EXCITED-STATE PHENOMENA ASSOCIATED WITH SOLVATION SITE HETEROGENEITY By Khader Ahmad Al—Hassan A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1981 To My Family ii ABSTRACT EXCITED—STATE PHENOMENA ASSOCIATED WITH SOLVATION SITE HETEROGENEITY By Khader Ahmad Al—Hassan Organic molecules that undergo a large change in dipole moment upon excitation may exhibit large shifts in absorption or emission spectra. Interesting excited- state phenomena involving solvent—solute interactions and solvent-cage relaxation may occur when such polar molecules are excited in a polar matrix. These phenomena include: temperature—dependent inhomogeneous spectral broadening, solvent-assisted electronic energy transfer, time—dependent spectral shift, lack of fluorescence depolarization, and excitation-wavelength dependent red shift of fluorescence and phosphorescence in rigid media (Red—Edge Effect). These phenomena can be explained in terms of the statisti- cal interaction between the polar solute molecules and their immediate polar environment. Thus, even in a homo- geneous condensed phase the polar solute molecules are Khader Ahmad Al-Hassan expected to occupy a variety of solvation sites at any given time giving rise to different absorption energies correspond- ing to the same electronic transitions. Such variation in solute-solvent local interactions introduces a significant source of broadening of both the absorption and emission spectra. We have examined the role of microenvironmental hetero- geneity in electronic spectra by studying the effects of medium polarity, rigidity, temperature, concentration and excitation-wavelength dependence of the spectra of pyridine merocyanine dye, 2-amino-7-nitrofluorene (ANF), A,A'- aminonitrodiphenyl (AND) and methyl- and tertbutyl-esters of 9-anthroic acid (9MA and 9TBA respectively). Pyridine merocyanine dye represents a class of organic molecules that undergo a decrease in dipole moment upon electronic excitation while ANF and AND represent a class of organic molecules that undergo a substantial increase in di— pole moments upon electronic excitation. 9MA and 9TBA represent a class of organic molecules that are flexible and undergo geometrical changes upon electronic excitations. The absence of excitation wavelength dependence of the fluorescence at room temperature in fluid polar media is explained in terms of orientational and translational relaxa- tions of various "solute-solvent conformations" that occur prior to fluorescence. The fluctuations in the interaction of solutes with different solvation sitesare~dynamic in fluid Khader Ahmad Al-Hassan media. Once the solution is made rigid (glass, polymer matrix, etc.) the dynamic character is lost. One may think of a viscosity—dependent barrier between these sites, the height of which depends on the rigidity of the medium. Under these circumstances, the lifetimes of various "solute- solvent conformations" are larger than the excited-state lifetime, and hence excitation wavelength dependent fluores- cence will occur (Red—Edge Effect). The presence of BBB in flexible 9MA and 9TBA in nonpolar matrix (3MP) is explained in terms of different conformers that have slightly different absorption energies. ACKNOWLEDGMENTS I would like to express my deep appreciation to my advisor, Dr. M. Ashraf El-Bayoumi, for his guidance, en- couragement and friendship during the course of this study. I would like to express my gratitude to the members of my Committee, Dr. Kathy Hunt, Dr. James Harrison, Dr. William ReuschaniDr. Andrew Timnick. Many thanks go to Mr. Ron Hass for his expert assistance in electronics and to Ms. Bev Adams for patiently drawing all the figures in the dissertation until they possessed unmatched perfection. To my laboratory colleagues, especially to Kamal Ismail and Nahid Shabestary, I thank them for their continuous support and friendship. Special thanks are extended to Yarmouk University in Irbid-Jordan for their financial and moral support during the course of this study. iii Chapter LIST OF LIST OF CHAPTER CHAPTER II. III. IV. TABLE OF CONTENTS TABLES. FIGURES I - INTRODUCTION. II - SOLVENT EFFECT ON ELECTRONIC SPECTRA. Solvent Effect on Absorption Spectra . . . . . . Effect of Hydrogen Bonding Solvents. . Theories of Solvent Spectral Shift . . . Solvent Shifts as an Aid in Characterizing Electronic States. . . . . . . . l. n + n* vs n + n* Transitions. 2. Locally—Excited States vs. Charge-Transfer States. 3. Electron—Transfer Transitions A. Singlet-Triplet Transition. Solvent Effect on Emission Spectra . . . . . . 1. Spectral Shifts 2. Viscous-flow Barriers 3. Solute—Solvent Relaxation in the Nano and Pico—second Range. A. Excited State Level Inversion iv Page vii viii 1U 17 29 29 31 3A 36 38 38 Al A5 52 Chapter CHAPTER CHAPTER I. II. III. CHAPTER I. II. III - THE RED EDGE EFFECT AND A RELATED PHENOMENON. Shpol'skii Effect IV - MOLECULAR SYSTEMS. Molecular Systems that Undergo a Large Decrease in Dipole Moment as a Result of Excitation 1. Merocyanine Dyes. 2. Alkyl Pyridinium Iodides. Molecular Systems that Undergo a Large Increase in Dipole Moment as a Result of Excitation. l. 2-Amino-7-Nitro Fluorene. 2. A,A' Amino Nitro Diphenyl Flexible Molecules that May Undergo a Change in Dipole' Geometric Configuration Upon Excitation. . . . . . . l. 9-Tertbutylanthroate: (tertbutyl- ester of 9-anthroic acid) . 2. 9- -Methylanthroate (methylester of 9- anthroic acid) . . V - RESULTS AND DISCUSSION. Molecular Systems that Undergo a Large Decrease in Dipole Moment as a Result of Excitation. A. Pyridine-Merocyanine Dye. B. Benzothiazole Merocyanine Molecular Systems That Undergo a Large Increase in Dipole Moment as a Result of Electronic Excitation. Page 57 67 73 7A 7A 75 79 79 8A 8A 8A 88 88 88 116 119 Chapter A. 2—Amino-7-nitrof1uorene (ANF) B. A,A'-Amino Nitro Diphenyl (AND) III. Flexible Molecules that May Undergo a Change in Dipole Geometric Configuration Upon Excitation. . CHAPTER VI - EXPERIMENTAL A. Materials . . . . . . . . . . I. Purification of Solvents. II. Preparation and Purification of Chemical Compounds III. Preparation of Polymer Films. B. Spectral Measurements . . . . 1. Absorption Spectra. 2. Emission Spectra. 3. Temperature Variation System. CHAPTER VII — CONCLUSION AND FUTURE WORK. I. Pyridine Merocyanine. II. ANF and AND . . III. 9MA and 9TBA. . . . . . . . . . . . . . . REFERENCES. . . . . . . . . . . . . . vi Page 119 133 1A0 151 151 151 152 15A 155 155 155 156 157 158 159 160 161 Table II LIST OF TABLES Page Optical Properties of Pyridine Merocyanine Dye in Various Solvents. . . . . . . . . . . . . . . . 77 Fluorescence Properties of ANF (2—Amino-7-Nitrofluorene) in Various Solvents . . . . . . . . . . 83 vii Figure 3a 3b 5a LIST OF FIGURES Solvent shifts due to hydrogen bonding. A) Hydrogen bonding is stronger in the ground state. B) Hydrogen bonding is stronger in the excited state. . . . . Room temperature absorption spectra Of p-nitroaniline in different sol— vents . . Room temperature absorption spectra of halogen ions in aqueous solution (D20) . . . . . . Room temperature absorption spectra of sodium iodide in acetonitrile, water and ethanol . . . Mechanical viscous-flow-barrier cage proposed by Dellinger and Kasha . . . . . . . . . Steady-state fluorescence spectra for ANS in n—propyl alcohol at the indicated temperatures. viii Page 15 33 35 35 U2 A8 Figure 5b 5c 7a 7b 8a Page Time-dependent fluorescence spectra for ANS in n-propyl alcohol at -90°C. A, O nsec; B, 2.5 nsec; C, 12.5 nsec; D, 21.5 nsec; E, 31.5 nsec. . . . . . . . U8 Time-dependent fluorescence spectra for ANS in n-propyl alcohol at -150°C. A, 2 nsec; B, 11 nsec; C, 68 nsec . . . . A8 Tor versus solution viscosity for rhodamine 6G in various solvents. . . . . 51 Effect of temperature on the fluorescence spectra of 5 x 10-5 M DEAB (p-diethylaminobenzonitri1e) in butylchloride-methylcyclohexane- is0pentane mixture (12:3:1 in volume). 293°K, ------ 23A0K, ————— 173°K, ----~-- 1A8°K . . . . . . . . . . . . . . 55 Excited state level inversion caused by mutual interaction between solute in excited state and polar solvent. . . . . 55 Diagram showing orientation of parallel and perpendicularly polarized emission with respect to plane of polarization of exciting radiation in the laboratory coordinate system . . . 59 ix Figure 8b 9a 9b lOa Excitation polarization spectra of indole in propylene glycol at -70°C. .I, 0.01m; A, 0.05 M; X, 0.1 911; 0.0.2 1’1; 0, 0-14 M. Fluorescence and 0-0 region phosphorescence spectra of 10-3 M indole in 1:1 ethylene glycol-water media at 202 and 77 K, respectively, demonstrating the shift between the emission excited at 280 and 295 nm. Plot of exciting wavelength de- pendence for the 10-3 M fluores- cence and phosphorescence spectra versus the temperature of the A:l glycerol-water medium. Fluores- cence and phosphorescence shifts were measured as the average of the red edge and blue edge dif- ferences between the spectra at 280 and 295 nm excitation Fluorescence spectra of mlO-l5 M TP (p—terphenyl) in ethanol at 77 K at different excitation wavelengths. 280 nm; ------ 310 nm; Page 59 62 62 6A Figure 10b 11a 11b 12 Page Fluorescence spectra of «.10.5 M QP (quaterphenyl) in ethanol at 77 K at different excitation wave- lengths. —————— 320 nm; ------ 330 nm; °°°°°° 335 nm . . . . . . . . . . . . . . 6A Fluorescence spectrum of coronene in a frozen n—heptane matrix. The sharp lines arise because of the unique conformation of the coronene in the matrix . . . . . . . . . . . . . . 68 Variation in the appearance of the A03 nm band system in the fluores- cence spectrum of benzOEaprrene in heptane matrices at 15 K under dye laser excitation as a function of excitation wavelength: (a) 385.5 nm, (b) 385.7 nm, (c) 385.9 nm, (d) 386.1 nm. . . . . . . . . . . . . . . . . . . . 68 The limiting cases of homogeneous and inhomogeneous broadening. In the homogeneous case (a) the spectrum summed over all molecules in the sample corresponds identically to that for any one of the individual molecules. In the inhomogeneous case, (b), the xi Figure 13 1A l5 16 17 Page spectrum summed over all molecules in the sample differs from the spectrum that would be seen for any single molecule. . . . . . . . . . . . . . . . . 71 Room temperature absorption spectra of pyridine merocyanine dye in dif- ferent solvents. DMSO = Dimethyl sulfoxide and EtOH = Ethanol. . . . . . . 76 Room temperature fluorescence spectra of pyridine merocyanine dye in different polar solvents . . . . . 78 Room temperature absorption spectra of ANF in different solvents: (1) 3MP, (2) Paraffin Oi1,(3) Polystyrene film, (A) Ethanol, (5) Polyvinyl alcohol film. . . . . . . . . . . . . . . . . . . 81 Room temperature absorption spectra of AND in different solvents: (l) 3MP, (2) Paraffin Oi1,(3) Polystyrene film, (A) Ethanol, (5) Polyvinyl alcohol film. . . . . . . . . . . . . . . . . . . 82 Room temperature absorption spectrum of (7 x 10'5 M) 9TBA and its fluorescence spectra at room tem- perature ( oooooo ) and at 77 K in 3MP . . . . . . . . . . . . . . . . . . . 86 xii Figure l8 19 20a 20b 21 22 Page Room temperature absorption spectrum of (7 x 10—5 M) 9MA and its fluorescence Spectra at room temperature ( ------ ) and at 77 K ( ------ ) in 3MP . . . . . . . . . . . . . 87 Room temperature absorption spectra of aqueous solutions of pyridine merocyanine dye (2 x 10—5 M) at different pH. . . . . . . . . . . . . . . 89 Resonance structure of pyridine merocyanine dye in its neutral and acidic form . . . . . . . . . . . . . . . 9O Photochemical cycle of pyridine merocyanine dye and its cis—trans isomerization . . . . . . . . . . . . . . 9O Fluorescence spectra of a dilute solution ( We: i.e., the hydrogen bODGijsstronger in the ground state than in the excited state as shown in Figure 1A. A blue shift in an absorption spectrum which exceeds we by a value Of Wg-We will result. In emission, the shift is less than W -We by w so that a blue or red shift may g g’ be observed depending on Wg-we and w However, the Shift g' is small compared to wg. If We > W i.e., the hydrogen bond is stronger in the g) excited state than in the ground state as shown in Figure 1B, a red shift in the absorption spectrum which is less than Wg-We by we is expected. In emission, the red shift exceeds Wg—We by wg. Well characterized hydrogen bonds have energies in the range of 1-7 Kcal/mole (350-2500 cm’l). According to 17 the above discussion, a blue shift in absorption may exceed the ground state hydrogen bonding energy, hence, the expected blue shift occurs in the range of 350-2500 cm-1 or larger than 2500 cm-1. But a red shift in absorption , should not be larger than should never exceed Wg, i.e. 2500 cm-1. For W* + n transition in hydrogen bonding media, a blue shift in absorption is usually observed.6’7 This is due to the decrease in charge density (M1 the lone pair atom as a result of lone pair promotion. Thus, the hydrogen bond is always stronger in the ground state. A red shift in absorption Spectrum indicates that hy- drogen bonding is stronger in the excited state. This may indicate an increase in the acidity or basicity depend- ing on the functional group at the chromophore involved in hydrogen bonding. III. Theories of Solvent Spectral Shift: The general theory which relates Spectral shifts to various interactions between the solute and the solvent molecule is incomplete. These Spectral shifts are at- tributed to (a) physical interactions between the solute and solvent molecules and (b) to some important specific effects like: hydrogen bond formation; proton or charge transfer between solvent and solute; and solvent- dependent aggregation; ionization or dissociation and iso- merization equilibria. 18 Ooshika8 presented a theory of solvent Shifts for polar and non-polar solvents. McRae9 published a similar 10 theory and applied it to dye molecules He used the 11 in his theory of the dielec- Same model used by Onsager tric constant of polar liquids. By Onsager's theory one cal» culates the polarization due to one molecule by representing it as a polarizable point dipole in a cavity electric field. All other molecules give rise to a homogeneous dielectric with a dielectric constant equal to the bulk liquid. Mc- Rae divided the polarization of the dielectric into two parts: one due to orientation and the other due to elec- tronic polarization of the solvent molecules. The inter- action between the polar solute and dielectric was cal- culated to second order in perturbation theory. Liptayl2 presented a theory of solvent sensitivity and discussed it qualitatively in terms of hydrogen bonding properties of the solvent. Kirkwood-Frohlich theory13'15 represents a modification of Onsager's theory that takes into account short-range order in the solvent. 16’ Ahen Several other theories by Marcus and Weigang and Wild18 have also been presented. The starting point of most theories of the general sol- vent spectral shift is the presumably known set of wave functions for the unperturbed electronic states of iso- lated solute and solvent molecules. Second-order per- turbation theory was used to develop expressions for the 19 change of the stationary electronic energy levels and of energy differences or transition energies of a single solute molecule which arise from electric fields due to a particular (but unspecified) configuration of an individual solvent molecule. The resulting Shifts were then averaged over all appropriate configurations of the solvent mole— cules of the solution. According to the scheme derived by Nicollg, the approp- riate solute—solvent pair distribution function was ex- pressed in terms of the wave functions 1K,i> = |K>Ii> for the K—th electronic state of an isolated solvent molecule and the unperturbed i-th electronic state of an isolated solute—chromophore. (Solvent index will always precede the solute index and the 0 designates the ground state of either molecule.) The Hamiltonian H = H + H' for this solvent-solute 0 pair was defined in terms of HO the sum of the separate, a unperturbed Hamiltonians for the isolated solvent and solute molecules, and H', the interaction operator. |K>, |i> and Ho were presumed to be known from vapor phase spectra H' was defined in terms of an electric field Em that the solvent medium produces at the site of the solute molecule. EKi was taken to be the sum of the unperturbed energies of the solvent molecule in state K and the solute in state i while EKi representstflmacorresponding perturbed energy including terms to second order in the perturbation. 20 If one deals with one particular transition of the solute from state 0 to state 1 then E(00-01) is the un— perturbed solute transition energy and equal to EOl — EOO while 3(00-01) is the corresponding energy for the per- turbed (solvated) solute and equal to €00 - 801, The Shift obtained for this particular configuration was: Au(cm-l) (hc)-1[AE(00-01)1 (hc)-l[€(OO-01) - E(OO-Ol)] (3) This has been averaged over the configurations of the solvent molecules to obtain expressions for the general solvent spectral Shifts. The perturbed energies of the solute chromophore- solvent pair were written in terms of quantities just defined: ll2 500 = E00 + <00IH'I00> + z KfiO EOO - EKO 2 . 2 + z 11 + z II (u) l#0 E00 - E01 13K#O E00 - EKI 21 2 | 01 H' K1 | 801 = E01 + <01|H'|01> + Z < I l > K50 E01 ’ EKl I . 2 2 + £1 [<211H j01>1, + 1&1 |<01|H'|Ki>l (5) i - E D E - E o 01 01 K#0 01 K1 The first term of the right hand side of equations (A) and (5) is the zero-order unperturbed energy for the iso- lated chromophore. The second term in both equations is the first-order term and represent the interaction of the permanent dipole moments of solvent and solute molecules. This contributes to the shift only if both molecules are polar. The three remaining terms in both equations arise in second order and represent respectively: (a) the third term is the interaction of the permanent dipole moment of the solute with the dipole moment induced in the solvent by the solute dipole. The contribution of this term is zero if the solute is non-polar. (b) the fourth term is the interaction of the permanent dipole moment of the solvent with the dipole moment induced in the solute by the solvent dipole. This term vanishes for a non-polar solvent. (c) the fifth term is the interaction of the mutually induced dipoles of the solute and solvent. This term is non-zero in all cases. The average shift of the transition energy due to 22 solute-solvent interaction, was obtained by subtracting Equation (A) from Equation (5) and averaging the resulting expression over the appropriate two particle distribution functions for the separation and mutual orientation of a solvent molecule relative to the solute. This was repre- sented19 as follows: AV. AV. _ Av. [(<01|H'|01>)AV. _ (<00|H'|00>)AV 1 (|<01|H'|K1>|2) (|<00|H'|K0>|2) + [ z { AV' - AV} 1 K50 E01 ’ EKl Eoo ‘ EKo (I<01|H'I0i>l2)AV (|<00|H'10i>|2)AV + [ z - E l lfii E01 - E0i 1%0 E00 — E0i . 2 . 2 (|<01|H'|Ki>l ) (|<00|H'|Ki>l ) + I ; Av _ 2 Av (6) K o E — E . K¢o E — E . 1%1 01 K1 i#0 00 K1 The interaction operator was approximated as follows: ' .- -> O -> H " TEm Usolute (7) + where Em is an electric field that the solvent medium pro— duces at the site of the solute molecule and Psolute is 23 the dipole moment operator on the solute coordinates. Wave— functions for the system that are Simple products of iso- lated molecule wavefunctions were introduced so that each matrix element of the perturbation in Equation (6) is ex- pressed as a product of three terms: a + + o <01IH'IKj> = = _ COSO (8) solutel where <0|Em|K> represents the electric field at the Site of the solute produced by the u K th permanent or 0 transition dipole moment of the solvent molecules. represents the appropriate permanent or transition dipolenmmmnm;of the solute; and cose represents the effect of mutual orientation of the solvent and solute. Each of the four terms in Equation (6) was decomposed and related to macroscopic properties: Only the final forms will be given without their mathematical details. A. Interaction between permanent dipoles of the solute and the solvent is represented by the first term of Equation (6) and is written as 2A 2 [<01IH'|0|>]AV - []AV = D[§}% _ “2‘:] (9) n+ where D = —t<2a-3>(ugolute>1 (10) e, n, represent the dielectric constant and the refractive l 0 index of the solution, respectively. “solute’ “solute are the magnitudescfi‘the permanent dipole moment of the excited and ground states, respectively, and B is the angle between them. d is the radius of a solute molecule obtained by ap- proximating it as a spherical particle. B. Interaction between the permanent solute dipole and induced solvent dipole is represented by the second term of Equation (6) and thus can be written as 2 2 (I<01IH'IK1>I )AV (|<00|H'|K0>| )Av _ n2-1 E - E ' E E 7 CE 2 J K#O 01 K1 00 - K0 2n +1 (11) where _ -3 l 2 _ o 2 C " '[d (“solute) (“solute) J (12) 25 C. Interaction between induced solute dipole and per— manent solvent dipole is represented by the third term of Equation (6) and has the following final form: . 2 . 2 (|<01|H'|Ol>| )AV _ (|<00|H'|01>| )AV _ 1%1 EOl - EOi 1%0 EOO - EOi 2 2 l - 2 + e (n+2) where 2 1 = _ 108 An (R/d) l 0 B E R3 kT (asolute - OLsolute)] (1A) h 0 d 1 ' b‘lit' f ere “solute an asolute represent the polariza 1 1es o the initial (ground) and final (excited) states of the solute respectively. R represents the radius of the sol- vent shell taken from the center of the solute. k, T are the Boltzmann constant and temperature in Kelvin scale. D. Interaction between mutually induced dipoles of solute and solvent if; represented by the fourth term of Equation (6) has the following final form: (|<01|H'|Ki>|2)AV] _ (IOOIH'IKi>'2)Av] z [ z = K#O K O 131 E01 ' EKi 1:0 Eoo ’ EKi 2 n -l A[-—§——1 (15) 2n +1 26 where 11 solute Oi )2)1 (16) 2 ) )-( Z A0i(usolute 1 A = -[—-1[( Z A .(u d3 11 ifio i#l where Ali and AOieuwaweighting factors9 depending upon the transition frequencies of both the solute and solvent. The final form of Equation (6) would be then: n2—1 2 1 2n +1 AV = (A + C) [ l [(e-n2)(2e+n2)] +B e(n+2)2 2 6‘1 n -1 The parameters A, B1, C and D may be viewed either as phenomenological constants to be evaluated from experimental results or as parameters to be calculated from Equations (l6, l3, l2 and 10) respectively. Equation (17) is still approximate and similar to Equation (18) found by other authors.9’20’21 n2—l e—l n2-1 Av = dispersion term + B( ) + C(——— - ) + Stark 2n +1 6+2 n2+2 effect term (18) 27 Either equation is still approximate and describes only Shifts resulting from interactions between the solute and solvent molecules. They do not take care of any specific interactions like, hydrogen-bonding, proton- transfer, etc. The ability of different terms in Equation (18) to explain solvent shifts in different types of solvent system has been examined for both n-n* and n-n* transitions of 2220321 n-n* transition organic molecules, including dyes. energies of C=O and C=S groups in different solvents are found to vary linearly with the stretching frequencies in the same solvents, indicating the importance of ground- state stabilization by solvents.2o’21 The London dispersion term in Equation (18) (causing red shifts with respect to the gas phase) also involves the function (n2-l)/(2n2+l). A plot of spectral shifts of non- polar solutes like aromatic hydrocarbons against (n2-l)/ (2n2+1) is found to be linear. The linearity is strictly expected for non-polar solvents (aliphatic hydrocarbons and so on). It is not surprising, therefore, that different linear plots are found for different families of solvents, particularly when some of them, like ketones, are quite 18’19 It is noteworthy that solvent Shifts of polar. polar solutes in non-polar solvents are also accounted for by the (n2-l)/(2n2+l) term, although the observed solvent shift would be due to the combined effect of the first 28 two terms in Equation (18). One could, in principle, rationalize Spectral shift data in different types of solvent by incorporating a dielectric constant or Stark effect term. This has been done by Nicol19 who derived Equation (17) and found a linear plot of Shifts in the absorption maxima of aromatic hydrocarbons in a variety of solvents. This equation also Shows that the dispersion effect on the refractive index terms of the solvent alone cannot account for spectral shifts in polar solvents. Equation (18) can be simplified to study the effect of the dielectric constant term (third term) alone by measur— ing band shifts in two polar solvents of nearly the same refractive index but different dielectric constant.20’21 From such a study, one can obtain estimates of the excited- state dipole moments of solute molecules. Diethyl ether (n = 1.356 and s = A.3) and acetonitrile (n = 1.3AA and e = 37.5) seem to make a good pair of such solvents. Basu22 has given a detailed quantum mechanical treatment of frequency Shifts in solutions by considering Onsager's reaction field model. Basu evaluated the stabilization energy of electronic states due to solute-solvent inter- nd order perturbation theory. action in terms of 2 Finally, I would like to mention that Kosower23 has given a solvent polarity scale based on the effects of solvents on the intramolecular ChargeeTransfer band of pyridinium iodide. This scale is defined in terms of Z 29 values of solvents given by the energies in Kcal/mole of the absorption maxima of A—methoxycarbonyl-l-ethylpyridinium iodide. The solvents listed in this scale vary from non- polar to highly polar and hydrogen bonding solvents. The change in charge-transfer absorption spectra are so large and their measurement so readily made the Z values have been preferred by physical organic chemists over the Y valueszu obtained from solvolysis kinetics of t-butyl- chloride. Charge transfer to solvent Spectra of I' and other systems have been correlated with Z values.25’26 While such empirical parameters may be useful for correla- tions, they do not provide the exact mechanism of solvent effects, considering the wide variation in the nature of solvents included in obtaining the scale. IV. Solvent Shifts as an Aid in Characterizing Electronic States Measurement of the effect of different solvents on absorption bands is one convenient way of characterizing the electronic states involved. As an example, solvent effect on different types of electronic absorption transi- tions can be classified as follow: 1. n + n* vs n-+n* Transitions In an n + n* transition in a carbonyl group the oxygen atom has less electron density in the excited state, and 30 the carbonyl group, as well as the molecule as a whole, becomes less polar. In a hydroxylic solvent such as ethanol, hydrogen bonding between the carbonyl group and the solvent is stronger in the ground state than in the excited state.7 The net result is that n-n* absorption bands are blue- shifted in hydroxylic solvents relative to non-polar, hydro- carbon solvents. A typical shift may be on the order of Moo—800 cm’l. A more extreme case occurs when a molecule with a non-' bonding orbital is dissolved in acid solution. Under these conditions, the atom with the n electrons is protonated, and the energy difference between the ground and excited states becomes much greater than in alcoholic solutions; the solvent shift to shorter wavelengths may be so large that the n-w* transition appears to vanish completely. Transitions in ketones that are classified as n-n* generally show a slight increase in polarizability in the excited state; the result is a small red shift (usually less than 600 cm-1) of a w-n* transition in a polar solvent relative to a nonpolar solvent.27 Transitions in aromatic hydrocarbons which contain no heteroatoms are considerably 28 less affected by polar solvents. Intramolecular charge transfer statesixlaromatic ketones, by contrast, become much more polar relative to the ground state. Red shifts 1 of 2000 cm' or more in polar solvents are common,27 for intramolecular charge transfer bands in aromatic ketones. 31 The general rule for solvent Shifts is that if the excited state is more polarizable than the ground state or has an increased permanent dipole, the Spectrum is red- shifted in the more polar solvent; if the reverse applies, the Spectrum is blue—shifted. 29 applied the criteria of solvent shifts in Kuboyama assigning the lowest—energy, singlet-singlet transition in fluorenone, which had long been considered to be an n + n* transition, as a n + n* transition. His results were later substantiated by Yoshihara and Kearns.3O 2. Locally—Excited States vs Charge-Transfer States One may classify the electronic states of substituted molecules as being locally excited (L.E.) or charge-trans- fer (C.T.) states in the zeroth order approximations. A charge-transfer (C.T.) state involves the transfer of an electron between the hydrocarbon and the substituent. Com- paring the spectra of a substituted benzene molecule taken in hydrocarbon solvents with those taken in polar solvents will offer a means to identify "C.T." bands. For substi- tuted benzenes,especially those containing a strong donor and a strong acceptor substituent, the dipole moment increases in the same direction in the CT state and thus a red shift is observed whose magnitude depends greatly on the polarity of the solvent and the change in dipole moment during excitation. One Should also mention that 32 the CT bands of chloro, bromo and iodobenzenes are ex- pected to shift to the blue since their dipole moments de— crease or may change direction as a result of excitation to CT states.31’32 Relatively large red shifts are usually characteristic of intramolecular C.T. transitions of sub- stituted benzenes. Room temperature absorption spectra of p-amino-nitro- aniline in methylcyclohexane MCH, acetonitrile CH3CN and ethylalcohol EtOH are Shown in Figure (2).33 It is clear that the first absorption band is more sensitive than the 2nd band, the first band (CT band) undergoes a red shift 1 in going from hydrocarbon solvent to ethanol. of NA000 cm- As a result of charge—transfer migration from the amino to the nitro group, the dipole moment of p-aminoaniline is expected to increase in the CT state but remain in the same direction as that of the ground state. Larger solva- tion energies are therefore expected in the CT state com— pared to the ground state due to the increase in the dipole moment. Also larger hydrogen-bond energies are expected due to the increase of the acidity of the amino group hydrogen atoms and the increase of the nitro group basicity. This explains the red shift observed in polar solvents, particularly hydrogen bonding solvents. The spectrum of p-nitroaniline in alcohol (Figure 2) 1 shows a shoulder at ~29A and 305 nm corresponding to the B211 Locally Excited (LE) benzene transition. Its frequency is 33 I ' I I j .r A I r 1 I ’\ .0 . a." no r- [I ‘- ' __ —:nm ALcoroL .’ A. --------- Miran cvcmrcum: ze+ f 3/ \\ .__m____ ' 1 : } \. ' ‘ l \ . ~£rncn \ /' // \ \ I as” ./}/ ,‘ \. \ L__._L._.J_- 1’ J -e; 1-- 11°C; I. -_ "i ------ _~._-\_-;— in» 2000 2400 2.00 3200 3600 0000 4400 4.00 ‘ IAVE LENGTH IN ANGSYROHS Figure 2. Room temperature absorption spectra of p- nitroaniline in different solvents.33 3A not much affected by change of solvents. 3. Electron-Transfer Transitions Halogen anions, 1’, Br’ and Cl’ show strong absorption 373' Their Spectra are shown bands in the ultraviolet region. in Figure (3a);fl”)where one can see that the I' and Br" exhibit two absorption bands. The spectra of halogen anions were ascribed to electron transfer absorption,35’36 where an electron was transferred from the halide anion to the solvent. Strickler and Kasha37 studied the Spectra of halogen anions in acetonitrile, water and ethanol. The Spectra of iodine in these solventsenwzshown in Figure (3b).37 They38 have studied the electronic transitions in N03 and N05 ions in solvents of different polarities and differentiated between electron transfer and other internal transitions (n + W*, n + v* . . ., etc) in these ions. 9,A0 studied the electronic Kosower and co-workers3 spectra of N-alkyl-pyridinium iodide in various solvents and found that the spectra are very sensitive to solvent polarity and a large blue shift occurs with increasing sol- vent polarity. Each absorption band was assigned to the charge transfer band in which an electron is promoted from the iodide ion to the pyridinium ring and therefore the CT state is characterized by a small dipole moment which lies in the plane of the ring. 35 I'M!) 79 1.0 - I ‘ a 'l - SOLVENT U o " 0-0 ' a '1 as» t - (puma: ‘ E E ‘0 x 1 ‘."~_;L._ \4 1 “> 55 a: 45 40 ,5 D (KKJ Figure 3a. Room temperature absorption spectra of halogen ions in aqueous solution (D20).3Ab ANGST ROMS Figure 3b.. Room temperature absorption spectra of sodium iodide in acetonitrile, water and ethanol.37 36 u = 13.1D p = 8.6D This is a case where the solvation energy of the ground state is larger than that of the excited state. Moreover, the dipole moment changes its direction as a result of excita- tion and therefore a large Franck-Condon orientation strain is expected. Both effects contribute to the large ob- served blue shift in more polar media. A. Sihglet-Triplet Transition The probability Of Tl + SO transition in organic com- pounds is a highly sensitive function of the presence of heavy atoms which enhance spin-orbit interactions and in- creases the intensity of the transition. The Spin-orbit interaction is treated quantum mechanically by introducing into the Hamiltonian operator, a term HSO for each electron of the form H80 = K a (L - s) (19) 37 where L is the orbital angular momentum operator, S is the spin angular momentum operator and g is a factor depend- ing on the nuclear field. g and therefore HSO is propor- A because of the reciprocal rela- tional to Z/r3, 343;, to Z tion between Z and r. (Z is the nuclear charge and r is the distance between the electron and the nucleus. Per- turbation theory shows that if mg and w% are the wave- functions of "pure" singlet and triplet states, respec- tively, then the triplet state produced under spin—orbit coupling can be written!41 in the form of the following equation < 0 [H W°> SK SO T wT = w% + z - ng (20) K ET ’ ESK A Similar expression can be written for the singlet state. Thus the effect of Spin—orbit coupling is to mix a small amount of singlet character into the triplet states and vice-versa, so that "pure" Singlet and triplet states no longer exist. The probability of S + T transition increases rapidly with the atomic number and thus solvents with a heavy atom such as iodine will enhance Singlet-triplet absorption bands; this will help in their identification. 38 V. Solvent Effect on Emission Spectra 1. Spectral Shifts A solute molecule in its ground state is surrounded by solvent molecules in equilibrium in solution. The geom- etry, charge density and dipole moment of the solute mole- cule may be different in the excited state; therefore, the equilibrium excited-state configuration of the solvent cage will also be different. The solvent configuration around the excited solute molecule immediately after the electronic transition does not correspond to the equilibrium-excited configuration, but to a configuration geometrically identi- cal to the solvated ground state, i.e., a Franck-Condon state configuration. The relaxation of this configuration occurs to the excited state equilibrium configuration. The time required for solvent reorientation is around 10—11 sec in fluid media. Since the life time of an excited singlet state is of the order of 10'8 sec, there is enough time for excited state equilibrium to be reached before deactivation occurs if the solvent is not viscous. Also the ground state configuration after fluorescence is not the equilibrium ground state configuration but a state of strain whose energy is higher than that of the ground state equilibrium configuration. If the dipole moment of the solute changes (in magnitude 39 and/or direction) upon excitation and the solvent is polar, reorientation of solvent molecules occurs before emission. However, if the solvent is rigid, relaxation times are several order of magnitude larger than the excited state life time, and emission occurs before solvent rearrangement takes place. If however, the polar solvent is fluid, relaxa- tion is much more rapid and emission may occur from the equilibrium excited state where dipole reorientation is completed. Therefore, absorption will occur to the meta- stable Franck-Condon state (Va = absorption frequency in wave numbers cm-l) and emission will occur from the equilib- rium state (;a = fluorescence emission frequency in wave -1). The quantitative expression for A? in numbers cm absorption is different from that in emission and the 0-0 band will not coincide. The difference (Stokes' shift) is 2 2 2 D—l n -1 2 = (u _u ) [— _ _—:l + T a3hc e g D+2 n2+2 a ho n2-l n2+2 [<3u§ - 5n: + 2ugue>1£%}% - 1 (21) where a is an effective cavity radius appropriate for the solvent, pg, “e, D, n, “e and a have their usual meanings. g The second term originates from the dipole-induced A0 dipole interaction, in many cases can be considered as a second order interaction term and makes a negligible con- tribution to the shift and the equation is Simplified as follows: 2 _ = I " — 2 2 (D—l) M A v v - v - ———— (u —u ) [T___I - l (22) The dipole moment difference in ground and excited state can be obtained directly from the Stokes' Shift. An esti- mate of the excited state dipole moment can be made from experimental absorption and emission shift data and the known ground state dipole moment using Equation (22). In highly viscous solutions this solvent relaxation is slow and fluorescence occurs, but from a non-equilibrium, Franck—Condon state. Since the Franck-Condon state is always higher in energy than the equilibrium excited state, fluorescence in highly viscous solutions is blue Shifted with respect to that in solutions of low viscosity. Another consequence of increasing the viscosity of the solution is the increase in the intensity of emission. In fluid media radiationless transitions are very fast in these relaxing systems.“2 One method of increasing viscosity is to use solvents which form rigid glasses at liquid nitrogen temperature; A1 a blue Shift of the low—temperature fluorescence Spectra compared with those at room temperature in the same sol- vent occurs. 2. Viscous-flow Barriers“3 Molecules in condensed media always are surrounded by a solvent cage. The cage may be a liquid solvent, a macromolecular enclosure (enzyme site), a lipid membrane, a multi-layer lamellar system, or may be a crystal or surface-adsorption cage. Although photochemical and other kinetic studies have long taken cognizance of the effect of solvent cages on recombination rates, Spectroscopic studies of the mechanism of solvent cage action have been neglected in comparison. The mechanical Viscous- Flow Barrier solvent cage has been analyzed on a quantum—mechanical basis by Dellinger and Kasha.uu’145 Figure (A) indicates schematically the basis of their model. If a molecule upon excitation requires a rather large distortional motion for relaxa— tion, then a large volume of solvent in the cage around the solute must be displaced. For example, in trans-stilbene the torsional relaxation about the double- bond requires that the phenyl groups sweep out a large volume of solvent to reach an equilibrium excited state configurationus (upper part, Figure A). On the other hand, A2 ISOMERIZATIONAL EXCITATION O O Q 0 0 0C? 000%”) Q 000 g°©o°3o 0m ' ‘oW 9 03,00 s00“? 0098898%39 00 Q00Q§00 DISSOCIATIVE EHCITATIPN . 0 00 0 A 07230 ‘ 0 00 00 o o€§8§fiég on 0693978937100 90 @009 690$?) U&@0%9%;08% '00 0’ 0 0Q O - SOLUTE GROUND STATE . SOLUTE EXCITED STATE Figure A. Mechanical viscous-flow-barrier cage proposed by Dellinger and Kasha.A3 A3 a hydrogen-bonded molecular pair, such as 9,10—diazaphen- anthrene complex with t-perfluorobutyl alcohol: 0 N-N .. .. CF3 H\ / O-IC-CF3 CF3 would require displacing a large volume of solvent mole- cules in the photo-dissociative excitation of this molecular complex.”5 According to the Dellinger—Kasha model, the solvent cage imposes a barrier to the molecular motion, which ap- pears as a Gaussian-flow barrier added to the potential function where the latter become horizontal, igeg, in the dissociative case, at the dissociative limit. The phenom- enological reality of these viscous-flow barriers has been tested by low-temperature spectroscopic studies. Thus, luminescence phenomena reveal that the molecule is trapped in a ground state configuration in rigid glass solvents. Recently, Mohammadi and Henry”6 have verified the viscous—flow barrier to molecular motion by an entirely AA different route. They studied the infrared overtone an- harmonicity of C-H vibrations in methylbutanes, and were able to show that the decrease in anharmonicity predicted by the Dellinger-Kasha model was experimentally verified and could be calculated by the Lennard-Jones potential as a quantitative perturbation of Morse potential.“6 An interesting case of excited state proton transfer spectroscopy was discovered recently in flavones by Sengupta and Kasha.u7 In 3—hydroxy flavone it was discovered that in hydrocarbon solution at room temperature a green fluorescence was observed, unrelated to the ultraviolet absorption of the molecule. It was deduced that an intramolecular proton-transfer had occurred within the internal H-bond to the carbonyl group, and that a pyrilium-like excited state tautomer yielded the green emission. Upon freezing in a rigid glass solvent, the normal violet-ultraviolet fluorescence could be inhibited by the solvent cage,u7 so that the viscous-flow barrier prevented tautomerization. A5 3. Solute—Solvent Relaxation in the Nano and Pico— second Range Molecular relaxation occurring in the "molecule-solvate Shell" system after optical excitation is governed by the correlation between radiation lifetime If and environment relaxation time TR. The main cases possible here are as follows:2 1) Tf >> TR - all molecular relaxations are completed within the time Tf, 143;, the molecule at the moment of deexcitation is in thermodynamic equilibrium with all modes of the "molecule-solvate Shell" system; 2) 1f m TR - molecular relaxations are realized only in part with the result that at the moment of emission the system is not in thermodynamic equili- brium; 3) If << TR - the case of extreme nonequilibrium, for the relaxations (orientational and transla— tional in particular) do not proceed at all. Picosecond and nanosecond time resolved spectra are used to measure these relaxation processes. For a solute that undergoes a large change in dipole moment upon excitation, the study of time dependent spectral Shifts will reveal information regarding solvent relaxation. Time dependent emission spectra taken at different delay A6 times after the excitation pulse are measured. An example of a time-dependent spectral Shift given by the polar molecule ANS (1-Anilino-8-naphthalene sul- fonate) in a polar solvent, n-propylalcohol, is shown in Figure (5).”8 Figure (5a) shows the steady-state fluores- cence Spectra of ANS at room temperature, intermediate tem- perature and liquid nitrogen temperature. A blue shift of A1 nm is observed. Large time-dependent spectral shifts were observed only in the temperature range from -700 to —l70°C in n-propyl alcohol. At higher or lower tempera- tures the emission spectra are essentially time independent over a time range of 70 nsec. (the lifetime of ANS of ap- proximately 20 nsec places a limit on the length of time available for time—resolved Spectral measurements). Fig- ures (5b) and (SC) Show the time-resolved spectra in n- propyl alcohol at -90° and —l50°C. The indicated delay- time for each spectrum is given relative to the initial rise of the nanosecond flash lamp. It is observed from the figures that the time required to approach the relaxed spectrum (corresponding to the spectrum at room tempera- ture) increases with a decrease in temperature or an in- crease in viscosity of the solvent. Wareu8'50 interpreted the phenomenon of time-dependent spectral Shift as being due to solvent reorientation about the excited molecule that is required to accommodate the change in dipole moment and moment direction that occurs A7 Figure 5. (a) Steady—state fluorescence spectra for ANS in n—propyl alcohol at the indicated temperatures. (b) Time-dependent fluorescence spectra for ANS in n—propyl alcohol at -90°C. A, 0 nsec; B, 2.5 nsec; C, 12.5 nsec; D, 21.5 nsec; E, 31.5 nsec. (c) Time-dependent fluorescence spectra for ANS in n-propyl alcohol at -l50°§. A, 2 nsec; B, 11 nsec; C, 68 nsec. FLUORESCENCE INTEWTY A8 a )- p. Q E Z A 8 496 450' (D . Id § 42:» ~29 d " v. 550 3&1 wwaimnnmm) c ' V ‘( ). c U) 2 E E LIJ O 2 U U , 3 c § ; d A Kb 130 its 400 WAVELENGTH («rd 1 Figure 5 l _ 450 500 WAVELENGTH (nm) 550 A9 upon excitation. The extent to which time-dependent spec— tral shifts are observed is a measure of the degree of relaxation of the excited molecule towards its equilibrium configuration at a given temperature. Azumi51 related the time—dependent Spectral shift to the edge excitation red Shift and interpreted the two phenomena as being due to the same mechanism, which was given by Ware. Naturally the Azumi model based on the above mechanism cannot explain the lack of time-dependent spectral Shift and the appearance of REE at 77 K. We believe that the two phenomena have different mechanisms. The time-dependent spectral shift represents spectra corresponding to dif- ferent stages of relaxation of the bulk continuum at a par- ticular temperature, while the red edge effect represents the spectra corresponding to different solvation sites (different local environments) when excited monochromatically at the red edge of the bulk continuum. The picosecond light pulse method was used to measure the molecular orientational relaxation times, and time dependencecfi‘solute rotation. The principle idea of this laser technique is to induce an anisotropy in the orienta- tional distribution of the solute molecule with a pico- second excitation polarized pulse, and to monitor the return of the system to isotropic distribution with an attenuated 52 picosecond pulse. Due to the induced anisotropy, the absorption of the probe pulse is polarization dependent. 50 The decay of this dichroism with time, due to thermal molecular motions is determined by measuring the relative transmitted intensities III/Ii of the probe light; I||(t) and 11(t) are the components, respectively, of the probe light polarized parallel and perpendicular to the excitation light at the time t after the excitation pulse. For rota— tional motion describable by the rotational diffusion equa— 52 tion, Eisenthal et al. obtained a relation I 2n Till .. expE-(6D + mm (23) where D is the rotational diffusion constant and T is the excited state lifetime. The orientational relaxation time is given by (6D)'l. Eisenthal et al.52 reported the orientational relaxa- tion of rhodamine 6G in a series of normal alcohols, ethylene- glycol, chloroform and formamide together with the effect of hydrogen bonding and the structure of liquids on the molecular rotational motion using picosecond light pulses. It was found that even though methanol forms a stronger hydrogen bond with rhodamine 6G than chloroform, the relaxation times are found to be equal. The insensitivity of the orientational relaxation of rhodamine 6G to its formation of hydrogen bonded complexes with the molecules of the solvents was explained in terms of the orientational freedom of the hydrogen bond and the fact that the complex 51 3o 1 v 1 I I T I T ' ' 28 P I l—Undecanol 'l 26 t ‘ 24 ~ ‘ I l—Decanol 22 t I ‘ Eth lane 20 - Glycol E 18r 7 O .0 16 __ I—Octanol .1 ha 14 e " 12 _ ‘ lO - ‘ Formam'd 8 ' l—Pentanol .1 6 e l—Butanol ‘ 4 _ l—Propanol - 2 +- Ethanol Methanol q Chloroform 1 1 1 L 1 1 1 1 0 2 4 6 8 10 12 14 16 18 20 21 nlcentipoises) Figure 6. T versus solution viscosity for rhodamine 6G or 52 in various solvents. 52 is dynamic in that the solute-solvent hydrogen bonds are breaking and reforming. The same results was found by 53.514 Levshin et al., for the same solute in formamide and pentanol solvents. Those observations are clear in Figure (6)52 and show that the relaxation times of rhodamine 6G in the liquids through octanol vary linearly with the solution viscosity. The departure from linearity in case of decanol and un- decanol has been interpreted in terms of greater linear dimension of solvent relative to solute. The deviation of rhodamine 6G in ethyleneglycol may be due to extensive solvent—solvent aggregation by hydrogen bonding interactions. A. Excited State Level Inversion If two different excited energy states S1 and 82 which have quite different electronic structure lie in close proximity, there is a possibility that the solvent effect will bring about the inversion of two different energy states because the solvent effect on these two ex- cited states may be considerably different. When this kind of inversion occurs in some solvents, a consider- able change of fluorescence Spectra, intensity, polarization and lifetime is invariably observed, though absorption spectra will remain almost unchanged. The dual fluores- cence character of p-cyano-N,N-dimethylaniline (DMAB) in different solvents first pointed out by Lippert 53 et al.55,56 is Shown in Figure (7a).55 Such dual fluores— cence caused by environmental perturbations has been ob- 61 62 served also in indole,57'6O 1-naphthylamine and l-naphthol. Suzuki et al.63 have observed fluorescent level inversion of dual fluorescence in alcoholic solutions of l—naphthol by a change in temperature only. These results can be interpreted in the following way. These molecules have two electronic states lLb and l 1 L o all’l L the region of their lowest absorption band. The a state lies above lLb in free molecules, but these states do lie close to each other. In fluid solutions at room temperature, the solvent molecules around the excited solute molecule will generally have time to reorient themselves before light emission occurs and hence they relax to their preferred equilibrium configuration which is of lower energy. In non p018? solvents, the 1L states of these molecules a still lie above 1Lb even in the preferred equilibrium con- 1L b state. In polar solvents, the interaction between dipole figuration and hence fluorescence occurs from the moments of such excited molecules and solvent molecules low- ers the energy of the 1La state below that of lLb in their equilibrium configuration and hence fluorescence occurs from the 1La state. The above sequence is shown in Figure (7b).6u At sufficiently low temperatures and high vis- cosities the relaxation processes which lead to an equil— ibrium configuration will not occur to any appreciable 5A (a) Effect of temperature on the fluorescence spectra of 5 x 10-5 M DEAB (p-diethylaminobenzo- nitrile) in butylchloride-methylcyclohexane- isopentane mixture (12:3:1 in volume). ——————293°K, ______ 23A°K, -._._ 1730K, -.._..— 1A8°K.55 Figure 7. (b) Excited state level inversion caused by mutual interaction between solute in excited state and polar solvent. 55 IF J 1 J J 1 20.3 22.9 24.0 25.6 27.4 29.9 7°10”[cm"] —_. IL increasing lorlly o ——————E: F‘.‘ § ‘Lb . 5 3 1.1—1— 3. Obi floor. lluor. obsorpllon fluorosconoo in in nonpolar polar-solvonlo solvonl Figure 7 56 extent during the lifetime of the excited state and there- fore emission will take place from an unrelaxed configura- tion. Mataga measured the time-resolved fluorescence spectra and fluorescence decay curves of the 1L and lLb fluores_ a cence of DMAB and demonstrated clearly the relaxation process forming the lLa—solvated state. The 1Lb fluores— cence approaches its intensity maximum after A nseconds and 1La fluorescence after about 7 nseconds after the end of a short excitation pulse. CHAPTER III THE RED EDGE EFFECT AND A RELATED PHENOMENON Fluorescence and phosphorescence of some organic mole- cules exhibit a red Shift when excited at the long wave- length edge of the first absorption band. This phenomenon has been termed the red—edge effect65 (REE), or edge—excita- tion red shift66 (EERS) which actually is the difference in wave numbers (cm-l) between the emission maxima ob- tained on shorter wavelength excitation and excitation at the red edge of the first absorption band. Narrow band excitation is one of the necessary condi- tions to be able to observe the REE. This can be accom- plished by using a good quality excitation monochromator or recently by using tunable dye lasers. Another condi- tion to observe this phenomenon is the rigidity of the medium such that the life time of the excited state would be in the range of solvation site life time. This condition can be satisfied by lowering the temperature of the solution or by dissolving the material in a polymer matrix (e.g., poly (vinyl alcohol) or polystyrene) at room temperature. A review of observations of the phenomenon made by dif- ferent authors during the last two decades is now given. 57 58 In 1960 G. Weber reported67 that concentration depolariza— tion in rigid solutions, which results from singlet—singlet intermolecular energy transfer, failed to occur upon excita- tion into the red edge of the chromophore absorption band. The emission polarization experiment is represented dia- gramatically in Figure (8a). The polarization P is defined by the following equation: Ill - II P=I||+Il (2“) where III is the number of photons per unit time (intensity) emitted with their electric vector parallel to the electric vector of the exciting radiation in the laboratory frame of reference and Ii is the number of photons per unit time (intensity) emitted with their electric vector perpen- dicular to the direction of the electric vector of the exciting radiation. Polarization reflects anisotropic character and hence lack of energy transfer while de— polarization reflects isotropic character and hence an energy transfer. Figure (8b)67 shows Weber's study of the concentration depolarization of indole in thin propylene glycol films at 220°K. Figure (8b) shows that excitation at the red edge of the absorption band (above 290 nm) resulted in enhanced polarization which indicates that little or no energy transfer occurred under these conditions 59 5 F-—‘ Perpendicular w tP/larized (IL) 5" /l ’1‘--. . Parallel Polarized Sample Cell Emission (In) Polarized Excitation Figure 8a. Diagram showing orientation of parallel and perpendicularly polarized emission with respect to plane of polarization of exciting radiation in the laboratory coordinate system. 01111111LJ 230 240 250 160 170 280 190 300 3‘0 limp) Figure 8b. Excitation polarization spectra of indole in propylene glycol at -70°C. I, 0.0l M; A, 0.05 M; x, 0.1 M; 0.0.2 131; o, 0.14 131.67 60 even in concentrated solutions. Other studies68 by the same author have indicated that this is a general phenomenon in aromatic molecules. Galley and Purkey69 measured the excitation wave- length dependence for emission spectra of indole in rigid media. Figure (9a)69 displays the fluorescence and 0-0 region of the phosphorescence spectra of indole in a 1:1 ethyleneglycol-water glass excited near the center (280 nm) and the red edge (295 nm) of the absorption band. Clearly the emission spectraenwanot independent of the exciting wavelength as is generally assumed, for the Spectra gen- erated at the longer exciting wavelength are red shifted by about 900 cm.1 in fluorescence and 90 cm"1 in phos- phorescence. The emission Shift decreases with increasing concentration, at 0.5 M the shift in the fluorescence of indole is 600 cm’1 for 280 and 295 nm excitation. An exciting wavelength dependence was not observed for the fluorescence spectra at room temperature. This depen— dence of the phenomenon on solution "rigidity" is. depicted in Figure (9b).69 The differences in emission wavelength for both the phosphorescence and fluorescence spectra of indole excited at 280 and 295 nm in A:l glycerolzwater were plotted as a function of temperature. The exciting— wavelength dependence is lost in either case, but the tem- perature at which the transition occurs for the phosphor- escence is much lower than that for fluorescence. Figure 9. (a) (b) 61 Fluorescence and 0-0 region phosphorescence spectra of 10‘3 M indole in 1:1 ethylene glycol-water media at 202 and 77 K, respectively, demonstrating the shift between the emission excited at 280 and 295 nm. Plot of exciting wavelength dependence for the 10'3 M fluorescence and phosphorescence spectra versus the temperature of the A:l glycerol-water medium. Fluorescence and phosphorescence shifts were measured as the average of the red edge and blue edge differences between the Spectra at 280 and 295 nm excitation.59 62 wavenumhets ism") 3 relative emission Intensity 313 311 2]? fig: 25 151.5 251.0 241.5 x10 lNlllllE nunnescsuc: mum: Pnnspunns’scenca exclling, beam l l l l l l l l 300 325 350 375 ' 395 400 405 410 wavelength lnml (a) 0007— “I" — ’ 3m 0 r:-—_.I—:’. L ......... 0 non-- - .’ ‘ - Ill 2 I, -0 g sun — . l.’ - Ell é " 400,. llumescence ': phosphmescenca _. 40 i \ ,1 / E200— /. I" - 20 ' ’o/ I ’I ’5 n 1. o -—-o -..-o' o —1 o .L_ l i _1 i i l) -4ll -80 -I20 - 160 —200 tempevatme (“Bl (b) Figure 9 63 K. Nagvi et al.70 showed that the shape of the fluores- cence spectrum is sensitive to the exciting wavelength in rigid media. Figures (10a and 10b)70 shows better resolu- tion in the spectra of P-terphenyl (TP) and P-quaterphenyl (QP) in rigid ethanol when excited at the red edge of the absorption band. Similar observations to the above were found by us for different systems and are discussed in detail in Chapter V. The phenomenon has been confirmed experimentally by many other authors.71-79 Earlier interpretations of the phenomenon will now be given. Weber7l’72 interpreted the phenomenon present in naphthyl amine derivatives as due to the existence of out- of—plane transition moments in absorption and emission, normal to the plane of the aromatic rings,which he thought arose as a result of coupling between out-of—plane nuclear vibrations and in—plane transition moment. His interpre- tation was based on the measurement of polarization at different wavelengths of excitation and calculation of the rates of in-plane and out-of-plane rotations in aromatic compounds.71 Galley69 interpreted the REE in terms of heterogeneity of solvation sites and emphasized the composite nature of both absorption and emission in rigid solution and pro- vided a simple rationalization for the red edge failure of concentration depolarization obtained by Weber. Excitation 6A INTENSITY l J l l L 1 L 320‘ 3‘0 300 WAVELENGTH (hm) Figure 10a. Fluorescence spectra of «.10'5 M TP (p-ter— phenyl) in ethanol at 77 K at different excita- tion wavelengths. ——————-280 nm; ----- 310 nm; ..... 315 nm. INTENSITY 340 300 ~ 300 400 ‘30 4‘0 WAVELENGTH (nm) Figure 10b. Fluorescence spectra of "’10"5 M QP (quater- phenyl) in ethanol at 77 K at different excita- tion wavelengths. —————— 320 nm; ------ 330 nm; ...... 335 nm.70 65 into the red edge of the chromophore absorption band selects a certain subpopulation of the solute molecules of small absolute concentration, which by virtue of their par- ticular local environments are characterized by relatively low 0—0 singlet electronic transition energy. Because the bulk of the chromophore molecules in the solution possess 0-0 transitions of higher energy than the subclass of molecules at the red edge, members of the latter subclass have, in general, only neighboring molecules with higher energies. Singlet-singlet energy transfer and, as a result fluorescence depolarization, thus cannot occur for those molecules selectively excited at the red edge of the chromo— phore absorption band. On the other hand, at Shorter exciting wavelengths, molecules of generally higher transi- tion energies are excited and readily undergo energy trans- fer and fluorescence depolarization. This model of dif- ferent local environment given by Galley explained the excitation wavelength dependence of the fluorescence in Figure (9a).69 The results depicted in Figure (9b)69 were explained by Galley as follows: at -110°C the mobility of the glassy solution is sufficient to randomize the chromophore solvation sites within the triplet—state life— time of indole (”7 seconds), whereas a temperature of -53°C is necessary to render the solution fluid enough to average the chromophore local environments within the nanosecond excited singlet state lifetime. 66 Nagvi7O interpreted the REE observed in P-terphenyl and P-quaterphenyl as a result of emission from different conformers. Chen73 interpreted the observation in quinene and 6—methoxyquinoline as a result of emission from more 7A than one electronic states. Tomin and Pavlovich75 gave an interpretation similar to Galley's model in terms of inhomogeneous spectral broadening. Azumi et a1.51 noted that the mechanism of the REE is similar to the one proposed by Ware et al.79’SQ in interpreting the time— dependent spectral shift. It may be convenient to distinguish two cases where the solute may assume different conformations: An example of this has been observed in the case of P-terphenyl and P-quaterphenyl in ethanol at 77°K7O and in the case of the esters of 9-anthroic acid in 3MP at 77°K, Chapter V, observed by us. The other case involves polar solute mole- cules interacting with various solvation sites in a polar solvent, iLQL, a distribution of different solvation sites; an example of this is the case of indole in ethylene glycol- water media at low temperatures69. Another example is our recent observation of a large REE in the case of mero— cyanine dye in polar solvents.79 67 A large number of polyaromatic hydrocarbons exhibit vibronic spectra consisting of a large number of qgasilines (i.e., a series of narrow bands which in most cases can be called lines) when present in a frozen n-alkane matrices at 77°K and lower temperatures. This phenomenon has been called "SHPOL'SKII EFFECT" after the Russian physicist E. V. Shpol'skii who discovered it in 1952.80 The crystal- linity and the transparency of the matrix (Shpolskii matrix) is one of the essential conditions to obtain the 80-8A quasi-linear spectra. Temperature as well as the rate of cooling of the matrix are other significant factors in observing Shpolisktieffect. This effect has become a general phenomenon in molecular spectroscopy.85:86 A typical example of the Shpolwfldi.effect is represented by the fluorescence spectrum of coronene in a frozen n- heptane matrix shown in Figure (11).86 The excitation wave— length dependence of the fluorescence Spectrum of benzoEaJ- pyrene in n-heptane matrix is also shown. Certain ali- phatic alcohols can also be used as solvents for the develop- ment of this effect. To explain the line-like structure one presumes that the aromatic hydrocarbon replaces an aliphatic hydrocarbon at a lattice site and thereby exists in a state that can be described as an oriented gas molecule. Thus the spectra are of unperturbed systems corresponding to the free Figure 11a. Figure 11b. 68 a (nun) Fluorescence Spectrum of coronene in a frozen n-heptane matrix. The sharp lines arise be- cause of the unique conformation of the coronene in the matrix. ,1 l "V— f -— CO! ‘00 CO! 0.. 000 ‘0' Variation in the appearance of the A03 nm band system in the fluorescence spectrum of benzo[a] pyrene in heptane matrices at 15 K under dye laser excitation as a function of excitation wavelength: (a) 385.5 nm, (b) 385.7 nm, (c) 385.9 nm, (d) 386.1 nm.86 69 molecule. The ideal situation results when the dimensions of the solute and solvent molecules are comparable so that the solute molecule can occupy a definite position in the host lattice with minimum deformation; for example, the emission spectrum of naphthalene in pentane shows quasi- linear structure, but is more diffuse in hexane and heptane. If the solvent molecule is too small, a similar situation arises. The broader, more diffuse Spectra that can result tend to be similar to those obtained in clear, rigid glass solvent systems. These can be a variation in the number of lines and their intensity depending on the solvent, ex— citation wavelength and temperature. The variations in part appear to originate from the fact that there are local differences in the crystal field surrounding different molecules. Thus, the multiplet structure appears in Fig- ure (11)86 is explained in terms of a number of distinct sites in the host crystal, in which the guest molecule can reside. Each Site will perturb the energy levels to a different extent and so a number of component Spectra, each with its own electronic origin, will be produced. Also, the rate of freezing of the mixture is important. Slow freezing results in the loss of some short wavelength com- ponents and weakening of some long wavelength components. This probably results from formation of rotational solvent isomers.82 Under slow freezing it is more likely that a greater proportion of the solvent molecules will be in the 70 most stable conformation and the spectrum will be simple. Such fine-line spectra can be used to determine structural details of the emitting molecule.82 In an amorphous medium (2;8;a solution), guest mole- cules occupy many different microenvironments or "sites" in a low-temperature matrix. Thus, the purely electronic energy levels of different molecules of the same solute are no longer the same and will be shifted to different 87-91 extent. Alternatively, it can be said that guest molecules reside in different "sites" in their "0-0" (purely electronic) absorption frequencies are different.87 The band width over which these different 0-0 absorption frequencies are found is a measure of the extent of "in- homogeneous broadening" which may contrast with the intrinsic or homogeneous band width of a single molecule. This contrast is clear from Figure (12).87 Upon monochromatic excitation near the 0-0 absorption band (where the density of vibronic state is small), only those molecules whose absorption bands overlap the laser line will absorb and consequently fluoresce. The resulting emission will therefore exhibit much narrower bandwidths than if a con- ventional broad-band lamp excitation source is used. This effect is called "site selection", "energy selection", or "fluorescence band narrowing".87"91 One of the experimental observations of site selection spectroscopy is shown in Figure (11).86 It shows the \ Figure 12. 71 0) AA A MOLECULES ENSEMBLE bl MOLECULES ENSEMBLE The limiting cases of homogeneous and inhomo- geneous broadening. In the homogeneous case (a) the spectrum summed over all molecules in the sample corresponds identically to that for any one of the individual molecules. In the inhomogeneous case, (b), the Spectrum summed over all molecules in the sample differs from the Spectrum that would be seen for any Single molecule. 7 72 excitation wavelength dependence of the benzo[a1pyrene emission band in a solid matrix at low temperature. Other examples in the literature Show the same concept as being useful in proving either Shpol'skii spectra or the Red Edge Effect discussed in the previous pages. CHAPTER IV MOLECULAR SYSTEMS The goal of our research is to study solvent effects on the absorption and luminescence properties of mole- cules that undergo large changes in their dipole moment (p) due to electronic excitations. One can study the change in fluorescence yield and energies as the polarity of the solvent is changed. Such changes can be examined in terms of specific interaction and bulk medium effects. One can also study the effect of solvent cage relaxation on the emis— sion spectra using nanosecond and picosecond time resolved Spectrosc0py, in fluids as well as in viscous media. Since different solvation cages may exist, different excitation energies may occur in dilute solutions of polar molecules. Thus, it was interesting to study the depen- dence of luminescence energy, lifetime and spectral resolu- tion on excitation wavelength at low temperatures. The molecular systems which we have studied are also potentially useful as fluorescence probes of biological systems. Three categories of molecular systems can be defined: 73 7A I. Molecular Systems that Undergo a Large Decrease in Dipole Moment as a Result of Excitation - i.e., pe < u g where Hg is the dipole moment of the ground state and He is the dipole moment of the excited state. Examples: 1. Merocyanine Dyes + / CTfli-' \\ I _ \ 6 (l-methyl-A—hydroxystyryl)pyridinium betaine or 1-methyl- A-((oxocyclohexadienylidene)-ethylidene-l,A—dihydropyridine. 3"” 3-methyl—2-((A-oxocyclohexadienylidene)-ethylidene- benzothiazol. 75 2. Alkyl pyridinium iodides :[ III methyl pyridinium iodide. For Dye I the static dipole moment decreases by 5.5D upon excitation to the first excited singlet state.92393 The room temperature absorption spectra of pyridine mero- cyanine (Dye I) in solvents of different polarities are shown in Figure (13). The absorption maximum in chloro- form occurs at 620 nm while in water the maximum occurs at l AAO nm, a blue Shift of 6500 cm- The half-band width of the absorption band increases with the polarity of the solvent. Table (I) shows the absorption maxima, Stokes' Shift A5 (cm-l) with respect to chloroform and the half band width (r l/2 cm-l) in some solvents. The fluores- cence spectrum of this dye undergoes a much smaller Rige- shlit as the solvent polarity is increased as shown in Figure (1A). These observations are consistent with a dipole moment decrease in the excited state. Methyl pyridinium iodide is another example which 76‘ .Hocmcpm u mowm one oofixomHSm szpoefim u omza .mpco>H0m pcoeom IMHU CH who mcficmmoonoe ocfiofinma mo mnpoodm coflpanmnm onsummoQEou Eoom .ma oesmflm IIIAES £o§o>o>>nlll 00» com com own com 09 00? Con _ q _ _ _ q _ l A .203 1 10m 8:5 1 <— aauquosqv 6333.5 77 Table 1. Optical Properties of Pyridine Merocyanine Dye in Various Solvents. Solvent rl/2 cm'l Amax’ nm AC, cm"1 Water 1900 AAO 6600 Ethanol 1500 508 3500 Dimethyl sulfoxide 1000 570 1AOO Chloroform A00 620 __-- Tl/2 is the width of the absorption band at half absorbance. A; (cm-l) is the shift of the absorption maximum with chloroform taken as reference. 78 3 '7. C Q) .‘s' 2 g m E m 8 L) 5 a L) A 2 LI. l 500 550 600 650 ——.Wovelength (nm)—-> Figure 1A. Room temperature fluorescence spectra of pyri- dine merocyanine dye in different polar sole vents. 79 undergoes a large blue shift ( 20000 cm-l) in polar sol- vents.9u The first absorption band corresponds to a charge transfer transition where 1’ acts as the electron donor and the pyridinium acts as the electron acceptor. In the ex- cited state, the intermolecular distance between the neutral component is increased. Dramatic relaxations involving the solvent cage as well as the two components solute will occur. In this category (“e < pg), we have studied in detail the pyridine merocyanine (Dye I). II. Molecular Systems that Undergo a Large Increase in Dipole Moment as a Result of Excitation - i.e., he > pg Examples 1. 2-Amino-7-Nitro Fluorene HN . N0 ANF 80 2. A,A' Amino Nitro Diphenyl H N N02 AND 95 The static dipole moment increases upon excitation to the first excited Singlet state by 18D for the case of ANF and by 12D for the case of AND. The room temperature absorption spectra of ANF and AND in solvents of different polarities are shown in Figures (15) and (16), respectively. The absorption maximum for ANF in 3MP (3-methyl pentane) occurs at 365 nm while in ethanol the maximum occurs at 385 nm, a red Shift of 1 1A00 cm- The red shift in the AND spectra is about 1 2300 cm- in going from 3MP to ethanol. The fluorescence spectra of these molecules undergo a larger red Shift as the solvent polarity increases as shown in Table 11.96 These observations are consistent with a dipole moment increase in the excited state. Both of these molecular systems were studied in detail. 81 .Hocoeom ASL ”muco>Hom econommfio CH mz< mo mnuoodm soapanomnm oeSpmnoQEop Eoom .mH osswfim 0000mm .sHHe Hocooeo asee>seom Amy .sHec oeoasomsaom AmV.Heo eeeeoaom Amv .m2m AHV AIIAEE 505.903 00m one 00v 0mm 0mm 0mm . _ — < l l l l 1 l 0.0 <—aouoq105qv— 0.0 82 .saae Honooao esca>seom Amy .Hoeoaom Any .saee oeoasomsaom Amv.aeo anemones ANS .azm AHV ”mp:o>aom psoLoMMHp :H 02¢ mo mhpooam coapanompm onSmeoQEou Boom .mH onswfim IAES 505.925 \ 80 0mm com one 00v 0mm 00» 0mm com ._fi...—...-_uq14_.1q—....A.._._...ad.un_4 e . . aauoqmsqv 1, 83 Table II. Fluorescence Properties of ANF (2—Amino—7- Nitrofluorene) in Various Solvents.9 Amax’ nm Solvent Absorption A0 (cm-l) Fluorescence A§(cm’l) Benzene 388 —-— 515 _-_- Diethylether 39A A00 532 620 l-Propanol A00 800 687 A800 Methanol 397 600 752 6100 8A III. Flexible Molecules that May Undergo a Change in Equilibrium Geometric Configuration Upon Excitation - Examples: 1. 9—tertbutylanthroate: (tertbutylester of 9- anthroic acid) 9-Tertbutyl Anthroate (9TBA) 2. 9-methylanthroate (methylester of 9—anthroic acid) 9-Methyl Anthroate (9MA) 85 These molecular systems undergo geometric relaxation to different equilibrium positions in the excited state depending on the size of the ester group, viscosity and polarity of the solvent matrix. The room temperature ab- sorption and fluorescence at room temperature and at 77°K of 9TBA and 9MA in 3MP are shown in Figures (17) and (18). respectively. The excitation wavelength dependence of the fluorescence Spectra in rigid media has been studied for 97 both anthroate esters. 86 .EZm cs AIIIIV x um um Mam A....V magpwpoqup Eoos pm mppooom oocoomonosam mufi pew ‘ Figure 22. FluoresCence Spectra of a dilute solution of pyridine merocyanine dye in l-propanol at 77 K as a function of excitation wavelengths. 98 P 069 089 0&9 099 099 099 _ 029 009 (N9 009 0th 099 _W J l O 03 Mysuaiul aauaosaiom :1 550 600 650 700 1———>. max. (nm)—-> 500 Fluorescence spectra of a dilute solution of pyridine merocyanine dye in l-hexanol at 77 K as a function of excitation wavelengths. Figure 23. 99 90— 5. L §§§§§§ e P \ er 9 E . Q) U E; - 3 22 - C) 2 “7 . oo- 111111111111111111111L] 450 500 550 600 650 —>\ mox(nm)—-> Figure 2A. Fluorescence spectra of pyridine merocyanine dye in PVA (polyvinyl alcohol) thin film at 77 K as a function of excitation wavelengths. 100 l 5607 E - v 540- S - In} T '5 520- 3‘: _ “5 .c 500” 3+ ‘2‘ . 2 9 480- := 3? N s . t m I ll 1 I 1 l 1 I 1 I 1 460 500 520 540 560 580 600 —Wovelengih of emission (nm)—-> Figure 25. Temperature dependence of fluorescence edge_ex¥ citation red Shift of pyridine merocyanine dye in ethanol. 101 Figure 26. Another example of viscosity dependence, the magnitude of REE in l-propanol at -108°C and in glycerol 1 in both solvents) proving the at 0°C are equal (2A00 cm- Significant role of viscosity in this phenomenon as Shown in Figure (27). Careful inspection of Figures (21—23) and (26) shows that the Shape of the fluorescence spectrum is highly dependent on the excitation wavelength. As the excitation wavelength is changed toward the red edge of the absorption band a shoulder is observed. The viscosity of the medium as well as the length of the alcohol chain affect the Shape of the spectrum. In a longer aliphatic chain alcohol solvent a more defined shoulder appears in the spectrum. We have also studied the REE as a function of the con— centration of the dye. In Figure (28) the emission of 10-5, A, 10‘3 and 2 x 10'3 M of the dye in ethanol at 77 K, 10' is shown; all solutions are excited at the same wavelength, namely the absorption maximum A60 nm. A red Shift of 21000 cm'1 is observed when the concentration is increased from 10'5 M to 2x10"3 M. lfimalOOO cm.l Shift represents the decrease in the magnitude of the REE as the concentration is increased, since no REE is observed for the concentrated solution (2 x 10—3 M) and its emission is identical to less concentrated solutions excited at the red edge. Besidesifluaobservation of a REE at low temperatures, we 102 Xexc. (om) Fluorescence Intensity l 0.0- 1J1114111111111Ll 500 - 550 600 650 ——>\ max (nm)--> Figure 26. Fluorescence spectra of a dilute solution of pyridine merocyanine dye in ethanol at -135°C as a function of excitation wavelengths. 103 g i °$° 5.3” glycerol i - Proponol I —'-— AREE Cm"——> A E? i l l. lOO 150 200 250 300 ——1"K——> Figure 27. Temperature variation of REE (Red Edge Effect) for pyridine merocyanine dye in two different solvents. 10A .AEC Dene x as am ECEmeE CoHpaCom lam onv CamcoHo>m3 COHpmpHoxo pome um x an em HOCmeo CH mCoHmep 1CooCoo pCoComme mo one oCHCmmooeoE oCHoHLHQ mo wepooow ooCoomoeosHm .wm oCCmHm \ IHES :Hocm_o>o>>II 00k 000 000 Can 000 On? _ H H H H H H H H H _ H H 1H H H H H H H _ 1H H H H H H 40.0 L .1 HI.— m 1 a x - m «U 1 M «a .. m. .M r/ 1 r a we a. a 1 m... .. Ha... _m 10m P b b - - - 105 have also observed a REE in rigid media at room tempera— ture. The results for the pyridine-merocyanine dye in PVA polymer matrix is shown in Figure (29). The magnitude of the REE at room temperature was about :A50 cm"l compared to .1200 om’1 at 77 K. The fluorescence spectrum of the dye undergoes a blue Shift of 2100 cm’1 as the temperature is lowered to 77 K. This observation is demonstrated in Figures (30, 31) for ethanol and glycerol solutions, respectively; all solu- tions were excited at the absorption maximum. The absorption spectra of the dye in ethanol (10'5 M) as a function of temperature are Shown in Figure (32). A blue shift of z2100 cm—1 in going from room temperature to 77 K is observed. Only a small (2500 om’l) blue Shift in absorption is observed for the dye in PVA thin film as shown in Figure (33). So far we have made the following experimental obser- vations: figst, the absence of a REE in fluid media; second, a large REE as the medium becomes rigid; tMlpg, the ap- pearance of Shoulders in the emission Spectrum when the media become more rigid or when working with long chain alcohol solvents; fourth, the appearance of REE at room temperature in polymer films; gigth, the gradual decrease of the REE as the concentration is increased; slgpg, the blue Shift of the emission spectrum as the temperature is low- ered, and seventh, the blue Shift of the absorption spectrum 106 90— g .. £§§§§E - "“\ Z" .7, .. .5 E . GI 8 I- Q) 8 e .. .3 U. H 0.0- 11l1111|1111|1111l11111 450 500 55a 600 650 —Wavelength (nm)—> Figure 29. Fluorescence Spectra of pyridine merocyanine dye in PVA (polyvinyl alcohol) thin film at room temperature as a function of excitation wave— lengths. 107 I I I I 90- - {IE5 .3 %’ i s. Q 3; g l- 2 E _ E r L) a) g F E . 00- 1 1 141 I 1 1 1 1 l 1 1 4,1 I 1 1 1 I I 1, 450 500 550 600 650 Wavelength (nm)——> Figure 30. Fluorescence spectra of (<10'5 M) pyridine mero— cyanine dye in ethanol at different tempera- tures. (The excitation wavelength was the ab— sorption maximum at each temperature.) 108 9O - >~ .— 15 _ C Q) .— E . m 8 ' r- 8 U) 2 .. O 2 u. . 0.0 - 450 Figure 31. —Wovelength (nm)—> Fluorescence spectra of a dilute solution of (<10-5 M) pyridine merocyanine dye in glycerol at different temperatures. (The excitation wavelength was the absorption maximum at each temperature.) 109 LO - "40 h -73 \\ -42 0.8 - § 0.6 - o 13 8 3 <); 0.4 - 0.2 - \. I 1 1 x 350 400 450 500 550 600 Wavelength (nm) Figure 32. Absorption spectra of pyridine merocyanine dye in ethanol («:10-5 M) as a function of tempera- tures. The temperature in °C is indicated on each spectrum. 110 Absorbonce 0.0 - 1 11 1 11 11 1 11 1] 1.11 1J 1e11 1 450 500 550 600 --Wovelength (nm)—- Figure 33. Absorption spectra of pyridine merocyanine dye in PVA (polyvinyl alcohol) thin film as a func- tion of temperature. \ 111 as the temperature is lowered. In order to interpret most of these observations we79 assume a statistical distribution of the solute among dif- ferent solvation sites. Thus, even in a homogeneous con- densed phase the polar solute molecules are expected to occupy a variety of solvation sites at any given time giving rise to different absorption energies corresponding to the same electronic transitions. A qualitative potential energy diagram representing the interaction of the dye molecule in its ground and excited states with a particular solvation site a is shown in Figure (3h). The interaction energy is assumed to depend on two coordinates, one representing the orientation of the polar solvent molecules with respect to the solute dipole symbolized by e. The other is a transla- tional coordinate, R, representing the intermolecular separation between the solute and the solvent molecules. The dependence of the interaction energy with respect to 6 and R if; drawn in two perpendicular planes. In the excited state, the magnitude of the interaction energy of the equilibrium configuration is smaller than that of the ground state reflecting a decrease in the dipole moment of the solute upon excitation. Excitation results in a sudden decrease of the solute dipole moment giving rise to a strained Franck-Condon state which will relax along the translational coordinate, R, to the equilibrium excited- state configuration followed by emission. Emission leads 112 ‘qk .-'7 1‘ .-j.°-°] : fifi'l P I l ‘\. ‘ :lfyflflflfl \ ‘V'l \, i/// .————m--— —-0 +,, Q/ / / ‘ I .3" ;/3}.'5 c b Figure 3“. Ground and excited state potential energy curves for a molecular system in which the dipole moment decreases as a result of excita- tion. Solvation sites, a, b, c and d differ in the orientation of the solvent molecules, 6, but are assumed to have the same inter- molecular separation R for simplicity. Solva- tion site a represents the most stable orien- tation. 113 to a sudden increase in the dipole moment of the solute giving rise to a strained Franck-Condon ground state which will relax along R to the equilibrium ground state. The solute may occupy different solvation sites differing in the orientation of the solvent molecules, 0, as well as the intermolecular separation R. In Figure (3“), we have repre— sented different solvation sites a, b, c and d having dif- ferent 6's and the same R for simplicity. Solvation site a represents the most favorable orientation with the largest interaction energy while site d represents the least favor- able orientation. In a fluid medium the lifetime of the various solvation sites is short compared to the excited- state lifetime and a dynamic equilibrium exist. One may think of viscosity-dependent barriers between these sites“3 the heights of which depend on the rigidity of the medium. Excitation of the solute dye molecule in site d will lead to a fast relaxation along the R coordinate as well as the a coordinate and emission originates from the equilibrium excited-state configuration corresponding to the most stable solvation site, i;e;, site a. Emission will occur at the same wavelength independent of the excitation wave- length. This accounts for the lack of REE in fluid media represented by the first observation where a dynamic equilibrium exists among the various solvation sites. If the medium is made rigid so that the relaxation rate constant along 6 and R is smaller than the fluorescence 11“ rate constant, emission will occur from a Franck-Condon excited state corresponding to the solvation site specific- ally excited. In rigid media the dynamic equilibrium between different solvation sites is lost, and excitation of the solute in a given solvation site will lead to an emission of the solute in that solvation site. In this case, an excitation wavelength dependence of the fluores- cence energy results, i.e., a REE is observed and is ex- perimentally observed in Figures (21-2u). The observation of shoulders in the emission spectra as shown in Figures (21-23 and 25) as the medium becomes rigid and as the alcohol chain becomes longer is explained in terms of inhomogeneous broadening and is similar to the Shpol'skii effect.80 The inhomogeneous broadening results from the statistical distribution of solute molecules among the different solvation sites that become static in rigid media. The presence of different local environments gives rise to the excitation wavelength dependence of the emis- sion and the appearance of shoulders in the emission spectra particularly in long—chain alcohols. The appearance of REE at room temperature in PVA polymer film, Figure (29) is explained in terms of a matrix that is rigid enough at room temperature to pre— vent complete equilibrium between different solvation sites. However the REE will be less than that observed at 77 K (Figure 24). 115 The decrease in the emission dependence upon excitation wavelength at higher concentration as shown in Figure (28) is anticipated as a consequence of singlet-singlet non- radiative transfer from molecular subclasses of higher transition energies to those of lower energy emissions. As a result, emission obtained by excitation into the middle of the chromophore absorption band becomes red—shifted at higher concentrations, which decreases the observed excit- ing wavelength dependence. In the high concentration limit it is expected that only long wavelength emission (cor— responding to solvated dye molecules with lowest transition energies) would be observed at all exciting wavelengths. The blue shift of the emission spectrum upon lowering the temperature (Figures 30 and 31) is explained in terms of the relaxation rates of the solvent. In fluid media, solvent-shell relaxation around solute molecules occurs fast enough before chromophore emission, lifiia emission will occur from an equilibrium excited state configuration. In a rigid media, solute molecules emit radiation long before solvent cage relaxation, i;g;,from a non-equilibrium excited state configuration, the strained Franck-Condon state. Hence emission will be blue shifted relative to the room temperature emission. The lifetime of the excited solute and the relaxation time of the solvent cage become comparable at intermediate temperatures, iggg, at inter- mediate viscosities. Hence, emission will occur from 116 intermediate states which will be red shifted relative to emission in rigid media, or blue shifted relative to room temperature (fluid) emission. The blue shift of the ab- sorption at low temperatures (Figures 32 and 33) is explained in terms of temperature—dependent inhomogeneous spectral broadening. Lowering the temperature will change the statistical distribution in favor of solute molecules occupying the most stable solvation sites which absorb at shorter wavelengths. This leads to an apparent blue— shift of the absorption as one lowers the temperature. B. Benzothiazole Merocyanine S 4. \ o“ I CH3 This system undergoes a decrease in dipole moment upon excitation. We have prepared this particular merocyanine dye and have measured its absorption and emission in sol- vents of different polarities. Its absorption spectra was already measured by McRae.10 Like that of pyridine merocyanine, its absorption spec- trum is pH dependent and it probably undergoes cis—trans iso- merization. The longest wavelength absorption band of the 117 dye in aqueous solution occurs at u95rm1in basic media. The acidic form of the dye absorb at “10 nm in H2O. Both forms are present at intermediate pH values. We have observed the disappearance of the color of the dye solutions after leaving them overnight, and another absorption at 335 nm was obtained. Adding acid generates the “10 nm band absorp- tion which gives the “95 nm absorption when excess base was added. Our experiments show that the 335 nm absorption band is a.result of chemical reaction of the dye with the base. Private communication with Professor Reusch106 helped in the elucidation of the basic hydrolysis of the dye at the C-S bond: Scheme I in water Acid Base 0" 1410 nn 118 Scheme II 0H acid 8 £2.) u\ "— CH 3 335 nm The photochemical reactions that occur in this system made it difficult to study its emission and optical properties Since many authors have worked with this dye, it seems im- portant to report our findings especially because similar dyes are used in membrane probe studies in aqueous media. 119 II. Molecular Systems That Undergo a Large Increase in Dipolg Moment as a Result of Electronic Excita- tion As an example of these molecular systems we have studied the absorption, emission and excitation wavelength de- pendence of the fluorescence of 2-amino-7-nitrofluorene (ANF) and A,A'-amino-nitrodiphenyl (AND). These molecules undergo a dipole moment increase92 by 18D and 12D, res- pectively upon electronic excitation. A. 2-Amino—7-nitrof1uorene (ANF) ANF The room temperature absorption spectra and emission maxima of ANF in solvents of different polarities are shown in Figure (15) and Table (II).92’96’107 The absorption spectrum of ANF exhibits a red shift in media of increasing polarity. The absorption maximum in 3MP (3-methylpentane) occurs at 365 nm while in ethanol the maximum occurs at 398 nm, a red shift of 22300 cm-1. The room temperature fluorescence spectra )92,96,l07 in Table (II show a much larger red shift as 120 the solvent polarity is increased. It undergoes =6000 cm-l red shift in going from benzene to methanol. These observa- tions are consistent with a dipole moment increase in the excited state. Earlier investigationscfl7ANF include the work of 92,107-109 who has made observations of the Lippert, et al. fluorescence shift as a function of temperature and sol- vent mixture. Their conclusions were that in viscous media (or alcoholic solvents at low temperatures) the fluorescence spectrum undergoes a blue-shift because the rate constant of dielectric relaxation becomes comparable to the fluorescence decay time. Thus, the fluorescence originates from progressively less relaxed states as molecu- lar motion is inhibited. The work of Topp, et al.96’110’lll indicatestflufi:the shape of the spectra and the fluores- cence quantum yields also depend on the solvent indicating the presence of some dynamic processes which are in the picosecond range. Thus, they used special techniques to obtain the fluorescence time profiles and emission spectra of the molecule in different states of solvation. This is similar to the nanosecond time resolved spectroscopy technique developed by Ware, et al.“8 for ANS and has been d96’llo’lll that the discussed in Chapter II. They showe Stokes'shift of ANF can be time—resolved in a picosecond laser experiment andgfives information about the rate of solvent polarization by the excited state dipole. They 121 have shown that this rate is faster in both 2-propanol and 0—dichlorobenzene than the calculated T2 relaxation time measured by microwave absorption. They also measured the time profile at different places in the fluorescence spectrum of ANF by varying the relative concentration of benzene and 2-propanol in mixed solutions. The delay pro- files measured at U90 and 680 nm were found to be exponen- tial. What they found also is that in hydrogen bonded solvents such as 2-propanol, the rate of dielectric polariza- tion is fast enough that at room temperature they could not resolve the shift from their pulse duration. However, when 2-propanol was diluted by benzene the rate of the shift decreased proportionately and the extent of the shift could no longertmeexplained by thetnflUcdielectric properties of the solvent. In conclusion, they showed that the en- vironmental Franck-Condon relaxation is complete before relaxation decay occurs, the differential in rates being about one order of magnitude. They proved that the actual position of the fluorescence spectrum observed by time- integrated measurements does not result from a competition between the two processes as suggested by Lippert.l The room temperature fluorescence of fluid dilute solutions of ANF in different polar solvents is hardly fluorescent. Our excitation wavelength dependence of the fluorescence was studied in rigid media at 77 K and in polymer matrices at 77 K and room temperature. Under 122 these circumstances we have observed a large red edge effect (REE) for ANF. The fluorescence spectra of dilute solutions of ANF in ethanol, PVA (poly(vinyl)alcohol) film and polystyrene film at 77 K as a function of excita- tion wavelength are shown in Figures (35—37), respectively. A large red shift of 2900 cm-1 in ethanol, 21300 cm'1 in 1 in polystyrene film was observed PVA film and 21600 cm- as the excitation wavelength was increased from the ab— sorption maximum at 77 K (the number at the top left of figures cited) to the red edge of the absorption band (the number at the top right of figures cited). Such observa- tions were accompanied by the appearance of shoulders in the emission spectra as the excitation wavelength was moved to the red. As in the case of pyridine merocyanine dye, the magnitude of this REE is viscosity dependent. It is decreased by warming up the solutions and it disappears in fluid media at room temperature. The magnitude of the l at room temperature compared REE in PVA film is z600 cm- to z1300 cm.1 at 77 K. Similarly the REE in polystyrene film at room temperature is 2850 cm_1 compared to 21600 cm-1 at 77 K. These observations are illustrated in Figures (38 and 39) respectively. The absorption spectra of ANF in different polar sol- vents undergo red shifts when the temperature is reduced. Figure (A0) shows the absorption spectra of dilute solu- tion of ANF in ethanol at different temperatures. A 123 .mnuwcmfim>m3 :ofipMpHoxm mo coapoczm a mm x um um Hocmnpo CH mz¢ mo soapSHom mpzafio m no mmuomom mocoomoLOSHm IAES 59.2263 00m Own 000 On? a q q 1 — u n J d u q — u d u q — q (1110) mo Y - l- h - .mm mesmfig 0.0 H n m a x a w a M a u S. .M om 12a .mcuwcmao>m3 coapmpfloxm mo cofipocsm m mm x we as EHHm cfinp ¢>m CH m2¢ mo mppomom mocoommpOSHm IIIAEE £82225 0mm 08 can con .._+...—q...d....- .wm ossmflm Ausuam aaueosaJonH 125 .mnpwcmam>m3 cofipmufioxm mo coapocsm m mm x up up Eaflm CH5» ocopzpmmfioo CH mz< mo mspooom mocoomoposfim .nm opzwfim IIAEE 50cm.m>o>>ll 00$ 000 can 000 09» — a u u d — q u q q — d u d d — u d a .- — d a 10.0 1 u n Au 1 .J .6 s a 1 w .d .. M 1 m. S \ .w... eono’rl, . .. v 9.!2 v\ 1 muwmmm m m. 1cm . . _ . . 126 \ .mnpwcoao>m3 coapmpfioxo mo :oauocsm m mm manpmstEop Eoog pm EHHM Gaza <>m cfl mz< mo wspooom cocoomoLOSHm 1.155 5829951! on» ook. 00m 00m , 08 con . _ . . q u _ .. fi,u . _ . . . a _ .. 019 .mm opswfim 0.0 Ausuawl eauaosaJon] :1 om www.m— (mu, any 127 90— E r 3§§§§ Fluorescence Intensity 0.0 - Wavelength (nm)—> Figure 39. Fluorescence spectra of ANF in polystyrene thin film at room temperature as a function of ex- citation wavelengths. 128 09 Absorbance l I l I l 350 400 450 500 550 —Wavelenglh (nm)—-> Figure H0. Absorption spectra of ANF in ethanol (3 x 10-5 M) as a function of temperature: (1) 24°C, (2) 0°C, (3) -36°C, (A) -78°C, (5) —1ou°c, (6) -115°c. 129 red shift of 21500 cm—1 was observed upon lowering the temperature to z-115°C. So far we have made the following experimental ob- servations: the absence of a REE in fluid media, a large REE in rigid solvents at 77 K and in polymer matrices at room temperature and 77 K, the appearance of shoulders in the emission spectra when exciting at the red edge, and the red shift of the absorption spectrum as the temperature is lowered. In comparing these observations with those in the case of pyridine merocyanine dye, one should recall that for ANF the dipole moment is increased in the excited state. A statistical distribution of the solute among different sol- vation sites is assumed. Thus even in a homogeneous con- densed phase, the polar ANF molecules are expected to occupy a variety of solvation sites at any given time, giving rise to different absorption energies corresponding to the same electronic transition. A qualitative potential energy diagram representing the interaction of the ANF molecule in its ground and excited states with a particu- lar solvation site a is shown in Figure (Al). In the excited state, the magnitude of the interaction energy of the equilibrium configuration is greater than that of the ground state,reflecting an increase in the dipole moment of the solute upon excitation. Excitation results in a sudden increase of the solute dipole moment, 130 Figure Al. Ground and excited state potential energy curves for a molecular system in which the dipole moment increases as a result of excitation. Salvation sites, a, b, c and d differ in the orientation of the solvent molecules, 6, but are assumed to have the same intermolecular separation R for simplicity. Salvation site a represents the most stable orientation. 131 giving rise to a strained Franck-Condon state which will relax along the translational coordinate, R, to the equilib- rium excited state configuration followed by emission. Emission leads to a sudden decrease in the dipole moment of the solute giving rise to a Strained Franck-Condon ground state which will relax along R to the equilibrium ground state. The solute may occupy different salvation sites differing in the orientation of the solvent mole- cules, 6, as well as the intermolecular separation R. In Figure (“1) we represented different salvation sites a, b, c, and d having different 6's and the same R for simplicity. Solvation site a represents the most favorable orientation with the largest interaction energy and the least transition energy while site d represents the least favorable orientation with the highest transition energy. In fluid medium the lifetime of the various solvation sites is short compared to the excited-state lifetime and a dynamic equilibrium exists. One may think of a viscosity- dependent barrieruu’u5 between these sites, the height of which depends on the rigidity of the medium. Excitation of the solute molecule in site d will lead to a fast relaxation along the R coordinates as well as the e coordinate and emission originatingiflwmithe equilibrium excited—state configuration corresponding to the most stable solvation site, 143;, site a. Emission will occur at the same wave- length independent of the excitation wavelength, i.e., no 132 REE is observed in fluid media. If the medium is made rigid such that the relaxation rate constant along 6 and R is smaller than the fluorescence rate constant, emission will occur from a Franck-Condon excited state corresponding to the salvation site specifically excited. In rigid media the dynamic equilibrium between different salvation sites is lost, and excitation of the solute in a given salvation site will lead to an emission of the solute in that salvation site. In this case, an excitation wavelength dependence of the fluorescence energy results, ELEL’ a REE is expected and is experimentally observed, of. Figures (35,37). The observation of shoulders in the emission spectra as shown in Figure (35) as the medium becomes rigid and as the wavelength of excitation move to the red are ex- plained in terms of inhomogeneous spectral broadening which results from the statistical distribution of solute mole- cules among the different solvation sites that become static in rigid media. The appearance of REE at room temperature in PVA poly- mer film and polystyrene film of Figures (38 and 39) is explained in terms of matrices that are rigid enough at room temperature to prevent complete equilibrium between different salvation sites. However, the REE is less than that observed at 77 K (Figures 36,37). The red shift of the absorption at low temperatures (Figure U0) is explained in terms of temperature—dependent 133 inhomogeneous spectral broadening. Lowering the temperature will change the statistical distribution in favor of solute molecules occupying the most stable solvation site which ab- sorb at longer wavelengths. This leads to an apparent red- shift of the absorption as one lowers the temperature. B. MLA'-Amino Nitro Diphenyl (AND) "2" N02 The room temperature absorption spectra of AND in differ- ent solvents are shown in Figure (16). Its absorption ex- hibits a red shift as the case of ANF in media of increasing polarity. The absorption maximum in 3MP (3-methyl—pentane) occurs at 339 nm while in ethanol the maximum occurs at 377 nm, a red shift of 33000 cm-1. The absorption spectra of AND in different solvents are blue shifted relative to the ANF absorption spectra in the same solvents as shown in Figures (16 and 15), respectively. Such observations are ex- pected from ground state geometries of these molecules. Con- jugation and hence more W-interaction is expected in the ground state of ANF relative to AND because of the carbon bridge in the former making it more planar. Thus ANF ab- sorbs at lower energy than AND in different solvents. In the excited state the two phenyl groups of AND will be at a smaller angle relative to each other and the Stokes shift of the emission is expected to be large even in 13M hydrocarbon solvents. However, the excited state dipole moment is increased and a large red shift in the emission is expected to occur as the polarity of the solvent is increased. Inspection of Figures (15 and 16) reveals the absorption spectrum of ANF is more resolved than that of AND. This is consistent with a planar rigid ANF and a flexible AND molecule. We have observed a large REE for AND in different polar solvents at 77 K as well as in polymer matrices at room temperature and at 77 K. The fluorescence spectra of dilute solutions of AND in ethanol, l-propanol, PVA polymer film and polystyrene polymer film at 77 K as a function of excitation wave- length are shown in Figures (A2-A5), respectively. A large red shift of 21200 cm.1 in ethanol, 2750 cm-1 in l- propanol, zlllOO cm'1 in PVA film and 21300 cm“1 in poly- styrene film were observed as the excitation wavelengths were increased from the absorption maximum at 77 K (the number at top left of figures cited) to the red edge of the absorption band (the number at the top right of figures cited). Such observations are accompanied by the appear- ance of shoulders in the emission spectra in ethanol as the excitation wavelengths were moved to the red or by changing the solvent to l-propanol. As is the case of ANF,the magni— tude of these REE are viscosity dependent. At -l26°C in ethanol, the REE is z200 cm—1 as shown in Figure (U6). 135 Xaxc (nm) 400 450 490 90 *- .>~ — 2: (D c h- 0) 4— C: - b 0) U C: _ Q) 8 93 _ C) 2 LL _ , 0.0 4 Figure 92. —Wave|englh (hm )—-> Fluorescence spectra of a dilute solution of AND in ethanol at 77 K as a function of excita- tion wavelengths. ' ‘ 136 .mzpwcoam>m3 coflpmpfioxo mo cofipocsm m mm x 55 pm Hocwoopola CH Qz< mo cofipsfiom opSHHp w no whoooom mocmowopozam I565 £32925 005 000 00$ 000 80 0.0? - d d d d - d 1 q u n J u I. a - a d d u q q a u d a III I I t z/. r 1 mmwmm m ., m .. p p n n - — .m: madman 0.0 Ausualul SOUGDSSJOUH 0m 137 . .mcpmcoam>m3 coapmpfimxo mo COHpoczm m mm x up no Eafim cflnp <>m CH Qz< mo appooam mocoommmosam IAES EocflgsslI 00m 00¢ Own 000 .qu..—-.q_.q.._... \J \\\. \ 1 \ x \ 1 x \. i \\ s \ 1 x s ~ 1 \\ t0». . 6mm"'V\ 1 “03“.". m .1 .2: ogswwm Ausualul aauaosaiamd 138 .mnpmcmao>m3 gowpmpfioxo mo coapwcsm m mm x we no EHHm GHQ» encampmmaoq CH 02¢ mo oppooam oocoomohosam I225 cficm_m>o§.ll 000 000 can con . 09V _ . _ .. . . 4. fl . . . . — . . 1 0.0 Ausuelul BDUSDSBJOHH \ .m: asswfim 139 .mzpmcoao>m3 cofipmpfloxm mo coapocsm m mm ooomal um Hocmzpo :fl 02¢ mo Cofipzaow opSHfip m mo oppooom mocmomohosam .w: mgsmfim IIIAES 50cm_m>os>Il 00b com 00m own con 09» u . . . . _ q . . . _ a .,.. q . . q . . _ . q a . q . Ausualul aauaosaionH mm (1110":an 1U0 The valuescfi‘AND REE's in PVA film and polystyrene film at room temperature are :500 cm"1 and 21400 cm'l, respec- tively. Their corresponding spectra are shown in Figures (A7 and U8). The absorption spectra of AND in different polar solvents undergo red shifts when the temperature is reduced. Figure (A9) shows the absorption spectra of a dilute solution of AND in ethanol at different temperatures. A red shift of 21500 cm.1 was observed upon lowering the temperature to -l20°C. Since AND undergoes an increase in dipole moment upon excitation and also has the same functional groups as ANF, one can explain the above experimental observations as in the case of ANF. III. Flexible Molecules That May Undergo a Change in Equilibrium Geometric Configuration Upon Excitation This kind of molecular system is exemplified by the methyl and tertbutylesters of 9-anthroic acid I and II whose ground and excited state geometries were studied earlier.97 Their room temperature absorption and emis- sion spectra together with their emission spectra at 77 K in 3MP (3-methylpentane),are shown in Figures (17 and 18) respectively. It was concluded that the ground state 1A1 .mnpwcmam>m3 coapmpwoxo go coapoczh m we angumgooEop 5009 pm EHHM cflcp <>m :fi Qz< mo mgpooam mocoomopoSHm IAES £82m>o>>nll Ooh CON Omm 00$ own 000 , . u d . dun-q-qufi-_-—- 006 069 out 02b (um) '38. y p l- P v- p .5: opzwfim AusualuI eaueasaionH Om 1A2 Fluorescence Intensity I Acne. (um) d \400 470 490 \ I 1 1 1 1 .I 1 1 L41 I 1 1i 1 L I LL 1 1 1 I Figure M8. 450 500 550 600 650 —fWavelength (nm)—-> Fluorescence spectra of AND in polystyrene thin film at room temperature as a function of ex- citation wavelengths. 1A3 -—Absorbance —-—- 0'0 - I J I J I 250 300 350 400 450 500 550 -——Wavelength (nm)—-> -A Figure “9. Absorption spectra of AND in ethanol (~10 M) as a function of temperature. (1) 2A°C, — (2) —23°C, (3) -86°C, (A) -120°c. iuu geometries of these molecules are similar to each other 9-Methyl Anthroate (9MA) 9-Tertbuty1 Anthroate (9TBA) with the carboxyl group out of plane, perhaps of 90° with respect to the anthracene ring, indicating no inter- action between the two moieties.112 A greater interaction between the carboxyl group and the ring in the excited state is expected. Such interactions depend on the size of the ester group, viscosity, structure and polarity of the solvent matrix. While a near-coplanar configuration between the carboxyl group and the ring in the excited state of 9MA is expected, a less planar configuration is expected in the excited state of 9TBA. It was shown97 that 9MA relaxes to its equilibrium excited state con- figuration even in a rigid glass hydrocarbon (3MP), while 9TBA does not relax under the same conditions. The 1A5 branched tertbutyl group is interlocked in the rigid sol— vent matrix while the methyl ester group can undergo some relaxation in the same solvent matrix. It should be noted that rigid parafin oil at -80°C does not prevent excited state relaxation of 9-anthroic esters. The structure of the solvent matrix around the anthroic ester molecule is such that there is enough free volume for the ester group to relax during the excited state lifetime. An excitation wavelength dependence of the fluorescence was carried out for 9MA and 9TBA solutions in a 3MP matrix element at 77 K. Figure (50) shows a red shift of $600 cm"1 when the excitation wavelength was increased from 385 nm to the far red edge (2397 nm) of the first absorp- tion band of 9TBA in 3MP. Exciting at the far red edge is accompanied with a highly resolved fluorescence spec- trum. The shifts observed in the case of 9MA in 3MP matrix were very small. As Figure (51) shows, a red shift of only 2200 cm’1 was observed when exciting at the far edge of the first absorption band of 9MA. Such shift was not ac- companied by more resolution as in the case of 9TBA. The above experimental observations were interpreted in terms of a red edge effect. The flexibility of these molecules allow the presence of different conformers which are unstable in fluid media. Upon freezing the solution, some of these conformers are trapped in certain configura- tions. Their lifetimes in the excited state will be 146 1 I I I l eo-jg £0008 " A $333 - 2 e . , ’ £2 £5 - 8 4 C: h I I g \ o .. 3 EC . 1 i 1 1 1 1 I 1 4 1 1 I 1 1 111 J 1 1 1 1 l 400 450 500 550 600 —Wavelength (nm)—> Figure 50. Fluorescence spectra of (7 x 10"5 M) 9TBA in 3MP at 77 K as a function of excitation wave- lengths. (Spectrum # l correspondstx>365 nm excitation and #u correspondstx>397 nm excita- tion.) 1H7 I l I I I 90- E S on - zxz§ E m - if» S I- d.) 8 - O) U (D g - E _ 0.0- 1 I 1 1 1 1 1 1 I 1 1 1 1 1 1 I 1 1 l 1 1 I 1 400 450 500 550 600 +Wawlength (nm)—> Figure 51. Fluorescence spectra of (7 x 10"5 M) 9MA in 3MP at 77K as functions of excitation wavelength. 1&8 dependent on the degree of the medium rigidity as well as the size of the ester group. Different conformers of 9TBA are being trapped in certain configurations which can't relax in their excited state, lLQL, they have a long lifetime compared to the fluorescence life time. The large tertbutyl ester group does not have enough free volume to rotate in the solid “3 "if a matrix. As proposed by Dellinger and Kasha, molecule upon excitation involves a rather large distor- tional motion for equilibrium relaxation, then a large volume of solvent in the cage around the solute must be displaced". 0ur interpretation is similar to the "Shpol'skii Effect"80 where heterogeneity in the local environment of rigid matrices gives rise to solute molecules absorbing at slightly different energies. Irradiation at the red edge of the first absorption band excites a subclass of molecules (conformers) which are more planar and hence absorb at lower energy than other solute molecules. The emission from this subclass will be red shifted relative to the bulk emission. Selective excitation produces a highly resolved emission spectrum result. However, the extent of resolution of the emis- sion spectrum of 9TBA in 3MP at 77 K did not change regularly with excitation wavelength. Careful inspec- tion of Figure (50) shows that the emission spectrum 1A9 excited at 390 nm was less resolved than the emission spectrum excited at 385 nm. More detailed results shown in Figure (52) confirm this observation. One must point out that it is only in the case of red-edge excitation that we select certain conformers that absorb at that energy. At shorter wavelengths one will excite a mixture of conformers, and the proportion in the mixture varies irregularly with excitation wavelength. 150 1 IA14J 1 II 1 l 1 I 1 I l 1 l 360 380 4CD 420 440 460 480 500 —Wavelenglh (nm)—> Figure 52. Variation of fluorescence intensity and resolu-' tion of vibronic band of (7 x 10"5 M) 9TBA in 3MP at 77 K at different excitation wavelengths which are: (l) 350 nm, (2) 355 nm, (3) 360 nm, (A) 365 nm, (5) 370 nm, (6) 375 nm, (7) 383 nm, (8) 386 nm. CHAPTER VI EXPERIMENTAL A. Materials I. Purification of Solvents l. Ethanol: 200 proof ethanol was fractionally distilled at a rate of five drops per minute. Portions of about 50 ml were collected and the absorption spectrum taken in a 10 cm cell to check for benzene. Distillation was continued until the characteristic benzene UV absorption was no longer ap- parent. Ethanol was then distilled and used as needed. 2. 3—methyl Pentane (3MP) The Philips pure grade 3MP was mixed with nitric acid and sulfuric acid, then stirred for two days. Then it was separated and washed with base and water. After drying with sodium sulfate for one night, the solution was refluxed over sodium wires for two days and distilled. The vapor passed through a four foot vacuum jacketed column and condensed at a speed of two drops per minute. The purity 151 152 was checked by obtaining the absorption spectrum. 3. Water was doubly distilled in our laboratory. A. l-Propanol, 1-butanol,gglycerin and DMSO (Dimethyl Sulfoxide) - from Fisher Scientific Company and chloroform from Mallinckrodt were used without further purification. 5. Paraffin Oil from Mc/B, Piperidine and Polystyrene from Aldrich Chemical Company, l-hexanol from Eastman and Poly(vinylalcohol) [PVA] from Polyscience, Inc. (Cat. No. 2815, 99-99.8 Mole %) were also used without further puri- fication. II. Preparation and Purification of Chemical Compounds 1. Merocyanine Dyes: l—Methyl-U—hydroxystyryl)peridinium betaine whose struc- ture is 153 was prepared using the general procedure described by lOO and others.lou’ll3 The merocyanine was puri- Brooker fied by repeated recrystallization in methanol and dried under vacuum. The benzothiazol merocyanine dye was syn- thesized in the same general procedure used by BrookerlOO and described in detail by Kampas.98 2. ANF(2-Amino—7-Nitro Fluorene) ANF was obtained from K & K Laboratories, Inc. and was purified by repeated recrystallization in ethanol. 3. AND (A—Amino-U'-Nitro Diphegyl) AND was obtained from K & K Laboratories, Inc. and was purified by repeated recrystallization in ethanol. A. 9MA and 9TBA (Methyl and Tertbutyl Esters of 9- Anthroic-Acid) 9MA and 9TBA were prepared from the purified 9-anthroic 11A acid according to the method of Parish and Stock and purified by repeated recrystallization from ethanol. 15“ III. Preparation of Polymer Films 1. Poly(vinylalcohol) PVA Films PVA (molecular weight l33,000-99% mole hydrolyzed Polyscience Cat. #2815) was combined with water in a ratio of 0.125 gram/ml. The mixture was heated over a steam bath and was stirred occasionally (to prevent dehydration at the surface for 20 minutes. The solution was then per- mitted to cool to room temperature in a moist environment (a desiccator with water in it works well). The dye solu- tion (1 ml of 10'“ M in ethanol for each 1.5 gm PVA) was added to the cool PVA water solution, mixed very well, and put back in a moistened environment for about one hour. The dye solution in PVA was then transferred into a small flexible centrifuge tube and centrifuged at the speed of 15,000 round per minute for about 10-15 minutes (this will remove any dehydrated polymer). The dye solution was then poured into clean, dry molds (molds were prepared by pouring melted paraffin in glass dishes to a depth of 8 mm. After solidification, A cm x A cm squares were cut out with a razor blade to form the molds). The films were then dried by directing air at them at a very slow rate for about 70 hours at room temperature. Then the films were removed from the molds and the needed measurements made. 155 2. Polystyrene Films A few billets of polystyrene were dissolved in the chloro- form solution of the dye. The contents were then poured into a flat dish and let to dry slowly in a clean atmos- phere. These films were very thin and had an absorbance of about 0.1. B. Spectral Measurements: 1. Absorption Spectra All reported absorption spectra were run on Cary 15 and Cary 17 spectrophotometers. 2. Emission Spectra Fluorescence spectra were recorded on a multicomponent system used in this laboratory consisting of a 500 W Xenon light source, 500 mm Bausch & Lomb excitation monochromator (which provides a narrow excitation band width), McPherson Model 235 Emission Monochromator and EMI Model 9558QB Phototube. Noise reduction and amplification of the PMT signal is achieved by using a Princeton Applied Research Model HR-8-LOCK IN Amplifier and appropriate chopping ap- paratus. Some of the room temperature emission spectra were recorded by using Aminco—Keirsard.Aminco-Bowman spectrofluorometers. The phosphorescence spectra were 156 obtained with these instruments equipped with a rotating can phosphoroscope. 3. Temperature Variation System A quartz dewar with flat quartz excitation and emission windows and a narrow glass tube for the samples were used. (A 0.1 cm absorption cell connected to a long pyrex tube was used for low temperature absorption). The temperature of the sample was controlled by boiling liquid nitrogen using a power resistor and allowing the N2 gas to flow into the sample dewar. The temperature of the sample was moni- tored through a thermocouple (copper, constantan) dipped inside the solution in the sample tube immediately above the point of excitation or absorption. The lowest stable temperature which can be reached by such a system is around -l60°C. Low temperature absorption or emission studies with polymer films were performed by hanging the films vertically in the dewar through a wire and a glass tube. For 77 K measurements the films were dipped directly in the liquid nitrogen. CHAPTER VII CONCLUSION AND FUTURE WORK One of the interesting applications of merocyanine dyes in particular and dyes in general is their extensive use in biology and photography. In biology, merocyanines have been found to be useful in measuring membrane potentials.115 Membrane potential sensitive dyes have been useful for studying rapid electrical activity in single nerve and muscle cells, in collections of individual cell bodies and in single cells in tissue culture. Dyes have also been used to determine changes in the membrane potential level of cells, organelles and vesicles in suspension and of single cells in conjunction with fluorescence activated cell sort- 116 used with lipid ers. Merocyanines have been recently bilayer membranes in which the trans-membrane potential can be rapidly and accurately controlled. Potential-dependent changes in the fluorescence and absorption spectra of these dyes were obtained along with the polarization dependence 117,118 of these properties. In photography a stable posi- tive image which is not thermally erasable below 70-80°C 119 was produced from merocyanines when irradiation of a recording paper took place. It leads to the reprography 157 158 of transparent documents and also produces continuous tone printing. This process requires no subsequent development, is dry, is an autoprocessor and can be realized with a relatively simple technology. Besides their importance in measuring membrane poten- tials and this application in photography, these dyes (merocyanines, ANF and AND) could be used as probes to study the dynamic behavior of liquids and could be po- tentially used in photography storage. Since these molecular systems (Chapter IV) are sensi- tive to polarity and structure of the solvent found from their absorption, emission and the significant role of REE, they could be used to further investigate the following important phenomena in the previous chapter: Shpol'skii effect, site selection spectroscopy, inhomogeneous spectral broadening, time dependent emission spectroscopy and viscous flow barriers. Some of the studies that could be suggested are now summarized. I. Pyridine Merocyanine: l. The room temperature emission Spectra of pyridine merocyanine undergoes a gradual red shift upon increasing its concentration in ethanol. Such observations in other 120-123 systems were interpreted in terms of energy migra— tion between molecules whose energy levels differ in position 159 as a result of inhomogeneous spectral broadening. Careful investigation of the concentration effect in the pyridine merocyanine dye is needed. 2. The REE phenomenon we observed requires further study. Thus polarization experiments at different excita- tion wavelengths at different temperatures in different solvents may reveal polarization enhancement at red edge excitation. Lifetime measurements in the picosecond range could be performed at room temperatures in different sol- vents and at different excitation wavelengths at low tem- peratures should be made. Excitation wavelength dependence could be performed using a tunable dye laser which will enable us to select solute molecules with the same sol- vent environment. II. ANF and AND l. The suggested experimental investigations of pyridine merocyanine should be extended to these dyes. 2. We have observed a large REE in the phosphores- cence spectra of these molecules in paraffin oil, poly- styrene and polyvinyl alcohol at low temperatures. De- tailed steady state and time-resolved study of the solva- tion of the solute in its triplet state would be very interesting. 160 III. 9MA and 9TBA It would be very interesting to study the lifetimes of 9MA and 9TBA in hydrocarbon at 77 K using different excitation wavelengths. REFERENCES 11. 12. 13. 1A. 15. 16. 17. REFERENCES N. S. Bayliss and E. G. McRae, 1; Phys. Chem., _8, 1002 (195“). N. G. Bakhshiev, "Luminescence of Crystals, Molecules and Solutions", Ed. F. William, Plenum Press, 1973, page 78. N. S. Bayliss, J. Chem. Phys. 18, 292 (1950). N. S. Bayliss and L. G. Rees, J. Chem. Phys., 8, 377 (19AO). G. C. Pimental, J. Am. Chem. Soc., 19, 3323 (1957). M. Kasha, Faraday Soc., 9, 1U (1950). G. J. Brealey and M. Kash, J. Am. Chem. Soc. _1, uu62 (1955). Y. Uoshika, J. Phys. Soc., Japan, 2, 59A (1954). E. G. McRae, J. Phys. Chem., 61, 562 (1957). E. G. McRae, Spectrochimica Acta, 12, 192 (1958). L. Onsager, J. Am. Chem. Soc., 58, 1A86 (1936). W. Liptay in "Optische Anregung Organisher'systeme", Verlag Chemie, Weinhelm (l9 6). J. G. Kirkwood, J. Chem. Phys. 1, 911 (1939). G. Oster and J. G. Kirkwood, J. Chem. Phys., 11, 175 (19u3). H. FrBhlich, Theory of Dielectric (Oxford University Press), London, 1950). R. A. Marcus, J. Chem. Phys. 13, 1261 (196“). (a) T. Abe, Bull. Chem. Soc., Japan, 38, 131A (1965). (b) T. Abe, Y. Amake, T. Nishioka and H. Azumi, Bull. Chem. Soc. Japan, 39, 8A5 (1966). 161 18. 19. 20. 21. 22. 23. 2A. 25. 26. 27. 28. 29. 30. 31. 32. 33- 3A. 162 O. E. Weigang and D. D. Wild, J. Chem. Phys., _1, 1180 (1962). M. F. Nicol, Appl. Spec. Rev., 8, 183 (197A). A. Balasubramanian and C. N. R. Rao, Spectrochimica Acta, 18, 1337 (1962). M. Ito, K. Inuzuka and S. Imanishi, J. Am. Chem. Soc., 81, 1317 (1960). S. Basu in "Advances in Quantum Chemistry" ed. P. O. Lowdin, Vol. 1 Acad. Press, N.Y., 196A. E. M. Kosower, J. Am. Chem. Soc., 88, 3253, 3261 (1958). G. Grunwald and S. Winstein, J. Am. Chem. Soc., 70, 8A6 (19A8). ’— (a) P. C. Dwivedi and C. N. Rao, Spectrochim. Acta, 26A, 1535 (1970). (b) C. N. Rao, Chem. Soc. Rev., 8, 297 (1976). M. J. Blandamen and M. F. Fox, Chem. Rev. 18, 59 (1970). G. Porter and P. Suppen, Trans. Farad. Soc., 81, 166A (1965). H. McConnel, J. Chem. Phys., 18, 700 (1952). A. Kuboyama, Bull. Chem. Soc., Japan, 31, 15A0 (196A). K. Yoshihara and D. R. Kearns, J. Chem. Phys., 88, 199 (1966). L. J. Andrews and R. M. Keefer, J. Am. Chem. Soc., 13. “500 (1952). J. A. A. Kerelaar, J. Phys. Rad., 18, 197 (195A). M. El-Aaser, Master Thesis, Alexandria University, Egypt, 1966. (a) G. Scheibe, Z. Elektrochem. 3A, A95 (1928); Z. Physik. Chem. (Leipzig)7BS 355—T1929). (b) N. Mataga and T. Kubota, "Molecular Interactions and Electronic Spectra", Marcel Dekker, Inc., New York, 1970, Chapter 6. 35. 36. 37. 38. 39. A0. A1. A2. A3. AA. A5. A6. A7. A8. A9. 50. 51. 52. 163 E. Rabinowitch, Rev. Mol. Phys., 18, 112 (19A2). J. Frank and G. Scheibe, Z. Physik. Chem., Al39,22 (1928). S. J. Strickler and M. Kasha, J. Chem. Phys., 3A, 1077 (1961). “— S. J. Stricker and M. Kasha, J. Am. Chem. Soc., 85, 2899 (1963). E. M. Kosower, J. Am. Chem. Soc., 89, 3253, 3261 (1958). E. M. Kosower, J. A. Skorcz W. M. Schwarz and J. W. Patton, J. Am. Chem. Soc., 8g, 2188 (1960). S. P. McGlynn, T. Azumi and M. Kinosnita, "Molecular Spectroscopy of Triplet State", Prentice-Hall, New Jersey (1969), p- 1999- M. Ashraf El-Bayoumi, J. Phys. Chem., 92 2259 (1976). M. Kasha, B. Dellinger, C. Brown,“Spectroscopy of the Solvent Cage. ”Generation and Characteristics of the Excited States, Florida State Univ., Preprint, 1980. B. Dellinger and M. Kasha, Chem. Phys. Lett., 38, A10 (1975). B. Dellinger and M. Kasha, Chem. Phys. Lett., 38, 9 (1976). M. A. Mohammadi and B. R. Henry, Proc. Nat. Acad. Sci., in press. P. K. Sengupta and M. Kasha, Chem. Phys. Lett., 88, 382 (1972). S. Chakrabarti and W. R. Ware, J. Chem. Phys., _8, 5A9A (1971). W. R. Ware, P. Chow and S. K. Lee, Chem. Phys. Lett. 3, 356 (1968). W. R. Ware, S. K. Lee, G. J. Brant and P. Chow, 1. Chem. Phys., EA, A729 (1971). T. Azumi, K. Itoh and H. Shiraishi, J. Chem. Phys., T. J. Chuay and K. B. Eisenthal, Chem. Phys. Lett., 11, 368 (1971). 53. 5A. 55. 56. 57. 58. 59. 60. 61. 62. 63. 6A. 65. 66. 67. 68. 69. 16A L. V. Levshin and D. M. Akbarova, J. Appl. Spec. (USSR) 3, 326 (1965). V. G. Bacharov and L. V. Levshin, Bull. Acad. Sci. (USSR) Phys. Series 27, 590 (1963). E. Lippert, W. Luder and H. Boos in "Advances in Molecular Spectrosc0py", Ed. by A. Mangini, Pergamon Press, Oxford (1972), p. AA3. E. Lippert, W. Lfider, E. Moll, W. Magele, H. Boos, H. Brigge and I. Seibold-Blankenstein, Angew. Chem., 3, 695 (1961). _- H. Zimmermann and N. Joop., Z. Electrochem., 65, 61 (1961). —_ H. V. Schutt and H. Zimmermann, Ber. Bensenges, Phys. Chem., 81, 5A (1963). P. Soon-Song and W. E. Kurtin, J. Am. Chem. Soc., 21. A892 (1969). N. Mataga, Y. Torihashi, K. Azumi, Theor. Chim. Acta, 8, 158 (196A). N. Mataga, Bull. Chem. Soc. Japan, 38, 65A (1963). S. Suzuki and H. Baba, Bull. Chem. Soc., Japan, 38, 2199 (1967). M. Suzuki, T. Fujii and K. Sato, Bull. Chem. Soc. Japan. 52, 1937 (1972). N. Mataga, T. Kubota, "Molecular Interactions and Electronic Spectra", Marcell Dekker, Inc., N. Y., 1970, Chapter 8. G. Weber and M. Shinitzky, Proc. Nat. Acad. Sci. US, 92, 823 (1970). K. Itoh and T. Azumi, Chem. Phys. Lett., 83, 395 (1973). G. Weber, Biochem. J., 75, 335 (1960). s. R. Anderson and G. Weber, Biochem., 8, 371 (1969). W. C. Galley and R. M. Purkey, Proc. Nat. Acad. Sci, USA, 81, 1116 (1970). 70. 71. 72. 73- 7A. 75. 76. 77. 78. 79. 80. 82. 83. 8A. 85. 86. 87. 165 K. R. Naqvi, J. Donatsch and U. P. Wild, Chem. Phys. Lett., 38, 285 (1975). B. Valeur and G. Weber, Chem. Phys. Lett., A5, 1A0 (1977). —— B. Valeur and G. Weber, J. Chem. Phys., 83, 2393 (1978). R. F. Chen, Anal. Biochem., 13, 37A (1967). A. V. Adamushko, I. M. Gulis, A. M. Rubinov, B. I. Stepanov and V. I. Tomin, Opt. Spektroskopiya, 88, 6A (1979). V. S. Paulovisch and L. G. Pikulik, Izv. Akad. Nank. 888R. 93,, 539 (1978). I. M. Gulis, A. I. Komyak and V. I. Tomin, Izv. Akad. Nank SSR (Ser. F. Z.) 83, 68 (1978). W. Klopffer, Chem. Phys. Lett., 11, A82 (1971). E. Leroy and H. Lami, Chem. Phys. Lett., 81, 373 (1976). K. A. Al-Hassan and M. A. El-Bayoumi, Chem. Phys. Lett, 18, 121 (1980). E. V. Shpolskii, A. A. Ilina and L. A. Klimova, Dokl. Akad. Nauk SSSR, 81, 935 (1952). ““ Shpolskii, USP. Fiz. Nauk, 11, 321 (1962). E. V. Shpolskii, USP. Fiz. Nauk, 11, 215 (1960). E. V. Shpolskii and T. N. Bolotnikova, Pure Applied Chem., 31, 183 (197A). E. V. Shpolskii, Zhurnal Prikladnoi Spektroskopii, J. C. Wright and M. J. Wirth, Anal. Chem., 31, 988A (1980). J. R. Maple, E. L. Wehry and G. Mamantov, Anal. Chem., 83, 920 (1980). B. E. Kohler in "Chemical and Biochemical Application of Lasers", Ed., 0. B. Moore, Acad. Press, New York, Vol. A, page 31 (1979). 88. 89. 90. 91. 92. 93- 9A. 95. 96. 97- 98. 99. 100. 101. 102. 103. 10A. 166 J. C. Brown, C. M. Edelson and G. J. Small, J. Anal. Chem., 88, 139A (1978). J. R. Maple and E. L. Wehry, J. Anal. Chem., 53, 266 (1981). —_ J. C. Brown, J. A. Duncanson and G. J. Small, J. Anal. Chem., 88, 1711 (1980). — Y. Yang, A. P. D'Silva, V. A. Fassel and M. Iles, Analyt. Chem., 88, 1350 (1980). E. Lippert, Z. Electrochem., 81, 962 (1957). F. J. Campas, Ph.D. Thesis, Stanford University, 197A. G. Briegleb, 88_81., Chem. Phys. Lett., 3, 1A6 (1969). H. H. Jaffe' and M. Orchin, "Theory and Application of Ultraviolet Spectroscopy", John Wiley and Sons, Inc. (1962). L. A. Hallidy and M. R. Topp, J. Phys. Chem., 83, 2A15 (1978). Khader Al-Hassan, Master Thesis, Michigan State University (1978). F. J. Campas, Ph.D. Thesis, Stanford University, (197A). U. Steiner, M. H. Abdel-Kader, P. Fischer and H. E. Kramer, J. Am. Chem. Soc., 100, 3190 (1978). L. G. S. Brooker, G. H. Keyes and D. W. Heseltine, J. Am. Chem. Soc., 73, 5350 (1951). S. Hunig, D. Rosenthal, J. Liebig, Ann. Chem., 592, 161 (1955). E. Lippert and F. Moll, Ber. Bensenges. Phys. Chem., 29.. 718 (19514). N. G. Bayliss and E. G. McRae, J. Am. Chem. Soc., 15, 5803 (1952). H. G. Benson and J. N..Murrell, J. Chem. Soc., Far. Trans., 8, 137 (1972). 105. 106. 107. 108. 109. 110. 111. 112. 113. 11A. 115. 116. 117. 118. 119. 120. 121. 167 F. J. Campas, Chem. Phys. Lett., 88, 33A (197A). Private communication with Dr. W. Reusch, Depart- ment of Chemistry, Michigan State University. B. Gronau, E. Lippert and W. Rapp, Ber. Phys. Chem., 18, A32 (1972). Bunsenges. E. (195A). Lippert and F. Moll, Z. E1ectrochem., 58, 718 E. Lippert in "Organic Molecular Photophysics", Vol. 2, J. B. Birks, Ed. Wiley, New York, N.Y. 1975, page 1. L. A8, A. Hallidy and M. R. Topp, Chem. Phys. Lett., A0 (1977). L. A. Hallidy and M. R. Topp, J. Phys. Chem., 82, 2273 (1978). T. 2005 (1969). B. u_7, 9 (1970). R. 21). L. 0. Werner and D. M. Hercules, J. Phys. Chem., 75, K. Tak and J. P. Saxena, J. Indian Chem. Suc., C. Parish and L. M. Stock, Tetrahedron Letters, 1285 (196A). M. Loew, S. Scully, L. Simpson and A. S. Waggoner, Nature, 281, A97 (1979). A. S. Waggoner, Ann. Rev. Biophys. Bioeng., 8, A7 (1979). R. Guglielmetti, J. Photographic Science, 88, 77 (197A). R. Steiger, R. Kitzing, R. Hagen and H. Stoeckli- Evans, J. M. Photographic Science, 88, 151 (197A). LeBaccon and R. Guglimetti, J. Photographic Sci. 21. 112 (1979). N. V. A2, L. Z. P. Senatorova, B. D. Ryzhikov, L. V. Levshin and D. Blazhin, Izv. Akad. Nauk SSSR. Ser. Fizich, 313 (1978). V. Levshin, A. M. Saletskii and V. Prik. Spect., 38, A1 (1980). I. Yuzhakov, 168 122. V. T. Koyava, V. I. Popechits and A. M. Sarzhevskii, Z. Prik. Spect., 38, 1023 (1980). 123. V. D. B1azhin, Z. Prik. Spect., 38, 667 (1979).