mode of perylene in n-hexane. For this experiment, Am... = 432 nm and two... = 462 nm. The line through the data is the fitted result See Table 4.1 for best fit results. 61 The difference signal, AS(t), (V7. excitation response - 0-0 excitation response) is of the form (Equation 3 - Equation 1), 112 4.1% k4A0 ] [k7 4111.1 k4A0 ] AS! = 3 —B ex ——lrt + " —B ex -lrt () (k3-/‘1 [Qt-kl 0 p( I) k3-k1 k3-k4 0 p( 3) [kl/1”“! + k2 + k3 - k4 )] exP("k4l) ‘ (k4 _ k1 )(k-l ‘ k3) [4] ln Figure 4.5a both the 0-0 and V7. excitation time scans for probing the V5 mode in n- hexane and the difference signal, AS(t), in Figure 4.50 The T. (=k3") and If (=k4") population relaxation times for V5, V7 and V7. in the three n-alkanes are presented in the Table 4.2. Table 4.2. Vibrational population relaxation times for the ground state and excited state V7 mode of perylene in n-alkane solvents. solvent 'l',( v5) 711/ V') TIY V5.) (1)3) ‘ (ps) (1)5) n-pentane 160:3 7 276:46 141:2 n-hexane 300: l 00 281:177 150:17 n-heptane 308:4] 355:100 129:71 These data Show several interesting features. We note that v5 exhibits a solvent- dependent T. time where V7 does not. In our previous work we have found that several different perylene vibrational modes exhibit unique solvent-dependent relaxation properties. and the data we report here on V5 and V7 are consistent with this trendm 12 10 intensity (a. u) dillerence signal (an) O Figure 4.5. 62 _ 0.," excitation a '_ / * ”Wm ' ' “We )- / 01-1111 _. 1 0-0 excitation 'Vltf'kii...}1.u l- 1 Q1 1 1 I 1 1 1 3 1 ’ b — F—' M‘ _J ‘_— I q .— — '— l 200 400 600 800 1000 delay time (ps) (a) Experimental stimulated response for the 0-0 and V7* excitation of perylene. For these scans the ground state V5 mode of perylene in n-hexane is probed; Mp = 432 nm for 0-0 excitation and 409 nm for v7* excitation and Am... = 466 nm for both excitation conditions. See Table 4.1 for the best fit results. (b) Difference signal, S(t), for the two scans shown in (a), with the best fit function shown as a solid line through the data. 63 It. is important to place the data we present here in context with the data we have reported in the following chapter. The T. values we present for the perylene V7 mode in n-pentane, n-hexane and n-heptane are significantly longer, by a factor of ~10, than the T. time we have reported for the V7 mode in tt-octane.["5' As discussed in Chpater 5, the perylene v7 mode relaxes anomalously fast in n-octane, compared to its relaxation in the other 11- alkanes, because of efficient V-V resonance coupling to the n-octane terminal CH3 rocking mode (1378 cm"). The perylene V7 mode relaxes with a time constant of 298 : 102 ps in tt-CgD.g, demonstrating that the dominant coupling mechanism for the anomalously fast T. time seen in n-CgH... is resonant V-V energy transfer. The solvent vibrational receptor mode is significantly localized at the termini of the alkyl chains and ‘ resonant V-V coupling is extremely sensitive to the spatial proximity of the “donor” and “acceptor” modes, i.e., the spatial relationship between the alkane terminal CH3 groups and the perylene v7 vibrational coordinate determine sensitively the efficiency of T. relaxation for this systemm The T. relaxation times of the perylene V7 mode in the n- alkanes indicate the presence of short range order in solution, and we have discussed this point in detail in Chapter 5. Rotational diffusion measurements of perylene in the n- alkanes Show that solvent organization exists on a length scale much less than 10 A, the “length” of the perylene molecule.‘6| The state-dependence of the relaxation times measured for the perylene V7 mode is discussed below. T. for V7 is the same for all three n-alkanes to within the experimental uncertainty and that T.’ is the same for the three solvents, but T.‘ < T. for V7. It is 64 important in and of itself that the T. times for both the V7 and V7. modes are solvent- independent for these three n-alkanes, but perhaps of more importance is that the relaxation times for V7 and V7. are different. There are two possible reasons for the difference between T. and T.‘ for these modes. The first is that the state-dependence arises from intramolecular changes in the coupling between the vibrational modes of interest and lower energy modes. A difference in anharmonic coupling between modes in the two electronic states is possible, as indicated by the ~21 cm'l blue shift of V7 on 121 excitation. Indeed, there are several modes, both IR and Raman active, in Close energetic proximity to V7 and V7., and the state-dependent frequency shifts seen for these modes are not, in general, the same as those seen for V7 and V7.12“ Any coupling between these modes will necessarily vary with the anharmonicity of each mode and their frequency differences. Because the measurements were performed in a room temperature liquid and the spectral resolution of the system is ~15 cm", one cannot separate cleanly anharmonic coupling effects from the non-selective simultaneous excitation of several nearly degenerate modes. The second possible reason for the difference in T. and T.° is that the intermolecular relaxation pathways available to V7 and V7. are different. If such a state- dependent intermolecular process is dominant, then it must be due to a V-V resonant, i.e. non-collisional. interaction between the perylene modes and the n-alkane solvent. A prerequisite for the existence of this mechanism is. of course. that there is a solvent vibrational resonance in the vicinity of the perylene 1375 cm'l mode. The n-alkane solvents possess a vibrational resonance at 1378 cm'l corresponding to a rocking motion of the terminal the CH; groups. Other recent data on ground State T. times for perylene in a broader series of n-alkanes indicate that this V-V coupling can be strong under certain 65 conditions, and may contribute to the state-dependent relaxation that is seen in this work”! V-V processes are believed to dominate the intermolecular contribution to the, observed relaxation because a state-dependent change in the inelastic collisional rate would require substantial local heating on excitation. For these experimental conditions the transient temperature rise for the perylene molecule should not exceed several K at most. It is likely that neither the intermolecular nor the intramolecular processes by themselves account completely for the measured difference in T. and T.‘ for V7, but, rather, that both factors combine to produce the observed result. 66 4.5. Conclusions The feasibility of measuring vibrational population relaxation times of complex organic molecules in dilute solution for both their ground and excited electronic states has been demonstrated. Specifically, we have examined the state dependent T. relaxation time of the perylene V7 mode in three n-alkanes and find that for a given electronic state the T. times are solvent independent. The excited state (V7.) mode relaxes approximately twice as rapidly as the ground state (V7) mode. This difference in relaxation times are attributed to either state-dependent changes in the anharmonic coupling Ofthis mode to other modes of equal or lower energy or to changes in the efficiency of intermolecular coupling to the solvent bath modes. If the latter mechanism is operative, then the intermolecular coupling must be predominantly through a near-resonant V-V channel. Further experimental work is in progress to elucidate the dominant relaxation pathway for this mode. 67 46. Literature cited 1. S. A. Hambir; Y. Jiang; G. J. Blanchard; J. Chem. Phys. , 28 6075 (1993). 2 S. Matsunuma; N. Akamatsu, T. Kamisuki; Y. Adachi; S. Maeda. C. Hirose, .l. Chem Phys, 88 2955 (1988). 3. M. A. Kovner, A. A. Terekhov; L. M. Babkov; Opt. Spectrosc. Lett. 28 (1971) 4. S. J. Cyvin; B. N. Cyvin; P. Klaeboe; Spectrosc. Lett, E 239 (1983). 5. Y. Jiang; G. J. Blanchard; J. Phys. Chem, 28 9411 (1994). 6. Y. Jiang; G. J. Blanchard; J. Phys. Chem, 28 6436 (1994). CHAPTER 5. VIBRATIONAL POPULATION RELAXATION OF PERYLENE IN n-ALKAN ES - THE ROLE OF THE LOCAL SOLVENT ORGANIZATION IN LONG RANGE VIBRATIONAL ENERGY TRANSFER Summary The vibrational population relaxation times of the Raman active v7 mode (1375 cm") and (v7 - v.5) combination mode (1733 cm") of perylene in eight liquid n-alkanes were measured using ultrafast Stimulated emission spectroscopy. The vibrational population relaxation time of the perylene V7 mode ranges from ~300 ps to < 10 ps depending on the n-alkane solvent chain length, but there is no Simple correspondence between alkane length and T. for V7. Energy transfer from the perylene V7 vibrational mode to a specific "-alkane solvent vibrational mode is dominated by long range resonance coupling. The Perylene (v7 + V. 5) combination mode exhibits additional efficient relaxation pathways for difI‘erent length n-alkanes. These data point collectively to short range order in the n- a”(fine solvent surrounding perylene molecule. 68 69 5. 1. Introduction In low pressure gas phase experiments, where an excited molecule is comparatively isolated from its neighbors, vibrational energy will dissipate slowly within the molecule to lower energy modes according to the extent of anharmonic coupling between the vibrationalmodesln For probe molecules in liquids, intramolecular relaxation is often less important than direct intermolecular relaxation because the number of collisional interactions with the surrounding medium is large and individual molecules are in closer spatial proximity to one anotherml In solution, excess vibrational energy within a molecule can be dissipated directly into the surroundings, sometimes very efficiently. Intermolecular vibrational population relaxation processes include energy transfer from solute vibrational modes into the translational, vibrational and rotational degrees of freedom of the solvent. The transfer of energy into these degrees of freedom can proceed by short range inelastic collisionsli's' or, for vibrational to vibrational (v-v) processes, can also proceed by long range polar coupling between excited solute Vibrational modes and the Vibrational states of solvent molecules.'6'9' Long range v-v processes occur because of dipole - dipole, dipole - quadrupole, dipole - induced dipole or other polar interactions. FOF collisional (short range) v-v energy transfer, the acceptor and donor molecules must necessarily be separated by fractions of an angstrom during the collision, and for long range processes, the efficiency of energy transfer will scale with r'". where it depends on the nature of the coupling and varies according to the model used.Ill For any given Vibrationally excited molecule, there is a non-zero probability that it will relax according to any Of the aforementioned processes. The interests of this work is in determining 7O experimentally which process(es) dominate relaxation in room temperature liquids. There is very little information available in the literature on understanding in detail the pathways of vibrational population relaxation in complex organic systems, and one must necessarily assume, as a starting point, that the physics of the problem are essentially the same as those of gas phase vibrational relaxation processes. The vibrational population relaxation rate constant is proportional to the probability of the transfer process. The probability of collisionally mediated, short range v-v energy transfer ranges up to ~10'3 depending on the difference in vibrational resonance frequencies of the donor and acceptor species and the frequency of collisionsm The probability of long range polar V-v process ranges up to ~l, depending on the difference between the donor and acceptor vibrational frequencies as well as the nature of the polar interaction responsible for the energy transfer. An important and distinctive feature of long range polar v-v processes is that the probability of a relaxation event depends on the distance between the donor and the acceptor and the alignment of the species. The probabilities of collisional v-t,r processes (up to ~10'5) are Wpically much smaller than either of the v-v processes.'” Separating the contributions of each of these processes in a given chemical system requires the examination of several relaxation pathways available to a given vibrational excitation, as well as the ability to vary eXperimental conditions in a controlled manner. To gain predictive control over vibrational energy flow in complex chemical systems, one needs first to determine the relative efficiency of each vibrational relaxation mechanism in a comparatively simple organic system and identify the chemical reasons for these relative effiCiencies. Much work has been done on small and medium size molecules in the gas 71 phase.“"” where the density of species available to accept vibrational energy is low, and in low temperature solid matrices,“°'”] where the spatial relationships between donor and acceptor are well defined and fixed on a time scale that is long compared to the dissipation of energy. Significantly less work has been focused on understanding vibrational relaxation of medium size molecules in solution at room temperature.[2'3'”’ It has been demonstrated previously that the efficiency of each relaxation pathway available to the excited molecule is determined by the specific chemical system and vibrational mode excited, and it is the purpose of this work to understand the dominant processes responsible for vibrational pOpulation relaxation of the v7 and v.5 modes of perylene in a series of n-alkane solvents. 5 .2. Experimental Ultrafast stimulated spectroscopy: The spectrometer used for Stimulated measurements Of perylene has been described in detail in Chapter 2““ Stilbene 420 dye (Exciton) is used for both dye lasers. The pump laser is operated at 434 nm, corresponding to the 0-0 transition for perylene in n-pentane and n-hexane or at 437 nm to pump the 0-0 transition OF Perylene in the longer n-alkane solvents. The probe dye laser is operated in the range of 460 to 480 nm. with the exact wavelengths depending on the mode accessed and the 0- 0 transition energy. Steady state spectroscopies: The steady state absorption spectra of all perylene solutions Were measured using a Beckman DU-64 spectrophotometer, with ~l nm resolution. Fluorescence spectra were recorded using a Perkin-Elmer model LS-5 fluorescence 72 spectrophotometer, with ~1 nm resolution. These data were used to estimate the spectroscopic origin of perylene in each solvent. Chemicals and sample handling: Perylene and all the n-alkane solvents were purchased from Aldrich Chemical Company and used as received. The concentrations of the perylene solutions used for the time resolved stimulated measurements were ~10 .184. To minimize thermal lensing contributions to the Signal, sample solutions were flowed through a 1 mm pathlength flow cell. The sample temperature was controlled at 30.0 : 0.1 K using a thermostatted bath. For experiments using n—octane-dm. Stimulated measurements were performed in a 1 cm pathlength cuvette equipped with a magnetic stirrer As a check, perylene relaxation in n-octane-hm was measured in the same cuvette and results were identical to those obtained using the flow cell. 5.3. Results and Discussion The absorption and emission spectra of perylene in n-octane are presented in Figure 5.1. The absorption and emission spectra of perylene in other n-alkane solvents are similar, with only minor solvent -dependent band shifts. The excitation wavelength is indicated by Ihe arrow and the vibrational modes accessed are contained within the emission feature 73 origin _ ©© 1.0 - ©©© lnlensrty(a u ) 0 L11 1 > r11 0.0 t 4 ‘ t - : .L c . 350 400 450 500 550 wavelength (nm) Figure 5.1. Steady-state absorption and emission spectra of perylene in n-octane. l \ x ‘ ’ y I ‘ A V \ V V I 1 A A \ A V V A A , V ‘ U x ’ 7 p < i 3“ re 5 .2. Exaggerated atomic displacements associated with the v7 and V15 normal modes of perylene. 75 boxed in Figure 5.1. While there are a variety of Raman active vibrational modes accessible to this measurement scheme,“7°’9' the focus has been on the relaxation dynamics of two particular vibrational modes in the ground electronic state of perylene; the V7 fimdamental mode at 1375 cm’1 and a combination mode composed of the V7 and the v.5 (358 cm") modes, at 1733 cm'1 (Figure 2). The V7 mode is an in-plane ring deformation mode, and the v.5 mode is the dominated by the in-plane stretching motion of the center ring between the two naphthalene moieties. Both of these modes are of ag symmetry and are Raman active. The (V7 + V .5) combination mode was used instead of accessing the v.. mode directly because spectral overlap of the Steady state absorption and emission bands prevents separation of ground and excited electronic state responses in the region of the (v01. - v.5) resonance. The relaxation times measured for both the V7 and (V7+V15) modes depend sensitively on the identity of the n-alkane solvent, and this aliphatic chain length- dependence provides direct evidence for solvent structure around the perylene molecule. The origins of the solvent dependent response will be discussed in below. In Figure 5.3 an example of the experimental stimulated signal, S(t), for perylene in n- octane excited at 437 nm, and probed at 472 nm to access the (V7 + V..) mode is shown. The spectroscopic selectivity is sufficient to distinguish between individual perylene -1 [16] Vibrational modes in liquidsm The frequency width of the probe laser pulse is ~4 cm While the vibrational linewidths of the perylene resonances are significantly broader at 300 K- The time resolved stimulated Signals were fitted using Equation 3. The vibrational AT/T (a.u.) Figure 6.3. 0.8 0.6 0.4 0.2 0.0 76 T1 = 60:14 p5 7, = 2180214 ps \ 0 200 400 600 800 1000 delay time (ps) Stimulated response of the (V7 + v..) mode of perylene in n-octane, presented with the cross correlation response function. 77 SUU — 400 — . + 300 - _L 0 Z _ 0 I375 om'l mode s: 0 I733 an" mode e— 200 — 1734 an“=1375cm"+359cm" ,. L T IUH r- d I . § § 1 § 1) O 1 l I l O 4 6 8 10 l2 I4 16 Number of C's in n-alkane Figure 5.4. T. relaxation times for v7 (0) and (w + v”) (0) modes in perylene as a function of solvent chain length. 78 relaxation times, T 1 = k;", in the eight n-alkane solvents we have studied, are presented in Figure 5.4 and Table 5.1 for the v7 and (v7+V15) modes. Table 5.1. T. times for the v7 and (v7 + V”) modes of perylene in n-alkanes. solvent v- T, (v- - V15) T1 (ps) (ps) n-Csle 276:46 286:.36 n-C6H14 281:177 < 10 n-Cva 355:100 316:9 n-CgHis 30:3 60:14 n-Cgqu 125:26 56:17 n-Cmsz 376:92 273:183 Il-Clezo 410:106 41:10 II-C16H34 < 10 < 10 II-Cngs 298:102 ----- Each of these data values can be reproduced at will and are determined by regression of the average of at least 15 individual time scans. The large uncertainty in these values is reflective of the small fractional contribution of the k; process to the overall transient gain resDonse (Equation 3). For the v7 mode, T1 ranges from 30 : 3 ps in n-octane and <10 ps i n N~hexadecane to ~3OO ps in the other n-alkanes. The majority ofthe T1 times are found ‘0 be much longer than the 10 ps instrument response time, and thus no attempt at deconvolution was made to determine the exact T. values for the few conditions where the relaxation proceeds in < IO ps, .The unexpected dependence of T. on solvent identity provides an important indication of the mechanism by which vibrational energy relaxes frOm perylene. For the (v7+vl5) combination mode, the measured T. times are 79 qualitatively similar to those for the v7 mode, but exhibit additional efficient relaxation pathways, which must necessarily be associated with the V” mode. There are at least two questions regarding these data which require an answer at some level, and, indeed, the information contained in the T1 data provide insight into vibrational population relaxation at a molecular scale. The first consideration lies in determining the dominant relaxation mechanism, and the second question is centered on any local molecular structural information the solvent chain length dependence provides. For the dilute solutions that were investigated, intermolecular energy exchange between perylene molecules will contribute negligibly to the observed response, and the focus is on the interactions between perylene and the surrounding n-alkane solvents. While anharmonic coupling in perylene is significant, as is apparent from the prominence of the (V7+V15) combination mode, it is expected that intramolecular energy transfer to low energy modes Will contribute little to solvent chain length dependence in the T1 times that were "'1 easured, because the linear optical response of perylene appears to be virtually solvent- independent. The intramolecular relaxation processes for perylene were approximated to be solvent-independent. It should be noted that the separation of intermolecular and intramolecular relaxation processes is more likely to be accurate for a rigid molecule such as perylene than for a labile molecule, where photoisomerization or other large amplitude molecular motions can influence the efficiency of intramolecular decay pathways.'2°‘2” 1 h principle, intermolecular energy relaxation from a solute molecule to a solvent molecule can proceed by collision-mediated short range v-t,v,r, processes or by long range resonance polar v-v coupling. Except in carefully chosen cases, the probability of 80 collisional v-t,r processes is much smaller than that for collisional v-v processes. For. most liquid phase systems, it is expected that the dominant relaxation pathway will be through v-v coupling, and the central mechanistic question is whether the coupling is short range, collisionally mediated or long range, through-space in nature. For the chemical systems we have chosen to examine, we believe that the dominant relaxation processes must be v-v because, in n-alkanes, the terminal CH3 rocking mode resonance occurs at 1378 cm'1 and this resonance exhibits negligible variation among the different n-alkanes. The perylene v7 mode is centered at 1375 cm", functionally degenerate with the solvent CH; rocking bath IT) ode. Since there is no applicable theoretical treatment of different v-v processes in room temperature liquids, gas phase treatments were used in attempting to distinguish between long range polar and short range collisional relaxation processes. The probability of a v-v energy transfer event for short range collisional processes betheen the donor and acceptor species at exact resonance is given by / all/PH" 2 WhEre Uir is a vibrational matrix element for the collisional interaction, u is the reduced mass of the system undergoing the collisional interaction, and L is the length scale over which the collisional interaction can take place, Le. L is the effective intermolecular di 31 ance of the colliding species when v-v energy transfer takes place. The value of L will depend on the chemical identity of the colliding species, and is typically taken as L = 0.2 A f0? both v-t and v-v gas phase collisional interactions. While the value of L may not be precisely the same in liquids, it is not likely to be much different because the fundamental 81 nature of the interaction responsible for the energy transfer is the same in both the liquid and the gas phases. For typical experimental conditions, a maximum value of <

> for collisionally mediated short range energy transfer is ~10'3. It is important to note that, for short range collisional processes, the probability of energy transfer does not vary smoothly with the distance between molecules. For an energy transfer event to occur, the donor and acceptor molecules must be in intimate contact. Long range energy transfer processes are fimdamentally different than collisionally mediated energy transfer. Long range interactions involve polar coupling between donor and acceptor species. 4yC3 -—-— [51 / 1)‘: \< >2 file/VT There are a number of different formulations for long range v-v energy transfer,l("8' and all yield qualitatively the same result. The term C contains information on the matrix elements for the donor and acceptor vibrational transitions involved in the energy transfer Process as well as geometric alignment terms. <

> for long range energy transfer depends inversely on the distance, d, between the acceptor and donor species, with the ex act distance dependence being determined by the type of interaction (i.e. dipole-dipole, Clipole-quadrupole, dipole-induced dipole, etc). Thus the length scale over which this er‘ergy transfer takes place depends on the chemical system, but, for exact resonance COTlditions and in a condensed phase system, where d is small, <

> ~ 1‘” This latter mechanism is significantly more efficient than short range, collision-mediated v-v energy trE‘lnsfer for our experimental conditions. In addition to the predictions of the above referenced models, the experimental data indicate the dominance of long range energy tr'cansfer. 82 The data in Figure 5.4 show that T. for the v7 mode does not change smoothly with the length of the aliphatic solvent. If collisional interactions dominate the relaxation process sensed by the T. measurements, then we would expect a smooth progression of T. times that is proportional to the frequency of collisions between perylene and the terminal methyl groups of the alkane solvent molecules. The frequency of collisional interactions between the solvent and the solute should vary with the solvent viscosity and density, both of which are well-behaved functions of aliphatic chain length: If short range v-v energy transfer were responsible for the data shown in Figure 5.4, then collisional interactions between perylene and n-octane would have to be a factor of ~l0 more frequent than they are for either n-hexane or n-decane, and this possibility is physically unreasonable. The dominance of long range polar v-v energy transfer implies necessarily that the solvent exhibits local structure about the perylene molecule. Equation 5 shows that the P rObability of a long range energy transfer event is related to the separation distance of the donor and acceptor, d, and the term C in Equation 5 also contains a geometric factor for aligriment of the species. The solvent acceptor mode is significantly localized on the t erI'hinal methyl groups on the alkane chains, and thus it is expected the efficiency of v-v transfer to be proportional to the distance between perylene and the terminal methyl grOUps of the solvent. The observed change in T. for the perylene v7 mode in n-octane and n-hexadecane indicates that the terminal methyl groups of these solvents are, on aVerage, in closer proximity to the perylene molecule than are the terminal methyl groups 0f the other n-alkane solvents. -l m-__~.__ awn... ‘M_~_ We expecte the response for the (v. + v.5) combination mode to be a superposition of the. responses of each fundamental constituent mode. The same response for the (v. + v.5) mode was observed as for the v7 mode, and in addition, efficient coupling to the solvents n-hexane 'and n-dodecane, and to a lesser extent for n-nonane were observed. The efficient coupling of the v.5 mode to the surrOUnding bath modes is likely due to v-v long range resonance processes, but, in contrast to the v7 mode, the solvent and solute resonances are not at exactly the same frequency and the length scale of the coupling may be different than for the V7 mode because coupling may proceed from the perylene Raman active mode to either Raman (AOL) or infrared (Au) active modes of the solvent. Despite these possible differences in the nature of the coupling, the qualitative information content ofthe data on the (v. + v.5) combination mode is expected to be the same as that for the v7 mode. The enhanced coupling of the v., mode to the solvents n-hexane and n- dodecane arises from arrangements of the solvent around the perylene molecule that are sensitive to the motions ofthe v.5 mode. Because these motions are significantly different than those for the V7 mode. a different solvent-dependence is expected for the relaxation of this mode. To establish the dominance of v-v processes in our measurements, especially for the cases where efficient relaxation occurs, T. for the v7 mode in n-octane-d... (II-C3D...) was measured. For this system T. = 298:102 ps was obtained, in contrast to 30:3 ps in n- octane-h,,.. For n-octane-d... the terminal CD3 rocking mode resonance occurs at 1050 84 cm", Am = 325 cm'1 for v-v relaxation from the perylene v7 mode to this mode. The detuning dependence of <

> is given by [71 27r2C 2150,11 pu'fl/ P = , exp — << >1 r-IthdsU [:7 2k?" where [6] . (ZdAaflch '1’3 u = y The experimental T. values, in conjunction with Equations 5 and 6 allow us to estimate the average perylene-solvent methyl group spacing. For a frequency difference of ~300 cm", for long range resonance energy transfer, <

> was estimated to be ~ 0.07 for d=l A and <

> ~ 0.0005 for d = 2 A. These values of <

> for perylene in n-octane-d... yield T. ~ 430 ps for the perylene v7 mode in n-octane-d... for d=l A and T. ~ 4.2 ns for d=2 A, based on the experimentally'observed time of T.=30 ps for perylene in n-octane- In... The assumption is made that the deuteration of n-octane does not substantially alter its solvation characteristics, and, if this assumption is valid, one can estimate the perylene- methyl group spacing, d ~ 0.9 A in the n-octanes from the ratio of the experimental T. times, i.e. <

> smucm-l ~ lO<

> x... 3......m-l (see Figure 5.5). From this estimate ofd for the n-octanes and n-hexadecane, a T. of ~ 300 ps for the perylene v7 mode in other H- alkanes suggests that the average perylene-methyl group separation for these solvents is ~l .7 A. There is also stoichiometric uncertainty involved in the interpretation of these numbers. At this point it is important to note that the value of d we report correspond to points of cloest contact, not intermolecular distances. 85 1.0 — Aw=0 / .. 0.8 [- 1 , Aw=300 crn'l L / . /\ 0.6 - A a. v v 0.4 ~ \ \\ 02 ~ “ \ r \\ 0.0 K - l 4 1 . d (K) Figure 5.5. Calculated probability, <

>, for long-range energy transfer for exact donor-acceptor resonance, 0) = O (eq.5) and a) = 300 cm", as a function of donor-acceptor separation, d. It is not clear to what extent the terminal CH3 rocking mode in n-alkanes behaves a collective motion of both CH3 groups, as opposed to acting as a doubly degenerate but spatially separated mode within an individual molecule. Despite this uncertainty, the above estimates of d seem entirely plausible for a liquid phase system. Also for n-octane- d... there is another vibrational mode, the CD2 scissors motion at 1080 cm'1 (Aw = 295 cm' I) which can contribute to the measured relaxation time. It is likely that both of these solvent modes act as acceptors for the perylene v7 mode. For a frequency difference of Am ~ 300 cm", <

> for collisionally mediated transfer falls to ~ 105, indicating that, for perylene in n-octane-d,,,, the dominant relaxation mechanism remains long range resonant v-v coupling. For gas phase systems, the cross-over point between v-v long range and collisional process dominance has been estimated to occur for A0) ~ 250 cm",‘81 Clearly the density of the bath medium has a significant effect on vibrational relaxation. [20.21] and 3 Mode- and solvent-specific intermolecular interactions have been seen before central question in all such work is the nature of solvent organization around the solute. The observation of mode-dependent .coupling to different n-alkane solvents invites speculation on geometric arrangements of the solvent about the solute. Such a practice is, of course, extremely speculative, and should be taken as such. In this context, we offer only two observations. The dominant motion of the v7 mode is a distortion of the individual naphthalene moieties, and this mode is found experimentally to couple efficiently to n-octane and n-hexadecane. The “length” of n-octane, if it were in an all- trans conformation, is quite close to that of the perylene long axis, which spans both of the naphthalene moieties. In contrast, the dominant motion of the v.5 mode is an in-plane 87 moving together and apart of the individual naphthalene groups, and this mode is observed to couple strongly to n-hexane and n-dodecane. Because the solvent resonances to which the v.5 mode couples are not as localized as those to which the v7 mode couples, any relevant geometric constraints are less well defined, but we note that the “length” of n- hexane is close to that of naphthalene. For both modes the coupling efficiency exhibits what appears to be a periodic effect, i.e. v7 couples to n-CgH... and n-C.6H3..; v.5 couples to "-C.;H... and n-C.2H26. The origin(s) of this effect are, at present, unclear, but suggest a regularity in the way aliphatic chains organize around a solute molecule. These postulations are reminiscent in some sense of Shpol’skii‘s work on perylene in n-alkane crystals at 77 Kim Shpol’skii used steady state emission and absorption measurements to observe perylene spectral line narrowing in frozen n-alkane matrices. For such measurements it is reasonable to expect that local structure will persist for the lifetime of the emitting state (several ns). In liquids there is a significant body of information that 1220-231 (vide infra), albeit with a persistence time points to short range solvent structure much shorter than for solids. It is possible that the mode-specific short range transient order detected in liquid n-alkanes is related to the structural effects detected by Shpol’skii, but the connection between these bodies of data remains unclear at present. The T. data for both the V7 and (V7 + v.5) modes indicate that local solvent structure is important to vibrational energy relaxation. The length scale over which such structure persists is not clear from the T. measurements alone, but other dynamical measurements can place bounds on this persistence length. The rotational diffusion dynamics of perylene in these same n-alkane solvents show that, while the boundary condition changes between 88 the solvent and the solute at ~n-octane, there is no discontinuous change in the viscosity '2‘” A discontinuous response is expected only if dependence of the reorientation time. there is a substantial solvent chain length dependent change in solvent ordering in the vicinity of the solute. The reorientation data are sensitive to changes in the relative hydrodynamic volumes of the solute and the solvent, but show no evidence of comparatively long range solvent structure. Thus the local structure sensed with the T. measurements persists on a length scale much shorter than the perylene molecule (~10 A). Finally, the need for a better model for vibrational energy transfer in liquids, particularly for v-v long range resonance coupling vibrational energy transfer in liquids, is needed so to aid the prediction and interpretation of these experimental results. 89 5.4. Conclusion The aliphatic chain length dependence of vibrational population relaxation for the v7 fundamental and (v. + v.5) combination modes of perylene in dilute solution have been measured. The measured T. times do not vary smoothly with solvent aliphatic chain length. For certain solvent alkane chain lengths, vibrational energy in the perylene molecule couples efficiently to the bath modes of the surrounding solvent. The dominant mechanism for this vibrational population relaxation is long range resonance v-v energy transfer. The observation of efiicient solvent-solute coupling for specific solvent aliphatic chain lengths demonstrates the existence of persistent local structure in this chemical system. Data from rotational diffusion measurements on perylene in the n-alkanes shows that the local solvent ordering exists on a length scale significantly shorter than the length of the perylene molecule (~10 A). The sensitivity of individual solute vibrational modes to different components of local solvent structure offers the ability to interrogate selectively the presence of structure in the solvent cage of a variety of condensed phase systems. 5.5. Literature cited 1. J. T. Yardley; Introduction to Molecular Energy Transfer, Academic Press, 1980. to . S. A. Hambir; Y. Jiang; G. J. Blanchard;J. Chem. Phys, 98 6075 (1993). . Y. Jiang; G. J. Blanchard; J. Phys. Chem, 98 9417 ( 1994). b) 4. J. T. Yardley; C. B. Moore; J. Chem. Phys, 4_6 4491 (1967). 5. C. B. Moore; Adv. Chem. Phys, Q 41 (1973). 6. B. H. Mahan; J. Chem. Phys, 4_6 98 (1967). 7. R. D. Sharma'. C. A. Brau; J. Chem. Phys, _SQ 924 (1969). 8. J. C. Stephenson, R. E. Wood; C. B. Moore; J. Chem. Phys, 48 4790 (1968). 9. J C. Stephenson, C. B. Moore, J. Chem. Phys, 52 2333 (1970). 10. T. C. Chang; D. D. Dlott‘, Chem. Phys. Lett, £2 18 (1988). l l. J. R. Hill; D. D. Dlott', J. Chem. Phys, 8_9 830 (1988). 12. J. R. Hill; D. D. Dlott;J. Chem. Phys, 8_9 842 (1988). 13. T. C. Chang; D. D. Dlott,J. Chem. Phys, _9_O 3590 (1989). 14. H. Kim; D. D. Dlott,./. Chem. Phys, % 8203 (1991). 15. T. Elsaesser; W. Kaiser; Ann. Rev. Phys Chem, Q, 83 (1991). 16 Y. Jiang; S. A. Hambir, G. J Blanchard. Opt. Commun, 92 216 (1993) 17. E. V. Shpol‘skii; R. l. Personov; ()pt. .S'pectrosc., 8 172 (1960). 18. S. J. Cyvin, B. l\'. Cyvin, P. Klaeboe, .Syiec'tt'osc. Lett, 1_ 239 (1983). 19. S. Matsunuma; N. Akamatsu, T. Kamisuki; Y. Adachi; S. Maeda; C. Hirose, J. Chem. Phys, E 2956 (1983). 20 W. L. Weaver; L. A. Huston; K. lwata; T. L. Gustafson; J. Phys. Chem, % 8956 (1992). 21 R. M. Butler; M. A. Lynn; T. L. Gustafson;J. Phys Chem, 91 2609 (1993). 91 . D. McMorrow; W. T. Lotshaw; J. Phys. Chem, fl 10395 (1991). . D. McMorrow; W. T. Lotshaw; Chem. Phys. Lett., 29; 369 (1993). . Y. Jiang; G. J. Blanchard; J. Phys. Chem, 28 6436 (1994). CHAPTER 6. ROTATIONAL DIFFUSION DYNAMICS OF PERYLENE IN n- ALKANE - OBSERVATION OF SOLVENT LENGTH DEPENDENCE CHANGE OF BOUNDARY CONDITION Summary Orientational relaxation dynamics of perylene in both its ground electronic state and its first excited singlet electronic state in the series of n-alkanes n-pentane through n-decane, n-dodecane and n-hexadecane were investigated. A curvilinear relationship between orientational relaxation time and solvent viscosity was observed, and these data were interpreted in terms of a solvent length-dependent change in the nature of solvent-solute interactions. 92 6. 1. Introduction Understanding molecular-scale interactions between solvents and solutes has been a subject of long standing interest because these interactions play a deterministic role in many chemical processes and reactions. The central focus of many of these investigations has been the nature of the local structure induced in the solvent by the presence of the solute, and the mechanisms that are operative in the interactions between the two chemical species. Indeed, these are questions that, for a given chemical system, have answers that depend on the length and time scale of observation. One of the more widely used experimental approaches aimed at understanding these complex interactions has been the measurement of the rotational diffusion dynamics of a probe molecule dissolved in selected solvents, where some property of the solvent such as bulk viscosity, dielectric response, static dipole moment or molecular structure is varied in a regular mannern’m The utility of rotational diffusion measurements stems from the fact that the length scale of the measurement is relatively well defined by the size of the probe molecule, and that- comparatively straightforward means exist for the theoretical treatment of the data. While it is difficult to elucidate specific molecular interactions between solvent and solute using rotational diffusion measurements except in special circumstances, such measurements detect the solvent-solute interactions over a time scale where a large number of molecular collisions occur. thereby providing insight into the average environment experienced by the solute. 94 Many probe molecules in a wide range of solvent systems have been examined using rotational diffusion measurements. There are essentially two classes of experiments; those where polar probe molecules in polar solvents are used and those where non-polar probes and solvents are used. The reorientation of the probe molecule in each type of system is mediated by different types of intermolecular interactions. For polar systems, dielectric “81 and even the formation of comparatively long-lived frictionm’, dipolar interactions solvent-solute complexes“9‘2°] have been shown to contribute to the observed response. The multitude of comparatively long range interactions in polar systems yields experimental results that are often ambiguous. Studies of molecular reorientation in non- polar systems have demonstrated that, despite the absence of long range electrostatic interactions, the data can contain contributions from a number of weaker interactions, with at least a portion of the response being controlled by solvent “structure” on a length scale comparable to the reorienting probe moleculem‘m In this chapter the rotational diffusion results are presented for the non-polar probe molecule, perylene, in a series of n-alkanes, where the size ofthe solvent molecules is varied in a regular manner. The data show that, when the solvent molecules are significantly smaller than the solute, near stick-limit hydrodynamic behavior is seen. and when the length of the solvent approaches that of the probe molecule, a change to slip-limit behavior is seen. Further, as the solvent molecular size is increased, perylene changes its effective rotor shape from essentially a spheroid to a prolate ellipsoid. 5.2. Background The rotational motion of a solute imbedded in a solvent has been treated theoretically by a~ number of workerslzmsl A common thread to these treatments is the approximation that the solvent surrounding the probe molecule is essentially a continuous medium, 1‘. e. there is no explicit consideration of the intrinsically molecular nature of the solvent surrounding the probe molecule. A well established starting point for most treatments of rotational [271 diffusion is the Debye-Stokes-Einstein (DSE) equation, 1 tyl' _ _ l TOR if] kl. [ l where a... is the orientational relaxation time constant, n is the solvent bulk viscosity, V is the hydrodynamic volume of the solute, and D is the rotational diffusion constant. The assumptions implicit in Equation 1 are that the solute is spherical and that the solvent is a continuous medium. Both of these assumptions are necessarily limited in their applicability, but for cases where the solute is much larger than an individual solvent molecule, the DSE model holds quantitatively. Both of the assumptions in the DSE equation realize their limits in chemical systems where the solvent and solute have similar hydrodynamic volumes. For polar solutes of V~300~500 A3, the DSE equation can predict ton to within a factor of ~2,'°' despite the fact that this type of chemical system possesses the largest number of intermolecular interactions likely to cause deviations from the simplistic DSE picture. For non-polar systems, the difference between experimental data and DSE predictions is substantial, often greater than a factor of two different, with the experimental values of TOR being faster than the DSE prediction. To account for these discrepancies, several groups have modified the DSE model to consider the boundary 96 condition for the solvent-solute interface. These works do not consider the molecular nature of the solvent, but instead treat the solvent-solute interaction as purely frictional, with a variable friction coefficient that depends on the solute shape. These corrections enter multiplicatively into the DSE equation. 771' f TOR =17; [ where f is a friction term to account for the solvent-solute interaction. The value off can range from near zero in the slip limit to unity in the stick limit, depending on the shape of 1341 the efiective rotating ellipsoid. S is a shape factor, determined from Perrin’s '281 which accounts explicitly for the non-spherical shape of the solute. The use equations, of these correction factors typically brings theory into significantly closer agreement with experiment for many non-polar systems. While this model does not account for the inherently molecular nature of rotational diffusion dynamics, it does provide a useful and nearly quantitative basis for the interpretation of our experimental data. 5 . 3. Experimental Spectrometer. The picosecond pump-probe laser spectrometer used to measure both the ground state and excited state rotational diffusion dynamics of perylene in several n- alkanes is same as the one used for vibrational relaxation T. measurements and has been described in detail in Chapter 2 The excitation or pump dye laser was operated at 437 nm using Stilbene 420 laser dye (Exciton). The output of this laser was ~ 60 mW average power with a 5 ps FWHM autocorrelation trace at 8 MHz repetition rate. The probe dye 97 laser was also operated with Stilbene 420 laser dye, and was set to either 429 nm for ground state recovery experiments to obtain the data for ground state reorientation dynamics or 470 nm for stimulated emission experiments for excited state reorientation dynamics. The probe laser wavelengths were chosen to coincide with the non-overlapped regions of the perylene absorption and emission spectra (Figure 6.1). Chemicals and sample handling. Perylene (99%) and all n-alkane solvents were purchased as their highest purity grade available from Aldrich Chemical Co. and were used as received. All perylene solutions were ~1x10°5 M and were flowed through a temperature controlled flow cell to minimize thermal contributions to the experimental signal. For all rotational diffusion measurements, the temperature of the sample was maintained at 300 : 0.1 K. 6 4. Results and Discussion Both the ground state and first excited singlet state rotational diffusion dynamics of perylene in n-alkane solvents n-pentane through n-decane, n-dodecane and n-hexadecane were measured. The data for R(O), ton. R°(O) and TOR. are presented in Table 6.1. 98 lntensity(a. u.) 0.0 ¢ + : T c . : fi 350 400 450 500 550 wavelength (nm) Figure 6.]. Normalized absorption and spontaneous emission spectra of perylene in n-octane. 99 Table 6.1. Experimental zero time anisotropies and reorientation times for perylene in several n-alkanes. The asterisks indicate an excited electronic state measurement. solvent 77 (cP)" 13(0) 1’10 10.. 1'10 R°(0) $10 {OR :16 (ps) (ps) n-Can 0.24 0.18 : 0.08 8 : 1 0.12 : 0.06 9 : 2 n-CaH... 0.33 0.32 : 0.03 14 : 3 0.33 : 0.05 11 : 2 n.c.H.6 0.41 0.26 a: 0.08 15 : 3 0.22 : 0.06 14 : 2 n-CgH... 0.54 0.33 : 0.03 16 x 1 0.21 : 0.08 19 : 4 "£911... 0.71 0.21 : 0.05 19 : 3 0.24 : 0.06 21 : 5 "-00sz 0.92 0.19:0.05 21 :3 0.30:0.09 19:3 new... 135 0.27 : 0.03 28 : 3 0.25 : 0.06 32 : 7 ”emit... 3.34 0.27 : 0.04 49 : 7 0.21 : 0.03 49 : 8 ‘1 Data from CRC Handbook of Chemistry and Physics, 7lst Edition, D. R. Lide, Editor, CRC Press. 1990. The rotational diffusion time constants through individual time resolved scans were determined, where the polarizations of the pump and probe laser pulses were set to be either parallel or perpendicular to one another, and present representative data sets in Figures 6.2a and 6.3a. The induced orientational anisotropy function was produced from the individual time scans according to Equation 3 (Figures 6.2b and 6.3b). [MU-II“) l“(l)+2]i(l) R(t): The ground state and excited state reorientation times for perylene are found to be identical for a given solvent, as is expected for a non-polar system. The number of exponential decays contained in R(t) is determined by the effective rotor shape of the reorienting species and the relative orientations of the pumped and probed transition 100 AT/I (a.u.) N o 10 ~ 0 O 100 200 300 400 delay time (ps) 0 4i . b 0.3 ~ 0.2 P E . 0.1 ~ 0.0 ——“WW—— 0 100 200 300 400 delay time (ps) Figure 6.2. (a) Time scans for ground state recovery response of perylene in n-octane. (b) R(t) signal produced from the experimental data shown in (a) using Equation 3. 101 1.0 " I.(I) AT/T (a u.) O 100 200 300 400 dela time s) 0.4 ~ y (p 0.3 - 0.2 ~ ‘1 R(t) 95"” 0.1~ \ delay time (ps) Figure 6.3. (a) Time scans for ground state recovery response of perylene in n-hexadecane. (b) R(t) signal produced from the experimental data shown in (a) using Equation 3. 0 0 WWW - _ . . {4W ‘ I 102 dipole moments in the probe molecule.133‘ In principle, R(t) can contain up to five exponential decays, but a single exponential decay is encountered for most systems. The “"11” and in at least rotational diffusion properties of perylene have been reported before, two cases, a double exponential decay of R(t) was found.'“‘391 There are several differences between those experiments and what are presented here. For the earlier experimental works, a double exponential decay is an expected result. For these experimental conditions, where the first excited electronic singlet state was accessed spectroscopically, it is expected to observe a single exponential decay functionality for R(t). The viscosity dependence ofthe data presented in Table 6.1 is shown in Figure 6.4. These data demonstrate a change in the effective boundary condition between the solvent and solute as a function oftheir relative sizes. 1n the limits where the solvent is significantly smaller or larger than the perylene probe molecule, a semi-quantitative agreement was obtained with the modified DSE equation in the stick and slip limits, respectively, (Equation 2). The change over between these two limits apparently occurs at n-octane The agreement of our data with the different limits of the hydrodynamic model is discussed below. The focus ofthe discussion is on the slopes ofthe two clearly linear regions shown in Figure 6.4. The quantity TOR/1’] is related to the hydrodynamic boundary condition as described in Equation 2. Because the same probe molecule for all of the measurements was used, the 103 DO _ excited state 40"- T .,.-'i’.’.”’ A O " N . ' I ’ a 3 .. , \ o , . ’ ground state P 20 .. _I’ 7 IO — ‘11“ 1: O + J I l A: I 0 l 2 3 n (GP) Figure 6.4. Ground state and excited state reorientation times for perylene as a function of n-alkane solvent viscosity. For all measurements the ground state and excited state reorientation times are the same to within the experimental uncertainty. 104 quantity [’5 must necessarily be the solvent-dependent variable. From Pern'n’s equationslm when modeling the probe as a prolate ellipsoid, the axial ratio of ellipsoid p = a/b = (3 + 7.4) / 10 = 0.52 (< l), where b is the long molecular axis and a is the short molecular rotational axis. (See diagrams in Figure 6.5) = . ~ — 141 (2__p2) _:_£____h, __p2 1 “l-le p , where S is the shape factor in Equation 2. For an oblate ellipsoid, p=wb=uo+2h/3=29pr) '7 _ -i- — ' p - 151 3 (2—p2) (p2 5) arctan «£02 — l) —p2 \“P' l s .1 These equations gave S = 0 69 if perylene is modeled as a prolate ellipsoid and S = 0.70 if perylene is modeled as an oblate ellipsoid. Therefore, because of the probe molecule shape, the only quantity that can contain a measurable solvent-dependence isf Despite the fact that S is virtually shape-independent for perylene, f contains information on effective rotor shape, and we discuss this point below. ‘ g b - axial axis Prolate ellrpsord a,c - other axes K\ i ,\ p = [(a +c)/2]/b f = [(3 + 7.4)/2]/IO w k = 0.52 < l Oblate ellipsoid =[ (a +c)/2]/b [(10 + 7 41/21/3 F1gure 6 5 Illustration of prolate and oblate ellipsoid rotor shape and the axial ratio, p. 106 1t.should be noted at the outset of this discussion that any changes in the effective rotor shape of perylene are manifested as changes in the relative values of the Cartesian components of D, the rotational diffusion constant, and not in the actual shape of the molecule. For the four lowest viscosity solvents that were examined, n-pentane through n-octane, the relationship between TOR and n is linear to within the experimental uncertainty, with a slope of 40 : 3 ps/cP. The hydrodynamic volume of perylene is calculted to be 225 A3140] For 5 = 0.7, T = 300 K, and f = 1 (stick limit), calculated TOR/T1 = 77 ps/cP. The difference between experiment and the prediction of the model is slightly less than a factor of two, but the data are not consistent with slip limit hydrodynamics (vide infra). For perylene reorienting as an oblate rotor, slip hydrodynamics predicts TOR/T] = 5 ps/cP and for a prolate rotor. the slip limit prediction is IOR/q = 16.5 ps/cP. Thus for n-alkanes C5 through C3 the observed reorientation behavior is intermediate between the slip and stick limits. The weak point in any such analysis is knowledge of the effective hydrodynamic volume V and thenon-spherical rotor shape correction, S. for the probe molecule. If perylene is assumed to reorient as a sphere (S=l) then the stick limit DSE prediction is IOR/n = 54 ps/cP. in excellent agreement with the experimental data. This nearly quantitative agreement is viewed as fortuitous, but indicative that the effective rotor shape of perylene in short chain alkanes is only weakly anisotropic. For the four longest chain n-alkane solvents, n-nonane, n-decane, n-dodecane and n- hexadecane. the slope of the best fit line for TOR/I] = 12.7 : 1.5 ps/cP, a factor of three 107 different from the value of IOR/n for the shorter n—alkanes. Clearly there is a fundamental change in nature of the interactions between perylene and n-alkanes as a fimction of alkane, length, and this change occurs for solvents longer than n-octane. The slope of TOR/T] for the longer n-alkanes is in excellent agreement with slip hydrodynamic predictions for perylene acting as a prolate rotor, i.e. reorientation predominantly along its x (long) axis (Figure 6.6). The slip prediction for perylene reorienting as an oblate rotor is significantly less than that observed experimentally. If we neglect the non-spherical shape correction (S=1), the slip limit for a prolate rotor is predicted to be 11.6 ps/cP and the oblate rotor is predicted to be 3.5 ps/cP. Thus, regardless of the extent of anisotropy in the shape of the perylene molecule itself, the long chain aliphatic solvents constrain its motion to be predominantly about its long axis. The experimental TOR/T] data in the longer n-alkanes do not provide the only evidence for perylene behaving as a prolate rotor. A supporting, but inconclusive, piece of evidence lies in the observed fiinctionality of the experimental R(t) decay curves. The Chuang and Eisenthal formulation for reorientation of an anisotropic probe molecule includes treatments for conditions where the measured transition dipole moments lie along arbitrary ~~ lé-‘l angles with respect to the molecular long and short axes. For the purposes of this discussion the x axis is assigned as the perylene long axis in the molecular plane, the y axis as the short in-plane axis and the z axis normal to the molecular plane, as indicated in Figure 6.6. “1‘ Ha- 108 7.421 10A 3A(z) Figure 6.6. Dimensions and Cartesian axis assignments for perylene. 109 Because perylene is planar the full Chuang and Eisenthal expression simplifies to R(t) = 0.3(,6’+ a)exp(—(6D + 2A)t) + 0.3(,6— a)exp(-(6D - 2A)t) [6] where a and B are terms relating the values of the Cartesian components of the rotational diffusion constant and the relative angles of the excited and observed transition dipole moments with respect to the Cartesian axes. D is the average of the Cartesian components of the rotational diffusion constant, and A is a term describing the anisotropy in the Cartesian components of D. Equation 6 might be taken to suggest that R(O) can exceed its theoretical maximum value of 0.4, but limits on the values of or and B preclude this possibilitym' Equation 6 is a general expression for a planar molecule, and additional restrictions on the orientations of the pumped and probed transition dipole moments and the anisotropy in D allow the prediction of the number of exponential decays in the experimental R(t) function. In principle, perylene can reorient either as an oblate rotor or as a prolate rotor. For an oblate rotor, Dz > DX=Dy and for a prolate rotor, D. > Dy=Dz. For a symmetric molecule such as perylene, the transition dipole moments will lie along the Cartesian axes defined for the rotational diffusion constants. The S. <— S.. transition accessed experimentally is polarized along the long (x) molecular axis. R(t) exhibits a double exponential decay if perylene reorients as an oblate rotor and to decay as a single exponential if perylene reorients as a prolate rotor. oblate: R(t) = 0.3-exp(-(2/),.+4/)_.)1) + 0.1.exp(- 613,1) [7] prolate: R(t) : 0.4-exp(— 61):!) A single exponential functionality was observed for R(t) providing support for, but not proof of, our assertion that the reorientation dynamics of perylene in the longer n-alkanes 110 is consistent with those of an effective prolate rotor. The practical limit on the ability to use the R(t) functionality to determine effective rotor shape lies in the finite signal to noise ratio of the data and the unknown relative values of D., Dy and D... The measured values for R(O) and R.(O) do not achieve the theoretical maximum of 0.4. Previous work on a large number of polar systems show that the theoretical maximum value for R(O) is difficult to obtain experimentally, with a typical maximum being ~ 0.33 for slowly reorienting molecules. The reasons for this experimental limitation are not understood fully, but polarization scrambling by the flow cell face(s) or the finite extinction ratio (~50: 1) of the pump and probe electric field polarizations could contribute to the observed behavior. For very fast reorientation, it is expected that the instrumental response function will serve to obscure the early time response and potentially reduce the regressed R(O) value. Previous reports on perylene reorientation also report an inability to achieve the theoretical maximum R(O) value."” In those reports, a double exponential decay of R(t) was reported. For their experimental conditions, where the excited transition was the S; (— So y-axis polarized transition and the So (— S. x-axis polarized transition was monitored, a double exponential decay indicates that perylene behaved as an effective oblate rotor. There is not discrepancy between earlier work on perylene indicating an effective oblate rotor shape and our work indicating an effective prolate rotor shape in longer n-alkanes because of the differences in the solvent systems examined. It is entirely likely that perylene exhibits an environment-dependent effective rotor shape. 111 6.5. Conclusion The rotational diffusion dynamics of perylene in a series of n-alkane solvents were presented. The reorientation time does not depend linearly on the solvent viscosity, but, rather exhibits two distinct linear regions. For shorter chain n-alkanes C5 through C3 the solvent-solute boundary condition lies close to the stick hydrodynamic limit, and for longer solvents C9, C..., C .2 and C... a Slip boundary condition applies. Further, the data for the longer chain solvents suggest that perylene behaves as an effective prolate rotor. The abrupt change in boundary condition occurs at n-octane. Solvents n-pentane, n- hexane and n-heptane are shorter than the perylene long molecular axis, while solvents longer than n-nonane are distinctly longer. n-Octane, the solvent at which the change in boundary condition is seen is approximately the same length as the perylene long axis; for this solvent-solute size ratio one of the basic assumptions of the DSE model is clearly violated. While there is obviously not enough information contained in these measurements to determine structural information about the solvent “cage” surrounding perylene, we notethat for solvents of length greater than C.., the solvent cage can, in principle, be comprised of molecules that span the entire solute length 112 66. Literature Cited Ix) b) 9. 10 . H. Labhart; E. R. Pantke; Chem. Phys. Lett., Q 482 (1973). . G. R. Fleming; J. M. Morris; G. W. Robinson; Chem. Phys, 17 91 (1976). . T. J. Chuang; K. B. Eisenthal; Chem. Phys. Lett., 11 368 (1976). . P. E. Zinsli; Chem. Phys, 29 299 (1977). . A. von Jena; H. E. Lessing; Chem. Phys, 40 245 (1979). . D. P. Millar; R. Shah; A. H. Zewail; Chem. Phys. Lett., _6_6 435 (1979). . U. K. A. Klein; H. P. Haar; Chem. Phys. Lett., Q 40 (1979). . A. von Jena; H. E. Lessing; Chem. Phys, £187 (1981). D. H. Waldeck; G. R. Fleming; J. Phys. Chem, Q 2614 (1981). . E. F. Gudgin Templeton; G. A. Kenney-Wallace; J. Phys. Chem, fl 2896 (1986). . R. L. Christensen; R. C. Drake; D. J. Philips; J. Phys. Chem, 99 5960 (1986). . J. Blanchard; M. J. Wirth; J. Phys. Chem, 9_0 2521 (1986). . Ben-Amotz; T. W. Scott; J. Chem. Phys, fl 3739 (1987). . J. Blanchard; J. Chem. Phys, 8_7 6802 (1987). . Ben-Amotz; J. M. Drake; J. Chem. Phys, _8_9 1019 (1988). . J. Blanchard; C. A. Cihal; J. Phys. Chem, 2 5950 (1988)”. . J. Blanchard; J. Phys. Chem, 22 6303 (1988). . J. Blanchard; J. Phys. Chem, 9_3 4315 (1989). C) C) C) O C C3 U C) . J. Blanchard; Anal. Chem, 61 2394 (1989). U . S. Alavi; R. S. Hartman; D. H. Waldeck; J. Phys. Chem, 9_5 6770 (1991). 113 21. R. S. Hartman; D. S. Alavi; D. H. Waldeck; J. Phys. Chem, 2; 7872 (1991). 22. M. Lee; A. J. Bain; P. J. McCarthy; C. H. Han; J. N. Haseltine; A. B. Smith;III; R. M. Hochstrasser; J. Chem. Phys, 85 4341 (1986). 10 L.) . G. J. Blanchard; J. Phys. Chem, 9; 5293 (1991). 24. R. S. Hartman; D. H. Waldeck; J. Phys. Chem, 3 1386 (1994). 25. D. Kivelson; K. G. Spears; J. Phys. Chem, _8_9 1999 (1985). 26. R. S. Moog; D. L. Bankert; M. Maroncelli; J. Phys Chem, fl 1496 (1993). 27. P. Debye; Polar Molecular: Chemical Catalog Co: New York, p84 (1929). 28. F. Perrin; .1. Phys. Radium, Z l (1936). 29. L. D. favro; Phys. Rev, 119 Q (1960). 30. T. tao; Biopolymers, 8 609 (1969). 31. G. J. Weber;./. Chem. Phys, _5_5 2399 (1971). 32. G. G. Belford; R. L. Belford; G. Weber; Proc. Natl. Acad. Sci, U. S. A., Q2 1392 (1972). 33. T. J. Chuang; K. B. Eisenthal; J. Chem. Phys, 51 5094 (1972). 34. C.-M. Hu; R. Zwanzig; J. Chem. Phys, @ 4354 (1974). 35. G. K. Youngren; A. Acrivos; J. Chem. Phys, 63 3846 (1975). 36. R. Zwanzig; A. K. Harrison. J. Chem. Phys, 83 5861 (1985). 37. G. Weber; .1. Phys. Chem, 93 6069 (1989). 38. D. J. Kivelson; J. Chem. Phys, g 709 (1991). 39. J. T. Edward;J. Chem. Ed, Q 261 (1970). 40. D. W. Piston; T. Bilash; E. Gratton; J. Phys. Chem, 93 3963 (1989). A}.-. t.;._ ._-‘-*AI-g-A 1‘ CHAPTER 7. VIBRATIONAL POPULATION AND ORIENTATIONAL RELAXATION DYNAMICS OF l-METHYLPERYLENE IN n-ALKAN ES - THE EFFECTIVE RANGE OF DIPOLAR ENERGY RELAXATION IN SOLUTION The vibrational population and orientational relaxation dynamics of l-methylperylene in the series of normal alkanes n-pentane through n-decane, n-dodecane and n-hexadecane were studied Both the vibrational population relaxation time constant. T., of the l- methylperylene 1370 cm'1 mode and the orientational relaxation time constant(s), 10., were . found to depend sensitively and nonlinearly on the aliphatic chain length of the n-alkane solvent. The data show that the two relaxations are sensitive to solvent local organization on approximately the same length scale, and stand in contrast to the relaxation dynamics of perylene in the same n-alkane solvents, where the operative length scale of T. relaxation was found to be substantially shorter than the length of the perylene molecule. These differences were understood in the context ofthe different polar v-v coupling processes utilized by perylene and l-methylperylene. The rotational diffusion data for l- methylperylene indicate that the dominant reorientation axis of the chromophore changes with solvent aliphatic chain length. 114 115 7.1. Introduction Earlier investigations in this thesis have focused on the probe molecule perylene in both polar and nonpolar solvents, and the ability to measure relaxation in both the ground and excited electronic states of the “probe molecule using this technique has been demonstrated.” The relaxation of the perylene Raman-active 1375 cm'1 mode in n-alkane solvents has been used to explore the role of resonance energy transfer in non-polar liquidsm Normal alkanes exhibit an infrared-active terminal methyl group rocking motion at 1378 cm". The perylene/u-alkane systems were chosen to evaluate whether or not local organization could be detected in nonpolar solutions. The data indicated that relaxation of the perylene 1375 cm'1 mode in n-octane is ~10 times more efficient than in n-hexane or n- decane. While these data indicated clearly that solvent local organization is important to vibrational energy relaxation, the T. measurementsm did not correlate with orientational relaxation measurementsm suggesting that the local “structure” in the n-octane surrounding perylene persists, at most, over several Angstrdms. These data are informative because they reveal the presence of local organization, but the length scale which T. measurements probe remains unclear. In this chapter the investigation of both the vibrational population (time constant T.) and orientational (time constant(s) To.) relaxation dynamics of l-methylperylene in several n-alkanes are presented. It is found that, in contrast to perylene, these two relaxation processes sense local solvent organization over similar length scales for l-methylperylene. We understand the differences between the perylene and l-methylperylene T. dynamical responses in terms of the symmetry ofthe chromophores. 116 7.2. Experimental Spectroscopy. The spectrometer used for stimulated measurements of l-methylperylene has been described in detail in Chapter 2.“) Stilbene 420 dye (Exciton) is used for both dye lasers. The pump laser is operated at 429 nm, corresponding to the 0-0 transition for l-methylperylene in n-pentane and n-hexane, at 430 nm in n-heptane and n—octane, at 431 nm in n-nonane. u-decane and n-dodecane, and at 432 nm in n-hexadecane. The l- methylperylene vibrational mode that was focused on occurs at ~1370 cm". The probe wavelength is 456 nm for n-pentane and n-hexane, 457 nm for n-heptane and n-octane, 458 nm for n-nonane through n-dodecane, and 459 nm for n-hexadecane. For the T. relaxation measurements the probe laser polarization was set to 547° with respect to the pump laser polarization to ensure the absence of orientational relaxation effects in the data. Rotational diffusion measurements were made separately and, for these measurements, individual scans were taken for the probe laser polarization parallel and perpendicular to the pump laser polarization. The form of the experimental signal we obtain from these measurements has been presented in detail previously. Steady-state spectroscopy. The steady-state absorption and emission spectra of all 1- methylperylene solutions were measured with 1 nm resolution using a Spex Fluorolog2 model F1 1 lAT spectrometer. These data were used to estimate the spectroscopic origin of l-methylperylene in each solvent. CIT. '-.‘A“-_.A__ l'_‘. _L 1 1 H 4‘ 117 Iz'ibrational spectroscopy. The infrared spectrum of solid l-methylperylene on KBr was recorded on an FTIR spectrometer at 4 cm'1 resolution (Mattson Galaxy 3000 ) using a DTGS detector. The spontaneous (resonance) Raman scattering speCIrum of solid 1- methylperylene in a capillary tube was obtained using 363.8 nm excitation from an Ar' laser (Coherent Innova 200). The Raman detection equipment consisted of a SPEX 1877 Triplemate spectrometer equipped witha 1800 groove/mm grating, operating at ~ 6 cm'1 resolution. The CCD detector was an EG&G Princeton Applied Research Model Spectrum One. Chemicals and sample handling. l-methylperylene was synthesized by alkylation of perylene with methyllithium. (see Scheme 7.1)‘51 This reaction was reported to methylate perylene at the 1- position with >95% selectivity. Perylene, methyllithium and 10% Pd on C catalyst were purchased from Aldrich and used as received. Following purification by plate chromatography, the identity of l-methylperylene was confirmed by mass spectrometry, 1H NMR, and UV-visible and infrared absorption measurements. All it- alkane solvents were purchased from Aldrich in their highest purity grade and used without further purification. The concentrations of the l-methylperylene solutions used for the T. relaxation and rotational difiusion measurements were ~10 11M. Sample solutions were flowed through a 1 mm path length quartz cell to minimize thermal lensing contributions to the signal, and the sample temperature was controlled at 300 : 0.1 K using a thermostatted bath. 118 ©© H3C\ / CH3 @ . + CH3L1 + /NCHz—CH2N oo ““3 H 80°C, 24 hrs CC H o + ether 0‘ ,. methyldihydroperylene (I) (I) Pd/C (5%) 4 @@© H3 ©© 1 -methylperylene Scheme 7. 1. Synthesis of l-methylperylene from perylene 119 7 .3. Results and Discussion Both the motional and vibrational energy relaxation dynamics of l-methylperylene in the alkanes n-pentane through n-decane, n-dodecane and n-hexadecane were measured, and the solvent-dependence of these two dynamical responses are found to be correlated. This finding is different from the case for perylene in the same n-alkanes, and we understand at a qualitative level, the basis for this difference. First the vibrational population relaxation dynamics of l-methylperylene and then its orientational relaxation dynamics in the n- alkanes will be discussed. The comparison of these bodies of data to one another and to 12.31 our earlier reports on perylene provides an understanding of the intermolecular interactions and local organization responsible for our observations. l'ihranona/ population relaxation. The method for measuring vibrational population relaxation time constants (T.) of fluorescent probe molecules in solution“'”’ has been discussed in Chpater 3. In Chapter 5 on perylene in the n-alkanes, the perylene v. (1375 cm") vibrational mode was chosen to interrogate because of its degeneracy with the terminal methyl group rocking mode of the n-alkanes. Because the dominant motion of the n-alkane vibration is confined to the CH3 groups, this “acceptor” vibrational mode was well suited to sensing local organization over short length scales in liquids. The data indicated that, for this mode, the solvent n-octane organized in such a way as to place its methyl groups in closer spatial proximity to the perylene vibrational coordinate than the other alkanes, and that other vibrational modes revealed preferential organization of different n-alkanes along their vibrational coordinates. While these results were important 120 because they demonstrated the existence of local organization within n-alkane solutions, they were also perplexing because of their lack of correspondence with reorientation dynamics measurements, where the molecular length scale of the measurement is comparatively well defined. The investigation of l-methylperylene in the same n-alkanes was undertaken because it is expected for the structural aspects of its interactions with the n-alkanes to be qualitatively similar to those for perylene but with spectroscopic selection rules that are relaxed significantly compared to those for perylene. Before presenting the T. data for l-methylperylene, it is important to examine the electronic and vibronic spectral response of this probe molecule. The absorption and emission spectra of l-methylperylene in n-hexadecane are shown in Figure 7.1. Not surprisingly. the linear optical response of I-methylperylene is very similar to that of perylene, except that individual features for l-methylperylene are blue-shifted by ~5 nm (~300 cm") compared to perylene in a given solvent, in agreement with previous reportsn‘s' This blue shift can be understood in terms of the strain imposed on the perylene ring structure by the presence of the methyl group at the 1- position. Semi- empirical calculations of l-methylperylene indicate a dihedral angle in the range of 13° to '8' The 0-0 transition energy of I-methylperylene in 23° between naphthalene moieties each solvent is obtained from the linear response and these data are listed in Table 7. 1. These data were used to determine the pump and probe laser wavelengths for the T. relaxation measurements. 121 , 432 nm 1.0 - A 0.8 - L": . 3 “m’ 0.6 - Absorption Emission g 1 Q. Q . a 0.4 l" / Q) .E ..1 0.2 P 0.0 . . A 1 - - 1 4 350 400 450 500 wavelength (nm) Figure 7.1. Steady-state absorption and emission spectra of l-methylperylene in n- hexadecane. The arrow indicates the 0-0 transition energy, and the box the range over which the stimulated response, shown in Figure 7.3, was taken. 122 Table 7.1. Spectral origin of l-methylperylene in the solvents used in this work, determined from the linear optical response. Solvent 0-0 transition energy ‘ (001") n—pentane 233 10 n-hexane ' 23310 n-heptane 23256 n-octane 23 256 n-nonane 23202 n-decane 23 202 n-dodecane 23 202 n-hexadecane 23 148 While the electronic spectra of l-methylperylene are similar to those of perylene, the infrared and Raman spectra of l-methylperylene and perylene differ significantly. The origin of the difference in the vibrational responses of these two structurally similar chromophores lies in the reduction of symmetry by the addition of a methyl group. Perylene is of D2. symmetry. a point group containing a center of inversion. Any Raman- active vibrational mode is infrared inactive, and vice versa for a molecule belonging to a point group containing an inversion center. The addition of the CH3. group to the perylene molecule eliminates the center of inversion, and all l-methylperylene normal modes are both Raman and infrared active. The infrared and Raman spectra of l-methylperylene were present in Figure 7.2. The 1370 cm'l mode is present in both spectra. The 1370 100— a 9? 95~ 51 90- '_"=_'_ U (L) h E 85’ Intensity (cps) Rarrnn shift (on!) Figure 7.2. (a) Infrared and (b) Raman spectra of I-methylperylene. The asterisks indicate the vibrational resonance for which we determined T. times. 124 cm'1 of l-methylperylene mode is believed to be derived from the perylene 1375 cm'1 mode based on the experimental energies of the resonances as well as semiempirical calculations. Figure 7.3 shows the time resolved stimulated emission spectrum of l-methylperylene in n-hexadecane, where vpump was fixed at v0-.. and vpmb. was stepped over the range corresponding to v... between ~117O cm'1 and ~1920 cm". At each pump-probe frequency pair, a time scan was recorded, and the frequency-domain response(s) were reconstructed by normalization of the time scans to the spontaneous emission spectrum at long delay timelm The spectra are time slices of the reconstructed surface taken at delay times of 10 ps, 30 ps , 50 ps and 100 to 700 ps delay, in 100 ps intervals. Multiple ground state vibrational resonances were detected within one broad steady-state spontaneous emission band (boxed feature in Figure 7.1), reminiscent of the earlier work of perylene in polar solvents'“ Because the vibrational response of l-methylperylene has not been assigned, and the Franck-Condon factors for the vibronic transitions accessed are significantly different than those for the spontaneous Raman response shown in Figure 7.2, To assign the individual resonances in the time-resolved stimulated spectra shown in Figure 7.3. was not attempted. For perylene, several combination and overtone resonances that are weak in the spontaneous Raman response are significant in the stimulated spectrum. The similar resonances are expected to play a role in the 1- methylperylene data as well. Foo} 125 1.0 —- —I—10 ps —.—30 _ —A—50 —v—100 ‘ 9/ ° —o—200 . V, g A —1:1— \ / lA/.-\ _O_ / -A- ATfl (a.u.) o..- .9375 g. Il>0 \ OI OID l //OI ”a? 06 l 4 J 1 l 1 l I 21200 21400 21600 21800 22000 frequency (cm") Figure 7.3. The time-resolved stimulated spectra of l-methylperylene in n-hexadecane, from 21978 cm" to 21231 cm". This range corresponds to vibrational frequencies between 1 170 cm’l and 1920 cm". 126 While the frequency-domain spectra contain a great deal of information, the quantitative T. information of interest is contained in the time domain responses. Presented in Figure 7.4 are the individual time domain scans for the l-methylperylene 1370 cm'1 mode in each n-alkane solvent where T. is observed as a build-up in intensity at early delay times. The ' mode are presented in Table 7.2. time constants T. for the 1-methylperylene 1370 cm' The uncertainties in these data are derived from at least five individual determinations for each solvent. A typical single determination is itself the average of 10 to 15 data ,1 acquisition cycles (time scans). q3 Table 7.2. T. relaxation times for the l-methylperylene 1370 cm'l mode in the u-alkane solvents. solvent T. i l 0' (ps) n-pentane 14 : 3 n-hexane 14 : 3 n-heptane 18 : 2 n-octane 18 : 1 n-nonane 28 : 8 n-decane 70 : 8 n-dodecane 77 : 8 n-hexadecane 105 : 7 127 NWT (a.u.) 1 A J A 100 200 300 400 500 delay time (ps) Figure 7.4. The stimulated response of l-methylperylene (1370 cm'1 mode) in eight alkanes. 128 Shown in Figure 7.5 is the dependence of the measured l-methylperylene T. times on n- alkane solvent identity. These data are significantly different than those reported for perylenem Clearly, the addition of a methyl group to the perylene chromophore alters the coupling between the solvent and the solute significantly. For short chain solvents, n- pentane through n-octane, T. is fast (~ 20 ps) and changes little as the solvent aliphatic chain length increases. For the solvents n~decane through n-hexadecane. T. is significantly slower, and slightly more solvent-dependent. A transition in T. relaxation behavior occurs between n-octane and n-decane. The T. times for l-methylperylene are, on aggregate. faster than for perylene, indicating more efficient coupling to the solvent environment. Perhaps more telling is that the dependence of T. on aliphatic chain length is very different for the two molecules. The magnitudes of the T. relaxation times measured are determined largely by the mechanism of the relaxation. Any local molecular organization within the solvent surrounding the chromophore is reflected in the modest solvent-dependent variations in T. we detect experimentally. In designing these experiments, the l-methylperylene 1370 cm'1 mode and n-alkane solvents were deliberately chose to use because of the degeneracy of this chromophore vibration with the solvent bath mode at 1378 cm’I (CH; rocking mode). Because of this degeneracy, polar (non-collisional) v-v coupling processes dominate vibrational energy transfer between the l-methylperylene and the solvent.“°' lntramolecular relaxation processes can also occur, but their contribution will be the same for all of the solvents we study here, and thus any solvent-dependence measured in T. will 120 100 80 E3:- 60 t: 40 20 129 E II 35 E A l I J 1 l 1 l l 4] 6 8 10 12 14 16 Number ofC 's in alkane Figure 7.5. Solvent chain length dependence of the l-methylperylene 1370 cm'1 T. relaxation time. arise from intermolecular relaxation processes. Aside from probability arguments relating to the magnitude of T. that was presented in Chapter 6 and the solvent-dependence of the . experimental T. data indicate the dominance of polar coupling over collisional relaxation processes. If collisional energy transfer processes dominate, T. will vary continuously with increasing solvent alkane chain length because of the direct relationship between solvent-solute collision rate and solvent‘viscosity. Also, the decrease in fractional density of solvent CH3 moieties with increasing solvent aliphatic chain length will contribute to a smooth dependence of T. on solvent chain length. This trend was not observed experimentally, implying the dominance of polar v-v coupling. The T. times for the l-methylperylene 1370 cm'1 mode increase with solvent aliphatic chain length, but there is an abrupt increase in T. starting with n-nonane. This solvent- dependence in T. implies the existence of local solvent organization around 1- methylperylene. 1n solvents shorter than the l-methylperylene normal mode coordinate probed, both solvent CH; groups are likely in close spatial proximity to the probe molecule, permitting efficient intermolecular energy transfer. For longer chain solvents, where the length of a solvent molecule is similar to or greater than the maximum dimension of the l-methylperylene vibrational coordinate, it is possible for the solvent terminal CH; groups to interact with the probe molecule, but the average distance between probe molecule and solvent terminal CH3. groups will be greater, on average than they are for the shorter solvents. This argument is recognized to be qualitative, and in room temperature liquids there is a broad distribution of n-alkane molecular conformations. this 131 interpretation is presented as a qualitative, empirical explanation consistent with the experimental data. It is expected that the interactions between l-methylperylene and the alkane solvents are similar to those between perylene and the alkanes, and that the differences measured experimentally arise primarily from spectroscopic rather than molecular geometry considerations. This point is discussed below. A recent paper by the Topp group on the rotational coherence response of perylene/alkane complexes in a supersonic jetl'” suggests that there are significant interactions between perylene and the n-alkane(s) along the chromOphore long axis. The details of the connection between the Topp group’s data and ours remains to be made, but structured perylene/n-alkane complexes have been observed at low temperature. The T. relaxation behavior of l-methylperylene in n-alkane solutions is different than that for perylene/alkane solutions. The primary reason for this difference is believed to be the removal of the chromophore center of inversion by the addition of the CH3. group. The consequent change in vibrational selection rules is important because it allows access to vibrational modes in l-methylperylene that are both infrared and Raman active. For perylene, one can access and detect Raman active, infrared inactive vibrational modes. Raman-active modes exhibit a change in polarizability on vibrational motion, while infrared active modes exhibit a change in dipole moment on excitation. The Raman-active perylene modes will exhibit modulations of their quadrupole moment (or higher multipole moments) on vibrational excitation where, for l-methylperylene, without a center of inversion, its vibrational modes exhibit a change in dipole moment on vibrational motion. 132 The dominant solvent bath mode is the infrared-active n-alkane terminal CH3. rocking mode. For perylene the dominant relaxation is therefore via quadrupole-dipole coupling (interaction energy 0: r'7)“2] while for l-methylperylene the most important relaxation process is through dipole-dipole coupling (interaction energy cc r6)” Because these two coupling processes operate over different length scales, we expecte the local environment sensed by perylene T. relaxation measurements to be significantly more confined than that for l-methylperylene. The difference in the T. data for perylene and l-methylperylene indicates that, for perylene, anharmonic coupling between vibrational modes does not contribute significantly to the measured T. response. If anharmonic coupling of the perylene Raman-active modes to its infrared active modes was significant, we would expect its solvent-dependent T. response to be similar to that of l- methylperylene, and we do not observe this trend experimentally. Rotational diffusion measurements. In the absence of a comprehensive theoretical treatment of T. relaxation processes in liquids, and without a means to calibrate the length scales over which these T. relaxations operate, it is necessary to compare these T. data to a different dynamical response where the length scale of the dynamics is better understood. Rotational diffusion is a technique used widely for understanding the complex interactions between probe molecules and solvent molecules. Because rotational diffusion measurements sense the motion of the entire electronic chromophore, it is difficult to determine the existence of site-specific molecular interactions between the solvent and solute except in cases where the molecules contain the appropriate polar or reactive functionalitiesm'm' What is significant for the purposes of this work is that the length scale of rotational diffusion measurements is comparatively well defined by the hydrodynamic volume of the probe moleculelm This “benchmark” can provide insight into the Operative length scale of polar v-v T. relaxation processes. In a rotational diffusion experiment, excitation of an ensemble of probe molecules by a polarized light pulse photoselects an anisotropic subset of the ensemble. This induced orientational anisotropy relaxes to a random distribution with characteristic time constant(s) and functionalities. The time course of the re-randomization contains information on the shape of the volume swept out by the rotating probe molecule (its rotor 119-211 shape)‘”” and on the solvent-solute boundary condition. The induced orientational anisotropy function is extracted from experimental data according to Equation 2 1..(t)—l.(t) R(/) = 1"(l)+ 211(1) where I..(t) and l,(t) are the signal intensity for pump and probe electric field polarizations parallel and perpendicular to each other, respectively. In general, R(t) can contain up to five exponential decaysml depending on the shape of the volume swept out by the reorienting molecule and the orientation of the pumped and probed transition moments with respect to the Cartesian diffusion constant axes. Under most circumstances. a single exponential decay of R(t) is observed. and thus there can be significant ambiguity in the interpretation of the experimental data. In cases where only limited information is available about the probe molecule transition moment orientation(s) or where a single exponential decay of R(t) is observed. the viscosity-dependence of the R(t) decay time constant can be measured to extract information on the frictional interaction between solvent and solutem The modified Debye-Stokes-Einstein (DSE) equation is used frequently to relate the decay time of R(t) to the viscosity of the solvent and the volume of the solute molecule,“9’2‘~23l T — —l- : I]: . I. [ 0" 6D H s where T.,, is orientational relaxation time, n is the solvent bulk viscosity, V is the hydrodynamic volume of the solute, (243 A3 for 1—methylperylene)“7] and D is the rotational diffusion constant. The terms f and S are correction factors to account for the solvent-solute boundary condition and the non-spherical shape of the reorienting species, respectively. Before discussing the information content of the l-methylperylene reorientation data. we want to make clear that we will focus our attention on the effective rotor shape of the reorienting species, S, and not on the solvent-solute boundary condition, f In Chapter 5 the focus was on the solvent-dependent change in the solvent- solute frictional interactions. because the data were of a form amenable to this treatment” The reorientation data presented here on l-methylperylene in the same n-alkanes are of a significantly different functionality. This difference necessarily arises from the presence of a single methyl group, producing an effect which is a combination of the transition dipole moment orientation of l-methylperylene and subtle differences in the way this probe molecule interacts with the solvents. A portion of this difference may also be the result of the torsional strain introduced to the aromatic ring system by the addition of the CH3 group."" The rotational diffusion constant can be decomposed into its Cartesian components and Chuang and Eisenthal have related the anisotropy decay determined 135 experimentally (Equation 2) to the relative directions of the pumped and probed transition dipoles and the Cartesian components of the rotational diffusion constant (D) for a general ellipsoidml For l-methylperylene the z axis is taken to be perpendicular to the molecular 1: system plane, with the transition dipole moment(s) along the x (long) axis of the electronic chromophore. There are two general ellipsoidal forms used to describe the volume swept out by a reorienting probe molecule. These are an oblate ellipsoid and a prolate ellipsoid. For an oblate ellipsoid, the fastest reorientation occurs along the axis perpendicular to the molecular plane (Dz > DS = D...) and‘for a prolate ellipsoid the dominant reorientation axis lies within the molecular plane, usually along the longest in- plane axis (D“ > D. = D.) For l-methylperylene, an experimental R(t) functionality that depends on the effective rotor shape is expected?“ oblate: R(t) = (%..)exp(—(2Dx + 4D: )t)+(%o) exp(—6Dxt) [4] prolate: R(t): (V...) exp(—6D_,t) [5] The reorientation dynamics of l-methylperylene in the n-alkanes are such that one can extract significant information on the effective rotor shape of the probe molecule. The rotational diffusion data were presented in Figures 7.6 and 7.7. The data in Figures 7.6a and 7.7a are the tail-matched 1(1) and 1(1) signals in n-pentane and n-hexadecane, respectively. The data in Figures 7.6b and 7.7b are the anisotropy decays, R(t), synthesized from the data in Figures 7.6a and 7.7a according to Equation 2. 136 1.0 — 1.10 0.5 ~ AT/T (a u.) 0.0 100 200 300 400 500 0.4 - R(t) delay time (ps) Figure 7.6. (a) Tail-matched parallel and perpendicular intensity data and (b) anisotropy decay, R(t), of l-methylperylene in n-pentane. I37 1.0 - / mm a :i t l l- < .- 500 52’ J 1 L 1 L 1 I 1 I 1 1 0 100 200 300 400 500 delay time (ps) Figure 7.7. (a) Tail-matched parallel and perpendicular intensity data and (b) anisotropy decay, R(t), of l-methylperylene in n-hexadecane. 138 In n-pentane through n-octane, R(t) exhibits single exponential decay functionality, whereas in n-nonane through n-hexadecane, a double exponential decay is required to provide adequate agreement with the experimental data. The R(0) values as well as the decay times, 10,, are presented in Table 7.3. The viscosity-dependence of these data is shown in Figure 7.8. Clearly there is a significant difference between the motional dynamics of 1- methylperylene in short (5 C3) and long chain (2C9) alkane solvents. Table 7.3. Experimental rotational diffusion time constants and zero-time anisotropies. All uncertainties are reported as standard deviations (: lo). solvent viscosity R .( 0) r0,(1) R2( 0) 5,0) (cPf ' (as) (123) n-pentane 0.24 0.22:0.02 12:1 -- -- n-hexane 0.33 0.25:0.02 13:2 -- -- n-heptane 0.41 0.24:0.04 15:1 -- -- n-octane 0.54 0.28:0.03 16:1 -- -- n-nonane 0.71 0.21:0.03 14:2 0.07:0.04 49:14 n-decane 0.92 0.24:0.03 17:3 0.04:0.03 70:26 n-dodecane 1.35 0.28:0.01 21:2 0.05:0.01 126:29 n-hexadecane 3.34 0.21:0 04 25:3 0.09:0.02 152:23 a. Data from CRC Handbook of Chemistry and Physics, 7lst ed.; Lide, D. R., CRC Press; Boca Raton, FL, 1990 1 I39 175 — _— 130 r- —P T 12 0 '1’ D )— C16 .00 .. C1o __ g . C12 ‘6 75 — . H C9 T 50 - l. - C. 25 — L is II E 0 1 J J I L l 1 I 1 1 00 0.5 10 l.' 20 2.5 30 35 0 (GP) Figure 7.8. Orientational relaxation time, 10., as a function of solvent viscosity. For the long chain alkanes (n-nonane to n-hexadecane), two exponential decays are found in R(t). E" 140 While there are several ways in which this solvent-dependent change in the functionality of R(t) could be interpreted, these data were considered in the context of a solvent- dependent change in the effective rotor shape of I-methylperylene. It is not possible to extract information on the solvent dependence of the fictional boundary condition from these data because of the different information content of the two forms of R(t). As the solvent chain length increases the effective rotor shape changes from prolate to oblate. Using the preexponential factors from the data in Table 7.3 and Equation 4, the 1- methylperylene rotor shape for the longer chain solvents can be estimated. For short chain solvents, where l-methylperylene behaves as a prolate rotor, only D. from the experimental data can be extracted, and thus there is little or no “shape” information available. For the longer chain solvents, however, where there are two exponential decays, it is straightforward to extract Dx and D2, and thus the major-to-minor axial ratio 1 of the ellipsoid of rotation (Table 7.4). While the detailed information can not. provided on the effective (ellipsoidal) rotor shape of l-methylperylene in n-pentane through it- octane, it can be determined from the ratio (Dz ID.) that, for the longer chain solvents the anisotropy of the rotational ellipsoid increases with solvent length to a limiting value in n- dodecane. we interpret these results as representing a quasi-lamellar confinement of 1- methylperylene in the long-chain solvents. 141 Table 7.4. Cartesian components of the rotational diffusion constant extracred from experimental data using Equations 4 and 5. solvent B. D,- D/Dx (GHz) (GI-1:) n-pentane --. 13.9 : 1.2 (< l) n-hexane -- 12.8 : 2.3 (< 1) n-heptane -- 11.1 : 0.8 (< 1) n-octane -- 10.4 : 0.7 (< 1) n-nonane 3.40 : 0.75 16.2 : 1.9 4.8 : 1.3 n-decane 2.38 : 0.64 13.5 :1.9 5.7 : 1.8 n-dodecane 1.32 : 0.24 11.2 : 0.9 8.5 : 1.9 n-hexadecane 1.10 : 0.15 9.5 : 1.2 8.6 : 1.9 The solvent-dependent change of rotational diffusion rotor shape and the onset of the change in T. both occur between n-octane and n-nonane. The interpretation of the reorientation data in terms of solvent local organization are consistent with the solvent local organization implied by the T. solvent chain length dependence. In short n-alkanes, the l-methylperylene rotor shape is prolate, i.e. the dominantirotational motion is about the l-methylperylene long in-plane axis. The solvent molecules are small enough that both terminal CH; groups are in close spatial proximity to the probe molecule, and fast T. relaxation is expected. In the longer n-alkane solvents, I-methylperylene behaves as an oblate rotor, where the dominant rotational m0tion is around the axis perpendicular to the probe molecule molecular plane. We believe that this solvent-dependent change of rotor shape for l-methylperylene is due to confinement of the probe molecule between solvent .fl‘f‘:_;[, .1. g . . I42 “lamellae”. Alternatively, this confinement can be expressed in the context of the individual solvent molecules being long enough to span the l-methylperylene long axis, thereby significantly reducing the structural freedom of individual solvent molecules on the probe molecule length scale. For such an environment, where the solvent terminal CH3 groups are, on average further away from the probe molecule than in the shorter n- alkanes, one expects a longer T. relaxation time, consistent with the experimental findings. The change in both the reorientation and T. data between n-octane and n-decane suggests that the effective “length” of the l-methylperylene vibrational coordinate probed is in the same range as the average length of the ensemble of these solvent molecules. These data also indicate that dipolar v-v coupling processes responsible for T. relaxation occur over a ~10 A length scale. Recent work by the Topp group on the rotational coherence spectroscopy of jet cooled perylene/n-alkane complexes suggests intermolecular [111 In the isolated interactions at least qualitatively in correspondence with these data. perylene/n-alkane complex, the alkane chain lies parallel to the perylene long axis with a 3.6 A separation between the molecules, with the n-alkane located over the center of mass of the perylene molecule. As the n-alkane chain length is increased (from n-octane), a displacement from the parallel structure occurs. It is interesting that these data, for room temperature solutions, and the low temperature perylene/n-alkane complex point to similar intermolecular organization. “'1? 1‘5... 1___~—.__L_h'__._x 143 6.4. Conclusions The vibrational population and orientational relaxation responses of l-methylperylene in a series of n-alkane solvents were measured using ultrafast stimulated laser spectroscopy. The T. response for the l-methylperylene 1370 cm'1 mode is dominated by non-collisional dipolar v-v coupling to the alkane solvent 1378 cm '1 CH3 rocking mode. Both the T. and R(t) data for l-methylperylene in the n-alkanes differ significantly from the data of the earlier work for perylene in the same solvents. The difference in the T. response is understood for the two probe molecules. For perylene, the vibrational mode interrogated is infrared inactive and therefore the dominant polar exchange process between solute and solvent must be quadrupole-dipole coupling. For l-methylperylene the vibrational resonance we access is both infrared and Raman active and thus the dominant exchange mechanism is dipole-dipole coupling. These two coupling processes operate over different length scales and it is therefore expected closer correlation between T. and R(t) data for l-methylperylene than for perylene. For perylene in the n-alkanes, the R(t) decays presented in Chapter 4 are single exponential in all cases, while for l-methylperylene, a double exponential decay for longer alkane solvents was observed, This observation'is not signal-limited. but rather represents a fundamental difference in the way the two molecules reorient in the same solvent. More work is needed to understand the large differences in dynamics that arise from the addition of a single methyl group to the chromophore, but one possible basis for this difference is the torsional strain induced in the l-methylperylene ring structure by the presence of the CH3 group. For the l-methylperylene data, the correlation between the R(t) and T. dynamical responses can be understood in terms of 144 local solvent organization about the chromophore. Both of the dynamical responses point to the close spatial proximity of the solvent terminal CH3 groups to the chromophore in short alkanes and a greater average distance between these moieties in longer alkane solvents. Both sets of data point to a change in solvent-solute interaction between it- octane and n-decane. These data appear to be in excellent qualitative agreement with recent low temperature examinations of perylene/n-alkane complexes in a jet expansion using rotational coherence spectroscopy. 145 6. 5. Literature Cited 1. Y. Jiang; G. J. Blanchard; J. Phys. Chem, 98 9417 (1994). Id . Y. Jiang; G. J. Blanchard; J. Phys. Chem, 3 9411 (1994). Lo) . Y. Jiang; G. J. Blanchard; J. Phys. Chem, 28 6436 (1994). . Y. Jiang; S. A. Hambir; G. J. Blanchard; Opt. Commun, 9_9_ 216 (1993). J}. 5. H. E. Zieger; E. M. Laski; Tet. Lett., 32. 3801 (1966). 6. S. A. Hambir; Y. Jiang; G. J. Blanchard; J. Chem. Phys, 98 6075 (1993). 7. L. Lewitzka; H.-G. Lohmannsroben; M. Strauch; W. Luttke; J. Photochem. Photobiol. A: Chem, 61 191 (1991). 8. S. Grimme; H.-G. Lohmannsroben; J. Phys. Chem, 96 7005 (1992). 9. Y. Jiang; S. A. Hambir; G. J. Blanchard; Chem. Phys, 183 249 (1994). 10. J. T. Yardley, Introduction to Molecular Energy Transfer, Academic, New York, 1980. . l l. T. Troxler; J. R. Stratton; P. G. Smith; M. R. Topp; J. Chem. Phys, l___l 9219 (1994). 12. C. G. Gray; K. E. Gubbins; Theory of Molecular Fluids, Vol. I : Fundamentals Oxford Science, pp. 91-100, 1984. ' 13. G. J Blanchard; C. A Cihal; J. Phys. Chem, _9_2_ 5950 (1988). 14 G. J. Blanchard; J. Phys. Chem, 22 6303 (1988). 15. G. J. Blanchard; J. Phys. Chem, 93 4315 (1989). 16. G. J. Blanchard; Anal. Chem, _6_1 2394 (I989). 17. J. T. Edward; J. Chem. lid, 41 261 (1970). 146 1.8. F. Perrin; J. Phys. Radium, _7_ 1 (1936). 19. C. M. Hu.; R. Zwanzig; J. Chem. Phys, 60 4354 (I974). 20 G. K. Youngren; A. Acrivos; J. Chem. Phys, Q 3846 (1975). 21. R. Zwanzig; A. K. Harrison; J. Chem. Phys, 8_3 5861 (1985). 22. T. J. Chuang; K. B. Eisenthal; J. Chem. Phys, 57 5094 (1972). 23. P. Debye; Polar Solvents, Chemical Catalog Co., New York, p. 84, 1929. CHAPTER 8. SUMMARY AND FUTURE WORK This thesis work focuses on intermolecular interactions between solutes and solvents using ultrafast laser spectroscopic method. A We have developed a novel pump-probe laser spectroscopy scheme to study vibrational energy transfer between solute molecules and surrounding solvent molecules. We are able to pump two blue picosecond dye lasers synchronously with the third harmonic output of a Nd:YAG laser. Other vibrational relaxation measurement schemes involve the use of more expensive and technically more complicated picosecond infrared lasers. Our spectrometer, combined with our measurement scheme, gives us time resolution of a few picoseconds and spectral resolution of ~ 4 cm". Picosecond time resolution allows us to look at solution phase dynamics such as vibrational relaxation and rotational diffusion. High spectral resolution enables us to access the vibrational states of interest selectively. From our studies on perylene and l-methylperylene we find that T. is vibrational mode and solvent sensitive as well as state dependent for a given probe molecule. We are able to utilize the information from rotational diffusion measurements to aid the interpretation of vibrational relaxation data. In general, vibrational relaxation can occur through intramolecular and intermolecular energy transfer. For intramolecular energy relaxation, excess vibrational energy in high 147 148 frequency vibrational resonances transfers to low energy vibrational states within the molecule because of anharmonic coupling between vibrations, typical of organic chromophores. Intermolecular energy transfer includes vibrational energy exchange between the solute molecules themselves and their surroundings. In solution, due to the high density of the medium the energy exchange between solute and surrounding solvent becomes dominant while intramolecular? vibrational relaxation is less important or at least constant for a given mode. When there are solvent vibrational resonances at frequencies close to solute modes, v-v long range resonance coupling allows the exchange to be very efficient. It is difficult to separate the different channels for vibrational energy transfer in solution. For studies of intramolecular energy transfer, the ideal system has a low frequency of the collisions between solute molecule and the surrounding molecules and no vibrational mode of the surroundings is close to energetic proximity of the solute modes. For the systems we chose in this thesis work, we have focused on intermolecular vibrational energy transfer between solute and solvent through v-v long range resonance couphng. For v-v long range resonance coupling vibrational energy transfer, the rate of transfer is a function of the frequency difference between a solute mode (donor) and solvent mode (acceptor) (A00), the distance between donor and acceptor (d) and the nature of the interaction (u). T.xf(Ac0,d,u) 11] Ground state T. measurements of perylene in n-alkanes revealed the existence of local solvent organization. For perylene in n-alkanes, A0), is ~ 0 for our conditions, and u is 149 related to d'7. For the perylene 1375 cm'1 vibrational mode (v..) in alkane solvents, the CH3 end group rocking mode at 1378 cm" is the energy acceptor for v-v long range resonance coupling. The variation of the perylene v.5 T. times in different alkanes reflects different average distances between perylene v.5 vibrational coordinate and CH; groups. These differences are caused by variation of the local solvent organization around the solute. The dependence of T. on aliphatic chain length is not a smooth trend. There is a transition occurring around octane between short and long chain solvents and the chain length of all trans octane is close to perylene long axis length. In the future, T. times of perylene in branched alkanes can be measured and the results can be compared to those in normal alkanes. This information will provide insight into local solvent organization by deliberate disruption of the solvent environment. Among the three parameters, A0) and u, the latter two are relatively easy to determine, but d is not possible to control in solution. one possible way to control, or vary (1 is to use supercritical fluids instead of liquids. A supercritical fluid is a special phase between the gas and liquid phases, where the intermolecular separation (d) can be controlled by varying temperature and pressure. The term u in Equation 1 is determined by the properties of both solute and solvent. We observed a close correlation between energy transfer and rotational diffusion dynamics. for l-methylperylene in n-alkanes. this correlation is not seen for perylene in same solvents. We attribute this difference to variations in the nature of solute-solvent interaction for energy transfer. For perylene, v-v energy exchange occurs between a solute Raman ag mode and solvent IR mode. The intermolecular coupling is by quadropole-dipole interactions, where the interaction energy is proportional to r'7. The solvent local organization, reflected by the T. solvent chain length dependence persists on a shorter length scale than the probe molecule itself so that no correlation between T. and To. dynamics is found. For l-methylperylene, v-v long range resonance coupling occurs between a solute IR mode and solvent IR mode. The coupling in this case is dipole-dipole and the interaction energy is proportional to r‘. This length scale is similar to the size of the probe molecule itself and thus a correlation between two dynamics is observed. To continue exploring the dependence of T. on the interaction energy surface, experiments of perylene in benzene and toluene, and l-methylperylene in benzene and toluene, need to be performed. From these experiments the coupling between solute Raman mode with solvent Raman or IR mode and solute IR mode with solvent Raman or IR mode can be compared. As we learn more and more about T. relaxation of organic chromophores in solutions experimentally, there is an urgent need for developing a theory of v-v long range vibrational energy transfer processes in solution. Future collaborations will be necessary to relate the experimental data to a sound theoretical interpretation. The technique that we have developed and the knowledge we have gained about intermolecular interactions can be applied to more complicated systems such as proteins, where ultrafast relaxation processes have been experimentally observed, but where the role of vibrational relaxation remains only poorly understood.