THESI‘; LIBRARIES MlCHlGAN STATE UNIVERSIT: EAST LANSlNG, MICH. 4882 This is to certify that the dissertation entitled Excited State Proton Transfer presented by Nahid Shabestary has been accepted towards fulfillment of the requirements for Ph . D. _ degree in Chemistry Major pr essor Date 17/97 (of .1?ij MS U i: an Affirmative Action/Equal Opportunity Institution 0- 12771 MSU LlBRARlES _:_. RETURNING MATERIALS: Place in book drop to remove this checkout from your record. FINES will be charged if book is returned after the date stamped below; fiEEG~ ~_. . i-_ - ._Wl EXCITED STATE PROTON TRANSFER By Nahid Shabestary A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1982 ]—_i , ABSTRACT EXCITED STATE PROTON TRANSFER By Nahid Shabestary Excited state proton transfer represents one of the ‘ Simplest photoreactions that is expected to occur when the acidity or the basicity of a functional group is enhanced upon excitation. Phenomenologically, excited state proton fluorescence band which has a large Stokes shift. For ex— ample, methyl salicylate in hydrocarbon solvents exhibits l l l transfer manifests itself usually by the appearance of a new I l : two fluorescence bands, one at 3A0 nm and the other with a l large Stokes shift of 10,000 cm"1 at A50 nm. The latter fluorescence band results from a tautomer produced in the excited singlet state as a result of an intramolecular pro— ton transfer. The methoxy derivatives which cannot under— go excited state proton transfer exhibit a normal fluores— cence, i.e., a band at N340 nm. The first chapter deals with examples of excited state proton transfer in solution, the solid phase and the vapor phase. Several aspects related to Nahid Shabestary the mechanism of excited state proton transfer are sum— marized. The question of whether the emission results are explained in terms of an excited state equilibrium or in terms of equilibrium between different ground state con— formers is discussed. We have studied the absorption and luminescence prop- erties of p-amino salicylic acid (PAS) and N,N dimethyl p- amino salicylic acid (DPAS) and their methyl esters (PASE, DPASE), under various conditions. Effects of the medium, pH and substituents helped in the interpretation of the ab- sorption spectra which are dominated by "charge-transfer transitions. In the "charge transfer" states of these molecules the hydroxyl and amino groups act as electron donors while the carboxyl group acts as an electron ac- ceptor. The emission spectra clearly reveal the occurrence of proton transfer from the hydroxyl group to the carbonyl oxygen during the lifetime of the first excited singlet state. In the case of DPAS the charge transfer state where the hydroxyl group acts as an electron donor (O)CT, is lower in energy than the (N)CT where the amino group acts as an electron donor. The N,N dimethyl amino group has a lower ionization potential than the NH2 group. Thus in the case of DPAS an energy level reversal occurs which has an important consequence on the emission spectrum of DPAS. A unique excitation wavelength dependence has been observed. In the (N)CT state, the basicity of the Nahid Shabestary oxygen of the carboxyl group is expected to be larger than that of the ground state, but the acidity of the hydroxyl group is not expected to change appreciably. In this case no proton transfer is expected for the hydroxyl group as a result of excitation of the (N)CT state. In contrast, for the excitation of the (O)CT state, both the acidity of the phenol group and the basicity of the oxygen of the car- boxyl group are expected to be increased relative to their values in the ground state. Experiments show that in the case of DPAS a longer excitation wavelength (330 nm) leads to emission at 354 nm while a shorter excitation wave- length (300 nm) gives rise to an emission at mA6O nm as well as at 35A nm. In the case of DPAS, we are dealing with excitation of two different electronic states of the same conformer. In contrast, the emission of MSA is varied as a result of excitation of different conformers which are at equilibrium in the ground state. All these results have been discussed in detail in Chapter II. Our studies Show that PAS also exhibits excited state proton transfer in the solid phase. The crystal structure of PAS shows intra- and intermolecular hydrogen bridges. Studies of l-azacarbazole in the solid phase, also shows that excited state double proton transfer accounts for the observed emission. The crystal structure,as de- termined by x-ray diffraction methods shows that the crystal contains cyclic dimers only. A summary of these results is given in Chapter III. Nahid Shabestary Chapter IV summarizes experimental procedures and puri— fication of solvents and compounds and Chapter V gives a brief conclusion with some suggestions for future work. TO Yasha ii ACKNOWLEGMENTS I wish to express my deep appreciation to my advisor, Professor M. Ashraf El-Bayoumi for his guidance, encouragement and friendship during the course of this study. I am grateful to professor James L. Dye. for presiding as my second reader and for the editorial assistance in improving my writing skill. I also thank Professor Andrew Timnick and Professor Michael W. Rathke. for serving on my committee. To Dr. Paul Hunter goes my thanks for being such a good friend and coworker throuhg my many terms of teaching. I would also like to thank Dr. D.W. Ward for his great help in defining crystal structure. Many thanks go to Mr. Ron Hass and Scott K. Sanderson for their expert assistance in electronics. Special thanks to my good friends and colleagues, Khader Ahmad Al-Hassan and Kamal Ismail. The author wishes to express her thanks to the chemistry department of MSU for the assistantships providing finacial support. The author is particularly grateful to her husband, Abbas, for his constant encouragement and understanding. iii Chapter LIST OF LIST OF CHAPTER I. TABLE OF CONTENTS TABLES. FIGURES I - EXCITED STATE PROTON-TRANSFER Introduction. Theoretical Models of Hydrogen—Bond Proton Transfer in the Ground State Hydrogen-Bonding in the Excited State Excited State Acidity and Basicity. Proton Transfer in the Excited State. A. Intramolecular Proton Transfer. Salicylic Acid and Its Derivatives. Salicylamide and Salicylanilide Naphthoic Acid and its Derivatives. Intermolecular Proton Transfer. Naphthol-Triethylamine Systems. 3-Hydroxypyrene-TEA System. 7-Azaindole Dimers (Double Proton Transfer) . . . . . . . . 7-Azaindole-Alcohol Complexes Kinetic Isotope Effects in Proton Transfer Reactions. . . . . . IsotOpe Effect in Excited State Proton-Transfer in 7-Azindole (7AI) Dimers. . . . . . . . . . . . . . . Proton Tunnelling iv Page viii l3 17 27 28 28 A7 55 6A 6A 71 72 78 82 85 88 Chapter CHAPTER Page Experimental Evidence in 7A1 Dimers . . 9l Excited State Proton Transfer in the Vapor Phase, Methyl Salicylate. . . . . . . 92 Excited State Proton Transfer in the Solid Phase, Acridine Derivatives . . . . . 97 II - EXCITED STATE PROTON TRANSFER IN P—AMINO SALICYLIC ACID DERIVATIVES . . 103 Introduction. . . . . . . . . . . . . . . . 103 Excited State Proton Transfer in p-Amino Salicylic Acid (PAS) and its Methyl Ester (PASE). . . . . . . . . . . . . . . . 105 l. Solvent Effect on the Absorption Spectra of PAS and PASE . . . . . . 11A 2. pH Effect on the Absorption Spectra of PAS and PASE . . . . . . . I20 3. Emission Spectra of PAS and PASE as a Function of pH at Room Temperature. . . 126 A. Fluorescence Spectra of PAS and PASE in Different Solvents at Room Temperature. . . . . . . . . . . . l3l Excited State Proton Transfer in Solid Phase of p-Amino Salicylic Acid (PAS) . . . 1A0 Excited State Proton Transfer in N,N- Dimethyl p-Amino Salicylic Acid (DPAS) and its Methyl Ester (DPASE). . . . . . . . 1A5 l. Solvent Effect on the Absorption Spectra of DPAS and DPASE . . . . . . . 1A5 2. pH Effect on the Absorption Spectra of DPAS and DPASE . . . . . . . . . . 1A8 3. Emission Spectra of DPAS and DPASE As a Function of pH at Room Tempera- ture. . . . . . . . . . . . . . . . . . 156 Chapter E. CHAPTER CHAPTER A. A. Fluorescence Spectra of DPAS and DPASE in Different Solvents at Room Temperature. A Unique Excitation Wavelength Dependence of Excited—State Proton—Transfer. . . III - EXCITED STATE DOUBLE PROTON TRANS- FER IN l—AZACARBAZOLE HYDROGEN- BONDED DIMERS, ITS ACETIC ACID COMPLEX AND IN THE SOLID PHASE. Introduction. 1. Excited State Double Proton Transfer in 1-Azacarbazole Hydrogen-bonded Dimers. 2. Phosphorescence Spectrum of l- Azacarbazole. 3. Excited State Double Proton Transfer in l-Azacarbazole-Acetic Acid Complex A. Excited State Double Proton Transfer in l—Azacarbazole in the Solid Phase. Crystalographic Data. Room Temperature UV Absorption and Fluorescence Spectra of l-Aza- carbazole in the Solid Phase. IV - EXPERIMENTAL Experimentally Studied Molecules. l. Para Amino Salicylic Acid (PAS) 2. Methyl Para Amino Salicylic Acid (PASE). . . . . . . . . . 3. N,N Dimethyl Para Amino Salicylic Acid (DPAS) . . . . . . . . . . . A. Methyl N,N Dimethyl Para Amino Salicylic ACid (DPASE). 5. l-Azacarbazole (lAC). vi Page 159 163 .170 170 171 181 183 186 186 188 193 193 193 193 193 193 19A Chapter 6. l. 7IIIIIIIIIIIIIIIIII---E_________________ Nl—Methyl 1—Azacarbazole Tautomer B. Solvents. 3—Methylpentane (3MP) Ethanol Ether Water Cyclohexane Hexane. C. Spectral Measurements l. 2. 3. A. 5. Absorption Spectra. Emission Spectra. Infrared Spectra. Mass Spectra.l NMR Spectra CHAPTER V - CONCLUSION AND FUTURE WORK. REFERENCES. V1]. Page 19A 195 195 196 196 196 196 196 197 197 197 198 198 198 199 20A Table LIST OF TABLES * pKa and pKa Values for Selected Molecules Room Temperature Absorption Maxima (nm) and Spectral Shifts (cm-l) Relative to 3MP for PAS in Different Solvents. Room Temperature Absorption Maxima (nm) and Spectral Shifts (cm-l) Relative to 3MP for PASE in Different Solvents. Room Temperature Absorption and Emis- sion Maxima for PAS at Different pH Values, Ground and Excited States Species Room Temperature Absorption and Emis- sion Maxima for PASE at Different pH Value, Ground and Excited State Species Emission Maxima in nm for PAS in Dif- ferent Solvents Emission Maxima in nm for PASE in Dif- ferent Solvents viii Page 19 53 116 120 121 122 13A 13A Table Page 9 Room Temperature Absorption Maxima (nm) and Spectral Shifts (cm‘l) Rela- tive to 3MP for DPAS Acid in Different Solvents. . . . . . . . . . . . . . . . . . 1A7 10 Room Temperature Absorption Maxima (nm) and Spectral Shifts (cm-1) Rela- tive to 3MP for DPASE in Different Solvents. . . . . . . . . . . . . . . . . . 1A7 11 Room Temperature Absorption and Emis- sion Maxima for DPAS at Different pH Value, Ground and Selected States Species . . . . . . . . . . . . . . . . . . 152 12 Room Temperature Absorption and Emission Maxima for DPASE at Different pH Value, Ground and Excited States Species . . . . . 155 13 Room Temperature Emission Maxima in nm for DPAS in Different Solvents. . . . . . . 159 1A Room Temperature Emission Maxima in nm for DPASE in Different Solvents . . . . . . 161 ix LIST OF FIGURES Figure Page 1 Absorption spectra of Schiff base (PLP 7 x 10’5 M, n butylamine 2.5 x lo-Ll M) in various solvents; di- oxane, AAA DMF, ———— H2O, the dotted line represents decomposition of this spectrum. 2 Potential energy diagrams for the motion of the proton in a hydrogen bond A—H...B. 3 Hypothetical potential energy curves for the formation of a H—bond A-H...B in Ground and excited state, (1) wg 16 >w (2) wg (b) (c) symmetrical single minimum unsymmetrical double minimum x >~ a 0‘ b L 0 0 C C 0 O I5 E E E 1—‘ :1 2 . 3 :fi g o V82 0 V=2C 7 ° Val a V=' v.0 V30 ' A R(A-B)- B A R(A-B)-> B i Figure 2. Potential energy diagrams for the motion of the proton in a hydrogen bond A—H...B. strength of the base.(22) The same concept was used to interpret the splitting of the first overtone of the NH2 symmetric stretching vibration band in a number of intra— molecularly H-bonded ortho-anilines.(23) Some of the basic criteria for the existence of a double-minimum po- tential in an H-bonded system, with the second minimum at about the u = 2 level, are (l) a splitting of the 002 but not the v band; (2) an increase in the magnitude of 01 the splitting with an increase in the strength of the H— bond; (3) a reduction of the splitting upon deuteration; (A) a temperature independent intensity ratio for the pair of bands resulting from the splitting of v02, since both transitions originate in the ground state. The band splitting is concentration independent in an intramolecu— larly H—bonded system, since the effect originates in the monomer. Some of the strongest evidence supporting a double- minimum potential comes from neutron diffraction studies. For example, neutron diffraction studies of KH2P0u show the proton as an elongated shape between the two oxygen A.(2u) This distribu— atoms which are separated by 2.A9 tion could be interpreted in either of two different ways: as a centrally located proton with a very anisotropic motion or, more likely as a disordered distribution of the protons between two possible positions, one on each side of the mid—point between the oxygen atoms. In the case of ice the 0-0 distance is much longer, being 2.76 A, and the two possible hydrogen positions are much farther apart (about 0.7A A). Because of this the disordered pro— tons appear as separate well-defined peaks. Other indirect evidence of double-minimum potentials in H-bonded systems (25) from (27,28) comes from dielectric saturation experiments, (26) data on compressibility, thermal expansion, the change in the bond energy,(29) change of intermolecu— lar distance upon deuteration<30> and NMR studies.(3l) At this point it should be mentioned that the inter— pretation of the above experiments is not as decisive as might be desired. For example even one of the best pieces of evidence, i.e., the split of the IR bands, could be due to an overtone of the AH bending mode, enhanced by Fermi resonance with vStr°(32) One of the early views of H—bonding was the concept of "mesohydric tautomerism" according to which the H was thought to resonate very rapidly between two equally prob— able positions, one near the donor atom, and the other near the acceptor atom. This concept would require that the potential energy function for the H—atom either passes two equal potential minima with a low barrier between the tWO (Figure 2a) or be symmetrical, with one broad mini— mum (Figure 2c). (33) Hadzi and collaborators have classified hydrogen bonds of the O -H...O type into four groups. In group A B 12 A are included the symmetric single minimum H—bonds; in group B, the symmetric double minimum H—bonds with small potential barrier, in group C, the symmetric double mini— mum H—bonds with a potential barrier higher than in group B; and in group D, the H—bonds with an asymmetric double minimum potential curve. The H—bonds of type A and B are uniformly very short, with R (0A...OB) varying from 2.A to 2.6 A. Some typical single minimum H—bonds are observed in KH maleate, NaH diacetate, KH dibenzoate, KH bisphenylacetate. The IR spectra are characterized by the —1 J absence of a v (O—H) band in the region above 1800 cm 01 the NMR signals are narrow and weak and remain unchanged at low temperature. The H-bonded compounds belonging to type B (KHEASOM, NH2H2POM, CaHPOu, NaHCOB) exhibit two O—H bands in the 1900—3000 cm—1 region, separated by 300-500 cm—l° 5 the PMR Signals are strong and narrow at room temperature and only slightly broader at —l80°C.(3”’35) Group C includes various carboxylic acid dimers and group D includes various phenols. A very interesting investigation, through ab initio techniques, of the double well picture of the hydrogen (36) bond was performed by Clementi et al. The study was on the DNA base pair Guanine—cytosine and the motivation (37) on the possible was the theory put forward by wadin importance of quantum mechanical tunnelling for the inter— conversion between different tautomeric forms. Although l r'l 13 there was a noticeable shoulder where the second minimum might have been expected no second minimum was found. Clementi and coworkers gave various possible sources of error in their prediction of the lack of a double minimum. The most important is that no simultaneous motion of two hydrogen bridges was considered. To test this possible error source, Clementi and coworkers carried out calcula— tions on the formic acid dimer<36> which has two hydrogen bonds. First only the motion of a single hydrogen bridge was considered. Using the Guanine—Cytosine basis set only a single minimum was found. When a larger basis set was used, a very pronounced shoulder was predicted but no minimum. It was only when the coupled motion of the two hydrogen bridges was considered and a double—zeta basis set was used that the calculation yielded a double minimum. A. Hydrogen-Bonding in the Excited State A molecule in its first excited singlet state has a different electronic distribution than the molecule in its ground state. If the chromophoric portion of the molecule is involved in H—bonding one would expect changes in its strength. In this way, electronic absorption and emission spectroscopy may provide experimental evidence about H—bonding. Kasha<38> pointed out that absorption bands corresponding to n + n* transitions are "blue shifted" in H—bonding media. This is due to the decreases in charge 1A density on the lone pair atom as a result of lone pair promotion. Thus, the H-bond is always stronger in the (39) ground state. Bayliss and McRae considered H-bonding interactions to be a special case of dipole-dipole inter- (A0) actions. However, Pimental pointed out that dipole induced- dipole and dipole-dipole interactions produce small solvent shifts compared with those due to H-bonding. He discussed the importance of H—bonding effects compared with other solvent effects and pointed out the role of the Franck-Condon principle in H-bonding. Solvent shifts due to H-bonding can be formulated as follows (See Figure 3): Av = v — 0 W - W + w Avf = vf - v = W - W - wg where: Av is the energy shift in the absorption maximum, Av is the energy shift in the emission (fluorescence) maximum, v is the energy of the absorption maximum, is the energy of the emission (fluorescence) maximum, v is the energy of the absorption maximum in the gas phase, W is the enthalpy contribution in the ground state from H-bonding interaction of the solvent with 15 the solute, W is the enthalpy contribution in the excited state from H—bonding interaction of the solvent with solute upon excitation; and w and Wg are the energies implied by the Franck—Condon principle and are always positive. Since we and WE are always positive, the shift in ab— sorption and emission will depend on whether Wg is greater or less than We. When Wg > We, i.e., the H—bond is stronger in the ground state than in the excited state as shown in Figure 3, a blue shift Ava > 0 which exceeds Wg—We by we will be observed in the absorption spectrum. Similarly, in emission a shift less than Wg—We by wg will be observed (either a red or blue shift depending on the electronic magnitudes of Wg—we and wg). When We > Wg, i.e., the H—bond is stronger in the excited state than in the ground state as shown in Figure 3, both absorption and emission spectra will Show a red Shift, which is less than Wg-We by We for absorption and is less than Wg—We by wg for emission. The well characterized H—bonds have energies in the range 1-7 Kcalmol‘l (350—2500 cm—l). According to the above discussion, a blue shift in absorption may exceed the ground state H—bonding energy. hence: the blue Shift ENERGY Figure 3. 16 R(A..B) Hypothetical potential energy curves for the formation of a H—bond A—H...B in ground and excited state, (1) Wg > We, (2) Wg < We' 17 is expected in the range of 350-2500 cm-1 or even larger than 2500 cm‘l. But a red shift in absorption should never be as large as Wg, so should not be larger than 2500 cm‘l. A red shift in the absorption spectrum indicates that the H—bonding is stronger in the excited state. This may indicate an increase in the acidity or basicity de— pending on the functional group at the chromophore in— volved in H-bonding. The spectra of nitrogen heterocyclic bases in which the lowest transition is n + n* have been found to undergo a red shift upon H—bonding. This is due to an increase in the charge density of the H-bonded nitrogen, which increases its basicity and in turn in- creases the H-bond energy in the excited state. 5. Excited State Acidity and Basicity In 1931, Weber which were interpreted by F6rster as a function of pH, shows that at both low and high pH the indole fluorescence is quenced (Figure A). In the high pH region the quenching is ascribed to the ionization of the pyrrolic hydrogen according to the reaction: @ . on (1’) H20 H From the midpoint of the titration curve one gets pK* b 12.3. Exactly the same number was obtained by E. vander Donckt.<7l) The excited state ng = 12.3 should be com— pared to the ground state pr = 16.97(72) which shows that indeed the pyrrolic hydrogen acidity has considerably increased in the excited state. 25 ‘ 4 Avon: 3.03 E 0.03: 4| . : masmwm (mun Mmqu) Kigsualm aouaosaionH 26 In the low pH region, the quenching is considered to be due to protonation of pyrrolic nitrogen according to the reaction: mlb : an“? ‘ H11 H - 1.8, in very good agree- In the same way one gets a pK: - 1.7 obtained by Bridges and ment with the value pK: Williams.(73) c. Photopotentiometry Photopotentiometry<7u> is based on the measurement of the potential developed between one illuminated electrode and one dark electrode in a solution. This potential, which is primarily a function of the various species in solution, was used to produce data leading to pK: values. The values found were —2.89 and 12.3 for 2-naphthy1amine cation and anion, respectively. These values compare reasonably well with the values obtained (—1.5 and 12.2)(u7) by using the Fdrster cycle. Other values obtained using this technique include, 13.5 for the 1-naphthylamine anion, 4.37 and 9.5 for 3—pyridinol anion and cation, respectively. 27 6. Proton Transfer in the Excited State In the case of ground state proton transfer in ground state "tautomerism", both the enol and the keto forms may exist in equilibrium in solution. Excited state proton transfer "phototautomerism", may also occur giving rise to a tautomer with an emission shifted with respect to that of the original species. The relative intensities of these two emission bands depend on the efficiency of the proton transfer process and the quantum yields of the original molecule and its tautomer. For example, it has (75) been found that the shift of the fluorescence maximum of B-naphthol in 06H6 brought about by the addition of tri- ethylamine (TEA) is quite large (A000 cm-l) compared to the shift of the 1Lb absorption band of B-naphthol caused by hydrogen bonding with TEA which is onlyfifllcm-l. These results suggested strongly that ion-pair formation, due to a complete proton transfer rather than just stronger H-bond formation, occurs in the fluorescent equilibrium state. Proton transfer may occur intramolecularly as in (76-78) salicylic acid or intermolecularly as in the 7- azaindole hydrogen bonded dimer.(79’80) A detailed dis- cussion of various examples of molecular systems that undergo excited state proton transfer will be given. a. Intramolecular Proton Transfer Salicylic Acid (SA) and Its Derivatives — Studies of excited state pK values reveal a general trend in which aromatic alcohols become more acidic, while aromatic ke- tones become more basic in their excited states.(8l) This gives rise to an interesting effect when both groups are present in the same molecule and are situated in ortho positions. In such cases an intramolecular hydrogen bond in the ground state is formed in non—hydrogen—bonding solvents. The enthalpy of formation of such a hydrogen— bond is estimated from infrared data to be in the range of 7—8 Kcal mole—ZL indicating a strong hydrogen—bond and a red shift of emission even larger than in aromatic mole— cules with just one functional group is observed due to intramolecular proton transfer. For example, in methyl salicylate the fluorescence maximum is red shifted by about 10,000 cm—1 from the absorption maximum, while in phenol the corresponding energy difference is about A000 cm—l.(76’82) Weller had observed that although the maxima of the long wavelength absorption bands of salicyclic acid (I) and 2—methoxy—benzoic acid (11) differ in wavelength by only about 1000 cm—1, the maximum of their fluorescence bands differ by 5000 cm.1 (See Figure 5). The fluorescence maximum occurs in the blue at A3A nm for I but that of the ether II occurs in the ultra—violet at 357 nm. The large St0kes shift observed in the case of salicylic acid 29 w l 8 f/fi' [3:0 S‘ [W , xr: Met/7000! 20 °C l * 5 * -—-—l———-1a --~-——+ ; E 7 "\ IQ ’3: / \ I u“ / \ l \ 05 +— / i / ;\ / . // I \\ I l x 0 w I l * 8 [(/\I}/[ g0 l 3"" A\0CH3 1r) Methanol 20°C l l ‘ 5 4 (a. l ,1 13 . e7 . a :“W’” 9 LP / \ Cu / \ l \ \ /’j \ 14’ 7 0 77 25 35 q 45 jail 9770’3cm'9—- Figure 5. Absorption and fluorescence spectra of salicylic acid an 2-methoxybenzoic acid in methanol at 20°c.(7 3O C|>H on on ocu3 I 11 (76,83) was interpreted as arising from a tautomer formed in the excited state via an intramolecular proton trans— fer. To further confirm these results Weller made similar observations for methyl salicylate (MSA) in aprotic solvents where two fluorescence bands occurred at room temperature. The one at 3A0 nm is simply the mirror image of the lowest near UV absorption band, while the second fluorescence is strongly red shifted (m10,000 cm-l) with a maximum near A50 nm and is much stronger than the other component at 3A0 nm (Figure 6). Proton transfer is driven by a pK (8A,85) change of about —6 for the phenolic oxygen and a pK change which may be as large as +8 for the carbonyl (65) oxygen of the carboxyl group. Methyl 2-methoxy-ben_ zoate with no transferable proton emits a single band at 320 nm with a normal Stokes shift. Furthermore, the emission of methyl salicylate in 0.1 M NaOH/CH3OH exhibits (86) a maximum at A16 nm which corresponds to the anion emission. Thus the A50 nm emission is assigned to a zwitterion formed via an intramolecular proton transfer in the excited state. 31 Rebbve quantuniintensfiytfi fluorescence 35 30 25 I 20 x 103 cm—1 VVavenunlber FiSure 6. Absorption of fluorescence spectra of MSA in methyl cyclohexane (solid lines) and fluorescence spectrum of methyl 2—methoxy benzoate in the same solvent dashed); fluorescence spectra after Weller.< 3) 9C“: pan c \ C\ // 9 \\Oe é—y o g H 0” on: 9 A = 3A0 nm )em = 3A0 nm em 32 Intramolecular proton transfers have also been re— ported for the salicylanilid I,(81387’88) salicylamide II,(88’89) phenyl salicylate 111,<90) and salicylaldehyde IV(91> ¢ \k cpo new»; H t OH OH I II ¢H\céo H\c,/O Go“ on III N Several studies designed to investigate the mechanism of excited state proton transfer led to many interesting findings. Quenching experiments using 082 (carbon di- SUlfide)<76) demonstrated that at room temperature the 3A0 and A50 nm emission of methylsalicylate were equally This led to the speculation that the two forms the quenched. of the excited molecule were in equilibrium, i.e., proton transfer reaction in the excited state reached equilibrium in a much shorter time than the lifetime of the excited molecules. This set a lower limit of 108 S—1 33 for the intramolecular proton transfer rate constant. The relative contribution of the two components to the total fluorescence was strongly dependent on temperature. As the temperature was lowered, the intensity ratio of the A50 nm to the 3A0 nm emission bands increased. From the temperature dependence of this ratio in hydrocarbon solvents, Weller(9) calculated a difference in enthalpy of about AH* = -1.0iO.5 Kcal mole"l (—0.7 Kcal mole—l)(92) between the neutral form and the zwitterion, and hence I ) = 5100 cm- he from the fluorescence spectral shift (AU deduced a ground state enthalpy change of 13.5i0.5 Kcal The activation energy for proton transfer has 1 mole—1 (9A) to be E i 2 Kcal mole_ according to: been estimated k = v exp (-E/RT) PT OH 1A where vOH is the stretching frequency and is equal to 10 secIl, k > 1/TO (>108 8-1) and T = 77°K where proton trans- fer is still observed. From A°K experiments an even smaller limit of E i 0.1 Kcal mole-l is derived. A tunnelling mechanism was suggested as the actual transfer mechan— ism.(9u) In acetonitrile, the ratio of the A50-3A0 nm emission is decreased.(92) This was interpreted in terms of a shift in the excited state equilibrium to form the neutral species. Because the neutral form has a large dipole moment in the excited state compared to the 3A zwitterion, it is stabilized to a large extent in aceto- nitrile.* This interpretation was modified as will be seen later. (86) Investigation by Klbpffer's group showed that in hydrogen bonding solvents (methanol, ethanol, water, and acetic acid) there are two strong emissions with maxima at A50 and 355 nm, whose relative intensities depend on the excitation wavelength (Figure 7). At excitation wave- lengths greater than about 310 nm, the A50 nm band domin— ates, while at excitation wavelengths below 310 nm, the 355 nm fluorescence is the stronger emission. This sug- gests that different ground state species are being ex- cited. The excitation spectra of the two bands are shown in Figure 8. The excitation maximum of the blue fluores- cence is at 310 nm, that of the UV component is at 300 nm. The excitation maximum at 3A9 nm is due to traces of the phenolate ion generated by ground state dissociation of methyl salicylate. The study of multiple fluorescence of 2,6-dihydroxy- (93) benzoate showed that the two emissions for methyl salicylate arise from species which are not in equilibrium in the excited state and which originate from different ground state species. In the case of 2,6-dihydroxybenzoate * This strongly suggested considerable intramolecular charge transfer in the excited state of salicylic ester. fluorescence intensity Figure 7. f 35 300nm ' _ _' 313nm 350 1.00 1.50 nm 500 wavelength Fluorescence spectra (uncorrected) of MSA, 10'“ mol 2’1 in methanol; excitation wavelengths and emission slit width indicated; spectral excita- tion slit width 20 nm.(86) 36 7 w IRE—‘7 a ' 2 2 .E s _, . C m a S O 2 l 400 250 300 . wavelength Fluorescence excitation spectra of methyl Figure 8. 10_u mole t—1 in methanol.(86) salicylate, 37 which cannot undergo intramolecular H—bonding through ro- tation around the ester-ring carbon-carbon bond, exhibits only short wavelength emission in polar solvents, but in non-polar solvents it emits only at longer wavelengths. If there were an equilibrium between the neutral and the zwitterion forms, one would have expected to observe both forms, especially in polar solvents in which methyl salicylate shows two emissions. 0 OH“ ° °OCH3 \ OCH3 \ O 0’" o/H Methyl Salicylate 2.6-Dihydroxybenzoote Sandros<9u> found that the ratio of the UV to blue emission was very sensitive to the hydrogen—bonding ability of the solvent and the excitation wavelength (Figure 9). Contrary to the earlier findings,(92) this ratio is much greater in alcohol solvents than in cyclohexane. The DOSitioncxfthe short wavelength emission is seen to be solvent dependent. The fluorescence maximum of methyl— 2—methoxylbenzoate shows similar solvent shifts. The excitation wavelength dependence results from exciting different proportions of the ground state conformer. 38 Figure 9. Fluorescence spectra of 2 x 10’“ mole cm—3 methyl salicylate in cyclohexane ( ), ethanol (— — —), methanol (————), and 2,2,2-trif1uoro— ethanol (....). The spectra are shown in rela— tive numbers of quanta vs wavenumber in um“ . The excitation wavelength is 286 nm. 9”) 39 Sandros proposed that a ground state equilibrium between cis and trans conformers rather than excited state equilib— rium is responsible for the relative intensities of the two fluorescence bands. OR o I I: ll C§9 __f__x c‘OR A *— o’ "6 <3 cis trons H The excitation of the cis form gives rise to a rapid and virtually complete proton transfer, while the trans form does not undergo proton transfer. (95) performed a kinetic study at 25°C Yasunaga gt a1. by using ultrasonic absorption techniques and obtained the rate constants for the interchange of the conformers: 1 kf = 9.5 x 105, kb = 2.6 x 107 s" , x = 0.0365 1 and AH = 2.5 Kcal mole- Thus, the cis-trans interconversion may be slow on the time scale of singlet state deactivation. The greater in— tensity of the short wavelength fluorescence band in H— bonding solvents can be attributed to stabilization of AO the trans ground state by hydrogen bonding with the sol— vent. In non—H bonding solvents, the cis conformer is the preferred species due to the strong intramolecular H— bond. An interpretation of the dual fluorescence of methyl salicy1ate(87) in terms of an emission from an intra— molecularly hydrogen-bonded species and from a species in which the hydrogen-bond is ruptured in the excited state is very unlikely. The energy difference between the UV and blue emission bands is too great to be accounted for in terms of hydrogen—bond breakage. Kaufmann(96) studied the picosecond time resolved fluorescence from MSA and SA in order to provide additional information on the mechanism of proton transfer in the excited state. Methylsalicylate in methylcyclohexane was (96) (8 psec pulse) and the excited with a 26A nm pic0pulse appearance of A50 nm fluorescence was measured with a streak camera. It was hoped that they could determine the rate of intramolecular proton transfer. The re— corded fluorescence signal was a convolution of the ap— paratus function with the fluorescence from the sample. Deconvolution of the data indicated that the transfer rate constant must be greater than 1011 sec-l. The data did not change even where the temperature was lowered to A°K. Replacing the proton with a dueteron and cooling (96) to A°K also did not alter the data. Therefore, they could not confirm or deny whether the proton transfer pro— ceeds via a tunnelling mechanism.(97’98) The lifetime measurement of methylsilicylate showed that the neutral excited state Species had a fluorescence lifetime in methylcyclohexane three times larger and in acetonitrile 10 times larger than that of the excited state zwitterion at room temperature. The rapid formation of A50 nm emis— sion and the very different lifetimes of the fluorescence components would also indicate that two species are formed immediately after the absorption of a photon and that they originate from different ground state molecules, since the excitation spectrum of the two emissions is different (Figure 10). These results indicate that the ground state equilibrium rather than the excited state equilibrium is responsible for the distribution of the fluorescence intensity between two wavelengths. The case of salicylic acid is more complicated than that of methyl salicylate since it can exist in several ionic forms in the ground state and excited state. In addition, it can form dimers. Methyl—5—ethoxysalicylate I also shows two f1uores~ cences,<92) but in this case the neutral form is favored in the excited state. The enthalpy has been determined to be AH* = +0.9 Kcal mole—l. Thus, increasing the temperature now favors the zwitterion. Time resolved studies are needed to unravel the dynamics of such ethoxy derivatives. A2 Intensity (orbllrory Ul'llIS) Figure 10. Excitation Spectra of methyl salicylate in methylcyclohexane. Solid dots represent the A50-hm band, while Open circles represent the 350 nm fluorescence. A3 0430‘ c,0 OH WW I Recently, a quenching study of methylsalicylate by carbon tetrachloride<9o> showed that the quenching of the short wavelength band is much more marked, the quenching of the long wavelength band being barely detectable (con— >176) trary to the previous result Moreover the short wavelength emission may contain contributions from two "slowly" interconverting ground state conformers that are not equally quenched. Preliminary fluorescence decay measurements using single photon counting appear to verify these results. A single exponential decay with a life- time of 131:10 PS<9O) was observed for the long wave- length emission band of methylsalicylate in methanol (in methylcyclohexane the corresponding lifetime is 280 PS).(96) Therefore there exist at least three distinct ground state conformers: I, II and III. In non-H—bonding solvents such as cyclohexane, we would expect 11 to be the major ground state precursor of short wavelength emission, due to stabilization by c 2’ "’ 9 i I cs9 ccqz c”: ego/C": o’H 0’" <3 H the intramolecular H-bond while I is expected to be the precursor of the zwitterion emission at longer wave- lengths. In methanol, however, there will be a stabiliza- tion of structure 111 due to H-bonding with the solvent. Structure III contributes a significant component to the short wavelength emission in methanol: such a component is more quenchable by CClu (lifii’ has a larger Stern—Volmer constant). Structure 11 contributes also to the short wavelength emission band. One should notice that the relative intensity of the short wavelength emission in- creases as the solvent proton's ability to form H-bonds in— creases. (99) Mori gt g1. concluded from IR data that structure (I) was the most stable; a fact confirmed in a later (100) study which used the CNDO/2 method and gave the cal- culated stability order I > 11 > III. The calculated ener- gy differences of (II) and (III) as compared with the most stable ones, (III), are 1.8 and 12.7 Kcal mole-l reSpectively- In another study using the INDO<101) method a difference in energy of 18.3 Kcal mole"l between 45 I (cis) and III (trans) has been obtained which suggests that a strong intramolecular H—bond must be present. The effect of temperature on the fluorescence spectrum of methylsalicylate in cyclohexane excited at 280 nm also has been studied by Ford §E_§l-(9O) (Figure ll). The essential feature is that a decline in the intensity of the long wavelength band as the temperature is raised is accompanied by an increase in the short wavelength band. Weller<92> had earlier observed a similar but apparently smaller effect. In contrast to the results of Ford et 3;. Smith and Kaufmann reported that the intensity of the short wavelength band remains substantially constant with temperature in methylcyclohexane. Ford et al. have noted one possible explanation for this discrepancy. The two fluorescence bands overlap appreciably, so that if spectra are recorded with a large emission monochromator slit width then the bands will be less well resolved and an apparent insensitivity to the relative intensities will occur. The detailed interpretation of temperature effects is quite complex, because the nonradiative rates for each of the excited state species and the ground state equilibria between the three conformers are temperature dGPGNdent- Moreover, the quantum yields of different emitting species are different. It appears certain however, that the short . . ' ' in wavelength emission intenSity increases With increas g ‘ t I I'll! . .. , Illlll I . ..| ..! I ill-Ill] >b.m2.w.»..2.-_. Figure ll. M6 INTENSHY 4bo ' sbo x/NM Temperature variation of the fluorescence spectrum of methyl salicylate in cyclohexane (1.4 x 10' M). As temperature increases the short wavelength emission intensity raises and the long wavelength intensity decreases. Tem— peratures (°C): 21.2, 29.6, 39.1, “8.3, 57.3, 67.5. 47 temperature indicating a shift in the ground state equilibria, since the rates of radiationless processes seldom decrease with increasing temperature. The pH dependence of the intramolecular proton trans— fer in salicylic acid was examined,(85) the following ionization scheme was postulated for the ground and ex— cited singlet state. Very similar behavior was found for salicylamide and salicylanilide.(102) Salicylamide and Salicylanilide — In contrast to methyl— salicylate in cyclohexane, salicylamide in cyclohexane shows only a single fluorescence band at mMSO nm assigned to the zwitterion species with a Stokes shift comparable to the long wavelength band for methylsalicylate (ca. 10150 cm-l) (Figure 12a). This behavior is similar to that of salicylanilide which also gives a single band at “69 nm(8l) in non polar solvents. In ethanol the fluores— cence of salicylanilide consists of a single band at #20 nm attributed to the phenolate anion form. Salicylamide ex— hibits two emission bands one at 342 nm due to the HGUtPal form and the other at U26 nm which is a composite band due to zwitterion and phenolate emission generated by exciting different ground state conformers (Figure 12b)- In water at pH 5.6 where salicylamide exists in its neutral form, The a Single fluorescence band at 420 nm is observed. . - ' ra fluorescence decay curve for salicylamide in wate 3'. and flair 148 Ground state a" 0‘ I C 0: So +H’ gcso +H’ - _H§ _ O O ‘ OH H pK(So) = 14 pK(So) = 3'0 (I)?! (IN-I gcso +3. gcsén on 41* OH Lowest singlet state a" 9“ C [:1 §O +H’ gc§0 +H’ 0. -H O- -H’ pK(S,) = 160 am a“ @Csan .u. aka. 0‘ 'H OH pK(S,) = 7'0 Ionization Scheme for Salicylic Acid85 “9 INTENSITY x3 X3 ‘ f i 350 WAVE LENGTH I NM ‘50 Figure 12a. Fluorescence spectra of salicylamide: (a) 2.U x 10.“ M in ethanol, Aex = 265 nm; (b) 1.3 x 10"5 M in cyclohexane, Aex = 300 nm; (0) 1.0 x 10'“ M in water, A = 265 nm.(89) 8X ..‘v4 uhv‘ 50 ). t a) 2 w .- g b a 350 WAVELENGTH I NM 450 Figure 12b. Effect of excitation wavelength on the fluor- escence spectrum of 2.5 x 10'5 M salicylamide in egganolz (a) Aex = 290 nm; (b) Aex - 310 nm 51 obtained by observing the whole fluorescence band from 400 nm upwards by picosecond spectroscopy, with a streak camera, is shown in Figure 13. The semi—log plot of the data (upper data points) clearly shows the nonexpon— ential nature of the decay. The most obvious explana— tion of these observations is that the observed decay is the result of the simultaneous decay of the two species with lifetimes of l.87i0.20 ns and 0.11:0.01 ns which have overlapping fluorescence spectra. The fluorescence decay curves have been recorded by observing different wave— length sections of the emission, isolated by suitable filters (Table 2). The observation of the two lifetimes indicates two separately emitting species, while the fluorescence and excitation spectra in water give no indication that this results from excitation of different ground state species. Similar results were obtained for acidified solutions. The possibility that one of the two decay components observed with a lifetime of 6 ns is due to salicylate anion(—COO') has been eliminated. However, the phenolate anion(—0‘) of salicylamide gives a single exponential with a lifetime of l.9i0.l2 ns which coincides with the long lifetime component of the decay curve. This sug- gests that the phenolate anion is a major contributor to the long wavelength emission band of aqueous salicylamide. Furthermore, the trend in the value of the fraction of 52 L 1202!- 23 5 ‘5 C) Q E 8910» U) 5 L 408r r Z_ 'i'f‘r-‘;_.-.._E.“::i,:__i‘-_"_;;_;, _ . BD 6.4 8.21 TIME (NSEC) Figure 13. Fluorescence decay curve of 10—3 M aqueous salicylamide (all emission above #00 nm) ex- cited at 265 nm; sweep speed 16 PS channel-1. The upper curve is a semi—log plot.( 9) 53 Table 2. Wavelength Fraction of Band Fast Component 400 nm 0.3M i 0.09 400 nm 0.59 i 0.05 >095 nm 0.80 i 0.02 short-lived component with emission wavelength indicates that the zwitterion fluorescence lies to the red of that of phenolate anion. The most plausible explanation of the results in water is that following excitation there is an immediate loss of a proton from the phenolate group. Some of these protons are transferred to the carbonyl group to give the zwitterion species. The fast decaying fluorescence can then be attributed to the zwitterion species, while the slow component arises from the excited phenolate anion. Whether the transfer of the proton lost from the phenolate group to the carbonyl group depends on the ground state conformation of salicylamide molecule in water is not clear, although this appears reasonable. The fluorescence Spectra for salicylamide in D O 2 (PD 0.5) at various excitation wavelengths are shown in Figure 10. In changing from H2O to D20, a large 51I INTENSITY 300 . 460 E03 ' 600 A / nm Figure 1“. Fluorescence of salicylamide in D20 (1.7 x 10'“ M). Excitation wavelengths (nm): (a) 330; (b) 330; (C) 265.(88) 55 increase in the intensity of the 355 nm emission band relative to that of the long wavelength band is observed. The zwitterion concentration remains constant on deutera- tion. The intramolecular proton transfer is expected to be so rapid that even when slowed by deuteration, the zwitterion yield is unaffected. The increased intensity of the neutral molecule emission makes it possible to record the excitation spectra of the neutral and zwitterion species (Figure 15). As expected, the ground state trans conformer of excited neutral salicylamide absorbs at shorter wavelengths than the cis ground state conformer, the latter can lead to excited state zwitterion formation. Naphthoic Acid and Its Derivatives - 3—Hydroxy-2— naphthoic acid (I) (87’103—105),and the corresponding methoxy derivative (11) have been investigated. COOH COOH OH OCH3 I II The infrared (and NMR) data suggested a relatively strong (81) intramolecular hydrogen bond. The ultraviolet absorp— tion spectra of these derivatives show this bond to be red shifted compared with the corresponding methoxy— derivatives which lack the hydrogen bond.(105'107) 56 b a ). t w z u: f... Z I 1 I fi 260 300 340 K/nm Figure 15. Fluorescence excitation spectra for salicyl— amide in D20. Emission wavelengths (nm): (a) 350; (b) H3O.(88> 57 The fluorescence spectrum of methyl-3—hydroxy-2— naphthoate is the subject of some controversy. A fluores— cence band at ca. H20 nm, corresponding to a normal Stokes shift (3200 cm‘l) of emission is observed. Naboiken(105) and co-workers have observed a weak, long wavelength emis- sion (ca. 650 nm) in hydrocarbon solvents in concentrated solutions where there is a noticeable quenching of the short wavelength fluorescence band. The interpretation of this band at 650 nm as being due to impurities, dimers or excimers was refuted;(105) instead it was attributed to an intramolecular proton transfer. The fluorescence decay of this band<108) gave a single exponential lifetime of 60e6 PS. " i The 650 nm band has not been observed by some workers<77’lou) in either the methyl ester or in the free acid. More recently, a band at 605 nm was observed (Figure 16). As we see in Figure 16 this band is more ap— parent in concentrated solutions and at longer excitation wavelengths. The fluorescence excitation spectra (Figure 17) of the short wavelength emission has two maxima and lies at shorter wavelength than the featureless excitation spectrum of the 608 nm emission band. Neither excitation spectrum bears a close relationship to the absorption spectrum, but this is hardly surprising, since the absorption spectrum is a superposition of the spectra of at least two distinct 58 a ?\ ll “ (\b \ t f) I Q I I If 1’ I g I \ I \ I \ I “ I \ I, X 'I \ I \\ I I I l I I ’I 400 r 560 ' 6730 700 A / nm. Figure 16. Fluorescence of methyl 3-hydroxy-2—naphthoate U (a) 1 x 10'5 M, (b) l x 10‘ M, 390.(108) in cyclohexane: Aex = 260 nm, (c) l x 10 M, Aex INTENSITY 59 300 Figure 17. 350 360 aéo 460 X/nm Fluorescence excitation spectra of methyl 3-hydroxy-2-naphthoate in cyclohexane (a) short wavelength emission and (b) long wave- length emission.(10 60 ground state species, as evidenced by the excitation spectra. The long wavelength emission can be attributed to a zwitterion species formed by intramolecular proton trans- fer in the first excited singlet state. The ground state precursor of the zwitterion is an intramolecularly H- bonded conformer, III. The anion of the ester emits at 0-5 - by 0 + (:40 ——'~ “OH I I III 0 OH 0/ \ OCH3 18,200 cm-1 (550 nm) in alcoholic NaOH, corresponding to an absorption maximum near 25,000(105) cm_l. The short wavelength emission band of methyl—3-hydroxy_ 2-naphthoate in cyclohexane is characterized by a non— exponential and concentration-dependent fluorescence A decay.(108) At a concentration of 10_ M, the lifetimes that best fit a double exponential decay function are 23.3:0.2 ns and 5.1i0.9 ns. As the concentration is increased, the longer of the two lifetimes is progressively ’/ 61 reduced. This observation suggests that there are two distinct contributions to this short wavelength band. Quenching evidence for a dual contribution to the short wavelength emission of methyl salicylate has been (90) given. The interpretation is in terms of two interconvertible ground-state conformers. Interconversion is sufficiently slow that it does not occur during the excited state life- time. The suggested conformers are ones in which the hydroxylic proton is H-bonded to the "ether type" oxygen of the carboxyl group (IV) and an "open-ring" form contain- ing no intramolecular H-bonds (V). 'I' O 0 \” OCH C’OCHS c’ 3 II II 0 O N Y The 23.3 ns lifetime can be attributed to conformer IV and the 5.1 ns lifetime to V. A calculation of the follow- ing equilibrium constant from absorption data (AH = 5- Kcal mole-l) yields a value of approximately 5 x 103. The equilibrium should therefore be almost entirely in favor of III. On this basis then, the absorption Spec- trum would bear a close similarity to the excitation spectrum of the long wavelength emission band which is 62 'I O OxH O ————s : (3" (”*3 ‘:¢5CI g l ocu-I3 Y III clearly not the case. McTigue(109) has shown that H—bonds involving ether oxygen and carbonyl oxygens are of similar strength for intermolecular H bonding between H20 and various organic ketones and ethers in 001” solution. Thus conformers IV and III will be of similar energy. The absorption spectrum is virtually independent of tem— perature between 293 and 3280K in cyclohexane and between 160 and 280°K in methylcyclohexane. This is consistent with conformers III and IV being energetically similar, with conformer V making up only a very small proportion of the ground state mixture. The study of 3-hydroxy—2—naphthoic acid(77) (free acid) shows that an anomalous long wavelength fluorescence appeared only in basic solvents or proton accepting sol— vents such as acetone, acetonitrile, and in aprotic sol— vents doped with small amounts of basic solvents, such as a mixture of toluene and pyridine,(77’lou) whereas in benzene only normal fluorescence with its maximum near 23700 cm—1 (420 nm) is observed. Two fluorescence bands have been observed, one near 2M000cm”l (416 nm) and the 63 other at 19000 cm"1 (526 nm). These two bands have been interpreted by Hirota in analogy to those of methyl salicyl— ate. From the temperature dependence of the relative in— tensity AH* = —3.1 Kcal mole"l has been calculated.(77) Recently the hypothesis of true intramolecular proton trans- fer in the excited state of 3—hydroxy—2-naphthoic acid l.who employed fluores— has been refuted by Ware gt cence quenching and singlet—state lifetime measurement techniques in a variety of solvents to show thatphotoauto— merization in the lowest excited singlet state involves intermolecular H-bonding with the solvent in the ground (10A) electronic state. According to Ware, pyridine acts as static quencher by forming a ground state com- plex. The proposed mechanism of the anomalous fluorescence .H- ’\ C 0 Ce : V— H?— 0’“ ' o’ o- of this acid consists in the formation of the H—bonded complex, followed by absorption of a photon and intra— molecular proton transfer. Thus the enthalpy (AH = -3,1 Kcal mole—l) measured by Hirota should be attributed to the enthalpy of formation of this complex. At high con- centrations of this acid complex spectral behavior is T——"— 64 observed due to association and to quenching by H+ (103) ions. The effect of pH on the absorption and fluorescence spectra of 3-hydroxy—2-naphthoic acid has been investi— gated.(110) Although the ground-state behavior was similar to that of salicylic acid, the molecules differed in the prototropic reactions of the S1 state, phototautomerism occurred only partially in the anion and not at all in the neutral molecules. Similar experimental evidence has been obtained for l—hydroxy—2-naphthoic acid and 2—hydroxy—l— naphthoic acid.(lll) In these cases the neutral molecules undergo intermolecular phototautomerism to form the zwitterion but, unlike 3—hydroxy—2—naphthoatethe anions do not react in this way. b. Intermolecular Excited State Proton Transfer Naphthol—Triethylamine Systems — Nagakura and Gouter- man have demonstrated that both u- and B—naphthol form quite strong H—bonds with triethylamine (TEA). The change of absorption caused by H—bonding in the case of the B—naphthcfl-TEA system in cyclohexane is shown in Figure 18. The ground and excited state equilibrium constants for this system in cyclohexane have been estimated to be 180 and 3300 respectively. The dipole moment of naphthol, its hydrogen-bonding ability and its acidity are greater in the first excited singlet state than in the 65 0.5- (I) (3) 04-- 0.3- 0.2- 0.!- m I I f F f r 300 310. 320 330 340 350 Tn Figure 18. Change of the absorption spectrum of B- naphthol in cyclohexane produced by added TEA (N150C). Concentration of B—naphthaol: 2 x 10‘“ mol 2-1. Concentration of TEA: (1) 0 mole 2'1, (2) 1.12 x 10‘2 mole 2‘1 (3) 1.80 x 10-1 mole £‘l.(75) 66 ground state (pKa= 9.1 and pK: = 2.8).(llu) These facts have been ascribed to the increased migration of n-elec— trons on the oxygen to the vacant w—MO's of the naphthalene ring in the excited electronic state.<113’115’116) In View of the above facts, one might expect proton transfer in the excited state of the naphthol—TEA system in non- polar solvents. The change in the fluorescence spectrum of B-naphthol caused by the addition of TEA in methylcyclohexane<92) is shown in Figure 19. Up to about 0.1 M TEA, the change undoubtedly is due to hydrogen-bond formation. However, the broadness of the spectrum observed at and above 0.1 M triethylamine, when practically all excited naphthol mole— cules are complexed, indicates that this spectrum is due to more than one emitting species. Comparison with the fluorescence spectrum in pure TEA suggests that the con— tact ion—pair (III) is the other component. The shift in the fluorescence spectrum due to ion—pair formation is about 5500 cmfll. The observed phenomenon is interpreted as due to a strong charge—transfer from the nitrogen non— bonding orbital to naphthol in the Franck—Condon excited state, which is followed by almost complete proton trans— fer from oxygen to nitrogen in the excited equilibrium state. Fluorescence quenching with oxygen showed that (92) both components are equally quenched. This indicates that the proton transfer equilibrium between II and III ID I- h'b\ 67 Ind Figure 19. Fluorescence spectra of B-napthol in methyl- cyclohexane at different concentrations of tri- ethylamine (1) 0.000 M; (2) 0.002 M; (3) 0.00M M, (A) 0.008 M; (5) 0.020 M; (6) 0.10 and 0.15 M; (7) as solvent.(9 I'——7— 68 is fully established in a time shorter than the lifetime of the excited molecules. 1.5 + ArOH + NB + Ar*OH...NR Ar*O-...HNR 3 3 “H 03* W (l) I 11 III Lowering of the temperature, however, changes the fluorescence spectrum of the H-bond complex in favor of the ion—pair III, as shown in Figure 20. From the intensity ratio at separate frequencies (22,500 cm"1 and 28,000 cm-l) the equilibrium constant, K* has been found 23 to be 1.5 and from the temperature dependence of K53, the enthalpy, AHE‘3 of the excited state equilibrium is -0.9 Kcal mole—l. The interaction of B-napthol in the excited state with triethylamine was studied by measuring nanosecond time— resolved fluorescence spectra and by analyzing the decay fluorescence bands due to the free and complexed proton doners. The time—resolved fluorescence spectrum has a new band at 010 nm in addition to the band of the free B—naphthol molecule at 355 nm. The former corresponds to the band found in the steady—excitation fluorescence spectrum of the B—naphthol-TEA-benzene system and ascribed to ion-pair formation by Mataga and Kaifu.(75’ll7) This band decays more rapidly than the band at 355 nm (Figure 21).(118) (119) Using the Stern—Volmer equation, the K V '5’" 69 I I l Methylcyclohexane I-0 - ‘7 0-5 — 0 20 25 30 v003mnfi) FiSure 20. Fluorescence spectra of B—naphthol in methyl— cyclohexane +0.1“ M triethylamine at dif— ferent temperatures.(92 / >«.m..Cqu«C_ QUCWUmuachD ~ 1 , x 7O 0 'a °o 3 o 3 O '0 Intensity Fluorescence '400' 450 7((nm) Figure 21. Fluorescence Spectra of B-naphthol under various conditions. The time-resolved spectra of the B—naphthol—TEA system in toluene at 25°C ob— served 0 ns curve (a) and 10 ns (curve b) after the fluorescence of the B—naphthol reached the maximum intensity. The concentrations of B- naphthol and TEA are 1.2x10'u M and 0.705x10-2 M respectively. Curve 0 is the steady state a fluorescence spectrum of the tolgene solution of B—naphthol (0.705x10‘” ) ll ) 71 lifetime of B-naphthol and the rate constant for complex formation in the excited state were determined at 30°C to be 11.6 ns and (2.17i0.01) x 109 M-ls-l, respectively. From the temperature dependence of the rate constant an activation energy Ea and a frequency factor A were de- termined as follows: 1.6u¢0.11 Kcal mo1e"l U1 ll 10 —lS—l M 11> II (3.05i0.12) X 10 This Ea value is comparable to the activation energy for the diffusion of toluene as determined from the temperature coefficient of its viscosity.(120) The rate constants ob- tained from analysis of the fluorescence quantum when they examined the fluorescence spectrum of 7—azaindo1e (7AI) in 3-methylpentane (3MP) as a func— tion of concentration. It was known from earlier work by El—Bayoumifim op ompmznom zaflswspflnpw ope: mfimcwfiw one pouwspcoocOo esp pom com: me: coapwpfioxm oommLSm pcopm Im>m3 coflumpfloxm .AIII-V z muoa x o.H use A .conoL am one CH mpfimcmp .mHQEwm .s: mmm summed v z muoe x o.H em AZm ea oaoucfimmmus mo Apcwfipv oocmomoposam pew Apmoav cofipapomnmlmpzpmpodsop Eoom :27: Ikozw-_w>§§ 0mm. 000 On. v 00v Om m a a d \\ . a 7/1 ex / \ .t x, x C I x mm /, ~ \ C I m x m“ ,/ \ R / \ O z m“ ,x c. // \\ F. //l\\ N” T .A .L F. R .mm messes (E- 0| " I NOLLDNLLXB HV'IOW 76 tunnelling). By studying the effect of temperature on the relative intensities of the normal violet fluores— cence (F1) and the green tautomer fluorescence (F2) in 3MP, a value of 1.4 Kcalmole"l (500 cm‘l) was obtained as the potential energy barrier for double proton trans- fer. It was found, however, that the relative integrated intensities of the tautomer and dimer fluorescence are the same at 77 and H.2°K, If the double proton transfer is a thermal process in which the two proton move over the energy barrier, then by assuming an Arrhenius behavior with a pre~exponential factor of 1012 _ 1013 sec—1, characteristic of unimolecular reactions, an energy barrier of the order 20 - 30 cm-1 for the reaction at H.2°K is obtained.<80) Thus at this temperature thermal energy is not sufficient to allow ap— preciable transfer over the barrier. However, substantial F2 is observed. Clearly these observations pointed strongly to quantum mechanical tunnelling. At higher temperature the proton transfer can take place by both tunnelling and thermal activation in which the two protons move over the energy barrier from high vibrational levels populated ac- cording to a Boltzmann distribution, while at very low temperature tunnelling may be the only reaction pathway. 77 In order to obtain further support for the occurrence of proton tunnelling at low temperature, a sample of 7AI in which the hydrogen atom at the N1 position was replaced by deuterium was used.(79) At 77°K where proton transfer over the barrier is presumably negligible, there is a large effect on the F2/Fl ratio. It was noted that deuterium substitution enhances Fl at the expense of F2 indicating that proton transfer is less efficient in the deuterated sample. The ratio of proton tunnelling rates for the deuterated and nondeuterated compounds was 2.9. h e kPT _ F2 h b1 d —— — (—-) (-—) kd F1 F2 PT Recently Bulska and Chodkouska have studied the effect of temperature on the F2/Fl ratio by exciting the dimer specifically and they found that the ratio F2/Fl approaches zero at low temperatures where the phosphor— escence was not yet observable. At low temperature the tautomeric fluorescence overlaps the isoelectronic mono— meric phosphorescence. From these results they concluded that a tunnelling mechanism is not required to account for the observed experimental effects. The role of proton tunnelling in excited—state proton transfer is still an open question that requires further study. 78 The dynamics of double proton transfer in the excited state of the 7AI hydrogen—bonded dimer has been studied<80> in our lab by nanosecond time resolved spectroscopy, the rate constants for proton transfer in dueterated 7AI (Nl-D-7AI) at 77°K are k? = 1.9 x 108 sec-1 xi = 1.2 x 107 see‘1 A picosecond study of the 7AI dimer by Eisentha1(128’129) shows dependence of the rate constant for excited state proton transfer upon wavelength which offers an explana— tion of the relatively slow risetime (A ns) for the tauto- mer fluorescence observed in dueterated 7AI at 77°K. 7—Azaindole—Alcohol Complexes — The fluorescence spectra of 7AI in alcohols are composed of two bands com— pletely analogous to those in hydrocarbon solvents.(12u) Besides the normal fluorescence band of 7AI, Fl’ another broad (excimer like) fluorescence, F2, appears with a maximum at A50 nm (Figure 23). In the case of alcoholic solvents, the relative intensities of F1 and F2 depend only on temperature and not on concentration and excita— tion wavelength. The F2 fluorescence in alcohols is not due to excimer formation since the F2/Fl intensity ratio is independent of concentration. >~.U:roo:. to your. vast-sou... s 79 Fluorescence Intensity K I I I I I 3“) 350 400 450 500 550 600 650 A(nm) Figure 23. Corrected room temperature fluorescence spectra of 7AI in EtOH ( ) and 7AI in EtOD ( ————— ). 80 Taylor gt gt. attributed the F2 band to a tautomer resulting from a double proton transfer in the excited state between 7A1 and an alcohol molecule in an analogous way to the tautomerization of H-bonded 7—azaindole dimers in hydrocarbon solvents. The tautomer is the same in both cases and only the mechanism of its production is dif— ferent. k / I \ __f_. \ \ ‘—— \ 5‘?! 1‘ kb \Wq !! : | 5 fix I“ H'- [H Fl ('5 If IL"'2 Et fit The effect of pH, temperature and solvent deutera- tion on the fluorescence spectra and quantum yields of 7A1 and other model compounds in ethanol gives some evi- dence supporting the above mechanism of excited state pro— ton transfer between a 7AI molecule and an alcohol molecule. One may visualize two kinds of complexes between 7AI and alcohol: a 1:2 complex (I) or a cyclic 1:1 com— plex (II).(131) When small amounts of ethanol are added to a dilute (1.0 x 10_u M) solution of 7AI in 3MP at room temperature, 81 ’I \ ’I \ ‘\ ‘\ ~ 7 .. 1,1 o/H H (moi; é ’0—C2H5 I ”52 H CH the absorption spectra are perturbed in the same way as during 7AI dimerization (red shifts). The absorption at 310 nm is due to primarily to the complex and can be used to determine the stoichiometry and the association constant of the complex. The results confirm the assump- tion of a 1:1 complex and an equilibrium constant of 49 Mn1 is obtained. The association constant of the al— cohol complex in 3MP is much smaller than the self-associa— tion constant of 7AI in 3MP which is 2.5 x 102 M‘1 at 25°C. This has been attributed<79> to cooperative effects present in the self-dimer and is absent in the case of the 7AI— ethanol complex. 82 c. Kinetic Isotope Effects in Proton Transfer Re— actions Kinetic isotope effects are particularly common in the case of hydrogen isotope effects, partly because hydrogen is involved in so many reactions, and partly because such effects are much larger for hydrogen atoms than for the isotOpes of heavier atoms. It is interesting to note that the rate differences between reactions of hydrogen and deuterium compounds are sometimes so large that the use of deuterium compounds has been proposed as a practical means for slowing down harmful reactions. The starting point for most discussions of isotope effects is the implication of the Born—Oppenheimer approxi— mation that nuclear substitution leads to no appreciable change in electronic energy. Of the other factors lead— ing to energy differences between molecules it is a feature of primary hydrogen isotope effects that zero—point energy changes are usually much more important than differences in isotopic partition functions. The formalism of the kinetic isotope effect has been given by many workers.(l32) Within the framework of transition—state theory, kinetic isotope effects may be expressed in terms of differences and ratios of isotopic zero point energies and partition functions between reactants and the transition state, i.e., kH/kD = MMI-EXC-TUN°epr(A€R — A€#)/2KT} 83 where MMI (masses and moments of inertia) and EXC (vi- brational partition function) include reactant and transi— tion—state partition functions, and A55 is the zero-point energy difference between transition states containing H and D. TUN is the 'correction' for quantum—mechanical tunnelling through the energy barrier to reaction. There are two qualitative manifestations of the importance of zero—point energy changes in the primary isotope effect. The first is the wide applicability of Swain's relation— ship> kD, the preexponential factor is near unity. The factor that determines the isotope effect is then just the difference 84 between the zero point energies. The substitution of D for H is an elegant probe to use in the elucidation of chemical dynamics. In particular the solvent isotope effect is a powerful technique for investigating the role of the solvent. When H20 is re- placed by D2O it is inevitable that we are dealing with the effects of substitution at many different sites. The effects of these substitutions can be described by fractionation factor theory(l39) and, because many sites are involved, more information can be found if measure— ments are made in equimolar mixtures of H20 and D20 as well O and in pure D20.(140) The two forms of (1A1—1A3) as in pure H2 water differ in structure and differences exist as evident from differences in thermodynamic prOperties of the two liquids, in the heats of hydration and solu— (144,145) bilities of salts in the two forms of water. Swain and Bader 1, corresponding to a net increase in structure breaking (net decrease in librational frequencies) on going from reactants to transi- tion state, while effects <1 are expected for ion—destroy— ing reactions, and effects close to unity are expected for reactions which do not produce or destroy ions. Solvation isotope effects will generally be small, probably no more than 20—30% except for small ions which can have a con— siderable effect on the water structure (librational fre- quencies). Isotgpe Effects in Excited State Proton Transfer in 7- Azaindole Dimers - Let us consider the kinetic scheme of double proton transfer: kTD .1 (2T)* k = k + k «7 «:\ / \ .T D + th 2T + th 86 The kinetic description of the system under conditions of continuous illumination is as follows: d[D*]/dT = IO - (kD + kTD)[D*] + kDTI(2T)*l (1) d[(2T)*]/dT = kTD[D*l - (kT + kDT)[(2T)*l Steady state conditions give d[D*]/dT = d[(2T)*]/dT = 0. By using the definition for quantum yields ¢FD = kFD[D*]/IO and ¢FT = kFT[(2T)*]/IO (2) and since after correction for different instrumental sensitivity in the two different spectral regions F /F ¢FT/¢FD = 2 1 (3) we obtain from Equations 1, 2 and 3 F2/Fl = (kFT[(2T)*]/(kFD[D*]) = (kFTkTD)/) SO (F2/F1)H/(F2/F1)D = kQD/k D (kgT+k$>/(kgT+k$) (u) since isotopic substitution is not expected to change the 87 radiative lifetimes. Further, since kT + kDT z kT and . _ h . d Since kT — kFT + knr’ kT is expected to be equal to kT except in the improbable case of a very large and specific effect of one H/D atom on the non-radiative rates. Under these assumptions equation 4 can be Simplified to (F /F > /(F /F > = kh /kd (5) 2 l H 2 l D TD TD The experimental isotope effect measured as in Equation 5, as the ratio of F2/Fl for protonated and deuterated 7- azaindole in 3MP matrix at 77°K is 2.9. So, h . d _ (kTD/kTD>77OK ‘ 2'9 Using the result for k%D(8.6 x 107 s_% one can obtain ng = 2.5 x 108 S-l.(80) Further, as we discussed earlier, this kinetic isotope effect is connected to the zero—point energy difference (ZPE). So, kh /xd = exp(—ZPE/RT) TD TD where h -d _ kTD/KTD — 2.9. The zps thus obtained is around 60 cm’i. If the 88 activation energy for double proton transfer is 500 cm-1 then the activation energy for the double deuteron trans- fer is around 560 cm_l. This is in good agreement with the estimated activation energy of 500-600 cm-1 consider- ing the double deuteron transfer at 77°K as a thermal process. It appears that a mechanism that involves mo- tion of the deuterons over the potential barrier can ex- plain the experimental data at 77°K because of the very small energy barrier. d. Proton Tunnelling One of the most unusual and interesting results of quantum mechanics is the prediction of tunnelling, i.e., the ability of a particle to exist in, or pass through, a region of space where its total energy is less than its potential energy. According to classical mechanics, such a phenomenon is impossible. As early as 1927(1UB) Hund discussed the probability of intramolecular rearrange— (149) ments via tunnelling and Wigner in 1932 discussed tunnelling with a view aimed at chemical kinetics. A tunnelling mechanism was prOposed for the doubling of (150) When proton certain Spectral bands of ammonia. transfer is very fast, such as in acid-base reactions in aqueous solutions, there exists the possibility that the proton in the incipient H-bonded complex A-A...B tunnels through the potential barrier between the two minima. 89 This possibility was pointed out by Bell in 1935. Since then, a great deal of indirect evidence has been accumu- lated which indicates the occurrence of proton tunnelling (151,152) (153) in ultrafast proton transfer. Johnston (154) and Caldin have provided reviews of various aspects of proton tunnelling in ordinary chemical reactions, and the significance of tunnelling to an understanding of the hydrogen-bond has been well described.<155_157) L6wdin has discussed tunnelling as a possible mechanism of causing mutation and genetic error. The JWKB method was used to obtain the transmission co— efficient for an arbitrarily shaped potential. The JWKB solution has an oscillatory behavior in the "permitted" region and an exponential behavior in the forbidden region. In ordertmiderive a formula for the transmission co- efficient g (the probability of getting through the bar- rier) in general, we will consider a single incident wave hitting a barrier with two turning points (Figure 24). If the barrier is parabolic at its top (Figure 24), one can obtain for KO, K = (1/4w2fi/ha )(2m(V —E ))1/2 0 o 2 o The factor w/4 comes here from the shape of the barrier. If one measures the energy from the top as a fraction k of V0, so that E = V2 — kVO one obtains 9O Figure 24. Model barrier for proton tunnelling calcula— tion. 91 and 1/2 09 ll exp(-kfi2/hao)(2m(V2-EO)) Experimental Evidence in 7—Azaindole Dimers — By studying the effect of temperature on the fluorescence Spectra of 7AI under conditions of continuous illumination a value of 500 cm—1 was obtained<79) as the potential energy barrier for double proton transfer. It was found, however, that the relative integrated intensities of the tautomer and dimer fluorescence are the same at 77°K and 4.2°K, i.e., (kTD)77°K= (kTD)u.2QKii'one makes the reason— able assumption that the radiative and nonradiative (k ) rates do not change from 77 to 4.2°K. If the ID,kIT double proton transfer is a thermal process where the two protons move over the energy barrier then by assuming an Arrhenius behavior with pre—exponential factors of 1012 — 10l3 sec—l, characteristic of unimolecular reac- -l tions we obtain energy barriers of the order of 20-30 cm for the reaction at 4.2K. Clearly these observations point strongly to quantum mechanical tunnelling. At higher temperatures the proton transfer can take place by both thermally activated and tunnelling mechanisms, while at very low temperature tunnelling may be the only 92 reaction path. The effects of deuterium substitution of the pyrrolic hydrogen on the magnitude of F2/Fl ratio at 77°K are consistent with current theories of proton tun- nelling. However, recent studieél27) of the effect of temperature on F2/Fl ratio appears to indicate that the proton-tunnelling mechanism is not necessary to account for the observed results. Further studies are required to examine the role of tunnelling in excited state proton transfer. e. Excited State Proton Transfer in the Vapor Phase Excited state proton transfer in the vapor phase was observed for the case of methyl salicylate for N-heterocycles containing intramolecular H—bonds, to be the order of 1010s.1 and by a picosecond study for methyl salicylate in hydrocarbon sol— lls_l.(96) Proton transfer vent, to have a lower limit of 10 therefor effectively competes with other deactivation pro- cesses. The UV fluorescent species emit at shorter wavelengths than observed for the solvated form (355 nm), these species 9... (:§§c> <:§() I; :E;:__. (l) O ’ ('3 H IflSUKCL IISARD 96 are most probably the open structure without a hydrogen bond. The equilibrium constant for this reaction oc- curring in the electronic ground state is given by: K = [MSAOJ/[MSAC] Fluorescence obtained at different temperatures, using 290 nm excitation, are shown in Figure 26. Increasing the temperature strongly increases the short wavelength fluorescence intensity, whereas the blue fluorescence turned out to be nearly temperature independent. Lowering the temperature after heating again yields the spectrum characteristic of the lower temperature, only after an extended irradiation time do new peaks appear which can probably be attributed to irreversible photochemical reactions. The insert in Figure 26 shows that the intensity ratio of short—to—long wavelength fluorescence decreases exponentially with l/T. The slope of the semilogarithmic plot in Figure 26 gives the enthalpy of the reaction (1). This quantity is found to be 3.59 Kcalmole-l and corresponds to the energy needed for breaking the intramolecular hy— drogen—bond. It is substantially smaller than the value estimated for the strength of the H—bond in MSA based on Spectral shifts caused by C=O...H-O in stretching vibra_ -l>(161) tion bands of the 0:0 group (6.7 Kcalmole and in much better agreement with the results of a measurement 97 using ultrasonic relaxation in liquid MSA (2.5 Kcalmole— , K = 0.0365).<95) Since the entropy change, AS, of reac— tion (1) is most probably very small, we can estimate the equilibrium constant K to be 0.0025 at 25°C. F. Excited State Proton Transfer in Acridine Derivatives — In 1946, a reversible color and fluorescence spectral change were detected in acridine at —l80°C. The acridine was present as a minor constituent in a vacuum-sublimed crystalline organic acid such as oxalic, terphthalic, or succinic. The transition from the spectrum of the acridine cation to the spectrum of the neutral acridine molecule under illumination indicates clearly that a transfer of a proton occurred from the acridine cation back to the acid anion as a result of electronic excitation. The study of the fluorescence and absorption spectra of acridine and its derivatives at different pH values in aqueous media showed that acridine has a higher affinity for the proton in the excited state than in the ground state. These results were in contrast to the observation that the phototransfer of a proton occurs from an acridine cation to an acid anion. This may be attributed to the fact that the conditions were different; the acridine was con— tained in a crystalline acid at low temperature, and not in aqueous solution. Further investigation on other compounds and media, 98 condensed together at low temperatures in a solid matrix gave similar results intra and inter— molecular hydrogen bridges. We measured the fluores- cence spectrum of PAS in the solid phase. Information from the solution spectra of PAS helped us to interpret the emission band in the solid phase of PAS in terms of an excited state intramolecular proton transfer. At the beginning of the chapter the spectrum of PAS is related to parent and related molecules, ttgt, benzene, phenol and salicylic acid. B. Excited State Proton Transfer in P—Amino Salicylic Acid (PAS) and its Methyl Ester (PASE) Electronic absorption spectra of substituted benzenes are usually compared with the spectrum of benzene itself. This suggests that the differences in Spectra of substi— tuted benzenes reflect the perturbation effects of the various substituents. Such a picture is valid if or- bitals are localized either on benzene or on the sub- stituent and if no charge migration occurs between the substituent and benzene. However in order to interpret certain intense transitions which occur in the spectra of substituted benzenes but do not correspond to any benzene transition (B2u (lLb), Blu (lLa), E2u (1B), charge—trans— fer between the substituent and benzene orbitals must be considered. One may classify an excited electronic state 106 of a substituted molecule as either locally excited (LE) state or a charge—transfer (CT) state only as a zeroth— order description of the electronic state, since mixing between LE and CT states occur. A locally excited state involves an excitation between different orbitals of the hydrocarbon, or between dif- ferent orbitals of the substituent. Transitions involving LE states will appear in the substituted molecule at dif— ferent energies than the corresponding transition of the unsubstituted hydrocarbon. Their intensities are also different due to the influence of the perturbing potential of the substituent. A charge—transfer state involves the transfer of an electron between the hydrocarbon and the substituent. A CT transition gives rise to a CT band which is inherent to neither the hydrocarbon nor the substituent. Intense absorption bands have been observed frequently in com— plexes involving an electron-donating molecule and an electron—accepting molecule. A quantum—mechanical theory for these complexes has been developed by Mulliken<167). (Excellent reviews of this subject have been given by (168) (169) (170) McGlynn, Briegleb, Murrel and Mulliken (171)), which explains the observed band as an and Pearson intermolecular CT transition. One would expect similar bands to be observed in the Spectra of poly substituted benzene molecules, in which an electron—donor group and an 107 electron-acceptor group are parts of the same molecule. Transitions which result from CT interaction between these groups consequently are called intramolecular CT transi— tions. Many examples of intramolecular CT transitions are known. For example, the long wavelength absorption band of p-nitro—aniline which is solvent sensitive, is charac- terized as an intramolecular CT transition involving charge migration from the amino- to the nitro-group (Figure 29). Arylboron compounds provide an interesting converse ex— ample where intramolecular CT transitions are expected to occur, thus the spectrum of tri-l-naphthylboron shows an intense absorption band at 3525 A which was inter— (172) preted as an intramolecular CT band in which the naphthyl group acts as the electron—donor and the boron atom as the electron—acceptor. Comparison of a given arylboron compound in methylcyclohexane as solvent before and after bubbling dry NH3 gas through the solution shows that the band assigned as an intramolecular CT band disappears in the presence of ammonia. This was explained in terms of the formation of an additive compound with NH3 in which the lowest vacant orbital of boron,involved in the intramolecular CT transition,is now filled with the nitrOgen lone pair electrons (Figure 30). The localized—orbital model is the most preferred model for the treatment of substituted benzene. One of the 108 6 no" \J/R\I ,1 ‘ . . . 5 \\ // / \ \\ _. . —— —enm ALCOIOL ,o’ '- A- --------- METHYL cmorcx»: I/ \ -* ,__. _ .—. amen /'}:"- \- I / I": /)\~ . ' / ' \\ 2/ \ \ /"-., '\\ 2 \ss' ,./// i I." - \ : I I“ I I - I I .I"MI_A 2000 2400 2000 3200 3600 4000 4400 4000 WAVE LENGTH IN ANGSTROMS Figure 29. Room temperature absorption spectra of p— nitroaniline in different solvents: (...) methyl- cyclohexane, ( ----- ) ether, and (————) ethanol (concentration in alcohol = 6.52 x 10- gm moles lit-1.). 109 Wovemmbers x I0"3 315 40 4s 50 30 L ‘\~ I I / o 4/ 1 T ‘ ‘000 3500 3000 A, Angstroms Figure 30. Room temperature absorption spectrum of tri-l- naphthylboron in methylcyclohexane (dashed line); and same after bubbling in dry ammonia gas (solid line). 110 reason is that it allows a correlation between the ab- sorption bands of various substituted benzenes. The CT state energies can be calculated by using the equation: where ID is the ionization potential of the donor, EA is the electron affinity of the acceptor and C is the Coulombic attraction energy between the electron and the hole. In monosubstituted benzenes symmetric and antisym- metric CT states result, due to the degeneracy of the highest filled and lowest vacant benzene orbitals, their energies are different due to the difference in the electro— static term. In disubstituted benzenes of the type ID-<:)-A, one can use two modifications.(l73) The first is the use of the ionization potential of D—O rather than that of benzene. The second arises from the realiza— tion that the degeneracy of the highest filled benzene orbitals is removed as a result of substitution. The highest filled MO of benzene is doubly degenerate, but this is not the case in phenol or aniline (Figure 31). In para disubstituted benzene the transition to the symmetric state is allowed but it is forbidden for the antisymmetric state, due to the zero coefficient of the wave function on the carbon atom which is linked to the substituent. The CT band observed for the para isomer 111 .__JLI__. I ,0 I ‘Q’ Figure 31. Qualitatively molecular orbital energy level for benzene, phenol and lowest vacant orbitals of substituent. 112 corresponds to the symmetric CT state and it borrows most of its intensity from lBlu band. For meta and ortho di— substituted benzene both transitions are symmetry allowed. Let us consider the absorption spectra of salicylic acid (SA) in different media (Figure 32). For SA in its neutral form there exist two CT bands, one at 302 nm and the other at 237 nm. These bands are due to a transition to a CT state in which the hydroxyl group acts as an electron donor and the carboxyl group acts as an electron acceptor. The n—n and B2u bands of benzene are hidden under the intense CT band at 302 nm, but the higher energy CT band at 237 nm of the ortho and meta isomers borrows most of its intensity from the 1B LE band of benzene. lu The intensity of the lBlu band is approximately 10 fold greater than that of the 1B211 band and therefore the band at 302 nm has greater CT character. When we remove a proton from the carboxyl group, the two CT bands are shifted to the blue because of the lower electron affinity of the carboxylate anion compared to that of the carboxyl group. The removal of the phenolic proton causes a red shift for both CT bands, because of the lower ionization potential of the phenolate anion with respect to that of the phenol group. The red shifts for the doubly—Charged anion for the two CT bands are smaller than that of methylated salicylic acid. This reflects the electron repulsion of these two negative charges which forces the carboxyl group 4v.» 3.38:: 088059....“ \I // \x. 2 113 3 xIO xI03 I5 >\ E 5' :5... 4 n " IO 3- 8 2 ' \‘—' '5 ‘8 I \ ES J -1 I \\ ”L 250 300 350 400 Wavelength (nm) -0 -0 H03 \ ceo 296 302 306 358 Figure 32. Room temperature absorption spectra of salicylic acid (SA) in different media: N/lO— HCl; ( ..... ) N/20-NaOH; (-..— .—) 5—n NaOH; and methyl salicylate in (-—-—) N/20-NaOH. 114 to be out of plane. The spectrum in this case is very similar to the spectrum of the phenolate anion. l. Solvent Effect on the Absorption Spectra of PAS and PASE Environmental factors affecting the absorption spectra can be conveniently divided into three types: intermolec- ular hydrogen-bonding between solute molecules only; inter- molecular hydrogen-bonding between solute and solvent mole- cules; and environmental effects not involving the forma- tion of hydrogen bonds such as dipole—dipole effects. All of our experiments were performed for very low concentra- tions of the compound in order to eliminate the possibility of intermolecular hydrogen bonding between solute mole- cules. Absorption spectra of dilute solutions (5 x 10-5 M) of PAS has been taken in different media and are shown in Figure 33. The energies of the two near UV absorption bands in different media and their spectral shifts with respect to 3-methylpentane (3MP) are listed in Table 3. Depending on the solvent, an intramolecular hydrogen bond or intermolecular hydrogen bond can be formed. In 3MP, one expects that the predominant species will have intra- molecular hydrogen bonds involving the hydroxyl and car- boxyl groups. We have assigned the 275.5 nm absorption 200 115 0.8 ‘ ' - 0.7 - . 0.6 - - 0.5 ' " §04- - B .203- q <[ 0.2 - - 0.l - - l l 200 250 300 350 Wavelength (nm) Figure 33. Room temperature absorption spectra of PAS in different solvents: ( ) 3MP, (----) ether, ( ----- ) ethanol, and (....) H2O (pH — 2). 116 Table 3. Room Temperature Absorption Maxima (nm) and Spectral Shifts (om-l) Relative to 3MP for PAS in Different Solvents. Solvent 1(nm) A§(cm-l) 1(nm) A5(cm_l) (N)CT (O)CT 3MP 275.5 0 302 0 Ether 278 —327 302 0 EtOH 280 -584 303.5 -164 H20(PH=2) 279 —456 299 +332 — red shift + blue shift band to a transition to a charge transfer state in which the amino group acts as an electron donor and the aromatic ring as an electron acceptor, (N)CT band. The 302 nm band is assigned as a charge transfer band involving the hydroxyl as an electron donor and the carboxyl group as an electron acceptor, (O)CT band. Of course, this is a zeroth—order description of the electronic states Since mixing between locally—excited and charge transfer states occurs. If the transition involves mainly an electron transfer from the highest filled orbital of the donor group to the lowest vacant orbital of the acceptor, then one would expect that the band would disappear if the energy of the highest filled orbital of the donor group becomes much deeper in energy as a result of protonations. 117 Charge transfer transitions involving promotion of an electron from a w—orbital "lone pair"* which has n-sym— metry and can therefore conjugate with the benzene molecule, virtually disappear as a result of protonation. The ab— sorption spectrum of PAS at a low pH value 0.85) is con— sistent with our assignments since the 275.5 nm band at— tributed to a charge-transfer band involving the amino group, disappears when the amino group is protonated and the spectrum becomes nearly the same as that of salicylic acid (Figure 35). The fact that neutral salicylic acid exhibits an absorption band at m302 nm and that p—amino benzoic acid exhibits an absorption band at m270 nm sup— ports our assignment. In ether, where intermolecular hydrogen bonds in? volving the hydroxyl and amino groups are formed, the spectrum is very similar to that in 3MP where an intra— molecular hydrogen band is formed. Dipole—dipole inter~ actions contribute to the observed transition energies in ether. In ethanol, in addition to the previously mentioned interactions, there are intermolecular hydrogen bonds * . The 2p7rorbitals on the nitrogen atom of aniline and pyrrole are w-orbitals while the 2p orbitals on the aza-nitrogen of pyridine and on the oxygen atom of the carbonyl group are n—orbitals. 118 involving the proton of ethanol. The net result is that both absorption bands are shifted slightly to the red compared with those in ether. In water, at pH 2, the Spectrum shown in Figure 33 corresponds to the neutral form and is Similar to that in alcohol. However, the relative intensities of the two absorption bands are reversed and the bands occur at slightly higher energies. Hydrogen bonding involving the protons of water with the amino group lone-pair is probably responsible for these effects. In water the intensity of the (N)CT band decreases due to H—bonding with the amino nitrogen lone—pair which if carried to the ex— treme situation, would be the addition of H+ which would eliminate the amino CT band. In water we expect a larger red Shift, but the hydrogen bond between the amino group lone—pair and the alcohol proton is weakened as a result of CT excitation and will cause a blue shift which is superimposed on the red shift due to dipole-dipole inter- actions. This effect is more pronounced in water than in alcohol because of the higher acidity of water. Hydrogen bonds formed between the amino group hydrogen atoms and the solvent molecules become more stable in the CT state and lead to a red shift. The absorption spectra of PASE (5 x 10‘5 M) have been measured in different media and the results are shown in Figure 34. Absorption maxima and spectral Shifts are 119 C) e 8 Absorbmce _o o: (D A) DJ 200 250 300 W 350 Wavelength (nm) Figure 34. Room temperature absorption spectra of PASE in different solvents: ) 3MP, (-——-) ether; ( ----- ) ethanol; and (—--—--) H20. 120 summarized in Table 4. The results are very similar to that of PAS. Table 4. Room Temperature Abggrption Maxima (nm) and Spectral Shifts (cm ) Relative to 3MP for PASE in Different Solvents. 1(nm) _ —1. 1(nm) __ —l Solvent (N)CT Av(cm ) (O)CT Av(cm ) 3MP 273 0 302 0 Ether 281 —1043 303 -110 EtOH 285 -1542 306 —433 H20 276 —398 301 +110 2. pH Effect on the Absorption Spectra of PAS and PASE _..._—— We have studied the effect of pH on the absorption spectra of PAS and its ester (PASE) in order to determine the absorption characteristics of different species which may exist in the ground state. The results are summarized in Tables 5 and 6 and in Figures 35 and 36. For PAS at pH 0.85, the cation exhibits only one charge transfer band involving the hydroxyl group as an electron donor since the amino group is protonated at this pH value. 121 mmz" mmzn co: oom mom oa.mH Io Io owe/o- owe/o- mmz" mm 00:2 mam 3mm om.m 0 mo O“O/OI O“O/OI mmz“ mmz” 3H: mam msm oo.m Io mo owe/om owe/om mmz" mmz+ 02:2 mam III mw.o Io mo +mosoxom ouarom AEQV mofiooam opmpm popfloxm mofloodm mumpm pcsopw AECV ma meflxmz wefixez coflpdpomn< coemmfiem .mofloodm mopmpm woufloxm one pcsopw smosfim> ma pcosoeefim pm m mm pcmgm%%flm pm mm that the rate of intra— molecular proton transfer is greater than 1011 s"1 even at A°K, whereas the interconversion between cis and trans d(94> conformers is slow. The rate constants foun in neat MSA between cis and trans forms, at 25°C were k = f —l 9.5 x 105 and k = 2.6 x 107 s If excited MSA in b cyclohexane solution has a similar value of kb, the forma_ tion of the cis form, a prerequisite for the intramolecular proton transfer, cannot efficiently take place during the 136 OR 0 ' I: : = (II) b l H lifetime of the excited state trans form. In air saturated solutions this lifetime is probably less than 10-8 s. It might well be that the proton transfer in excited cis form molecules is very efficient, i;g;, that equilib— rium (I) (which is similar to that of MSA) is almost tot— ally displaced to the right. The temperature effect on equilibrium (II) should then be the main cause of the fluorescence temperature dependence in methylcyclohexane solution found by Weller. The lifetime measurement of MSA showed<96> that the neutral excited species has a fluorescence lifetime three times larger than the excited zwitterion at room temperature. Rapid formation of 450 nm emission and the very different lifetime of the fluores— cence components would also indicate that the two species are formed immediately after absorption of a photon. Per— haps they originate from different ground state molecules. In the case of MSA this is suggested by the observation of differences in the excitation spectrum of the two emis— sions in methylcyclohexane. One would expect in hydro— carbon solvents where there is an intramolecular hydrogen 137 bond (cis conformer) and not an intermolecular hydrogen bond between solute and solvent molecules, to get emis— sion at longer wavelengths due to the zwitterion, whereas for both MSA and PAS we see the short wavelength emission also. The interpretation is as follows: Fluores— (90) cence quenching measurements on the short wavelength emission band of MSA indicated that the short wavelength emission may contain contributions from two "slowly” interconverting ground state conformers that are not equally quenched and which upon excitation, give rise to an emission in the short wavelength band. The suggested conformers were ones in which the phenolic proton is H- bonded to the "ether" oxygen atom of the ester group (Ala), and a nonintramolecularly H—bonded, or "open—ring", trans conformer (Alb). The latter is stabilized in hydroxyl— ic solvents by solute—solvent H—bonding. Excited—state prototrOpism depends in part on an increase in the basicity upon excitation, of the carbonyl oxygen, no such change in basicity is expected for the ether oxygen. Prelimin— ary fluorescence decay measurements using single-photon counting appears to verify the results of the quenching (108) experiments in MSA. In a third conformer (Alc), cis form, in which the phenolic proton is H-bonded to the carbonyl oxygen, intramolecular proton transfer occurs upon excitation, giving rise to a zwitterion which emits at A50 nm. 138 l‘ Cx\ti' (D‘s CL\+' ,O- ‘ ,o 5 ‘CH3 El CH3 ('2’ ° 0 0. CH3 0 b C Figure U1. Ground—state conformers of methyl salicylate: (b) ”open—ring" (trans), "closed ring" (cis). (a) "ether bonded", and (c) The observation of a nonexponential fluorescence decay and two fluorescence lifetimes (T1 = 23.3:O.2 ns, and T2 = 5.1:O.9 ns) for the normally Stoke shifted emission of (12) methyl 3-hydroxy-2—naphthoate in cyclohexane also was interpreted in terms of more (originating from excitation ers) emitting within the two the observation of the short than one excited state of distinct ground state conform— fluorescence band. Therefore, wavelength emission in MSA and PAS could be mainly due to the existence of the ether bonded conformer in the ground state. In ether the results are very similar to those 3MP. There are two emission bands, one at m338 nm responding to the neutral form of PAS and the other stable in COP- at muu6 nm, corresponding to the zwitterion. In ether, intra— molecular hydrogen bonds are formed in PAS. In ethanol there are mainly intermolecular hydrogen 139 bonds between the hydroxyl group and solvent molecules in the ground state. Here we observed two emission bands, one at m416 nm and the other at m3AM nm. The short wave- length emission band at t3uu nm arises from the neutral form of PAS after excitation. The long wavelength emis- sion band can be attributed to a species produced by an intermolecular proton transfer from the hydroxyl group to the solvent ethanol molecule producing a singly-charged anion (phenolate anion). In order to prove this assign- ment we produced this species in the ground state by using sodium hydroxide and found that the emission occurs at the same wavelength. Therefore, in ethanol, the phenolate anion is created after excitation of a neutral molecule ground state rather than by direct excitation of a ground state phenolate anion. In water at pH 2, where PAS exists in its neutral form in the ground state, there is only one emission band at mulu nm, which again corresponds to a species produced via an intermolecular proton transfer between the hydroxyl group and solvent molecules. The shift in emission is due to a solvent effect. In ethanol and in water the ground state equilibria between different conformers play an important role in determining the emission properties. In the case of PASEZin all four solvents two emissions are observed. One occurred at shorter wavelengths and cor— responds to emission from the excited neutral form. The wavelength of this emission is solvent dependent as with an... 140 MSA.(94) The long wavelength emission occurs at nearly the same wavelength (m452 nm) and corresponds to a zwit— terion emission. The zwitterion results from intra- molecular proton transfer in the case of 3MP and ether and from intermolecular proton transfer in the case of ethanol and water. Again, one could consider the ground state equilibrium between conformers to be the origin of the two emission bands in different solvents. Depending on the solvent and which conformer is more stable in the solvent we have inter— or intramolecular proton transfer. C. Excited State Proton Transfer in the Solid Phase of P—Amino Salicylic Acid (PAS) The crystal structure of PAS has been investigated by x-ray diffraction. The packing of the molecules in the unit cell, the intermolecular distances, the bond distances and the positions of the hydrogen atoms were all consistent with the following scheme for the hydrogen bond system in PAS. 11: 2705 O-l-l r.“ fin ou-H—o 1H1 H The different band distances found for C-O and C-O” 1 II and the positions of the hydrogens show clearly that these groups are not equivalent. Ol forms two hydrogen bonds, an intramolecular one with the phenolic OH and an inter— molecular one with the carboxylic OH of a second molecule related by a center of symmetry. The intramolecular H— bonds are stronger than the intermolecular H—bonds (dOI'Olil= 2.62 K and dOl_OII = 2.70 A). The inter— molecular bridge is responsible for the relatively high melting point and low solubility of PAS, while the intra— molecular hydrogen bridge might well explain the fact that the dipole moment of p—amino-benzoic acid does not change appreciably when the OH group of PAS is introduced. by assuming the following contributions: I 58%, (III) 20%, (Iv) 13% and (II) 9%. The Kekulé formulasaccount for only about 60% of the 1H2 H H I _ '4. l I O 0 NHz NH2 I N H H | _ l H O O H‘ O 0 9° 0 buiz +IVH2 11 III total contribution, and this may explain several properties of this substance such as the high rate of decarboxylation, the low pK value, corresponding to the ionization of the amino group, and the marked variation of the absorption spectrum with pH. The existence of intramolecular hydrogen bonds in PAS in the solid phase encouraged us to look for intra— molecular proton transfer in the excited state. For this reason we measured the fluorescence spectrum of PAS in a KCl pellet. We observed a broad emission band with a maximum in the visible region (WANG nm) (Figure H2). Ex— citation at 300 nm and 370 nm gives the same emission spectrum. The observation of this emission band in the solid phase may be rationalized as due to proton transfer 143 poaaoa Hum SH m mm psosommflm pm m . o nob ma unapomMHQ pm mm.3C0.CH Ouguntg. a 160 Fluorescence Intensity 250 Figure 50. 4 " "nu-"1.4 1 1...”... 300 350 400 450 500' 550 Wavelength (nm) Room temperature fluorescence spectra of DPAS in different solvents: ( ) 3MP; (----) ether; ( ----- ) ethanol, (....) H20 (pH = 2.A). 161 Transfer from the hydroxyl group to a solvent ethanol molecule producing a singly-charged anion, and the long wavelength emission band at WA6O nm which is a shoulder and is attributed to the zwitterion produce via an inter- molecular double proton transfer between the solute and the solvent. One important difference between PAS and its N,N-dimethyl derivative is that in DPAS the carboxyl oxygen is more basic than in the case of PAS and tends to be protonated in the excited state. In the case of PAS in ethanol we have two emission bands, one due to the neutral form (“3AA nm) and the other due to the phenolate anion (“A16 nm). In water at pH 2.A, where DPAS exists in its neutral form, there is only one emission band at AlA nm which is due to a singly-charged anion as was observed in PAS in the same pH region. The fluorescence spectra and emission maxima of DPASE in different solvents are shown in Figure 51 and Table 1A. Table 1A. Room Temperature Emission Maxima in nm for DPASE in Different Solvents Solvent Emission Maxima in nm 3MP 350 A68 Ether 360 A5A EtOH 366 A92 H O 380 2 162 Fluorescence Intensity Figure 51. 300 350 400 450 500 550 600 Wavelength (nm) Room temperature fluorescence spectra of DPASE in different solvents: ( ) 3MP; (-——-) ether; ( ----- ) ethanol; and (....) H2O. 163 In 3MP, ether and ethanol there are two emission bands, one at short wavelengths with a normal Stokes' shift and is due to the neutral form. The other occurs at long wave- lengths with a large Stokes' shift due to the zwitterionic species. In 3MP and ether, intramolecular proton transfer occurs whereas in ethanol intermolecular proton transfer occurs. In water, as we mentioned earlier, there is only one emission band due to the neutral form of the molecule. E. A Unique Excitation Wavelength Dependence of Excited— State Proton—Transfer A comparison of the relative intensities of the Short (35A nm) and long (m A60 nm) wavelength emission bands of DPAS in 3MP indicates that the intensity of the 35A nm emission band is higher than that of the A60 nm emission band which is the reverse of the Situation with PASE and MSA. The high intensity of the short—wavelength emission band relative to that of long wavelength emission band prompted us to investigate whether the energy—level reversal Of the two charge transfer states in DPAS could be respons- ible for this observation. To verify this suggestion we studied the fluorescence spectra of DPAS in 3MP at two different excitation wave— lengths (Figure 52). Excitation at the 330 nm band gave mainly the emission at 35A nm, while excitation at 300 nm gave two emission bands, one at 35A, the other at mA60 nm. 16A T A I I l I l l l \ l >‘ I 3: m l #5. I /’ \\ 5 I / \ (D / \ 2 I I / \ / A ‘ / \ g ‘I [x \ g \\ // \\ u- "" \ \ __;____““~___ 1 J 1 J __ 350 400 450 500 Wavelength (nm) Figure 52. Room temperature fluorescence spectra of DPAS in 3MP at two different excitation wavelengths: (---~) Aex = 300; < ) *ex = 330. 165 It is most interesting that at a longer excitation wave— length (330 nm) the emission occurs at 35A nm while a shorter excitation wavelength (300 nm) gives rise to an emission at A60 nm as well as at 35A nm. It is important to note that in the case of MSA, shorter excitation wave— lengths gave mostly an emission at 3A0 nm while longer excitation wavelengths gave mainly emission at A50 nm. To interpret these data, we have constructed an energy level diagram for DPAS (Figure 53a). From the absorption maxima we calculated that the energy of the (0) charge transfer state lies at roughly 2600 cm_1 above that of the (N) charge transfer state. In the (N) C.T state where the amino group acts as the electron donor, the basicity of the oxygen of the carboxyl group is expected to be larger than that of the ground state, but the acidity of the hydroxyl group is not expected to Change appreciably. In this case, no proton transfer is expected for the hydroxyl group as a result of excitation of the (N) C.T state. In contrast, an excitation of the (O) C.T state where the hydroxyl group acts as an electron donor, both the acidity of the phenol group and the basicity of the oxygen of the carboxyl group are expected to be increased relative to their values in the ground state. Thus excited state proton transfer is expected to occur as a result of excitation to the (O) C.T state unless other deactivation processes are much faster than the rate of proton transfer. One can therefore expect two possible ——_—-———— 166 ““92 l (O)CT 1"- 12. M -" -- : '0 2600cn‘l' (N)CT 11:1 —T—- . —- - ' l l I 300 ' I 4§O '330 l l I ___L_ I 359 I I + _J_ 01......Hogc 0-Hmmno.c Figure 53(a). Energy level diagram for DPAS (change in em- ission due to excitation of different elec- tronic state of the same conformer). *I‘ O\\c/O\r 6° ‘\-- -- HO\ ‘ ..'H/-“ | : C I II H 0 I I l : I axe/10H 3'0: : 450 :9" 3?O| l '390 I l g I =E'1—J' - 1...... 600cm" ‘T" Figure 53(b). Energy level diagram for two conformers of MSA (change in emission due to excitation of different conformers). 167 pathways depending on the excitation wavelength. If the (0) CT state is excited radiationless transition will occur with a rate constant of 1012 sec.1 to the lowest excited singlet state, (N)CT; giving rise to emission at 35A nm. However, excited state proton transfer may occur during the lifetime of the (0) CT state giving rise to the zwitterion Species in the excited state and leading to an emission at A60 nm. The intensity of the A60 nm band indicates that the rate constant for proton transfer must be of the order of lo12 sec—l. Excitation of the (N)CT band will give rise to 35A nm emission only. Figure 53b also shows the energy level diagram of MSA. -1 The diagram shows an energy gap of 600 cm for the two ground state conformers. This value is obtained from (100) theoretical calculations In this case excitation to the lower level corresponding to the cis conformer leads to proton transfer with a rate constant of 1011 sec—1 resulting in an emission band at A50 nm. However, excita- tion to the higher energy level, i.e., excitation of the ether bonded conformer where no proton transfer may occur gives rise to an emission band at 3A0 nm. A comparison of the two energy level diagrams clearly shows that in the case of DPAS, we are dealing with excita- tion of two different electronic states of the same conformer. In contrast, the emission of MSA is varied as a result of excitation of different conformers which are at equilibrium in the ground state. 168 Consistent with our suggestion, we expected different excitation spectra for these two emission bands. The excita— tion spectra monitored at two different emission wavelengths are shown in Figure 5A. The excitation maximum corresponding to the 35A nm emission occurs at 310 nm while the excitation maximum corresponding to the A6A nm emission occurs at 300 nm. In the case of MSA<96), emission at 3A0 nm gives an excitation maximum at 308 nm and emission at A50 nm gives an excitation maximum at 311 nm. Our experiments also sup- port our earlier assignments of charge transfer states, and to our knowledge, this result is the first example of this unique excitation wavelength dependence. 169 T 1 ‘57 3IO >5 4- "I5 C .9.’ S Q) U C 8 U5 8 O 9.. LL 300 350 Wavelength (nm) Figure 5A. Room temperature excitation Spectra of DPAS in 3MP (uncorrected spectra) ( ) represents the 35A nm band, while (--—-) represents the A6A nm fluorescence. CHAPTER III EXCITED STATE DOUBLE PROTON TRANSFER IN 1-AZACARBAZOLE HYDROGEN BONDED DIMERS ITS ACETIC ACID COMPLEX AND IN THE SOLID PHASE A. Introduction Both l-azacarbazole hydrogen bonded dimers and the 1-azacarbazole-acetic acid complex were shown(175) to undergo photoinduced double proton transfer reactions in their lowest excited singlet state. An emission band with a maximum at 510 nm arises from a tautomer formed in the excited state as a result of the double proton transfer process. Studies of l-azacarbazole in the solid phase also Show that excited state double proton transfer accounts for the observed emission. The crystal structure, determined by x-ray diffraction methods(l76), shows that the crystal contains cyclic dimers only. Spectral measurements of l—azacarbazole in solution under various conditions provide further details regarding the mechanism of excited state proton transfer. The absorption and emission properties of l-azacarbazole- were reexamined under various conditions. 170 171 l. Excited State Double Proton Transfer in l—AZA Carbazole Hydrogen—Bonded Dimers Dilute solutions of l—azacarbazole in 3-methylpentane (3MP) exhibit a fluorescence band with a maximum near 367 nm. Concentrated solutions of l-azacarbazole in 3MP, in addition to the normal fluorescence (Fl) at 374 nm with a small contribution from the dimer, exhibit a second fluorescence band (F2) with a maximum at about 510 nm (Figure 55). The relative intensities of Fl and F2 depend on the concentration, excitation wavelength and temperature. The room temperature absorption spectrum of l-aza carbazole in 3MP (Figure 56) exhibits an absorption band covering the range from 300 to 350 nm. This band corresponds to the lAlI'lAl (lLa) transition of carbazole which has a transition moment along the short axis of the molecule. Such an assignment was obtained from polarized absorption of carbazole in a single—crystal matrix of fluorene at about 1 8 l5 K Lb > A. Analysis of the spectral a shifts due to change in medium leads to the qualitative conclusion that the acidity of the pyrrolic proton and the basicity of the aza—nitrogen of l—azecarbazole are both enhanced, particularly in the 1La state, the lowest excited singlet state. Figure 57 clearly shows the large red shift of 1La band of dilute solution of l—azacarbazole in 3MP at 77 K. Solvent effects on the emission spectrum of l-aza carbazole also are consistent with a larger dipole moment in the 1La state compared with that of the ground state. The ground and excited state dipole moments of carbazole were found(180) to be 1.7 and 3.6 D respectively. If equilibrium is established during the excited state * a a lifetime, then the pKa and pK; can be determined in a way analogous to the spectraphotometeric determination of the ground state pK's, by fluorometric titration. For this reason we have repeated earlier measurements of the fluores— cence and absorption intensity of l-azacarbazole in water as a function of pH. In the range from 7 to 13, neither the absorption spectra nor the emission spectra show any change. Above a pH value of l3, the absorption spectra remain the same; however, the emission spectrum exhibits the A00 nm band as well as a shoulder around the 500 nm region (See Figure 58). The new emission band can be clearly identified as the emission of the anion of l—aza 175 Absorbance O 1 250 300 350 Figure 57. Ultraviolet absorgtion spectra of l-azacar— bazole (2.5 x 10‘ M) in 3—methylpentane at room temperature (~—-)3 77°K (————). 176 Fluorescence Intensity l 1 1 L l 350 400 450 500 550 600 650 Wavelength (nm) Figure 58. Room temperature fluorescence spectra of l- azacarbazole at different pH values: ( ) “'13), (....) 13-59 ( """ ) 1307, ("’")>13-7- 177 carbazole which exists in pH Z 14 solution. From the above discussion, we can conclude that the acidity increases upon excitation to the La state. In other words, the proton will dissociate more easily in the lowest excited singlet state than the ground state. Other evidence that supports our idea is the dramatic decrease of the fluorescence quantum yield around pH 13.5 (see Figure 59) which is due to the proton dissociation of the pyrrolic nitrogen and production of the weakly emitting anion. From the midpoint of the fluorescence and absorption titration curves we * b b ground states respectively which show that indeed the acidity obtain a pK = 12.12 and pK = 13.A2 for the excited and of pyrrolic hydrogen increases by 1.3 pK unit. In the same way we get pK: = A.2 and pKa = 2.A at low pH due to protona— tion of the aza nitrogen which indicates that the basicity of the aza nitrogen increases by 1.8 pK unit. The actual magnitude of the pK value change due to excitation has been estimated (by using the Forster cycle) to be Abe = 1.6 and ApKa = “.9 for the pyrrolic hydrogen and aza nitrogen respec- tively. Hydrogen bonded dimer formation of 1—azacarbazole has been studied by comparing the ultraviolet absorption Spectra at different concentrations in 3MP at room temperature (Figure 60). A shoulder appears at 353 nm in the more concentrated solution which corresponds to the dimer. The association constant of dimer formation as obtained from absorption spectra at different concentrations is 178 .mpzpmmeEmu Eoop pm maowmnhmomwmla mo wLuomdm mocoommzosfim one no mo mo powwow .mm mpzmfim V. m_ N. : 9m m N m n v m N _ d d f Ausualul allumaa 179 F'TTT‘1 I l I I I 1 ~1- 1 “5 4 l 6xlO 2 8xl0‘5 3 l x IO”4 \N 4 2xl0‘4 8 c: 8 § 3 .0 2 <1 I l I ii I L_ 1 i4. .L 300 350 Wavelength (nm) Figure 60. Room temperature absorption spectra of l—aza- carbazole in 3—methylpentane at different con_ centrations. 180 -l (6.8:0.7)x102 M at room temperature (21°C). This corresponds to a AG value of 3.8 Kcal mole_l at room temperature (8.1i0.8)x103 M-1 at 0°C and W108 M-1 at 77°K. The association constant in 3MP as a function of temperature resulted in an estimate of -l8.25 Kcal mole_l for AH and -A9 eu for AS. The large negative value of the enthalpy change means the system is energetically favorable towards dimerization at low temperature. The excitation spectrum (mlO—uM) for the F band is 2 shifted to longer wavelengths compared with that for the F1 band indicating that the F2 fluorescence band arises from excitation of the hydrogen bonded dimer which absorbs at longer wavelength compared to the monomer. The F1 fluores— cence band arises mainly from monomer excitation. The assignment of the F band as being due to a tautomer 2 formed as a result of double proton transfer in the excited state of the '1—azacarbazole hydrogen-bonded dimer has been further confirmed by studying the emission spectrum of the N -methyl tautomer of l-azacarbazole (I) which we have 1 synthesized. CH3 I This molecule has the same w-electron system as the prOposed tautomer. Its emission spectrum in ethanol occurs at the 181 same energy and exhibits similar fine structure to the F2 band observed by excitation of the 1-azacarbazole hydrogen— bonded dimer. By using nanosecond time resolved spectroscopy(l8l), the rate constants of double deuteron transfer has been estimated at 77°K to be k? = 6.1 x 107 and k3 = 2.1 x 107 sec—l. The large rate of the double deuteron transfer points to a very low barrier energy for the process. The energy barrier for excited—state proton transfer in the l—aza carbazole hydrogen—bonded dimer can be estimated by studying the temperature effect on the F2/Fl intensity ratio. An 1 estimated energy barrier is about 3 Kcal mole— A potential energy diagram which qualitatively describes the energetics of the double proton transfer process is shown in Figure 61. Potential energy is plotted versus a reaction coordinate associated with the simultaneous motion of the two protons in the dimer. It is assumed that the tautomer is unstable in the ground state. -1 By using an activation energy of 3 Kcal mole and l k = 6.1 x 107 sec—1 one may obtain a value of 10 3 for the pt frequency factor A in the equation for the absolute rate constant, which is similar to the expected theoretical value 12 10 m 1013. 2. Phosphorescence of l—Azacarbazole In the case of 7—azaindole where photo-induced double (79) 3 proton transfer occurs in concentrated solution in 3MP 182 F?h+_++-———a- c 03"" W I? 33.. ._;J: : Dimer Tautomer Figure 61. Qualitative double minimum potential energy diagram illustrating excited-state double proton transfer in l-azacarbazole hydrogen- bonded dimers. 183 the low temperature fluorescence spectrum overlaps with the isoenergetic monomeric phosphorescence(127). This coinci- dence makes it difficult to remove ambiguities regarding the temperature dependence of the two fluorescence bands and makes the mechanism of proton transfer less certain. The phosphorescence Spectrum of 1—azacarbazole was measured at 77°K and is shown in Figure 62. The maximum of the phosphorescence band occurs at MAO nm while the fluores— cence band due to the excited state double proton transfer of hydrogen—bonded dimers occurs at 510 nm. Thus, in the case of l-azacarbazole no ambiguity is present. 3. Excited State Double Proton Transfer in the l—Azacarbazole-Acetic Acid Complex 1-Azacarbazole forms a complex with acetic acid. The room temperature absorption spectra of a dilute solution in hexane, 2 x 10"5 M of l—azacarbazole in the presence of various amounts of acetic acid are shown absorption at 353 nm is due primarily to can be used to determine the association complex. A value of A X 105M—1 has been association constant of the complex. It in Figure 63. The the complex and constant of the estimated for the should be noted that the association constant for the l—azacarbazole acetic acid complex is much larger than that for the mer (6.8 x 102M_1). l—azacarbazole di— The room temperature emission spectra of l—azacarbazole in the presence of excess acetic acid exhibits only the F2 184 Intensity J l n LJJI I 1 1 300 40 500 GOO Wavelength (nm) Figure 62. Phosphorescence spectrum of 1—azacarbazole (6 x 10‘ M) in ether at 77°K excited at 330 nm . l L \ 413—- 3£5- ,0 3.0 — O S? )< > 2.5— 3 l. o m 2x)-— .0 O L 3 0 L5 '— :5 LO-- 015—— Figure 63. 320 340 360 380 Wavelength Room temperature absorption spectra of l— azacarbazole (2 x 10_5 M) in hexane, in the presence of variable amounts of acetic acid, 0, 1.66 x 10‘6, 3.32 x 10'6, 4.98 x 10‘6, 6.6A x 10‘6, 8.30 x 10-6, 9.96 x 10‘6, 1.16 x 10‘5 M in 10 cm cell. 186 band as shown in Figure 64. Similarly, at 77°K only F2 emission is observed. This means that the proton transfer is much faster than the fluorescence rate of 1—azacarbazole The estimated value for kpt is 1.“ x 109 sec—l, so that the activation energy in the excited state must be less than I 0‘16 Kcal mole_ and consequently, the thermal energy at 77°K is enough to overcome the barrier. A- Excited State Double Proton Transfer in l—Aza Carbazole in the Solid Phase Crystallographic data The crystal structure of l-azacarbazole has been deter- mined(l76) by x-ray diffraction. Crystals of l-azacarbazole are monoclinic and belong to the space group P2l/n; the dimensions of the unit cell are: a = 13.516(A), b = 5.526 (2), c = 11.336(5)Z, B = 95.77(3)O; z = u, M = 168.20, pC = 1.326 g cm_3. Lattice dimensions were determined using a Picker FACS-I diffractometer and IVIOKOLl (A = 0.70926 A radiation. Intensity data were measured by using Mqu radiation (26 ax = 55°) yielding 1940 total unique data m points and, based on I>30(I), 1236 observed data points. The data were reduced(l82), the structures were solved by direct methods(183), and the refinement was by full-matrix least squares techniques<18u). The finalll.value was 0.035. The final difference Fourier map showed densities ranging 187 Relative Intensity /”\\ / \ \ / / \ l ’ J \.\ 300 400 500 500 Wavelength (nm) Figure 6A. The emission of 5 x 10‘ M of l-azacarbazole in 3—methy1pentane in the absence ( , presence of excess acetic acid (—-——). 188 from +.19 to -.28 with no indication of missing or incorrectly placed atoms. The planes of the three rings individually are flat to within $0.01A. The plane of the three rings together is parallel to the plane of the equivalent molecule joined by hydrogen-bonds and related by a crystalographic center of symmetry, but these two planes are not coincident. The distance between the two planes is’V0.5lA and the angle between either plane and the plane of A nitrogen atoms involved in the H-bond is 10.250. Figure 65 shows stereo- (185) D scopic View 01 a hydrogen-bonded dimer molecule. The stereoscopic views of the packing of the dimer molecules into the unit all are shown in Figure 66. The intermolecular distances, the bond distances and the positions of the hydrogen atoms were all consistent with doubly hydrogen— bonded '1—azacarbazole dimers in the crystal. Room Temperature Absorption and Fluorescence Spectra of l—AzaCarbazole in the Solid Phase The absorption spectrum of l-azacarbazole in a nujol mull at room temperature is shifted to the red similar to that observed for concentrated solutions in 3MP, which indicates the existence of hydrogen-bonded dimers in the solid phase of 1—azacarbazole.. The infrared spectrum was measured in the region of A000 Cm—l to 2000 cm—1 in two different media, a nujOl mull and 189 Stereoscopic View of a hydrogen-bonded dimer molecule. Figure 65. Figure 66. ./ j x”, ’§" / \ I ’ \ l I I- I-\ \cl 1 |'\ - ’7 'I \' \r‘ ’\1\/ ‘ l—\ \'| -\ l- ‘ L \ \" '\/\l \ ‘\ ‘-\ \;-\ "' ‘ .. ‘ ‘ \o\ 'V\I .\ ’\I\’ I." \ Q ; ’\/\ \’\ \\. 1‘ ‘1‘ , l \ - \ l “A sax \ I. ' \ f\ “ ’ l l .\,\ "\ .- t.- -: . u \r n l l.\ \\ \ul " ~ I-‘ \ it“ 1 \ \w ‘l " -‘J- . \ , I \ I :\ est}; ”‘0?\ \J \ )1“ \ g ‘ n J \ z‘ . \IJ -‘ . II _I:\ “V“ 7 I“ .\\ ”\a l ~\ \t- .i\e- -‘\- \ ‘ {\. \ \{\ \’\l \\ f\ \yf: ‘ rt . ‘ \ . l " '\'\. l \‘ x. 4 \ \ r '\ \l\' \J \. ‘ " ‘\ \’\’ - \ \ \J ‘\ \\, \.\ "\}\‘ ‘\\ J a a \J- \ \~ JJ \ \-‘ ‘\ l .1 \\ J ' \ or \ s u .1 ' \ Stereoscopic Views of the packing of dimer molecules into the unit cell. a. Viewed down towards the b. Viewed down towards the 0. Viewed down towards the the top the top the top a* axis with +b to the right and +9 of the page. b axis with +9 to the right and +3 of the page. 3* axis with +a to the right and +2 of the page. 191 a KBr’pellet. In both cases the free N-H stretching vibration (m3A80 cm-l) is absent. The room temperature fluorescence Spectrum of a single crystal of 1-azacarbazole which has been grown by slowly cooling the liquid is shown in Figure 67. The spectrum shows clearly that an emission band occurs at W500 nm which is similar to the band observed in concentrated solution of l-azacarbazole. The emission is therefore interpreted as being due to a tautomer formed via double proton transfer in the excited state. This result is what is expected from the fact that the crystal consists of hydrogen bonded dimers. 192 .mmmcq UHHom as» CH mHonopmomNth mo ESLpoodm oocmommposam apzpmpmoEop Eoom AES £32903 0mm 00m 09v 1 _ q u 4 d — .So opswwm Mgsualul aoueosetonl :1 CHAPTER IV EXPERIMENTAL A. Experimentally Studied Molecules 1. Para Amino Salicylic Acid (PAS) PAS was obtained from Aldrich Chemical Company, and was further purified by crystallization from ethanol M.W. = 153.1u, n = lAA-1A5 (dec.) 2. Methyl Para Amino Salicylic Acid (PASE) PASE was synthesized by using the method of selective es- terification of aromatic carboxylic acids which have other reactive groups by using a Boron Trifluoride EtherateeAlcohol (186) and was purified by crystallization from ethanol. reagent Confirmation of the identity of the compound was done by measuring the melting point (120°C), the mass spectrum (MW= 167.1“), and NMR and infrared spectra (in a KBr disk). The UV absorption spectrum is almost identical to that of PAS. 3. N,N Dimethyl Para Amino Salicylic Acid (DPAS) DPAS was purchased from Aldrich Chemical Company, and was crystallized from ethanol, MW = 181.19, MP = 1A0 (dec.) A. Methyl NyN Dimethyl Para Amino Salicylic Acid (DPASE) 193 19A DPASE was synthesized in the same way as PASE. Charac- terization of the compound was done by melting point 85°C, and by obtaining thermassspectrum (M.W. = 195.19), NMR and infrared (in KBr) spectra. The u.v. absorption spectrum is almost identical to that of DPAS. 5. l—Azacarbazole (lAC) lAC was synthesized by using the method of L. Stephenson and W. K. Warburton<187) with more attention to the effect of light, and purification of starting materials. It was recrystalized from ethanol several times. Confirmation of the identity of the compound was done by measuring the melting point (217°C) and from the mass spectrum. The pyrrolic hydrogen was identified by N.M.R. with .[= -l.79. The UV absorption spectrum is almost identical to that of carbazole. The expected rrflfli transition must be completely hidden by the more intense nem*' absorption band. 6. N -Methy1 1-Azacarbazole Tautomer 1 After dissolving 1 gm of lAC in toluene, we added 1.5 ml. dimethylsulfate. The solution was then refluxed for 10 hours, and cooled down. The solid was filtered and recryg stalized from acetone and methanol. The white crystals have been identified as d—carbolium methylsulphate. After dis— solving in acetone, it was neutralized by adding NaoH solution. The yellow precipitate was separated. The final 195 bright yellow crystals with m.p. 130 to 132°C have been identified as the Nl-methyl 1—azacarbazole tautomer by Mass, NMR, and UV, spectra. B. Solvents l. 3-Methylpentane (3MP) A modified version of the purification method of Potts (188) was used. Phillips pure grade 3-methylpentane (3MP) was mixed with a 50:50 mixture of concentrated sulfuric acid and concentrated nitric acid, then stirred for two days. Then it was separated and stirred with concentrated sulfuric acid for several hours until the dark red color disappeared and the acid layer became yellowish brown. After separation, it was stirred with a dilute solution of sodium carbonate for one hour until CO2 production ceased. The 3MP was then stirred several times with distilled water until the water remained clear compared to the initial yellow color it attained after the first stirring. After drying the 3MP overnight over anhydrous sodium sulfate, the solution was refluxed over sodium wires for two days and distilled. The vapor was passed through a four foot vacuum-jacketed column and condensed at a speed of two drops per minute. The purity was checked by obtaining the absorption Spectrum. Passing the distillate through a 1m column of activated silica gel did not alter its absorption characteristics so this was not required in the purification process. 196 2. Ethanol Absolute ethanol was fractionally distilled through a 1 meter vacuum jacketed column. The distillation rate was adjusted so that a very slow rate (about 5 drops per minute) was maintained. Distillation continued until the benzene—alcohol azeotrope was no longer present as deter- mined by an absorption spectrum of the distilled alcohol in a 10 cm cell. That is, the characteristic benzene UV absorption was no longer apparent. Ethanol was then distilled and used as needed. 3. Ether Ether was distilled over sodium hydroxide. A. Water Only doubly-distilled water was used. 5. Cyclohexane Aldrich Chemical Company (spectra grade). 6. Hexane This solvent was slowly refluxed over benzophenone and sodium wire until the blue color turned to white. Dry hexane was used after distillation. 197 C. Spectral Measurements 1. Absorption Spectra All reported absorption spectra were run on a Cary Model 17 spectrophotometer. 2. Emission Spectra Most of the fluorescence spectra were obtained with an Aminco—Bowman spectrofluorometer equipped with a high pressure Xenon arc lamp and an EMT 9781 R photomultiplier tube. To obtain better resolved emission spectra, a multi- component system was used. A high intensity Xenon lamp (500W) was used as the light source, the excitation wave- lengths were reflected by a Bausch & Lomb 10 cm grating glazed at 300 A in a B & L 500 mm monochromator (which pro- vides a narrow excitation band width). The excitation illumination was focused on the sample by using quartz lenses, and emission was detected at a right angle relative to excitation. The emission monochromator was a Spex 1700-11 which utilized a 10 cm B & L grating glazed at 5000A. The emission spectrum was detected with an EMI 9558 QA photo- multiplier tube. The tube's input voltage was maintained by a Fluke A12B power supply which was normally Operated at 1100V. Signals from the detected emission were fed to a Princeton Applied Research HR-8 Lock-in Amplifier whose reference was provided by a light chopper. Finally, the 198 amplified emission signal was displayed on a strip-chart recorder. Phosphorescence spectra were obtained with Amino SPF 500 equipped with a rotating can.(SLM Instruments Inc) chopper. 3. Infrared Spectra Infrared spectra were obtained by using a Perkin-Elmer A57 Grating Infrared spectrophotometer. Calibration of frequency reading was made with a polystyrene film. Samples were examined as KBr disks. A. Mass Spectra Mass spectra were taken with a Finnigan EI.CI gas chromato- graph—mass spectrometer. 5. NMR Spectra NMR Spectra were run on a varian T—60 NMR Spectrometer system. r—' CHAPTER V CONCLUSION AND FUTURE WORK Absorption spectra of dilute solutions of PAS have been measured in different media. Depending on the solvent, an intramolecular hydrogen bond or intermolecular hydrogen bonds are formed. The interpretation of the absorption spectra which are dominated by "charge—transfer” transitions has been obtained by studying the effect of medium, pH and substituents effect and comparison with the related compounds. The absorption and emission characteristics of different species which may exist in the ground and excited states have been obtained by studying the effect of pH on the absorption and emission spectra of PAS and its methyl ester (PASE). This study has revealed the occurrence of excited state proton transfer in both PAS and PASE. The fluorescence spectra of PAS in 3MP and ether consist of two bands, one at 3A0 nm which is due to the neutral form, and the other at A50 nm and is due to zwitterion species produced via an intramolecular PT in the excited state (81). In ethanol and water we have intermolecular proton transfer from hydroxyl group to the solvents molecule. In all solvents PASE exhibit two emission bands, the long wavelength emission band in 200 3MP and ether arises from intramolecular PT. The short wavelength emission band is solvent dependent. The observation of these two emission bands is probably due to an equilibrium between the ground state conformers. Further study is needed specifically for these compounds of three functional groups to account for ground state conformers. Excitation wavelength dependence in both hydrocarbon and hydrogen bonding solvents could confirm such possibility. To identify different species in the excited state and to measure the rate of excited state proton transfer, pico— second time—resolved studies are required. The crystal structure of PAS shows intra— and intermole— cular hydrogen bridges. We measured the fluorescence spec— trum of PAS in the solid phase. Information from the solu— tion spectra of PAS helped to interpret the emission band in the solid phase of PAS in terms of an excited state intra— molecular proton transfer. Further studies including the effect of temperature and time resolved technique may reveal interesting information about excited state proton transfer (EPT), in the solid phase. Absorption and luminescence properties of DPAS and its methyl ester DPASE also have been investigated under various conditions. The results of absorption Spectra of DPAS in different solvents are similar to that of PAS with a very important exception, namely, an energy-level reversal of the two charge—transfer states. The emission spectra clearly reveal the occurrence of PT in the excited state (81). Excited state PT occurs either intramolecularly or 201 intermolecularly depending on the medium. For DPAS in 3MP, the intensity of the short wavelength (35A nm) emission band is higher than that of the A60 nm emission band which is the reverse situationwith.PASE and MSA. Excitation wavelength dependence clearly demonstrates that at a longer excitation wavelength (330 nm) the emission occurs at 35A nm while a shorter excitation wavelength (300 nm) gives rise to an emission at mA60 nm as well as at 35A nm. In the case of DPAS, depending on the electronic state excited two pathways are possible. Excitation of the (0) C.T. state, may lead to radiationless transition to the lowest excited singlet state, (N) C.T., giving rise to an emission at 35A nm or an excited state proton transfer may occur giving rise to the zwitterion species and leading to an emission at A60 nm. Excitation of the (N) CT band on the other hand, will give rise to 35A nm emission only. In the case of DPAS, we are dealing with excitation of two different electronic states of the same conformer. In contrast, the emission of MSA is varied as a result of excitation of different conformers which are at equilibrium in the ground state. Further studies such as temperature effect and time-resolved Spectroscopy will be very helpful in revealing more information about the mechanism and the rate of P.T. in this molecular system. Solutions of l—azacarbazole in 3MP at concentrations where hydrogen-bonded dimers are formed (6 x lo'um) exhibit both F and F fluorescence. The relative intensities of Fl 1 2 and F depend on the concentration, excitation wavelength and 2 202 temperature. All these studies as well as the study of model compounds indicates that the record fluorescence band origin— ates from a tautomer formed by excited state double P.T. within a hydrogen bonded dimer. The enhanced acidity of the pyrrolic proton and basicity of theazarnitrogen in the lowest excited state of l—azacarbazole (lLa state) makes the proton transfer process favorable. A preliminary study for the rate of P.T. set 6.1 x 107 sec_1 for the rate of P.T. and the energy barrier is estimated to be 2.95 Kcal mole—l. The l-aza- carbazole—acetic acid complex undergo also an excited state proton transfer that is much more efficient than in l-aza— carbazole dimer. The phosphorescence spectrum of this compound occurs at NAAO nm which is far from F2 emission band. This system is ideal to study the temperature and deuterium—isotope effect in order to determine whether a tunneling proton trans— fer mechanism is operative at low temperature. A study of the dynamics of excited—state proton transfer in 1-azar carbazole using picosecond time resolves spectroscopy is also needed. The crystal structure as determined by x-ray diffraction methods shows that the crystal contains cyclic dimers only. We have also measured the UV absorption and infrared spectra of this molecule in nujol mull and in KBr pellets; both are consistent with the existence of hydrogen bonded dimer in this system. A preliminary study of the fluores— cence Spectrum of a single crystal shows an emission band at “500 nm which is similar to the band observed for concentrated 203 solution of l—azacarbazole. This indicates that excited state proton transfer occurs in the crystal. REFERENCES 10. 11. 12. 13. 1A. 15. l6. REFERENCES J. N. Brbnsted and K. J. Pederson, Z. Phys. Chem. 108, 185 (1928). M. Eigen and L. deMaeyer, Z. Electrochem. 59, 986 1955). ‘— For this theory, P. Mitchell has been awarded the Nobel Prize in Chemistry for 1978; see A. L. Lehninger, Biochemistry (Worth, NY, 1970). A. J. Kresge, Accts. Chem. Res. 8, 35A (1975). Isotopes in Organic Chemistry, E. Buncel and C. C. Lee, ed. (Elsevier, Amsterdam 1976). R. A. Marcus, Faraday Symp. 10, 60 (1975). M. Eigen, Angew. (Chem. Int. Ed. Eng. 3, 1 (196A)) and references therein. E. Grunwald and E. K. Ralph, Accts. Chem. Res. A, 107 (1971). A. Weller, Progress in Reaction Kinetics 1, 189 (1961). M. R. Loken, J. W. Hayes, J. R. Gohlke, and L. Brand, Biochem. J., 11, A779 (1972). G. Weber and K. Rosentieck, Biochem. Symp. 1, 333 (196A). G. Weber, D. J. R. Laurence, Proc. Biochem. J., 56, (195A). M. Deluca, L. Brand, T. A. Cebula, H. H. Seliger, and A. F. Makula, Time—resolved Proton—Transfer, 2A6, 6702 (1971). P. A. Kollman, and L. C. Allen, Chem. Rev., 72, 283 (1972). J. E. DelBene, J. Chem. Phys., §§, 3139 (1973). C. A. Coulson, and V. Danielson, Arkiv, Kysik. 8, 205 (1955). 20A 205 17. S. Bratoz, Symp. Forces Intermoleculaires, Bordeaux, 1955. 18. G. M. Barrow, J. Am. Soc., 8, 5802 (1956). 19. H. Baba, A. Matsuyama, and H. Kokubum, J. Chem. Phys., .51. 895 (196A). 20. M. Arrio—DuPont, Biochem. Biophys. Research Communica— tions 33, 653 (1971). 21. G. L. Bell and G. M. Barrow, J. Chem. Phys., 31, 1158 (1959). 22. G. M. Barrow, Spectrochim. Acta 18, 799 (1960). 23. Kreuger, Can. J. Chem., 83, 201 (196A). 2A. G. E. Bacon, in D. Hadzi (editor), Hydrogen Bonding, Pergamon London, 1959, p. 23. 25. A. Piekara, J. Chem. Phys., 6, 21A5 (1962). —— 26. A. R. Ubbelohde, and K. J. Gallagher, Acta Cryst., 8, 71 (1955). 27. Y. Imry, I. Pelah, and E. Wiener, J. Chem. Phys., 5;. 2332 (1965). 28. S. Ganesan, Nature, 188, 757 (1960). 29. A. Grimson, J. Phys. Chem., 81, 962 (1963). 30. C. C. Costain, and G. P. Srivastava, J. Chem. Phys., 31, 1620 (196A). 31. R. Blinc, and M. Pintar, J. Chem. Phys., 38, llAO (1961). 32. J. L. Wood, J. Mol. Structure, 18, 1A1 (1972). 33. R. Blinc, D. Hadzi, and A. Novak, Z. Elektrochem., 8A 567 (1960). 3A. R. Blinc, and D. Hadzi, Spectrochim. Acta, 18, 852 (1960). 35. D. Hadzi, Pure and Appl. Chem., 11, A35 (1965). 36. E. Clementi, J. Mehl, and W. von Niesen, J. Chem. Phys., 53. 508 (1971). 37. 38. 39. A0. A1. A2. A3. AA. A5. A6. A7. A8. A9. 50. 51. 52. 53. 5A. 55. 56. 58. 59. 60. 206 P. 0. wadin, Rev. Mod. Phys., 5, 72A (1963). M. Kasha, Disc. Faraday Soc., 8, 1A (1950). N. S. Bayliss and E. G. McRae, J. Phys. Chem., 88, 1002 (195A). G. C. Pimentel, J. Am. Chem. Soc., 12, 3323 (1957). K. Weber, Z. Phys. Chem., 818, 18 (1931). A. Weller, z. Physik. Chem., 3, 238 (1955). A. Weller, Z. Elektrochem., 88, 662 (1952). . Weller, Z. Physik. Chem., 18, A38 (1958). . Férster, Naturwiss. 88, 186 (19A9). . F6rster, z. Elektrochem. 83, A2 (1950). . Férster, Z. Elektrochem., 88, 531 (1950). Z. Lippert, Naturforsch, 188, 5A1 (1955). Z. Lippert, Elektrochem., 81, 962 (1957). :Dtill'flt—Eit-Ht-Jib . Weller and W. Urban, Angew Chem. 88, 336 (195A). Kokubun, Naturwiss, 33: 233 (1957). Weller, Z. Elektrochem., 88, 8A9 (195A). . Weller, Z. Elektrochem., 81, 956 (1957). Weller, Z. Physik. Chem., 18, 163 (1958). Q3>3>3>III . Jackson, and G. Porter, Proc. Roy. Soc., (London), A2 , 13 (1961). E. Vander Donckt, Progr. Reaction Kinetics, 5, 273 (1970). I J. F. Ireland, and P. A. H. Wyatt, Adv. Phys. Org. Chem., 13, 131 (1976). C. Parker, "Luminescence of Solution", 1970. Albert and E. P. Sargent, "Ionization Constants of Acids and Bases", Methuen, London, 1962. S. G. Schulman, P. T. Tidwell, J. J. Cetroelli, and J. D. Winefordner, J. Am. Chem. Soc., 88, 3179 (1971). 61. 62. 63. 6A. 65. 66. 67. 68. 69. 70. 71. 72. 73. 7A. 75. 76. 77. 78. 207 S. G. Schulman, and 1. Pace, J. Phys. Chem., 6, 1966 (1972). "— T. C. Werner, and D. M. Hercules, T. Phys. Chem., 15. 1030 (1970). E. L. Wehry, and L. B. Rogers, Spectrochim. Acta, 81, 1976 (1965). E. Lippert, W. Lfider and F. M611, Spectrochim. Acta, .15. 858 (1959). E. Vander Donckt, and G. Porter, Trans. Faraday Soc., 83, 3215 (1968). H. H. Jaffe and H. Lloyd Jones, J. Org. Chem., 88, 96A (1956). S. Udenfriend, "Fluorescence Assay in Biology and Medicine", S. Schulman, and H. Gershon, J. Phys. Chem., 13. 3297 (1968). C. A. Parker, "Photoluminescence of Solutions", Else- vier, London (1968). R. Argauer, and C. E. White, "Fluorescence Analysis", Dekker, New York (1970), Academic Press, NY (1962). Li. L. Yang, M. S. Thesis, Michigan State University, 197A. E. Vander Donckt, Bull. Soc. Chim. Belges, Z8, 69 (1969). G. Yagil, J. Phys. Chem., 11, 103A (1967). J. W. Bridges, and R. T. Williams, Biochem. J. 107, 225 (1968). D. D. Rosebrook, and W. W. Brandt, J. Phys. Chem., 70. 3857 (1966). N. Mataga and Y. Kaifu, Mol. Phys., 1, 137 (1963). A. Weller, Z. Elektrochem., 88, llAA (1956). K. Hirota, z. Phys. Chem. N. F., 38, 222 (1962). D. L. Williams and A. Weller, J. Phys. Chem., 15, AA73 (1970). 79. 80. 81. 82. 83. 8A. 85. 86. 87. 88. 89. 90. 91. 92. 93. 9A. 95. 96. 208 K. C. Ingham and M. A. El-Bayoumi, J. Amer. Chem. Soc. 88, 167A (197A). M. A. El-Bayoumi and P. Avouris and W. R. Ware, J. Chem. Phys., 23: 2499 (1975). W. Klépffer, Adv. Photochem., 18, 311 (1977). W. Bartok, P. J. Lucchesi, and N. U. Snider, J. Am. Chem. Soc., 88, 18A2 (1962). A. Weller, Naturwissenshaften, 88, 175 (1955). G. M. Gantz, and W. G. Summer, Textile Rev. J., 31. 299 (1957). P. J. Kovi, C. L. Miller, and S. G. Schulman, Anal. Chim. Acta, 81, 7 (1972). W. Klépffer and G. Naundorf, J. Luminescence, 8, A57 (1979). ' IU. V. Naboikin, E. N. Pavolva, and B. A. Zaboro- zhnyi, Opt. Spectrosc. (Engl. Transl.), 6, 231 (1959). G. J. Woolfe and P. J. Thistlethwaite, J. Am. Chem. Soc., 102, 6917 (1980). P. J. Thistlethwaite and G. J. Woolfe, J. Chem. Phys. Lett., 93. 401 (1979). D. Ford, P. J. Thistlethwaite, and G. J. Woolfe, J. Chem. Phys. Lett., 88, 2A6 (1980). K.BZ. Ismail, Ph.D. Thesis, Michigan State University 19 2. H. Beens, K. H. Grellman, M. Gurr and A. H. Weller, Disc. Faraday Soc. 88, 183 (1965). E. M. Kosower and Hanna Dodiuk, J. Luminescence, 11, 2“9 {1975/76). K. Sandros, Acta. Chem. Scand., 188, 761 (1976). T. Yasunaga, N. Tatsumoto, H. Inoue and M. Miura, J. Phys. Chem., 18, A77 (1969). K. K. Smith and K. J. Kaufmann, J. Phys. Chem., 83, 2286 (1978). 97. 98. 99. 100. 101. 102. 103. 10A. 105. 106. 107. 108. 109. 110. 111. 112. 113. H. Shizuka, K. Matsui, Y. Hirata, and 1. Tanaka, J. Phys. Chem., 88, 2070 (1976). H. Shizuka, K. Matsui, Y. Hirata, and 1. Tanaka, J. Phys. Chem., 81, 22A3 (1977). N. Mori, Y. Asano, T. Irie and Y. Tsuzuki, Bull. Chem. Soc., Japan, 88, A82 (1969). J. Catalan and J. I. Fernandez—Alonso, Chem. Phys. Lett . 19. 37 (1973). E. Canadell, J. Catalan and J. I. Fernandez—Alonso, Advance in Molecular Relaxation and Interaction Pro— cesses, 1g, 265 (1978). S. G. Schulman, P. J. Kovi and J. F. Young, J. Pharm., Sci . fig. 1197 (1973). Tu. V. Naboikin, and B. A. Zadorozhnyi, Opt. Spec- trosc. (Engl. Transl.), 8, 3A7 (1960). W. R. Ware, P. R. Shukla, P. J. Sullivan, and R. V. Bremphis, J. Chem. Phys., 88, AOA8 (1971). Iu. V. Naboikin, B. A. Zadorozhnyi, and E. N. Pavlova, Opt. Spectrosc. (Engl. Transl.), 6, 312 (1959). M. Oki, M. Hirota, and S. Hirofuji, Spectrochim., Acta. 88, 1537 (1966). E. D. Bergmann, Y. Hirshberg, and S. Pinchas, J. Chem. Soc., 2351 (1950). G. J. Woolfe, and P. J. Thistlethwaite, J. Am. Chem. Soc., 103, 38A9 (1981). P. T. McTigue, and P. T. Renowden, J. Chem. Soc., Faraday Trans. 1., 11, 178A (1975). P. J. Kovi, and S. G. Schulman, Anal. Chem., 88, 989 (1973)- S. G. Schulman, P. J. Kovi, Anal. Chim. Acta, 81, 259 (1973). S. Nagakura, and M. Gouterman, J. Chem. Phys., 88, 881 (1957). (a) E. Lippert, Z. Naturf., 188, 5A1 (1955). (b) E. Lippert, z. Elektrochem., 81, 962 (1957). 113. 11A. 115. 117. 118. 119. 120. 121. 123. 12A. 126. 210 (c) N. Mataga, Y. Kaifu, and M. Koizumi, Bull. Chem. Soc. Japan, 38, 690 (1955)- (d) N Mataga, Y. Kaifu, and M. Koizumi, Bull. Chem. Soc. Japan, 88, A65 (1956)~ (e) J. Czekalla, Z. Elektrochem., 88, 1221 (1960). A. Weller, Z. Elektrochem. 88, 662 (1952). (a) N. Mataga, Y. Kaifu, and M. Koizumi, Bull. Chem. Soc. Japan, 33, 115 (1956). (b) N. Mataga, Bull. Chem. Soc. Japan, 31, A81 (1958). (a) Th. Farster, Naturwissenschaften, 38, 108 (19A9). (b) Th. Fbrster, Z. Elektrochem., 88, A2 (1950). (0) Th. FCrster, Z. Elektrochem., 88, 531 (1950). (d) A. Weller, Z. Elektrochem., 88, 55 (1960). N. Mataga and Y. Kaifu, J. Chem. Phys., 38, 280A (1962). A. Matsuzaki, Nagakura, and K. Yoshihara, Bull. Chem. Soc . Japan, 81, 1152 (197A). 0. Stern, and M. Volmer, Phys. 2., 39, 183 (1919). "Handbook of Chemistry and Physics“, ed by C. Hodg— mon, Chemical Rubber, Cleveland, OH (1962), pp. 2258— 226A. A. Weller, z. Phys. Chem. N. F., 11, 22A (1958). M. V. Smoluchowski, Z Phys . . Chem., (Leipzig), 93, 129 (1917)- . R. Ware and J. S. Novros, J. Phys. Chem., 70 32A6 (1966) C. A. Taylor, M. A. El—Bayoumi, and M. Kasha, Proc. Nat. Acad. Sci. 0.8., 83, 253 (1969). M. A El Bayoumi, Ph.D. Thesis, Florida State Uni— versity, 1961. K C Ingham, M. Abu—Elgheit and M. A. El—Bayoumi, J. Amer. Chem. Soc., 83, 5023 (1971). 127. 128. 129. 130. 131. 132. 138. 139. 1A0. 1A1. 1A2. 1A3. 1AA. 211 H Bulska, and A. Chodkouska, J. Am. Chem. Soc., 102, 3259 (1980)- K. B. Eisenthal, K. Gnaedig, W. M. Hetherington, M Crawford, and R. H. Micheels, Springer Ser Chem. Phys. 8 (Picosecond Phenom.) 3A—9 (1978). W M. Hetherington III, R. H. Micheels and K. B Eisenthal, Chem. Phys. Lett. 66 230 (1979). a_) P Avouris, L. L. Yang, and M. A. El—Bayoumi, Photo- chemistry and Photobiology, 38, 211 (1976). C. A. Taylor, Masters Thesis, Florida State Uni— versity, 1969. W A. Van Hook in Isotope Effects in Chemical Re— actions, eds., J. C. Collins and N. S. Bowman, ACS Monograph 167, Van Nostrand Reinhold, New York (1971) Earlier reviews are cited in this reference C. G. Swain, E. C. Stivers, J. F. Reuwer, Jr J. Schaad, J. Amer. , and L. Chem. Soc. . 99. 5885 (1958). . Lewis and J. K. Robinson, J. Amer. a SA337 (1968) 90 M. J. Stern and P. C. Vogel, J. Amer. Chem. Soc 33, A66A (1971); M. J. ' 3 Stern and R. E. Weston, Jr Chem. Phys., 88, 2815 (197A). Chem. Soc., J. Bigeleisen, Tritium in the Physical and Biological Sciences, IAEA, Vienna, Vol. 1 (1962), p. 161. R. P Bell The Proton in Chemistry, 2nd Ed., Chapman and Hall (1973). E. F. Caldin, Chem. Revs., 83, 135 (1969) A. J. Kresge, Pure Appl. Chem., 8, 2A3 (196A). V. Gold, Adv. Phys. Org. Chem., 1, 259 (1969). P. M. Laughton, and R. E. Robertson, Can. J. Chem 35, 171A (1956). R. E. Robertson, Can. J. Chem., 38, 613 (1957). R. E Robertson, and P. M. Laughton, Can. J. Chem 35. 1319 (1957). F D Ross1ni, J. W. Knowlton, and H. L. Johnston, J. Res. Natl. Bur. Stand. 2A, 369 (19A0). 3 1A5. 1A6. 1A7. 1A8. 1A9. 151. 152. 162. 163. 16A. 212 W. M. Jones, J. Chem. Phys., 88, 207 (1968). . G. Swain, and R. F. W. Bader, Tetrahedron, 10, 182 (1960) C. G. Swain, A. D. Ketley, and R. F. W. Bader, J Am. Chem. Soc., 18, 2353 (1959). F. Hund, z. Physik, 53, 805 (1927). E. P. Wigner, Z. Phys. Chem. (Leipzig), B19, 203 (1932). D. M. Dennison, and G. E. Uhlenbeck, Phys. Rev., 81, 313 (1932)- E. Grunwald, in "Progress in Physical Organic Chem— istry". Cohen, Streitwieser, and Taft, eds. Wiley, NY 1965 Vol. 3. H. S. Johnston, Adv. Chem. Phys., 3, 131 (1961). E. F. Caldin, Chem. Rev., 88 . 135 (1969). C. Haas, and D. F. Horning, J. Chem. Phys. (1960). G. . 32. 1763 M. Barrow, Spectrochim. Acta. 19. 799 (1960). D. Hadzi, J. Chem. Phys., 33, 1AA5 (1961). P. 0. wadin, Adv. Quantum Chem. 9 a. 213 (1965)- J. Soc., 188, A18 (192A). W Klopffer and G. Kaufmann, J. Luminescence, 20, 283 (1979). K. Marsh, J. Chem. B. A Zadorozhnyi and I. K. Ishchenko, Opt Spectry (English. Transl.) 18, 306 (1965). H. Kokubun, Z. Physik. Chem., 13, 386 (1957); Natur— wissenschaften AA, 233 (1957);_Z. Elektrochem. 88, 599 (1958) A N Terenin and A. V. Karyakin, Doklady Akad Nauk SSSR 8, A25 (19A7); Nature 159, 115 (19A7). A. V. Sheblya and A N. Terenin, Optika. Spektr. 18, 617 (1961). 165. 166. 167. 168. 169. 170. 171. 172. 173. 17A. 175. 176. 177. 178. 179. 180. 181. 182. 213 Yu. N. Sheinker and 1. Ya. Partovskii, Zhur. Fiz. Khim. 33. 39A (1958). F. Bertinotti, G. Giacomello and A. M. Liquori, Acta Cryst. Z, 808 (195A). R. S. Mulliken, J. Am. Chem. Soc., :8, 811 (1952). S. P. McGlynn, Chem. Rev., 88, 1113 (1958). G. Briegleb, Elektronen-Donator—Acceptor Komplexe Springer—Verlag, Berlin-Gottingen—Heidelberg, 1961. J. N. Murrel, Quart. Rev., 18, 191 (1961). R. S. Mulliken and W. B. Person, Ann. Rev. Phys. Chem. 13, 107 (1962). B. Ramsey, M. A. El—Bayoumi, and M. Kasha, J. Chem. Phys., 38, 1502 (1961). J. N. Murrell, The Theory of Electronic Spectra of Organic Molecules, Wiley (1963). L. Lannelli, P. Giordano Orsini and G. Daniele, R. C. Accad. Napoli, 38, 1 (1953). C. Chang, N. Sharbestury and M. A. El—Bayoumi, J. Chem. Phys. Lett. Z8, 107 (1980). X—ray structure was determined by D. L. Ward, Michigan State University in 1978. A. Bree and R. Zwarick, J. Chem. Phys., 88, 3355 (1968)- K. R. Popov, L. V. Smirno, L. A. Nakhimovskaya and V. L. Grebneva, Opt. i Spektroskopiya 31, 363 (1971). H. U. Schuett and H. Zimmermann, Ber. Bunsenges. Physik. Chem. 81, 5A (1963). A. Marty and P. Viallet, J. Photochem., 1 (6), AA3 (1973). C. Chang, Ph.D. Thesis, Michigan State University, 1978. K. T. Wei and D. L. Ward, Acta Crystallographica, B32, 2768 (1976). 183. 18A. 185. 186. 187. 188. 21A G. G. Multan, P. Main and M. M. Woolfson, Acta Crystallographica, A27, 368 (1971). A. Zalkin, private communication (197A). C. K. Johnson (1965). "ORTEP. A FORTRAN Thermal- Ellipsoid Plot Program for Crystal Structure 11- lustrations", Report 0RNL—379A, Oak Ridge National Laboratory, Oak Ridge, Tennessee. P. K. Kadaba, J. Pharm. Sci., 83, 1333 (197A). L. Stephenson, and W. K. Warburton, J. Chem. Soc. (C). 1355 (1970). N. S. Potts, J. Chem. Phys., 28, 809 (1952).