THESIS This is to certify that the dissertation entitled VITAMIN B6 MODEL COMPOUNDS AND THEIR INTERACTIONS WITH METAL IONS presented by Kamal Zaki Ismail has been accepted towards fulfillment of the requirements for Ph . D . degree in Chemistry 2y. War/{W Majoi professo/ v M. Ashraf El-Bayoumi Date April 23, 1982 MS U is an Affirmative Action/Equal Opportunity Institution O~ 12771 ‘IV1E31_J RETURNING MATERIAL§z Place in book drop to LJBRARJES remove this checkout from .—:——. your record. FINES will be charged if book is returned after the date stamped below. ____ VITAMIN B6 MODEL COMPOUNDS AND THEIR INTERACTIONS WITH METAL IONS By Kamal Zaki Ismail A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1982 ABSTRACT VITAMIN B6 MODEL COMPOUNDS AND THEIR INTERACTIONS WITH METAL IONS By Kamal Zaki Ismail Nearly all reactions concerned with amino acid trans- formations are catalyzed by pyridoxal phosphate (PLP). Such reactions involve racemization, transamination, de- carboxylation, a,B-elimination, B, y-elimination and de- aldolization. Pyridoxal phosphate, PLP, (vitamin B6) is a substituted pyridine derivative with a hydroxyl group in position 3- and an aldehyde group in position H-. The bio- logically active form of pyridoxal phosphate, PLP, is its Schiff base with an e-amino group of a lysine residue in the active site of the enzyme. Salicylaldehyde derivatives have been used as model compounds of PLP since they contain its essential chromophore. We have studied both absorption and emission properties of four model compounds namely: salicylaldehyde, 2-methoxybenzaldehyde, 2-hydroxyacetophenone, and N,6-dimethyl-2-hydroxyacetophenone and their corresponding Schiff bases with an aliphatic amine, n-butylamine, and an amino acid, DL-valine,in different media and at both room Kamal Zaki Ismail temperature and 77°K. Intermolecular and intramolecular excited-state proton-transfer have been observed. The steric hindrance in u,6-dimethyl-2-hydroxyacetophenone model compound and its Schiff base clearly manifested itself in their spectral properties. The presence of monovalent cation is very essential for activity in some PLP-enzymatic reactions. In trypto- phanase both NH”+ and K+ are good activators whereas, Na+ and Li+ are not. Different techniques have been used ex- tensively to study monovalent cation interactions with dif- ferent enzymes. We have studied both absorption and emis- sion spectra of some monovalent cation (e.g., Li+, Na+, K+ and Tl+) salts of salicylaldehyde and three other para- substituted phenols in THF and 111 other solvents. Stronger cation-anion interaction (tighter ion-pair formation) was observed in Salicylaldehyde salts. Even in polar solvents, e.g., ethanol and DMSO an equilibrium between ion pairs and free ions was observed. Because of their intense and narrow emission, lanthanide ions, Ln(III), have been extensively used to probe the microenvironment at the active site in proteins. The number of coordinated water molecules in the first shell of Eu(III) are determined by following the fluorescence decay of the metal ion in HZO/DZO mixtures. Energy transfer 2+ 2+ in Ln(III) bound protein, substituting Ca or Mg , is used to measure the distances between different sites in proteins. Kamal Zaki Ismail To explore the potential of use of europium (III) in prob- ing PLP-enzymes active sites we have prepared the europium chelate of salicylidenevalinate Schiff base and studied its magnetic and optical properties. The complex seems to be in an octahedral configuration. Energy transfer from ligand to metal was observed and an energy level diagram revealing both absorption and emission properties of the complex was constructed. To my Mothejt, My Wondenfiul 011.69., and my Son ii ACKNOWLEDGMENT I would like to express my deep appreciation to my advisor, Dr. M. Ashraf El-Bayoumi for his guidance, en- couragement, and friendship during the course of this study. I also thank the members of my committee, Dr. Carl Brubaker, Dr. Thomas Pinnavaia, and Dr. Andrew Timnick. My thanks also go to Mr. Ernest Oliver for the mass spectrosc0pic measurements, Ms. Jo Kotarski for patiently drawing all the figures, and Ms. Peri-Anne Warstler for doing an excellent Job in typing this Dissertation. I thank my former colleagues, Dr. Khader Al-Hassan and Ms. Nahid Shabestary, for their friendship and support. My thanks are extended to my beloved country, Egypt, and to Alexandria University. I also extend my gratitude to all my friends in the Egyptian Club in East Lansing for making me feel at home with their friendship and brotherhood. Last, but not least, I would like to thank my wife, Eman, for her continuous encouragement and endless patience throughout this work. To my son, Kareem, I give my ever- lasting gratitude for waiting one entire week after my oral defense to be born, on Friday, April 30, 1982, thus allowing me a recovery period. iii Chapter LIST OF LIST OF CHAPTER CHAPTER I. II. TABLES. FIGURES TABLE OF CONTENTS I - INTRODUCTION. II - EXPERIMENTAL Systems Studied A. Model Compounds a. b. Aldehydes . Schiff bases. Preparation and Purification of Compounds . . . . . . Monovalent-cation Salts Potassium bis(salicylidene-DL— valinato) Europium(III) Chelate Materials . . . . . . . A. Purification of Solvents. O\U'l 12'me Ethanol . . . . . . . 3-Methylpentane (3MP) Water . . . . . . . . . . . Acetonitrile, P-Dioxane and N,N-Dimethylformamide (DMF) Dimethylsulfoxide (DMSO). Tetrahydrofuran (THF) Spectral Measurements U'l-DUOI'UH Absorption Spectra. Emission Spectra. . 77°K Emission Spectra Infrared Spectra. Mass Spectra. . iv Page vii . viii 0rd «4 -4 «a l4 (I) 10 ll 11 ll 11 12 12 12 l2 l3 l3 l3 l3 I“ IN Chapter CHAPTER II. III. CHAPTER II. III. CHAPTER III - OPTICAL SPECTRA OF SALICYLALDEHYDE ‘ MODEL COMPOUNDS . . . Introduction. . . . . . . . Room Temperature Absorption Spectra. . . . . . . . . . Emission Spectra and Excited—State Proton—Transfer in Salicylaldehyde Model Compounds . . . . . . . . . IV - ABSORPTION AND EMISSION STUDY OF THE ASSOCIATION OF MONOVALENT CAT- IONS WITH PHENOXIDE AND SALICYL— ALDEHYDE ANIONS. . . . . . . . Introduction. Monovalent Cation Interactions. Spectroscopic Studies of Ion-Pair Formation. . . . . . . . . . . A. Ultraviolet-Visible . B. Infrared and Raman. C. Electron Spin Resonance (ESR) . . . . . . . D. Nuclear Magnetic Resonance (NMR). E. Luminescence. . . . . . . . . . . . Results and Discussion. . . . . . . v - EUROPIUM (III) CHELATE WITH SALICYLIDENE-VALINATE SCHIFF BASE. . . . . . . . . . . . Lanthanide Ions as Luminescence Probes of the Structure of Biological Macromolecules. . . . . . . . A. Decay Lifetimes as a Measure of the Number of Metal Co- ordinated Water Molecules Page 15 15 16 38 67 67 68 71 73 89 93 95 97 98 125 125 132 Chapter Page B. Characterization of Individual - Binding Sites . . . . . . . . . . . 136 C. Inter Metal Ion Energy Transfer Distance Measurements . . . . . . . 138 II. Energy Transfer in Rare-Earth Metal Chelates . . . . . . . . . . . . lHO A. Paths of Energy Migration . . . . . 1H1 B. Rate of Energy Transfer . . . . . . 1&5 C. Mechanism of Energy Transfer. . . . 1H6 III. Europium Complex with Salicylidene- DL-Valinate Schiff Base . . . . . . . 147 A. Introduction. . . . . . . . . . . . I“? B. Results and Discussion. . . . . . . 151 CHAPTER VI - CONCLUSION AND FUTURE WORK . . . . . 160 REFERENCES. . . . . . . . . . . . . . . . . . . . 163 vi Table LIST OF TABLES Solvent Effect on the Absorption Spectra of Compounds I-IV . Absorption Spectra of Schiff base Model Compounds in Different Solvents Emission Spectra of Compound I in Different Media. Emission Spectra of Compounds (I-IV) in Ethanol and 3MP Emission Spectra of Compounds (I-IV) in Water . l . . . . . . . Log K, AH, and TAS for the Formation of 18-Crown-6 Complexes with Metal Ions in H2O at 25°C Rate Constants for Reaction a; 25°C in H2O, u=0.3: M+ + 18-Crown-6 EdM-Crown+ Absorption Bands of Alkali Metal Salts in Dimethylsulfoxide. Room Temperature Absorption and Emission Maxima of Para-Substituted Phenols in THF . . . vii Page 23 33 “5 53 53 87 87 91 123 Figure LIST OF FIGURES Pyridoxal phosphate (vitamin B6) related compounds The effect of pH on the spectrum of holotryptophanase. Room temperature absorption spectra of dilute solutions (5 x 10-5 M) of compound I (-———), II (-..-.--), III ( ------ ) and IV (-——-) all in ethanol. Room temperature absorption spectra of dilute solutions (5 x 10'5 M) of compound I ( ), II (------), III (—-----) and IV (-—--) all in 3MP Room temperature absorption spectra of dilute alkaline solutions (5 x 10"5 M) of compound I (————), II (-------), III (- ----- ) and IV (----) in ethanol Room temperature absorption spectra of dilute solutions (5 x 10-5 M) of compound I-b (————), II-b ( ------ ) and IV-b (----) all in 3MP. viii Page 17 22 25 28 Figure 10 11 Room temperature absorption spec- tra of dilute solutions (5 x 10-5 M) of compound I-b in different sol- vents: ethanol (————), dioxane (----) and DMSO ( ----- ) Room temperature absorption spec- tra of dilute solutions (5 x 10'5 M) of compound II-b in ethanol ( ----- ), acidic ethanol (---—), dioxane (....) and 3MP (————). Room temperature absorption spectra of dilute solutions (5 x 10"5 M) of compound I-v in different solvents: ethanol (----), dioxane ( ) and DMSO (- ..... ). . . . . . . . . Room temperature absorption spectra of dilute solutions (5 x 10‘5 M) of compound I-b in: ethanol (————), acidic ethanol (----), and alkaline ethanol (- ----- ) . . . . . . Room temperature absorption spectra of dilute solutions (5 x 10-5 M) of compound I-v in ethanol (---—), acidic ethanol ix Page 29 3O 32 3M Figure ll l2 l3 1” 15 ) and alkaline solution ( ethanol solution (- ----- ) Room temperature absorption spectra of dilute solutions (5 x 10"5 M) of compound II-v in ethanol (----), and in acidic ethanol solution (-—-—0. Room temperature absorption spectra of dilute solutions (5 x 10"5 M) of compound IV-b in: ethanol (----), acidic ethanol solution ( alkaline ethanol solution ( ----- ) ) and Room temperature emission spec— tra of a dilute solution (5 x 10-5 M) of compound I in ethanol at dif- ferent excitation wavelength Aexc = 330 nm ( 380 nm (----) ) and xex Emission spectra of a dilute 5: C alkaline solution (5 x 10—5 M) of compound I in ethanol, Xe 380, at room temperature ( and at 77°K (---—). XC ), Page 35 36 37 39 HO Figure l6 l7 l8 19 20 Room temperature emission spec- tra of a dilute solution (5 x 10‘5 M) of compound I in water at different excitation wavelengths, xexc = 320 nm ( <---->. Room temperature emission spec- ) and Aexc = 380 nm tra of dilute solution (5 x 10‘5 M) of compound I in 3MP xexc = 320 nm. Room temperature emission spec- tra of dilute solutions (5 x 10‘5 M) of compound II, A 320 nm in exc 3MP (------), ethanol (----) ). Emission spectra of dilute solu- u and water ( tions (5 x 10- M) of compound III in ethanol; neutral solution (Aexc = 320 nm) at room tempera- ture (---—---) and 77°K (----); alkaline solution (Aexc = 380 nm) at room temperature ( ) and 77°K (------) . . . . Room temperature emission spectra of dilute solutions (5 x 10"5 M) of compound III, in aqueous solution xi Page H2 an M6 “7 Figure Page Aexc = 320 nm (————), and in alkaline aqueous solution, Aexc = 3&0 nm (-----). . . . . . . . . . . . . . H9 21 Room temperature emission spectra of a dilute solution of compound IV in 3MP at xexc = 320 nm. . . . . . . 51 22 Room temperature emission spectra 4 of dilute solutions (5 x 10- M) of compound IV, Xe = 330 nm, in xc ethanol (—-—-) and in alkaline ethanol (------). . . . . . . . . . . . 52 23 Absorbing and emitting species in PLP-Schiff base systems. . . . . . . 56 2“ Room temperature emission spectra of a dilute solution (5 x 10-5 M) of compound I-b in ethanol at dif- ferent excitation wavelengths: Aexc = 320 nm (————), and Aexc = 380 nm (----) . . . . . . . . . . . . . 58 25 Emission spectra of dilute solu- tions (5 x 10-5 M) of compound I-b in ethanol at both room tem- perature and 77°K: Alkaline solu- tion Aexc = 380 nm at room tem- perature (----) and at 77°K (-..-.._) xii Figure 26 27 28 29 acidic solution Aexc = 330-380 nm at room temperature ( ) and at 77° (-'-°--). . . . Room temperature emission spectra of dilute solutions (5 x 10-5 M) of compound I-v in ethanol: in neutral solution Xe ( in acidic solution Aexc = 365 nm xc = 320 nm ) and Aexc = 400 nm (—..—..), (----), and in alkaline solution Aexc = 360 nm ( ----- ) . . Room temperature emission spectra of dilute solution (5 x 10-5 M) of compound II-b in ethanol Aexc = 330 nm ( ) and in acidic ethanol Aexc = 350 nm (----). . . Room temperature emission spectra of dilute solutions (5 x 10-5 M) of compound II-v in dioxane ( ----- ) and in ethanol (————) A c = 320 nm . . . . . . . ex Room temperature emission spectra of dilute solutions (5 x lO'Ll M) of compound IV-b in ethanol: neutral solution, A = 320 nm GXC xiii Page 59 61 62 6M Figure 30 31 32 33 0-———), acidic solution Xe = 320 nm (----) and alkaline solu- XC tion Aexc = 360 nm (------) Room temperature emission spectra of dilute solutions (5 x 10- of compound IV-v, Aex C u M) = 320 nm: in ethanol (-------), alkaline ethanol (—-----), acidic ethanol ( ) and in dioxane (----). Comparative effects of mono- valent cations on the spectrum of holotryptophanase at 21°. Samples contained 2.0 mg of enzyme per 1.0 ml of 0.02 M gimidazole-HCI buffer (pH 8.0), 2 mM mercaptoethanol, and ether 0.1 M NaCl, 0.1 M KCl, or 0.1 M imidazole-HCI (Im), as indicated. Correlation between wave number and the inverse of the cationic radius for contact fluorenyl ion pairs in THF at 25°C. Absorption spectra of fluorenyl salts in THF at -30°C for various counterions xiv Page 65 66 7O 76 76 Figure 34 35 36 37 38 39 Page Optical spectrum of contact and separated ion pairs of fluorenyl- lithium at 25°C in 3,“ dihydropyran, .; 3-methyltetrahydrofuran, -—--; 2,5 dihydrofuran, ; and hexamethylphosphoramide, -----. . . . . 77 Optical absorption spectrum of fluorenylsodium in THF at 25, -30, and -50°C. . . . . . . . . . . . . 80 Temperature dependence of the contact-solvent—separated ion- pair equilibrium of 9-(2-hexyl)- fluorenyllithium in 2,5-dimethyl- tetrahydrofuran.t . . . . . . . . . . . 80 Pressure dependence of the con- tact-solvent-separated ion-pair equilibrium of fluorenylsodium in THF at 25°C. . . . . . . . . . . . . 81 (I) Dibenzo-l8-crown-6. (II) Dicylohexyl-lu-crown-u; (III) monobenzo-lS-crown-S. . . . . . . . . . 83 Selectivity of l8-crown-6: log K values for reaction of lB-crown-6 with metal cations in H 0 vs. 2 ratio of cation diameter to l8—crown-6 XV Figure 40 Al 42 “3 an cavity diameter. Value for Ca2+ reported - To a dilute solution (10' to 10'5»M) of p- methyl, p-methoxy, or p-formyl phenol in THF was added an excess amount of the metal hydride. The solution was stirred under dry argon, then,the spectra were measured for a clear solution sample of this mixture. r. 0' M" .01 C. Potassium bis(salicylidene-DL-valinato) Europium(III) Complex To a clear solution of 2.2 mmoles of the Schiff base was added 2.2 mmoles of carbonate—free KOH in methanol. To this mixture was added 1 mmole of Eu(NO -5H20 solution 3’3 in methanol and the solution was refluxed for 30 minutes. The complex was obtained by evaporating the solvent under vacuum. On recrystallization of the crude complex from 11 methanol and ether, yellow crystals were isolated, de composition temperature >250°C K[(C N03)2Eu]°2H O, 12Hl3 2 analysis, Calcd: C, 43.31; H, 4.54; N, 4.21; Eu, 22.83%. Found: C, 43.17; H, 4.54; N, 4.36; Eu; 22.60%. II. Materials A. Purification of Solvents l. Ethanol Absolute ethanol was fractionally distilled through a 1 meter vacuum Jacket column, at a slow rate (about 5 drops per minute). Portions of about 50 ml were collected, and the absorption spectrum was taken in a 10 ml cell to check for benzene. Distillation was continuediuuflj.the charac- teristic benzene UV absorption was no longer apparent. Ethanol was then distilled and used as needed. 2. 3-Methylpentane (3MP) A modified version of the purification method of Potts(15) was used. 3MP (Phillips Pure Grade) was shaken for 30 minutes with a 50:50 mixture of concentrated sulfuric acid and concentrated nitric acid. It was then shaken 3 times for 30 minutes each with concentrated sulfuric acid. The solution was then treated with sodium carbonate solution until CO2 production ceased. The 3MP was then washed 12 several times with water until the water remained clear relative to the yellow color of the first wash. The 3MP was stored over sodium ribbon in a flask overnight. It was refluxed through a vacuum jacketed 1 meter column and distilled for use as needed. Purity was checked by obtain- ing the absorption spectrum. 3. Water Was doubly distilled in this laboratory. 4. Acetonitrile, P-Dioxane and N,N-Dimethylforma- mide (DMF) Spectroquality solvents from Metheson Coleman and Bell (MC/B) were used without further purification. 5. Dimethylsulfoxide (DMSO) It was stored overnight over NaOH and distilled at reduced pressure (m2-3 mm Hg, b.p. 50°) from NaOH pellets and stored over molecular sieve 4A. 6. Tetrahydrofuran (THF) Traces of peroxide were removed by refluxing a 0.5% suspension of cuprous chloride for one-half hour, followed by distillation before proceeding. Predried over KOH pellets l3 and finally refluxed over and distilled from lithium aluminum hydride as needed. B. Spectral Measurements 1. Absorption Spectra All reported absorption spectra were recorded by use of a Cary 17 spectrophotometer. 2. Emission Spectra Fluorescence spectra were recorded by use of a multi- component system consisting of a 500 w xenon light source. 500 mm Bausch & Lomb excitation monochromator (which pro- vides a narrow excitation band width), Spex 1700-11 emission Monochromator and EMI Model 9558QA Phototube. Noise reduc- tion and amplification of the PMT signal is achieved by using a Princeton Applied Research Model HR-8-Lock In Amplifier and an appropriate chopping apparatus. Some of the room temperature emission spectra were recorded by using an Aminco-Bowman spectrofluorometer. The phosphorescence spectra were obtained with these instruments equipped with a rotating can phosphorosc0pe. 3. 77°K Emission spectra A quartz dewar with flat quartz excitation and emission windows filled with liquid nitrogen, and a narrow glass l4 tube for the sample were used. 4. Infrared Spectra Infrared spectra were obtained by using a Perkin-Elmer 457 Grating Infrared Spectrophotometer. Calibration of frequency reading was.made by polystyrene film. Samples were examined as KBr disks. 5. Mass Spectra Mass spectra were obtained with a Finnigan EI-CI gas chromatograph-mass Spectrometer. CHAPTER III OPTICAL SPECTRA OF SALICYLALDEHYDE MODEL COMPOUNDS I. Introduction The purpose of this study is to investigate the ab- sorption and emission properties of four salicylaldehyde model compounds (I, II, III and IV) and their correspond- ing Schiff bases with n-butylamine (I-b, II-b and IV—b) and with potassium-DL-valinate (I-v, II-v and IV—v) under various conditions. More specifically, we studied: (1) The absorption and emission prOperties of pos- sible cations and anions. (ii) Solvent effects on the absorption and emission spectra of the various possible species. (iii) The possible occurrence of excited-state proton transfer either intramolecularly (in hydrocarbon solvents, e.g. 3MP) or intermolecularly (in hydrogen bonding solvents e.g., water, alcohol,and dioxane); in the latter case sol- vent molecules would be involved. (iv) The steric effects on the spectra, since in com- pound III and/or IV the carbonyl group may be forced out- of-plane with respect to the benzene ring. 15 16 This study should provide relevant information regarding the spectra of enzymes containing similar chromophores and possible transient intermediates occurring as a result of pH changes or during the course of enzymatic reactions. To achieve these goals we have investigated the absorption and emission spectra of the previously mentioned compounds at room temperature as well as at liquid nitrogen tempera- ture (77°K). The spectra were measured in water, dioxane, N,N-dimethylformamide (DMF), dimethylsulfoxide (DMSO), acetonitrile and 3-methy1pentane (3MP). In water and ethanol the spectra were measured in both acidic and alka- line solutions. The figures and tables shown in the sub- sequent pages provide a summary of our results. II. Room Temperature Absorption Spectra In Figure (3), room temperature absorption Spectra of dilute solutions (5 x 10-5 M) in ethanol are shown for compound I-IV, the electronic absorption spectrum of com- pound I, salicylaldehyde, exhibits two absorption bands in the near ultraviolet, one at 323 nm with an absorptivity (e = 4,000) and the second at 253 nm (e = 12,000). Con- sidering salicyaldehyde as a disubstituted benzene, its absorption spectrum may be interpreted, in the zeroth- order approximation, as resulting from transitions to either charge-transfer states or to locally-excited states. 1 l The locally excited states of benzene are B2u’ Blu and Absorbance 17 0.6 - Q A I Q N l 0.0 Figure 3. l l l j 250 . 300 350 400 Wavelength (nm) Room temperature absorption spectra of'dilute solutions (5 x 10‘5 M) of compound I ( ), II (-------), III (- ----- ) and IV (----) all in ethanol. 18 1 (16) E111 . responding to the In salicylaldehyde the absorption bands cor- lB2u benzene transition and carbonyl locally-excited n + n* transitions are probably buried under the more intense absorption at 253 nm. Accordingly, the first absorption band at 323 nm corresponds to an elec- tronic transition to an essentially charge-transfer state, where the phenolic group acts as an electron-donor and the carbonyl group acts as an electron-acceptor. Since the two highest filled orbitals in phenol do not differ greatly 1). in energy (As is about 4000 cm' One may also expect another charge transfer band at higher energies but in the same region. Thus, the second absorption band at 253 nm corresponds also to a transition to a charge-transfer state. In the first order approximation, this charge-transfer state lBlu locally-excited state of benzene. Such assignment is consistent with electronic-energy calcula- (l7) (18). is mixed with the tions as well as solvent and steric effects The charge-transfer energies can be calculated using the equa- tion(17), where ID is the ionization potential of the donor, EA the electron affinity of the acceptor,and C the coulombic attraction energy between the transferred electron and the hole it left behind. 19 (19) as a basis for The use of localized-orbital model a perturbation treatment of substituted benzene to cal- culate the energy of charge-transfer states is very useful because it is easily applied to polysubstituted benzenes, it allows a correlation with the absorption bands of various substituted benzenes and the model can account successfully for steric effects. Although the characterization of ab- sorption band as being charge-transfer (C.T.) or locally excited (L.E.) bands is correct only in the zeroth-order approximation, yet it is very useful in predicting changes in basicities, acidities and dipole moment as a result of (20) that steric effects forcing excitation. It is known the carbonyl acceptor group out of plane causes a decrease in the intensity of the charge-transfer band. The inten- sity of the charge-transfer band depends on the resonance integral across the substituent-hydrocarbon bond which is proportional to coso, ¢ being the angle of twist of the substituent. The intensity of a charge-transfer band diminishes gradually as o increases until it disappears when 9 = 90°. For the 90° twisted substituted-benzene molecule the spectrum approximated that of benzene together with low lying absorption bands involving local excitation of the substituent. The substituent in this case may influence the benzene transitions only by inductive pertur- bation. A charge-transfer state may change only very slightly in energy as a result of twisting the substituent 20 involved in the charge-transfer transition in question. Such small change is due to the possible variation of the inductive parameter of the substituent with the angle of twist. For example, the spectrum of 4-nitro-N,N-di- methyl aniline approaches that of nitrobenzene when the angle of twist of N,N-dimethylamino group approaches 90°. N,N-dimethylamino group which has (+Ifl) effect due to its H-lone pair electrons becomes less n—repelling as a result of twist. The inductive parameter of the nitro group which has (-Ifl) effect do not vary much with ¢~ The angle of twist of the substituent may be calculated from the ex- tinction coefficient of the charge-transfer band using the (18,20) prOportionality ea cos2¢. Steric effects show that the following absorption bands of substituted-benzene spectra are charge-transfer bands: the 322 nm band of p- nitroaniline, the 239 nm of acetophenone, the 233 nm of aniline and the 252 nm of nitrobenzene. In ethanol two different intermolecular hydrogen bondings may occur between the solvent and solute mole- cules depending on whether the ethanolic oxygen atom or proton is involved. In compound IV, the first absorp- tion band is blue shifted (shifted to shorter wavelength) and decreased in its intensity showing an absorption at 305 nm (e = 2,600). This blue shift as well as the reduc- tion in absorptivity demonstrate the steric effects due to the interaction between the acetyl methyl group and the ring methyl group in position 6 in compound IV. If 21 the carbonyl group was completely out of plane with respect to the benzene ring, one should expect compound IV to absorb near 280 nm like methylated phenol (p-methylphenol absorbs at 280 nm). In 3MP in dilute solutions, only intramolecular hydrogen bonding is expected between the phenolic and the carbonyl oxygen through the phenolic proton. Compound II without phenolic proton (methylated phenolic group) shows a blue shifted absorption band at 307 nm, Figure (4), compared to the rest of the compounds. The intramolecular hydrogen bond in compound I, III and IV caused the red shift in the first absorption band in 3MP. Contrary to its behavior in ethanol, compound IV, shows an intense absorption band at 335 nm reflecting the coplanarity of the carbonyl group with the benzene ring in 3MP. This implies that the energy gained through the formation of the hydrogen bond outweighs the repulsion energy between the sterically hindered methyl groups. The small red shift of the absorption band in compound IV relative to that of compound III is due to the methyl substitution in position 4 and 6. In Table (I), the solvent effect on the first absorption band of the four model compounds is summarized. The band maximum for compound IV in dioxane lies at a much longer wavelength compared to the band maximum in ethanol and in water; one may conclude that in dioxane the carbonyl group is relatively more coplanar compared to the situation Absorbonce 22 0.6 - 0.0 - 0.4 - Figure 4. 1 . 1 n I " 250 300 350 400 Wavelength (nm) Room temperature absorption spectra of dilute solutions (5 x 10-5 M) of compound I (———), II (-~----), III (- ----- ) and IV (---) all in 3MP. . 23 Table l. Solvent Effect on the Absorption Spectra of Compounds I-IV. Compd. (I) Compd. (II) SOlvent v-lo"3 *Av-lo‘3 v-lo‘3 *Av-10’3 -1 -1 -1 -l A(nm) (cm ) (cm ) A(nm) (cm ) (cm ) 3MP 328 30.5 0.0 307 32.06 0.0 Dioxane 323 30.9 +0.4 314 31.8 -0.8 Ethanol 323 30.9 +0.4 318 31.4 —l.2 Water 323 30.9 +0.4 322 31.1 -l.5 Compd. (III) Compd. (IV) Solvent v-lo‘3 *Av-lo‘3 v-lo'3 *Av-lo'3 A‘ -1 -l A -l -1 (nm) (em ) (cm ) (mm) (cm ) (cm ) 3MP 325 30.8 .0.0 335 29.8 0.0— Dioxane 321 31.2 +0.4 325 30.8 +1.0 Ethanol 324 30.9 +0.1 307 32.6 +2.8 Water 323 30.9 +0.1 3-5 32.8 +3.0 Av = vsolv. — v3MP 24 in protic solvents. Dioxane can form only an intermolecu- lar hydrogen bond with the phenolic proton. In an alkaline aqueous and ethanolic solution the ab- sorption spectra of compound I, III and IV are shown in Figure (5). The first absorption band of the anion is red shifted compared to the corresponding first absorption band in neutral solutions. In the case of compound II no change was observed as expected. Due to a smaller ion— ization potential energy of the phenolate compared with phenol one should expect the anion to absorb at longer wave- lengths. The increase in the intensity of this band in the anion is a result of a relatively greater charge migration in the charge-transfer state of the anion. The relative intensities of this band in compounds I, III and IV reflects the steric effect which is‘more dominant in the case of compound IV. In pyridoxal phosphate enzymes so far studied, the carbonyl group of the coenzyme is combined with the c- amino group of lysine residue to form a Schiff base<2l). A number of PLP-enzymes have an absorption band at 410- 430 nm. In addition to this band, most PLP-enzymes have a peak at 325-340 nm. Some enzymes display an absorption at 360—364 nm at high or at all pH values(21). These absorp- tion bands can be related to specific species in the acid- base equilibria of the Schiff base of PLP. The 410-430 nm band is assigned to the Schiff base with a protonated 25 I. I I I I -."\, .. ’ l ‘. l ; 0.4 J- ‘ - .09 Il _ \' . _ 8 It. ./ ‘\ s v. . -/ e \‘. x ="\ ,-’ ‘- 8 0'2 - 1‘ / .' '/ \ \ \- d a \I \ " y} \ ‘ q ,‘I. \ ./ /./ \\ \ '- \\ \ . //./ . \ \ - 0°. «.1 / ‘ \ \ ,__./ \ \ . *\\. \ \. ' -. \\\'\. H, _ (DC) I I 1 F" I ‘I -' I 250 300 350 400 450 Wavelength (nm) Figure 5. Room temperature absorption spectra of dilute alkaline solutions (5 x 10-5 M) of compound I (___), II (-..-..-), III ( ...... ) and IV (—---) in ethanol. 26 azomethine nitrogen and phenolate group (ketoenamine species), the 325-340 nm band to that of a phenol group without a proton on the azomethine nitrogen (enolimine species), and the 360-364 nm band to that with phenolate group and unprotonated azomethine nitrogen (anionic Species). The basis of this currently accepted assignment was provided by Heinert and Martell<22) who studied the absorption spectra of the Schiff bases of 3-hydroxy-4—formylpyridine and re- lated compounds. i 5+ 5 3 NH H NH H N H N ‘c’ \c¢ ‘c’ \c” O / O / OH / 0‘ I , «—» \ I \ I \ I N N N N (ketoenamine) (enolimine) (anion) 410-430 nm 325-340 nm 360-364 nm It is important here to note that there were no profound effects on the spectra of those model compounds when com- pared with the spectra of PLP Schiff bases as a result of the 2- and 55position substituents in PLP. A number of arguments for this conclusion have been presented from the spectral studies on solution equilibria of Schiff bases of pyridoxal and related compounds and on PLP- enzymes and from other physicochemical studies<21’23). 27 Figure (6) shows the room temperature absorption of compound I-b, II-b and IV-b in 3MP. In this solvent, intra- molecular hydrogen bond between the phenolic group and the azomethine nitrogen gives rise to an absorption at 317 nm in the case of Schiff base I-b. This absorption band can be assigned to the neutral nonpolar form of the Schiff base namely, enolimine form. Again if we treat these model compound Schiff bases as substituted benzene derivatives one should consider the smaller electron affinity of the azomethine group when compared with the carbonyl group in the model compounds. This explains the blue shift (1,060 cm'l) of the first absorption band of these Schiff bases in 3MP compared with the corresponding model compounds in the same solvent. In polar solvents, Figure (7), a new absorption band is develOped at 410 nm, the intensity of this band depends upon the polarity of the solvent. In DMSO, for example, the absorption band at 410 nm is assigned to an absorption by the neutral bipolar form of the Schiff base, keto- enamine form, this band does not exist in the absorption spectrum of the methoxy compound II-b in DMSO, ethanol or dioxane, Figure (8), because of the absence of the phenolic proton and accordingly no ketoenamine formation. The ob- servation that salicylaldehyde, compound I, does not absorb at N410 nm in DMSO demonstrates the lack of intramolecular proton transfer in the ground state in polar solvents, 28 Absorbance 250 300 350 Wavelength (nm) Figure 6. Room temperature absorption spectra of dilute solutions (5 x 10’5 M) of compound I-b ( ), II-b (-°-°-°) and IV-b (----) all in 3MP. 29 Absorbance ()‘D" 1 1 I 300 350 400 Wavelength ( nm) Figure 7. Room temperature absorption spectra of dilute solutions (5 x 10'5 M) of compound I-b in dif- ferent solvents, ethanol (-——), dioxane (---) and DMSO (-----). 30 8 § 0.2 - - ’5 U3 1: 'I>H mam mam :Hm mom * >IHH NH: we: can mos * >IH sflm * Ham mmm mam oI>H :om * mom mom mam QIHH me: * sam 00: Ham oIH AfiflvmeK AECvKMEK AEvaME< Agvxmfix Agvxmfix UCSOQEOO omzm oafipoficooooa ocmxofia Hocccom dam pco>Hom .mpco>Hom psoLoMMHmzfimmUQSOQsoo Hobo: ommm mafizom ho oppooam soathomn< .m manna 34 azaamxam can AIIIIV Hosanna ofiofiom .A .A ..... IV Hosanna v Hosanna ":H nIH ocsoasoo no A: mIoa x mv mcoauSHom ouzafio no mnpooam coauapomnm opspmpoQEop Boom .oa opswfim A85 £95663 00¢ _ 0 mm . OOM _ d l O. O I N O 11 eouoqiosqv 35 0.6 b " \ m0.4-1; - Q, . C: C) 13 _ 8 U) 8 0.2 ‘ (DE) I I I ll. - 300 350 400 450 Wavelength (nm) Figure 11. Room temperature absorption spectra of dilute solutions (5 x 10"5 M) of compound I-v in ethanol (----), acidic ethanol solution (———) and alkaline ethanol solution (------). 36 Absorbance l l l l 250 300 350 400 ' Wavelength (nm) Figure 12. Room temperature absorption spectra of dilute solutions (5 x 10-5 M) of compound II-v in ethanol (----), and in acidic ethanol solu- tion ( 37 .A ..... IV soapSHow Hosanna ocaamxam can .A v soapsfiom Hocmnuo ofipfiom .AIIIIV Hosanna ”ca nI>H ccsoasoo no A: mloa x mv mCOHpSHom audafio mo mauooam coapapomnm chaumquEop Boom .ma opswfim AEchficomSoS 00¢ 00m 00m . OmN Q 0 eouoqmsqv I l N O 38 III. Emission Spectra and Excited-State Proton Transfer in Salicylaldehyde Model Compounds Salicylaldehyde (Compound I) shows a medium as well as an excitation wavelength dependent emission spectrum. In a dilute ethanol solution excitation of the neutral molecule at 330 nm gives two emission bands as shown in Figure (14). The first band at 426 nm was assigned to an emission from the neutral nonpolar form. The longer-wavelength emission at 500 nm was assigned to an emission from a bipolar form resulting from a double proton transfer involving solvent molecules. One should point out that the phenolic and carbonyl oxygens of the solute in ethanol are hydrogen bonded intermolecularly to solvent molecules} Upon excita- tion, charge transfer from the phenolic to the carbonyl group results in an increaSe in the acidity of the phenolic (24’25) and an increase in the basicity of the car- proton bonyl oxygen. Accordingly, in the excited state, the phenolic group may loose a proton and the carbonyl oxygen may pick up a proton giving rise to a bipolar species. Excitation of the neutral solution at 380 nm,where the phenolate ion absorbs,resulted in an emission at 490 nm. Similar emission, Figure (15), was observed at 490 nm when an alkaline ethanolic solution was excited at 380 nm. These results suggest that a ground-state equilibrium between the neutral and anion forms of compound I exists in ethanol solutions. In a glass solution, at 77°K, the 39 Fluorescence Intensity Figure 14. 400 450 500 550 Wavelength (nm) Room temperature emission spectra of a dilute M) of compound I in ethanol solution (5 x 10'5 at different excitation wavelengths, A = _ . exc 330 nm ( ) and Aexc - 380 nm (----). 40 Fluorescence Intensity l I l J 400 450 500 550 Wavelength (nm) Figure 15. Emission spectra of a dilute alkaline solution (5 x 10-5 M) of compound I in ethanol, Aexc = 380, at room temperature ( ), and at 77°K (----). 41 emission of the anion is blue shifted to 434 nm. In such a rigid medium, solvent relaxation does not occur during the lifetime of the excited singlet state causing the fluorescence maximum to be blue shifted compared to the fluorescence maximum in a fluid polar medium. 0,5: H I .1. ' 3.. i‘ A _ ..- = C:O...H °’°" 9"" <—"" O t O-H: ’51 O' O-HD‘E' “O 3 l H H I H (:‘PCV’ go In an aqueous solution, the room temperature emission spectrum shown in Figure (16) exhibits two bands upon ex- citation at 330 nm, a shorter-wavelength emisSion band at 428 due to the neutral form and a longer-wavelength emis- sion at 505 due to the bipolar form. Excitation of an alkaline solution showed an emission at 500 nm and was assigned to the anion form. In 3MP, the phenolic proton and the carbonyl oxygen are intramolecularly hydrogen bonded. Upon excitation, 42 Fluorescence Intensity I 350 Figure 16. l . l l l 400 450 500 550 600 Wavelength (nm) Room temperature emission spectra of a dilute solution (5 x 10"'5 M) of compound I in water at different excitation wavelengths, Aexc = 320 nm ( ) and Aexc= 380 nm (----). 43 intramolecular excited—state proton transfer will occur resulting in the formation of the bipolar form which emits at 506 nm in this solvent, Figure (17). In Table (3), emis- sion band maxima observed in different media are summarized together with the corresponding emitting species. In compound II, 2-methoxybenzaldehyde, the phenolic proton is replaced by a methyl group and hence it cannot undergo proton transfer. Figure (18) shows its room temperature emission spectrum in 3MP, ethanol,and water. In 3MP an emission band at 350 nm was observed, this cor- responds to the neutral form. In ethanol, a single emission band appears at 376 nm. This band is blue shifted com- pared to the emission band of the neutral form of compound I observed at 426 nm. Such a large shift is attributed to the fact that in compound I the phenolic proton is hydrogen bonded to ethanol and such hydrogen bond becomes stronger in the excited state, in compound II this hydrogen bond is absent. In water, the larger dipole moment of water com- pared to that of ethanol led to further stabilization of the excited state resulting in a red shifted emission band in water at 440 nm. It is possible also that the 440 nm band in H20 is due to a cation formed by an excited-state proton transfer to the carbonyl oxygen. 'The emission spectra of ethanolic solutions of compound III,2-hydroxyacet0phenone,at room temperature and 77°K are shown in Figure (19). Intermolecular excited-state 44 Fluorescence Intensity I I l l 450 500 550 600 Wavelength (nm) Figure 17. Room temperature emission spectra of dilute solution (5 x 10'5 M) of compound I in 3MP Aexc = 320 nm. 45 Table 3. Emission Spectra of Compound I in Different Media. Ground State Excited State Fluorescence Medium Species Species max. (nm) )5? le..éo ,.H’° C H\ [p OH 426 Ethanol C IE, °~H---°\ , H I-I\C,OH GO.- 500 H 0 H 0 Ethanol/ \Cé ‘c" NaOH - (3 - Cr 6° + «a»? Mic/,0.)H 0" 3MP C) 506 H\c,0H 0 ° H/ O—H “\Céo .0° H28 H\c,o 60H H20 0‘ H... o’H + \ H\ [’0 H C O” 500 J: 46 Fluorescence Intensity I. Figure 18. 300 I l l l l 350 400 450 500 550 Wavelength (nm) Room temperature emission spectra of dilute solutions (5 x 10'5 M) of compound II, Aexc~= 320 nm in 3MP (---'-'), ethanol (----) and water ( ). Fluorescence Intensity 47 l I l l I 350 400 450 500 550 600 Figure 19. Wavelength (nm) Emission spectra of dilute solutions (5 x 10'“ M) of compound III in ethanol; neutral solution = 320 nm) at room temperature (-H - -) andx $7 °K (--—-); alkaline solution (Aexc = 380 nm) at room temperature ( ) and 77°K (- °°°°° ) 48 proton transfer in the neutral solution of compound III in ethanol gives an emission band at 512 nm. The small red shift of this emission band when compared with that of compound I, which occurs at 500 nm, may be attributed to a higher basicity of the carbonyl oxygen in compound III. In an alkaline solution, the 478 nm band is assigned to the anion emission. This band is blue shifted when compared with the emission of the anion of compound I. This is probably due to the lack of planarity of the carbonyl group in the case of compound III anion. At 77°K, the emission of the bipolar form occurs at 475 nm and that of the anion at 414 nm, these blue shifts compared with the correspond- ing emission maxima in ethanol at room temperature is due to the freezing of solvent relaxation in rigid ethanol at 77°K. The emission of compound III in water is shown in Figure (20), the bipolar form emits at 512 nm and the anion emission occurs at 490 nm. The results in water are similar to those observed in ethanol, small red shifts in water are attributed to its larger dipole moment. In 3MP where intramolecular hydrogen bonding occurs, com- pound III gives rise to an emission band at 500 nm cor- responding to the bipolar form resulting from an intra- molecular excited-state proton-transfer. In compound IV, 4,6-dimethyl-2-hydroxyacetophenone, the repulsion between the acetyl methyl group and the substituted methyl group in position 6- forces the acetyl 49 Fluorescence Intensity /'l \- /' \. /' \ '\ Figure 20. l I L I 400 450 500 550 600 Wavelength (nm) . Room temperature emission spectra of dilute solutions (5 x 10'5 M) of compound III, in aqueous solution, Aexc = 320 nm ( ), and in alkaline aqueous solution, Aexc = 340 nm <-----). 50 group to be out of plane with respect to the benzene ring. These steric effects manifest themselves in the absorption spectra as discussed before. The steric effect is expected to be more pronounced in protic solvents where intermolecu- lar hydrogen bonds with solvent molecules are found. In nonpolar solvents like 3MP absorption spectra indicate that the molecule assumes a nearly planar structure through an intramolecular hydrogen bonding between the phenolic and carbonyl groups. In 3MP compound IV exhibits two emission bands, one at 503 nm corresponding to the bipolar form and a shoulder at 440 nm corresponding to the neutral form as shown in Figure (21). The appearance of the neutral form emission is probably due to the existence of a ground- state equilibrium between intramolecularly hydrogen bonded species and species that are not hydrogen bonded, the latter cannot undergo excited state proton transfer. Figure (22) shows the room temperature emission spectra of compound IV in ethanol, an emission band observed at 422 nm is attributed to a non planar neutral molecule; no proton transfer occurs in the excited state in neutral ethanol. Band maxima of the emission spectra of compound I-IV in ethanol, water and 3MP are summarized in Tables (4) and (5). Pyridoxal phosphate dependent enzymes exhibit absorp- tion bands around 415 nm and at 335 nm. The studies made with simple model compounds have shown that the 415 nm band is due to a Schiff base with structure [1], bipolar form, which is stable in polar media (hydr0philic media). The 51 Fluorescence Intensity Figure 21. J, I ‘ l l 1 3150 400 450 500 550 600 Wavelength (nm) Room temperature emission spectra of a dilute solution of compound IV in 3MP at Aexc = 320 nm. Fluorescence Intensity 52 / \\ ./ \ / \./ \ I .ll \ I I ‘ \ I . \ I ’ \ \ l I \ \ I] ’ \ \ I f \ \ I j \ . I - \ . \ I .l .\ - [I / \~ \\ \ \ l 4350 400 4 5 0 550 600 Wavelength (nm) Figure 22. Room temperature emission spectra of dilute solutions (5 x 10' M) of compound IV, A C -330 nm, in ethanol (-——-) and in alkalIne ethanol (- ----- ).. 53 Table 4. Emission Spectra of Compounds (I-IV) in Ethanol and 3MP. Ethanol Ethanol/NaOH 3MP RT 77°K RT 77°K RT -, 77°K Compound A(nm) A(nm) A(nm) A(nm) 1(nm) 4(nm) 1(nm) neutral bipolar I 426 500 431 490 434 506 506 I II 376 450 350 450 III 512 475 478 414 500 485 IV 422 430 468 505 503 Table 5. Emission Spectra of Compounds (I—IV) in Water. H20 H O/NaOH Compound A(nm) 5(nm) I 428 505 500 II 440 440 III 512 490 IV 437 422 54 335 nm band was assigned to structure [2], nonpolar form(26), which is stable in nonpolar media (hydrophobic media). Ac— cordingly, the ratio between the absorbance of these two N" ., \ O \ 0 l , I IV 1 2 bands can be used as a measure of the polarity experienced at the pyridoxal chromophore's site(27). Solvent effects on the emission spectra of PLP-model compounds should also be helpful in interpreting enzyme emission data and predicting the microenvironment that the pyridoxal moeity experiences at the active site of the enzyme i.e., whether the site is a hydrophilic or hydro- phobic in character. For example, glycogen phosphorylase contains one PLP molecule per protein subunit, PLP is very tightly attached to the protein through a Schiff base bond with an e-amino group of a lysine residue of the (28) protein In the pH range of 5.0 to 9.5, at which the enzyme is stable, two absorption bands due to PLP are observed with peaks at 333 and 425 nm(29). These two ab- sorption bands indicate that an equilibrium exists between the enolimine and ketoenamine forms and the active site is not hydrophilic. The enzyme shows two emission bands: one occurs at 335 nm and corresponds to the emission of 55 tryptophane residues in the protein; the other band at 535 nm corresponds to the emission of the ketoenamine form of PLP<30’31). In the excited state the keto-enol equilibrium is completely shifted towards the ketoenamine form. For model systems, n-butylamine, valine, or n- hexylamine-PLP in various solvents, the observed properties of Schiff bases are very different. The Schiff base of PLP-n-butylamine in aqueous solution and the Schiff base PLP-valine in DMF exhibit two emission bands: one oc- curs at 510 nm from the ketoenamine form absorbing at 410 nm; the second emission at 430 nm occurs from the enol- imine form which absorbs at 325 or 333 nm(32). In polar solvents, excitation of the enolimine form at 325 nm re- sulted in an emission from the ketoenamine form indicat- ing that proton transfer occurred in the excited state. In hydrogen-accepting solvents like dioxane, the two emissions were observed from the n-butylamine Schiff base. These results can be summarized in Figure (23). In non polar solvents the ground state equilibrium is shifted towards form [1], whereas, in the excited state it is shifted towards form [2]*(32). In contrast to the nitrogen hetero-atom, the phenolic group plays a major role in the spectral behavior of Schiff bases of aromatic ortho-hydroxy aldehydes. Trans- fer of the phenolic proton causes tautomerism in the ground state and phototautomerism in the excited state. 56 R R I ' .. I a. c¢ "y \c// '5’. \ 0 4VI 1, \ o l , 330nm l , N L N . w [I] enolimine ' [|]* (330nm) (430nm) -x- llK ll K I ' I ‘* + + \ 0' I‘ll/2 I \ 0" I"! 4IO nm N/ [2] ketoenamine [2]* (420nm) . (5lOnm) Figure 23. Absorbing and emitting species in PLP- Schiff base systems. 57 Blocking of the hydroxyl group thus dramatically modi- fies the absorption and emission properties characteristic of these Schiff bases(33). Thus Schiff bases of o— methoxybenzaldehyde is expected to show little resem- blance to the optical and luminescence properties of the Schiff bases of the corresponding o-hydroxybenz- aldehyde. The spectral properties of PLP-model com- pounds depend on whether the phenolic group forms a hydro— gen bond with the azomethine nitrogen of the Schiff base or with solvent molecules. These considerations lead to our choice of the model compounds that we have studied. Room temperature emission spectrum of compound I-b in ethanol is shown in Figure (24). Two emissions were observed when the neutral solution was excited at 320 nm where the enolimine form absorbs. The shorter-wavelength emission with a peak at 425 nm corresponds to an emission from the enolimine form and the longer-wavelength emission at 500 nm corresponds to an emission from a ketoenamine form. Excitation at 420 nm where the ketoenamine form absorbs gave only the longer-wavelength emission at 500 nm. In acidic ethanolic solution, Figure (25), an emis- sion from the cationic form was observed at 472 nm. An alkaline solution of compound I-b in ethanol gave an emis- sion at 510 nm when excited at 380 nm where the anion absorbs, this emission belongs to the anion form. 58 Fluorescence Intensity I II I I 400 450 500 55 Wavelength (nm) Figure 24. Room temperature emission spectra of a dilute solution (5 x 10'5 M) of compound I-b in ethanol at different excitation wavelengths: Aexc = 320 nm ( ), and Aexc = 380 nm (----). Fluorescence Intensity 59 I I I I I l l 350 400 450 500 550 600 Wavelength ( nm) Figure 25. Emission spectra of dilute solutions (5 x 10'5 M) of compound I-b in ethanol at both room tem- perature and 77°K: alkaline solution Aexc = 380 nm at room temperature (----) and at 77°K (-H - --), acidic solution Ae xc = 330- 380 nm at room temperature ( ) and at 77° (-----). 60 Emission in both acidic and alkaline solutions were blue shifted to 415 and 408 nm respectively at 77°K because of the lack of solvent relaxation. Room temperature emis- sion spectra of compound I-v in ethanol are shown in Figure (26). Excitation at 320 nm led to an emission from the enolimine form at 450 nm, whereas, excitation of the ketoenamine form at 400 nm gave an emission at 495 nm. In acidic solution, an emission at 480 nm was observed as a result of the cation form excitation at 365 nm. In alkaline solution, excitation of the anion form gave an emission peak at 500 nm. In compounds II-b and II-v, there is no phenolic pro- ton. The emission spectra of compound II-b in ethanol are shown in Figure (27). Excitation of the neutral ethanolic solution at 300 nm resulted in two emission peaks. The first peak at 366 nm corresponds to the neutral form, and the second at 450 nm was assigned to an emission from a cationic form where the azomethine nitrogen accepts a proton from the solvent. This assignment was confirmed by studying the emission in acidic ethanol shown also in Figure (27). These results indicate that the basicity of the azomethine nitrogen is apparently greatly enhanced in the excited state. Figure (28) shows the emission spectra of a dilute solution of compound II-v in dioxane and ethanol. Emission from the neutral form was observed in dioxane at 352 nm. The 450 nm emission in neutral ethanol 61 O. ‘. \\, \ ~ I l l l l 400 450 500 550 600 Wavelength (nm) Fluorescence Intensity Figure 26. Room temperature emission spectra of dilute solutions (51:10'5 M) of compound I-v in ethanol: in neutral solution A = 320 nm ( ) and Aexc = 400 nm (----°- , in acidic solution Aexc = 365 nm (----), and in alkaline solu- tion Aexc = 360 nm (-----)._ 62 Fluorescence lntensity / l I l 1 1‘- 350 400 450 500 550 Wavelength (n m) Figure 27. Room temperature emission spectra of dilute solution (5 x 10‘5 M) of compound II-b in ethanol Aexc = 300 nm ( ) and in acidic ethanol Aexc = 350 nm (----). 63 Fluorescence lntensity ' 1 l I l I I 350 400 450 500 550 600 Wavelength (nm) Figure 28. Room temperature emission spectra of dilute solutions (5 x 10'5 M) of compound II-v in dioxane ( ----- ) and in ethanol ( . ) Aexc = 320 nm. 64 which results from excitation of the neutral form at 320 nm can be interpreted as resulting from a cation form ob- tained by proton transfer from solvent to the azomethine nitrogen. Room temperature emission spectra of compound IV-b in neutral, acidic and alkaline solutions in ethanol are displayed in Figure (29). Excitation of the neutral form at 320 nm produced a broad emission with a peak at 415 nm, this band is probably a composite band consisting of emis- sion of the neutral form at 415 nm and an emission at N460 nm due to the anion emission. In acidic solution, the cation emits at 450 nm whereas, in alkaline solution the anion emits at 465 nm. In the corresponding valine Schiff base, compound IV-v, excitation of the neutral solu- tion at 320 nm results, Figure (30), in a broad emission which can also be considered as a combination of two bands: the first around N420 nm which corresponds to the neutral form emission and the second at N465 nm which cor- responds to the anion form. In acidic solution an emission peak was observed at 435 nm. In alkaline solution, the emission of the anion form was observed at 463 nm. The lack of excited-state proton transfer in the case of compound IV and its Schiff bases IV-b and IV-v demonstrates the lack of planarity in these molecules because of steric effects. Fluorescence lntensit y 65 l J l l 400 450 500 550 Wavelength (nm) Figure 29. Room temperature Emission spectra of dilute solutions (53:10' M) of compound IV-b in ethanol: neutral solution, Aex = 320 nm ( acidic solution Ae xc = 320 nm (- line solution Aexc = 360 nm (- - - ). ---) and alka- 66 Fluorescence lntensitL, 400 480 500 550 Wavelength (nm) Figure 30. Room temperature emission spectra of dilute solutions (5 x 10' M) of compound IV-v, Aexc = 320 nm in:ethanol (-'-----), alkaline ethanol (-. ----- ), acidic ethanol ( ) and in dioxane (----). CHAPTER IV ABSORPTION AND EMISSION STUDY OF THE ASSOCIATION OF MONOVALENT CATIONS WITH PHENOXIDE AND SALICYLALDEHYDE ANIONS I. Introduction The presence of a metal atom as an essential consti- tuent of some enzymes, and the metal ion requirement of others for maximum activity, provide an obvious link between enzymatic reactions and coordination chemistry. The pos- sible functions of metal ions in enzyme substrate systems (34)' are easily visualized as follows: (i) The model may form a complex with donor atoms of either the enzyme or substrate and thereby enhance their tendency towards reaction; (ii) It may serve merely as a bridge through common coordination to bring the enzyme and substrate into proximity; (iii) While serving function (ii) it may provide as well a chemical activating influence; and (iv) While coordinated to either the enzyme or substrate it may appropriately orient groups undergoing reaction. 67 68 The most obvious property of a metal ion is its positive charge which makes it effectively a Lewis acid, and it will have a tendency to withdraw electrons from atoms and groups to which it is attached. The catalytic efficiency will depend on the effective charge of the metal ion. Monovalent Cation Interactions All tryptophanases so far examined require NHu+ or K+, and the concentration required for half-maximum activity (NHu+, 2—6 mM; K+, 8-20 mM) is similar for each enzyme(2). Na+ and Li+ are essentially without activating effects in each case, but they inhibit the E.coli enzyme when added together with NH“+ or K+(3). At optimal concentrations, NH“+ permits a level of activity at least equal to and usually slightly greater than K+, Rb+ also activates but less effectively than K+(u). T1+ also replaces K+ for the E. coli enzyme and has a greater affinity (0.35 mM gives ap- proximately half-maximum activity<35)). Equilibrium dialysis studies showed that apotrypto- phanase binds one mole of pyridoxal phosphate per 57,500 grams of protein<36> , that is four pyridoxal phosphate molecules per molecule of native tetrameric enzyme (mol- ecular weight 220,000). K+ and Na+ are equally effective in permitting the dissociation of tetrameric to dimeric apotryptophanase at low temperature<37), and it is unlikely therefore, that this effect serves as a basis for their 69 markedly different catalytic effects. In addition, distinct spectral differences are observed in the presence of K+ versus Na+. Holotryptophanase shows characteristic pH- dependent absorption maxima at 420 and 337 nm in the pres- ence of the catalytically essential K+ or NH: ions. At equiv- alent concentrations of Na+ or imidazole, only the 420 nm band is formed, Figure (31), at pH 8.0(5). In the presence of K+, at the same pH, the 337 nm band is formed. In the enzymatically inactive form with Amax 420 nm, intramolecular hydrogen-bond between the phenolic group and the azomethine nitrogen was proposed, whereas, in the active form with A x = 337 nm, the phenolic group lost its proton(5). ma (6,7) Recent studies ,using a rapid-scanning stopped-flow absorption technique, of the interconversion between the 420 and 337 nm absorbing species showed that the form absorbing at 337 nm is the inactive form but acts as a reservoir for the active form which absorbs at 420 nm. The activity as well as the spectroscopic behavior in the presence of K+ or Na+ ions demonstrate a difference in the tertiary structure of tryptophanase and in the mode of binding of pyridoxal-p in the two environments that is consistent with a more compact structure and tighter bind- ing of pyridoxal-p in the K+ environment. This difference is sufficient to explain the requirement for K+ (or NHu+) for catalysis, but does not elucidate the molecular basis for this behavior. In this regard, we have studied the 70 3 , , ' pH a0 w I" f i In .2 °: 0 Z ( m K 8 a: ‘.1 300 350 . 400 450 WAVELENGTH (MIL) Figure 31. Comparative effects of monovalent cations on the spectrum of holotryptophanase at 21°. Samples contained 2.0 mg of enzyme per 1.0 m1 of 0.02 M imidazole-HCI buffer (pH 8.0), 2 mM mercaptoethanol, and either 0.1 M NaCl, 0.1 M 5 KCl, or 0.1 M imidazole-HCl (Im), as indicated. 71 the monovalent cation interaction with some pyridoxal phosphate model compounds. In media favoring ion-associa— tion, e.g., tetrahydrofurane, we have studied both absorp- tion and emission properties in order to examine the affinity of ion-pair formation both in the ground and excited states. Before discussing our results, a short review of the spectroscopic methods of studying ion association is given. II. Spectroscopic Studies of Ion-Pair Formation The concept of ion pairs was introduced in 1926 by (38) Bjerrum It was known in those days that ionophores- compounds built up of ions and not neutral molecules — are completely dissociated in aqueous solution, and it was ex- pected that they should behave in the same way in other solvents. It therefore came as a surprise when Krauss<39) reported that sodium chloride, a typical ion0phore, behaves like a weak electrolyte, an ionogene, when dissolved in liquid ammonia. The electric conductance of such a solu- tion is given by the law governing the conductance of aqueous solution of acetic acid indicating that only a small fraction of the dissolved salt is dissociated into free ions. To account for these observations, BJerrum proposed that in liquid ammonia and in other nonaqueous solvents the oppositely charged ions are associated into neutral ion pairs which do not contribute to the electric 72 conductance. Solute-solute interactions result in the formation of ion pairs or of higher aggregates. The extent of ionic aggregation will depend not only on the dielectric constant of the solvent but also on its solvating ability (donicity) as well as on the nature of the ions. Ion pairing can result in contact pairs, solvent shared pairs or separated pairs(40). At higher concentrations and/or in solvents of low dielectric constant and low donicity ionic triplets, quadruplets, etc. may also form. It should be noted also that the introduction of a salt into a solvent may affect the solvent-solvent interaction especially in highly struc- tured solvents such as water or dimethylsulphoxide (DMSO). To solve this complex puzzle, we must have the means of identifying the various chemical species present in a given solution. Once all species are identified, we may then study their interactions and the equilibria that exist amongst them. Such data are the key for understanding chemical processes in solutions. For many years studies of electrolyte solutions were limited to electrochemical measurements or measurements of colligative properties of solutions. It was only since 1960 that the behavior and structure of ion pairs and ion-pair complexes in solution have been extensively investigated by using a variety of spectroscopic techniques such as electron paramagnetic and nuclear magnetic resonance, ultraviolet, visible, infrared, 73 and Raman spectroscopy. A brief summary of the type of results obtained using these various techniques is now given. A. Ultraviolet-Visible, One drawback of this technique is the inability to decide whether an observed solvent effect is due to a ground state interaction, an excited state interaction or both. The large band width of absorption bands observed in solution requires large spectral shifts if ion-pair (5) association are to be detected by this technique. In spite of these shortcomings, ultraviolet-visible absorption spectrOSCOpy is an attractive technique for studying ion- pair interactions due to the simplicity of its excecution and the straightforwardness of data analysis. Moreover, it is not limited to paramagnetic species like electron spin resonance methods, and does not require high concentrations of reagents which is often imperative in the nuclear mag- netic resonance studies. Warhust ep_al,(ul) were perhaps the first to observe shifts in optical spectra arising from ion pairing. They reported that the absorption peaks of ion pairs of alkali and alkaline earth salts of ketyls of benzophenone, fluor- enone and other ketones shift towards longer wavelengths when the radius of the cation increases. Similar effects were later observed by Hogen(42) for fluorenyl salts, Waak 74 for organolithium salts and by Zaugg and Schaefer Almpl Optical spectrum of contact and separated ion pairs of fluorenyllithium at 25°C in 3,4 di- hydropyran, ...; 3-methyltetrahydrofuran, ----; 2,5 dihydrofuran ; and hexamethyl- phosphoramide, ------.fi7 78 shifts are the result of a decreased influence of the cation on the anion, caused by specific cation-solvent interactions which led to a partial separation of ions. A pronounced effect was observed with solvents containing small amounts of reagents capable of coordinating strongly with the cation, for example, DMSO and hexamethylphos- phoramide(42’u7). The lithium salt was singled out for these studies because of the strong specific interaction of the small Li+ ion with solvent molecules permits studies of low polarity media in which the Na+salt exists only as a contact ion pair. Moreover, the overlap of the absorption peaks corresponding to the two types of ion pair (contact and solvent separated) is the smallest for Li+ salt, and this facilitates the calculation of equilibrium constants between contact and solventdseparated ion-pairs. The enthalpy change, AH, governing the equilibrium between the two kinds of ion pairs is a significant thermodynamic quantity, knowledge of which permits the discussion of energetics of ion-pair separation. It also plays an important role in determining the temperature dependence of the reactions in which the two kinds of ion pairs simul- taneously participate, each with its own rate constant found that lithium anthracenide in diethylether gave spec- tral features for the normal ion pair showing cation hyper- fine coupling together with a species showing no cation coupling, but having proton splittings that were different from the normal solvated ion values. A more definitive (71) for sodium naphthalenide study is that of Hofelmann et al. in THF. Contact ion-pairs [A(23Na) = 1.23 G] together with another type of ion pair, solvent separated, [A(23Na) = 0.38 G] were detected when tetraglyme was added to the solution. Excess glyme gave only the latter species. Triplet-ions formation has also been detected by ESR. The addition of sodium tetraphenylborate to sodium-2,5-di-t— butyl-pebenzoquinone in THF resulted in a spectrum charac- teristic of the triplet ion(72). 95 D. Nuclear Magnetic Resonance (NMR) In recent years the use of NMR in the study of electro- lyte solutions has become quite popular. Proton and 130 NMR have been widely used for the studies of ionic inter- actions in solutions. At low temperatures,exchange between bulk water molecules and those in the inner solvation shell ofzacation is slower than the NMR time scale and, therefore, two separate resonances are observed for the free and bound water. Hydration numbers for a number of transition (73) metal cations were obtained by using acetone-water sol- vent mixture, acetone was assumed to act only as a com- pletely inert diluent. Alkali metal NMR as well as halogen NMR has been par- ticularly fruitful in the elucidation of the structure of alkali salt solutions in nonaqueous solvents. A compre- hensive study of sodium iodide solution in nonaqueous <7u>_ solvents has been reported by Bloor and Kidd Twelve solvents were used and chemical shifts varied from +9.9 ppm (upfield) shift for acetic anhydride to -l3.l ppm (down- field) for ethylenediamine, the chemical shifts were re- lated to the aqueous pKa of these solvents. More detailed studies by Popov gt al.(75’76) of different salts indicated that the counter-ions also play an important role in determining the behavior of the 23Na chemical shifts. The concentration dependence of 23Na chemical shift in the sodium halides and thiocyanate solutions in different 96 solvents can be ascribed to the formation of contact ion- pairs in these solutions<77). The 7Li nucleus has a spin of 3/2 and thus is amenable to NMR studies. Although the nucleus has a quadrupole 7 moment, the Li resonance lines for Li+ ion are exception- ally narrow and the chemical shifts can be measured with considerable accuracy. (78) (79) Studies by Maciel E; El- and by AkittanuiDowns in several nonaqueous solvents showed that the frequency of the 7Li resonance is very sensitive to the environment. While a large number of 7Li, 23Na and 13303 NMR studies have appeared, 39K and especially 89Rb NMR investi- gations are scarce due to the low sensitivity in 39K and the broad resonance lines obtained with 89Rb. In dilute solutions of'lithium perchlorate in nitro- methane and THF the width at half-height of the 3501 resonance was 10-20 Hz. Considerable broadening was ob- tained with increasing the salt concentration in both sol- vents whereas, very little dependence was observed in acetone, methanol and acetonitrile solutions. This broad- ening was due to the change in the electric field gradient at the chlorine nucleus as the perchlorate ion interacts with Li+ to form a contact ion pair<80). 97 E. Luminescence In comparison with the other spectrOSCOpic methods, luminescence spectroscopy was little used in studying ion- pairs. Much information can be obtained using luminescence techniques. Due to charge redistribution in the excited state, ion-pair characteristics will be different in the excited state compared to the situation in the ground state. Excitation may cause a shift in the equilibrium between free ions and ion pairs i.e., a change in the dissociation constants. (81) used the phosphorescence lifetimescn?phenyl- McGlynn carboxylic acids and their saltstx>show that the lowest triplet state of benzoic acid is not an excitation localized on the benzene ring but involves considerable delocalization onto the carboxyl group. The large spin-orbit coupling effect caused by the heavy lead ion in the lead salt of benzoic acid was completely lost in the lead salt of A- phenyl-butyric acid, in the latter case the benzene ring is separated from the carboxylic group by three methylene groups. These observations also show that the observed effect of lead on the lifetime of the benzoic acid salt is not an external heavy—atom effect on the benzene ring. Recently, Carter and Gillispie<82) studied the luminescence of ion pairs of alkali-metal cations with dichromate anion. The influence of other alkali-metal cations on the emission of sodium dichromate was probed by adding their chloride 98 salts to the sodium dichromate solution. The wavelengths of the emission maximum in aqueous solution at 77°K de- creased continuously in going from Na+ to Cs+. The emis- sion was assigned as phosphorescence. The greatly en- hanced phosphorescence intensity when K+, Rb+ or Cs+ cations are present reflects the heavy-atom effect on the spin-forbidden process (triplet + singlet). The magni- tude of any heavy-atom effect is presumably sensitive to (46). If the ion whether the ion pairs are loose or tight pairs are loose the heavy-atom effect may be an external one whereas, in the tight ion pairs case it is better described in terms of an internal heavy-atom effect, i.e. because of orbital overlap between the already bended emitting species and heavy atom not through collision mechanism with the heavy atom. Accordingly,the emission at 77°K was assigned to ion-pair formation between the alkali metal cation and the dichromate anion. III. Results and Discussion The room temperature absorption spectrum of a dilute solution of sodium salt of salicylaldehyde in DMSO exhibits two absorption bands, Figure (“2). The band absorbing at 417 nm corresponds to the absorption of the free anion. A solution of the same concentration of tetrabutylammonium salt of salicylaldehyde in DMSO absorbs at A17 nm, the bulkiness of the tetrabutylammonium cation prevents 99 .A V ammmz z muoa w m co monomoaa ca omza 2H was AIIIIV Oman ca momnooamazodHMm mo pawn +mz no A: mica x my mcofiuzaom canaao no mppooom coaponomnm chapmnoQEmp+ Eoom .m: opsm«m “ES £32325 00¢ 00? 00m 00m 0. O N O eouquosqv 100 ion—pair formation. The other absorption at 323 nm is interpreted as corresponding to the absorption of the as- sociated metal cation with the organic anion. This inter- pretation is consistent with the increase in absorptivity of the 323 nm band upon addition of Na+ ions while that of the ul7 nm decreases as shown in Figure (”2). Similar behavior was observed for the thallous, Tl+, salt of salicylaldehyde, Figure (“3). In the presence of excess T1+ ions, the maximum of the A17 nm absorption band was blue shifted, this probably reflects a change in the solvent structure while the absorptivity of the 323 nm absorption band is enhanced, Figure (A3). In water the 323 nm ab- sorption band is weak and appears only as a shoulder. Figure (AU) shows the potential energy of an ion pair as a function of the interionic distance R. The absorption of the free anion is represented by AE. For the formation of an ion pair, the positively charged cation gets closer to the anion, the smaller the ionic radius of the cation, the closer the distance between anion and cation, resulting in a greater stabilization of the ground state of the ion pair. Because of the covalency in O-H bond, maximum stabilization corresponds to the formation of the phenolic derivative. In the excited state, the negative charge on the oxygen will have partially migrated to the ring, and will decrease the negative charge on the phenolic oxygen. Accordingly, less interaction between the cation and anion 101 8 ,C: 3 _ .8 .8 <1 0.0 b l 1 l 1 9 300 350 400 450 Wavelength (nm) Figure H3. Room temperature absorption spectra of dilute solutions (5 x 10'5 M) of Tl(I) salt of salicyl- aldehyde in DMSO (----) and in DMSO in presence of 5 x 10-3 M T1N03 ( ). 102 AB, ©—0'--~ M“ 2§J59'()"fi chation-anion interionic distance Figure “A. Potential energy of ion-pair formation as a function of the interionic distance. 103 will occur, resulting in an increase of R and less stabi- lization of the ion pair in the ground state. During the absorption,the interionic distance does not change and therefore a Franck-Condom energy contributes to the transi- tion energy. The result is a blue shift in the absorption spectra of the ion pair, with respect to that of the free ion. This blue shift increases with the decreasing cationic radius. The blue shift of the “17 nm band, Figure (A3), to MOO nm in presence of excess (5 x 10'3 M)Tl+ indicates that the solvated cation interacts with the solvated anion, i.e., solvent separated ion pairs absorbing at A00 nm are formed. Thus the following equilibria occur in solu- tion M+SSL- : M+L' M+S + L‘s —> . +. solvated free ions solvent separated ion contact ion pairs pair (417 nm) (MOO nm) .323 or 326 S: solvent M+z Na+ or Tl+ L‘: salicylaldehyde anion In ethanol, both sodium and thallium salts of salicyl- aldehyde show an equilibrium between the free ions which absorb at 380 nm and contact ion-pairs which absorb at 325 nm. Thus adding an excess of Tl+ shifts the equilibrium to the right; the relative concentration of the contact ion-pairs increased. In addition, excess Tl+ induces 10A structural changes in the liquid structure of the solvent. The enhanced order in the solvent structure favors the formation of solvent-separated ion-pairs at the expense of solvated free ions. In THF solution of the sodium salt of salicylaldehyde, the longer-wavelength absorption band dominates the spectrum as shown in Figure (H5). This indicates that in THF there is an equilibrium between the free ions and contact ion-pairs with the equilibrium shifted towards the free ion formation. This can be ex- plained in view of the high solvation of Na+. For a thal- lium salt in THF, Figure (N6), both absorptions of the ion pair and free ions are of comparable absorptivity. This reflects the lower solvation of the Tl+ ion with THF com— pared with Na+ because of the larger radius of Tl+. Figure (N7), shows a comparison between the absorption spectra of Na*and Tl+salts in THF. When the anion of salicylaldehyde is excited, charge transfer occurs from the phenolate oxygen to the carbonyl oxygen. As a result,the negative charge density on the phenolate oxygen will be smaller in the excited state than in the ground state. This weakens the interaction between the positive and negative ions in the excited state‘ i.e., the ion pair becomes less stable, and may lead to dis- sociation of the ion pair in the excited state. Figure (nu) shows the energetics of both excited and ground states of ion pairs. Figure (H8) displays the room temperature Absorbcnce 105 l l l l 250 300 350 400 Wavelength (nm) Figure #5. Room temperature absorption spectra of dilute solution of Na+ salt of salicylaldehyde: (H. 8 x 10-5 M) in water (----) (u. 5 x 10-5 M) in ethanol ( ----- ) and (u. 5 x 10-5 M) in THF 106 04—/I Q) U C £302. 3 (D 13 <1 - 0.0L 1 l l l 250 300 350 400 Wavelength (nm) Figure H6. Room temperature absorption spectra of dilute solutions of Tl(I) salt of salicylaldehyde: (u.5 x 10-5 M) in water (--——), ( .2 x 10-5 M) . inethanol ( ----- ) and (H.5 x 10- M) in THF 107 .ooscooaoflsofiaon co ooflon A . VmHa coo Alluuv+oz co mma an O A: mloa x m.zv mcofipzaom ousafic mo caveman Onwv “panomoa madcapoasop Boom .ha ohsmfim 3.5 508853 0mm 00m 0mm . _ _ 0. o 4 1 N O eouoqiosqv Fluorescence Intensity Figure H8. 108 l___l I 4 I 400 450 500 550 600 Wavelength(nm) Room temperature emission spectra of dilute solutions of Na+ salt of salicylaldehyde (5 x 10-5 M) in DMSO: lexc = 320 nm (--—-), Hexc = 380 nm ( ) and in DMSO in the pres- ence of NaBFu (2 x 10'2 M),A exc 8320 (- ----- ). 109 emission spectra of the sodium salt of salicylaldehyde in DMSO. Exciting at the absorption maximum of the free anion produced an emission from the free anion at H80 nm. Ex- citation at 323 nm, the absorption maximum of the contact ion-pair gave two emissions. The shorter-wavelength emission at H05 nm belongs to the associated pair and the longer-wavelength emission at H80 nm to the free anion. 2 M) Na+ ions, excitation at In the presence of (2 x 10- 323 nm gave only the H05 nm emission whereas, excitation of the same solution at 380 nm produced only the free anion emission. Thus, the production of two emissions when exciting at the absorption band of the contact ion pair indicates that the contact ion pair dissociates as a result of excitation i.e., the equilibrium shifts towards free ion formation in the excited state. Similar results were obtained in the corresponding thallium (I) salt. In a dilute solution of the thallium salt (5 x lo“'5 M) in DMSO in the presence of (10'3 M) T1 N0 Figure (H9), the 3, anion emission was observed upon excitation at the longer- wavelength whereas, fluorescence quenching occurred when at the absorption band of the contact ion-pair was excited, this may reflect the high degree of covalency in the metal cation-salicylaldehyde anion ion-pairs. Since in the same solution the observation of the intense anion emis- sion eliminates the quenching through collisional mechanisms by the excess T1+ ion present in solution, i.e., the 110 Fluorescence Intensity jf, l. 1 l 1 ° 'T' I 350 .400 450 500 550 600 650 Wavelength (nm) Figure H9. Room temperature emission spectra of dilute solu- tions (5 x 10-5 M) in DMSO of: Tl(I) salt of salicylaldehyde, Aexc = 320 nm (----) and H20 nm ( ); Tl(I) salt of salicylaldehyde in pres- ence of ~5 x 10-3 M T1N03, lexc = 320 nm (-----). 111 (“6). In quenching is via internal heavy atom effect solutions of Na+ and Tl+ salts of salicylaldehyde in THF, Figure (50), both gave the free anion emission at 500 nm when excited at 320 nm where the ion pair absorbs, or at 380 nmwhere the free anion absorbs. The red shift of the emission of the anion in THF compared with that in DMSO at H80 nm can be interpreted as a result of more stabilization of the ground state in DMSO than in THF. Excitation of the anion leaves a smaller negative charge on the phenolate oxygen in the excited state compared with the ground state charge distribution. In the presence of excess metal ions the emission spectra started to show ion pair emission peaks at ~H20 nm as well as the anion emission. The intensity of this emission was higher for the Na+salt than for the K*salt i.e., the shift in the equilibrium between the ion pairs and free ions is more shifted towards ion pair formation in the case of Tl+than Na salts. In aqueous solutions,complete dissociation,at least in the excited state,gave only the anion emission. In this section, the results of our study of the interaction of monovalent alkali metal ions with three phenols,name1y: p-hydroxybenzaldehyde, p-methoxybenz- aldehyde,and p-methylphenol are summarized. Para-sub- stituted phenols were chosen to contrast possible inter- action unique to ortho-substituted derivatives. It should (1“) be pointed out that Zaugg and Schafer studied the 112 Fluorescence lntenslty_ Figure 50. 400 500 660 Wavelength (nm) Room temperature emission spectra of dilute solutions (5 x 10'5 M) of Na+ and T1+ salts of salicylaldehyde in THF. Na+ salt, Hexc = 320 and 390 nm ( ); Na salt in presence of ~2 x lo-2 M NaBFu,A e . 320 nm (----); Tl(I) salt, Aexc = 320 and 390m W§ -); and Tl(I) salt in presence of ~5 x 10- M TlNO3, Hexc a 320 nm (-------). 113 cation and solvent effects on the ultraviolet absorption spectra of alkali metal salts of some phenols and enols. They reported a red shift in the absorption spectrum of alkali salts of a phenol in dimethoxyethane as the radius of the cation is increased. These results were inter- preted in terms of an increased interionic distance, and hence smaller interaction between the anion and cation forming the contact ion pairs. The absorption spectra of dilute solutions (5 x 10'5 M) of p-hydroxybenzaldehyde and its Li+,Na*and K*sa1ts in THF are shown in Figure (51). The phenol absorbs at 272 nm. When LiH was added to the phenolic solution, the obtained absorption spectrum showed two bands: one identical to that of the phenol and another band with maximum at 328 nm. The shorter-wavelength absorption band is due to unreacted phenol; while the absorption band at 328 nm is due to ion- pair formation. Addition of NaH or KH to the phenolic solution produced the sodium or potassium salt respectively, and the reactions were complete. The absorption maxima occurred at 33H nm for the Né’salt and at 3H7 nm for the K+ salt. The increase in the radius of the cation leads to progressive red shift of the absorption of the ion pair from 328 nm for Li" salt to 3H7 nm for the K"sa1t. The free anion of p-hydroxybenzaldehyde absorbs at 360 nm in THF as obtained from the absorption spectrum in the pres- ence of tetrabutylammonium hydroxide. When 18-crown-6 11H I f)? F? ':\\ .. U. U. U - H. !l H " o I; o ., C l} O ': _ .0 ‘.l L 2 O H m 2 .o l‘ <[ | " . l‘= l.‘= .. \‘-. .. \\ \I .. (1‘)]. l 1 ”."‘m2. .— 250 300 350 Wavelength (nm) Figure 51. Room temperature absorption spectra of dilute solutions (5 x 10'5 M) of alkali-metal salts of p-hydrOxybenzaldehyde in THF. P-hydroxy- benzaldehyde ( ); Li+salt (....); Na+salt (- ----- ) and K+salt (----'--). 115 was added to metal salt solution, no effect was observed for the Li+sa1t. Li+ does not form stable complexes with 18- crown-6. However, in the case of the Na*and KIsalts, red 1 and 900 cm"1 shifts of 500 cm' respectively,were ob- served. The stability constant of the K+ complex with 18-crown-6 is larger than that of Na+. These red shifts are interpreted in terms of the crown ether forming a complex with metal cation in the ion pair. Such inter- action will decrease the interionic separation,decreasing the perturbation influence of the positive charge of the cation,causing a red shift, i.e., the spectrum approaches that of the free anion. The room temperature emission spectra of p-hydroxybenzaldehyde salts in THF are shown in Figure (52). The emission maxima of the Lit Na+and K+ salts occur at 357, 361,and 365 nm respectively. It is clear that the emission maximum undergoes a red shift with the increasing radius of the metal ion. In the presence of l8-crown-6 both Na+and K+sa1ts produced the same emission maximum at 372 nm which is the same emis- sion of the free anion in THF. These results indicate that the ion-pair, weakened by cation interaction with 18-crown-6, dissociates upon electronic excitation giving rise to the excited free anion. Comparing the absorption and emission spectra of the salts of ortho and para hydroxybenzaldehyde leads to the conclusion that a tighter ion pair is formed with salicyl— aldehyde and that a metal chelate is formed with the former 116 Fluorescence Intensity Figure 52. Wavelength ( nm) Room temperature emission spectra of dilute solutions (5 x 10'5 M) of alkali-metal salts of p-hydroxybenzaldehyde in THF. Li*salt, 1e c = 335 nm (--—-), K+salt, He c = 3H5 nm ( 1 and K+salt in presence of 10-3 M 18C6, Ae c = 3H5 nm (--—----), Na+salt l = 3H0 nm (....)‘, Na+ salt in presence of 10‘3eMc18C6, Aexc = 3H0 nm (------ 117 where the ortho-substituted anion acts as a bidentate ligand with both phenolate and carbonyl oxygen atoms. Similar results were obtained for p-methoxyphenol, the absorption spectra of its metal salts are shown in Figure (53). Addition of LiH did not change the absorp- tion spectrum of the phenol, it appears that no salt for- mation occurred. The absorption maxima of the Na+and K+ salts indicate that ion pairs are formed in THF solu- tions. The corresponding emission spectra of the two salts are displayed in Figure (5H). The emission spectra of both Na+and K+ salts are blue shifted compared to that of the parent phenol. When l8-crown-6 was added to the Na+ salt in THF, a new emission was observed at HHO nm, but the original emission at 36H nm of the salt while dimin- ished in intensity still occurred. In the case of the K+ salt, addition of l8-crown-6 led to the disappearance of the 367 nm emission of the ion pair and only one emission band was observed at HHO nm. The latter emission cor- responds to the free anion emission obtained in THF in the presence of tetrabutylammonium hydroxide. These results are interpreted in terms of the change in the stability of the ion pair upon electronic excitation. In the excited state, as mentioned before, the charge density at the phenoxide oxygen is diminished due to the charge migration to the benzene ring. This weakens the interaction with the metal ion. However, this effect does not manifest Absorbance 118 0.8 - 0.6 - I 0.2 Figure 53. 2:50 360 350 Wavelength (nm) Room temperature absgrption spectra of dilute solutions (1.8 x 10' M) of alkali-metal salts of p-methoxyphenol in .THF. P—methoxyphenol (-———), Na+salt (~-- - -) and K+salt (-- - ). 119 l I l )5 3: 1 (n a. g [I \\ *- .’ / ‘ 3 .EE 3 I \\.H 0) \ ,3 I \ 1 o \ :' ’ ‘ g: . I / \'z § \ / \ \f\/ ‘{ e ' \ X \ O . / °\ \ . 3 )‘ '\ \\'"-. LL / .\. .\ \.\' / .\,\\. L l l N l 300 _ 350 400 450 500 Wavelength (nm) Figure 5H. Room temperature emifision spectra of dilute solutions (1.8 x 10' M) of alkali-metal salts of p-methoxyphenol in THF. P-methoxyphenol, lexc =- 290 nm ( ); Na+salt, xexc -- 320 nm (-'---); Natsalt in presence of 10-3 M 18C6, lexc - 320 nm (....); xtsalt, lexc = 33; nm (-°°-°--) and K*salt in presence of 10' M 18C6, Aexc = 335 nm (----). 120 itself except in the presence of the crown ether. Thus, in the excited state, the crown ether competes effectively with the phenoxide anion for the metal cation and is particularly true in the case of K+ which has a larger radius and forms a weaker bond with the anion and also because it forms a more stable complex with the crown ether compared with Na+. Thus, the two emission bands observed in the case of the Na+salt in the presence of 18- crown-6 are due to partial dissociation of the ion pair in the excited state. In the case of the K+sa1t in the presence of 18-crown-6, complete dissociation of the ion pair occurs. The absorption and emission spectra of alkali metal salts with p-methylphenol in THF are shown in Figure (55) and (56), respectively. Ion-pair formation occurs in THF and their absorption maxima occur at 310 nm for Na+ and 317 nm for K1 In the presence of l8-crown-6 a further red shift occurs for K+giving an absorption at 326 nm. The emission spectra gives rise to a maximum at 3HH nm and 355 nm for Na+and Kt respectively. This should be com- pared with 360 nm for the free anion emission in THF. Table (9) summarizes the absorption and emission spectral results of the three phenols in THF. From the previous discussion, it is clear that mono- valent cations form tight ion pairs in THF. The interionic separation increases with the size of the cation giving 121 k u 0.8“ ‘ 06* " a) o 8 F l - .{3 l o ‘ - (n 0.4 F l 5 : l <1 .- . \ q ° l = l 02" ' = ‘ '1 . ( \ - '\ \\ .- '\ '\\\ 0.0 L l \l‘ O. " 250 300 350 Figure 55. Wavelength (nm) Room temperature aberption spectra of dilute solutions (2.5 x 10' M) of alkali-metal salts of p-methylphenol in THF. P-methylphenol ( Na*salt (-----), K+salt (—~----) and K’salt in presence of 10'3 M 18C6 (----). ), 122 I Z.‘ "(5 c m 15 a) o c m 0 U) 9 8 LT. .. l l ." ..T 300 350 400 Wavelength (n m) Figure 56. Room temperature emission spectra of dilute solutions (2.5 x 10' M) of alkali-metal salts of p-methylphenol in THF p-methylphenol, Aexc = 290 nm ( ); Natsalt, Aexc = 320 nm ( ----- ) and Kisalt, Aexc = 330 nm (-------). 123 .mlczoholmfl ..u mOwH mmm mmm o=.o Nam x + .o . m zem mm o cam + z .+<‘ mom Hm.o «mm m + are 0:: eom amm om.o amm +m . 4 no one arm .zom mmm :m.o mmm +nz :4 0mm mm.o mam +m + a o . . mam mom smm mm.H eam +m a o . mam H m o m N.H mm oz o a w . a + ...2 O emm me.o mmm +Hq r u 3.0 New ...m oe 1: coma coma Aecvxw8< Ascvme< AEcvmeK aloaxxmso Aecvxms +2 unsoQEoo coamMHEm cofiuopomn< .mme ca modem saose.oco maococm condufiumozmlmsmm go mafixmz COHMMfiEm 62m :oHuapown< manpwpoQEoB Eoom .m canoe 12H rise to a red shift in going from Na+ to K+. In the pres- ence of 18-crown-6 it interacts with the cation resulting in a looser ion-pair formation. In the excited state the ion pair becomes weaker due to a decrease in the negative charge on the phenoxide oxygen and in the case of bigger cation (K+) in the presence of l8-crown-6 complete dis- sociation of the ion pair may occur. CHAPTER V EUROPIUM (III) CHELATE WITH SALICYLIDENE-VALINATE SCHIFF BASE I. Lanthanide Ions as Luminescence Probes of the Structure of Biological Macromolecules Calcium and magnesium occur as essential inorganic ions in living systems. A Ca2+ concentration gradient is maintained across many cell membranes with the concentra- tion within the cell lower than that outside. Ca2+ is involved in muscle contraction, blood clotting, K+ trans- port, neurotransmitter release, cellular adhesion, inter- cellular communication, microtubule formation, and hor- monal responses. Many of these Ca2+ related activities occur by means of interaction with proteins, which Ca2+ may stabilize, activate, and modulate. In this way Ca2+ plays a significant regulatory role in many biological processes. Because of the lack of suitable physical techniques for studying Ca2+, it has not received the attention it deserves. The electronic transitions of Ca2+ cannot be studied by conventional optical absorp- tion and emission spectroscopy, and the absence of un- paired electrons precludes the use of magnetic resonance 125 126 techniques in probing the chemical and structural nature of Ca2+ binding sites. Fortunately, about 12 lanthanide ions (Ln3+) possess properties which make them excellent probes for Ca2+(83). 2+ and Ln3+ In forming complexes, both Ca prefer charged or uncharged oxygen donor groups to nitrogen donor atoms. In aqueous solution, except for some multidentate ligands, hydroxo complex formation almost always occurs before amine (8“). Both Ca2+ and Ln3+ nitrogen coordination takes place display a variable coordination number and a lack of strong directionality in binding donor groups. Trivalent lan- thanide ions Ln3+ exhibit effective ionic radii that show a gradual contraction from one end of the series to the 3+ (atomic number 57) other, for example, from 1.16 K for La to 0.98 K for Lu3+ (atomic.number 71) in 8—fold coordina- tion(85). For a given coordination number, the decrease in the ionic radii in going from one end of the lanthanide series to the other (0.17 - 0.19 K) is about equal to the increase in ionic radii on going from 6- to 9-fold coor- dination. Thus, there are Opportunities for fine adjust- ment upon substitution of heavier Ln3+ for Ca2+ ions either by small decreases in ionic radii or by an increase in co- ordination number. As an example, the ionic radius of 7- coordinate Ca2+ is nearly equal to the ionic radius of 8- coordinate Eu3+ ion. The replacement of Ca2+ by Ln3+ has been demonstrated by x-ray diffraction to occur in two (86a) (86b). proteins, carp parvalbumin and thermolysin 127 In the latter case, the results have been reported with sufficient precision to allow some comparisons of bond distances and coordination numbers between Ca2+ and Ln3+ containing proteins. Accordingly, most Ln3+ should be able to substitute Ca2+ in proteins without causing serious structural modifications in the active sites. The unit charge difference between Ca2+ and Ln3+ is apt to be of 3+ for Ca2+. secondary importance for substitution of Ln In biological systems Na+ and Ca2+ of comparable ionic radii have been found to be competitive for sites as have K+ and Ba2+<87’88). Charge differences may assume more importance in rate phenomena which may be sensitive to net charge at a binding site. Except for La3+ and Lu3+, all 12 other available Ln3+ ions contain unpaired f-electrons, and both nuclear mag- netic resonance (NMR) and electron spin resonance (ESR) methods may be used to investigate their environment<89a’b). 3+ may be used as chemical shift probes in NMR. Several Ln cd3+ with half-filled Hf subshell is used as a broadening probe in NMR and also is employed in ESR studies. These 12 Ln3+ ions also exhibit absorption spectra due to inter- configurational f-f transitions which are easily acces- sible to study by conventional optical absorption tech- niques. However, the molar absorptivity associated with these transitions are generally so low (of the order of unity) that the study of Ln3+-protein systems in which Ln3+ 128 concentrations are smaller than 10'2 M, optical absorption spectroscopy can be performed only near the limits of instrumental sensitivity. 0n the other hand, luminescence due to interconfigurational f-f transitions in Ln3+ ions bound to protein systems remains well above detection limits 3+ even when Eu and Tb3+ are present at concentrations as low as 10.6 M. Thus a factor of at least 10Ll in sensitivity favors luminescence over absorption spectroscopy. In order for luminescence to occur, appropriate excited states of the emitting (luminescent) species must be populated. The use of Ln3+ 3+ ion emission spectra as a probe of Ln complexes requires, then, excitation of the Ln3+ f-f emitting states. Excitation may be accom- plished either by direct excitation of the metal ion chromOphore (the exciting light being absorbed directly by the Ln3+ ion) or by an indirect mechanism involving optical excitation of ligand chromophores follows by sensitization of Ln3+ emission via radiationless energy transfer (ligand to metal ion). Because of the weak absorptivities of Ln3+ transitions in the visible and near-ultraviolet spectral region, direct excitation of Ln3+ emission requires either a powerful excitation source (such as a laser) or relatively high concentration of the metal ion (>lO"2 M). 0n the other hand, indirect excitation of Ln3+ emission may be accomplished using somewhat less intense exciting light and at lower metal ion concentra- tions if the ligand environment includes a highly absorbing 129 chromophore capable of acting as an efficient energy donor to the Ln3+ acceptor-emitter species. Both the direct and indirect excitation methods have been used in emission studies of Ln3+-protein complexes. In the indirect excitation method, in Ln3+-protein 3+ systems, the Ln ion may be excited by energy transfer from nearby aromatic side chain chromophores which are excited directly by near-ultraviolet radiation. The over- all process is thus started by an absorption of an aro- matic side chain of phenylalanine, tyrosine, or trypto- phan in the 250-300 nm region; followed by a non-radiative energy transfer from an aromatic side chain to a nearby Ln3+; giving rise to a strongly enhanced (up to 105) Ln3+ emission in the visible region. The energy transfer step probably occurs by‘a Forster dipole-dipole reson- ance transfer mechanism<90> with a r.6 dependence on the distance between the donor and acceptor sites. A semi- quantitative analysis suggests that 50% probability of energy transfer occurs at a donor to acceptor distance of 5-10 3 for tyrosine-Tb3+ and tryptophane-Tb3+ pairs<9l). Using Tb(III) or Eu(III) as an acceptor, Horrocks gt gl,(9l) measured the distance between tryptophane as an energy donor and the calcium-binding sites in parvalbumin from codfish. The 11.8 and 10.2 K distances obtained using Tb3+ and Eu3+ as acceptor respectively, are in a good agreement with the 11.6 A distance estimated from x-ray structural results. 130 Direct excitation of Eu3+ and Tb3+ emission using a pulsed laser source was employed in the study of lumin- escence lifetimes for a variety of complexes in aqueous (92). solution The lifetime and emission decay charac- teristics of Eu3+ and Tb3+ with a variety of ligands (in- cluding the protein thermolysin) were measured as a func- tion of D20 versus H20 content of the water solvent. Such studies gave estimates of the number of water mole- 3+ cules bound to the Ln in each complex<92). This follows from the somewhat different perturbative influences of D20 versus H20 upon the rate of radiative and radiationless pro- cesses of Eu3+ and Tb3+ excited states(93). Eu3+ and Tb3+ are the most strongly emitting members 3+ of the Ln3+ series. Energy level diagram for Eu and (11). Tb3+ are shown in Figure (57) The 5D and 5D“ emis- 0 sion levels of Eu3+ and Tb3+ can be excited by 579 and H88 nm light, respectively. The use of visible light for the excitation eliminates problems of protein-ultra- violet photosensitivity. The excited state lifetimes of these ions are environmentally sensitive and lies in the conveniently long 100-300 us range. While ligand field splittings of f-electron levels are much smaller than for the d-electron levels of transition metal complexes, the absorption bands are generally much narrower, and small splittings can be readily detected. Individual emission bands can be examined under high resolution to yield Figure 57. Energy (cfi' X l0'3) 131 z:- Europium(III) “0*" "04” Tctbium (m) '05 _. v07 ’0: "b ., W l ’°‘ n— ..4 ’°° l \- "" Is- "3 __ v“ - '03 '02- : E 5 5 C C o It)h . 2 _ I v 5 ‘.’ " : e 2. e «l- 1 e 2 .2 - vol .3 2 5» "‘ 0’0 - — '0 "o ‘ 7 7'2 '5 o- 'r. __ 7" Electronic energy levels for europium(III) and terbium(III). The two upward-pointing arrows show the transitions which occur upon laser ex- citation at the wavelengths indicated. The two downward-pointing arrows label the most intense emissive transitions of the two ions. Radiationless energy transfer competes with the radiative processes through coupling of the emissive states to the O-H vibrational over- tones of coordinated H20 molecules. (The energy level diagram of Tb(III) has been displaced slightly to position the highest electronic acceptor state of the ground 7F term to coin- cide with the zero-point energies of the vibra- tional overtone ladders.)11 132 information about the splitting of ground and excited states. Excitation spectroscopy using tunable lasers also yields information of this type. A. Decay Lifetimes as a Measure of the Number of Metal Coordinated Water Molecules The early observation that hydrated crystals show a much reduced fluorescence intensity in the middle of the series of rare earth ions (Sm, Eu, Gd, Tb, Dy) and prac- tically no emission in those at the beginning and end of the series, led to the conjecture(9u) that the highly energetic O-H vibrations in the crystal lattice act as energy sinks. The role of OH vibrations in the quenching of electronic f-f transitions in the rare earth transitions ' (95) was demonstrated in crystals by measuring fluorescence lifetimes of deuterated and protonated hydrous crystals of EuCl and TbCl . Quenching occurs if the energy gap 3 3 between the lowest fluorescent and highest non-fluores- cent level is matched by a single high energy (OH or OD) vibration of one solvent molecule. The decisive parameter is R, the ratio between the electronic energy gap (E) )(96). and the vibrational quantum energy (hm R = E/hw Of all the lanthanides Gd is in a class by itself with 133 R m 10 for H O and m15 for D20. 2 There is a very low probability of radiationless transition to the very high vibrational quanta thus re- (16) quired. Accordingly addition of a single high-energy vibrational quantum due to deuterium substitution make no significant contribution to the radiationless dissipa- tion of electronic energy. Hence, no isotOpic effect is observed in the case of Gd. For Eu3+ (E m 12,200 cm’l) 3 the isotope effect is the greatest of all Ln(III) series and is due to the requirement of v = H for OH and v = 5 for OD. Radiationless transition probability is not small and will differ greatly depending on whether H O or D O 2 2 is interacting with Eu(III). For Tb3+, (E m 1H,800 cm‘l) , radiationless transition occurs to v = 5 for OH and to v = 6 or 7 for OD resulting in smaller probability for 3+ the transition compared to the case of Eu and hence longer lifetime for Tb3+ is observed, and the isotopic effect is smaller. The experimental reciprocal excited-state lifetime (exponential decay constant), Toisd’ consists of several terms T-1 = T-l + T--l + -1 obsd nat nonrad T0H where T-1 is the natural rate constant for the emission nat nonrad represents the rate constant for of photons, T 13H nonradiative deexcitation which does not involve OH oscil- lators, and 16% is the rate constant for nonradiative energy transfer to the OH vibrational manifold of OH oscillators in the first coordination sphere (e.g., oo- 16% is very significant. For instance, for Eu3+ (aq), 16% = 9.5 ms-l, whereas, 1-1 nat = 0.19 ms‘l, Thinrad = 0.25 ms’1(93’97a). ordinated water molecules); Replacement of OH oscillators by the OD variety causes the vibronic deexcitation pathway to become exceedingly inefficient(97a'c) because of the smaller orbital overlap as the vibronic quantum increases, and enables one to determine the number of OH oscillators in the first coordination sphere by carrying out experiments independently on H20 and D20 solutions. In D20 solutions TOH vanishes and Tobsd — -1 -1 Tnat + Tnonrad’ where the latter term includes any small deexcitation via DO oscillators. 13%8d varies linearly with the mole fraction of H20, XHZO’ in H20-D20 mix- tures(92). Figure (58)(11), shows a plot of Tobsd vs ngO for EuCl3. In pure H20, nine water molecules are co- ordinated to Eu3+. When 1 mole of EDTA (ethylenediamine- tetracetate) was added the number of coordinated water molecules dropped to six. This is in agreement with the tridentate nature of EDTA which substitutes water mole- cules in the first coordination sphere. The same tech- nique was used to estimate the number of coordinated water molecules in crystalline solids.(98a) Figure 58. 135 " Europium(III) Plot of the observed reciprocal luminescence lifetimes, ragga d, vs. the mole fraction of H20, XHZO: in D2 0- H 0 mixtures of Eu(III) solutions. The effect 2of ghe chelating ligands NTA and EDTA is to displace water molecules from the first coordination spheres of the aquametal ions to yield the approximate values indicated along the right-hand ordinate.1l 136 B. Characterization of Individual Binding Sites Eu(III) ion is unique in that both the ground (7FO) and the emissive excited (5DO) states are nondegenerate. Since neither of these levels can be split by a ligand field, the absorption band corresponding to a transition between these levels must consist of a single, unsplit line for a given Eu(III) ion environment. Different ionic environments can, in principle, yield transitionszuzslightly different frequencies. Because 5D + 7F transition is O 0 highly forbidden, study of this band by ordinary absorption spectroscopy on dilute solutionsijsnot feasible. However, since 5D0 is emissive, this transition can be studied via excitation spectroscopy by monitoring emitted photons 5 7 ( D0 I Fz’ scanned through the 5D0 + 7F0 The position and shape of the excitation band will depend 612 nm) while a tunable laser is continuously transition region (578-580 nm). on the microenvironment around the Eu3+ ion. Figure (59) shows the excitation spectra resulting from the titration of the Eu(III) aqua-ion with tridentate ligand, dipicolinate (DPA)(98b). Individual narrow peaks characteristic of Eu3+ (aq), [Eu(DPA)]+, [Eu(DPA)2]’, and [Eu(DPA)3]3' are apparent at the following respective frequencies: 17272, 17263, 172u8, and 17231 cm'l. Eu(III) excitation spectroscopy was used to characterize the in- dividual metal ion binding sites in calcium-binding pro- teins such as thermolysin. These various Eu(III) coordination Figure 59. 137 C02 DRQLW=<:E: co; [EU (CPA 1313- J "31" \ ‘llEu‘DPAlzl- "71‘. l Excitation spectral profiles of the 7F0 + 500 transition obtained during the course of the titration of an aqueous EuCl3 solution with the sodium salt of the dipicolinate anion (DPA). Eu(III) to ligand ratios are indicated to the left of each trace.9 9 138 environments were further characterized by their excited- state reciprocal lifetime which can be used to determine the number of coordinated water molecules in each site. C. Inter Metal-Ion Energy-Transfer Distance Measurements If there is a significant overlap between the emission spectrum of an energy donor D and the absorption spectrum of an energy acceptor A, energy transfer of a nonradiative (99) kind can take place between D and A. F6rster showed that the efficiency of dipole-dipole energy transfer, E, is inversely proportional to the sixth power of the D-A separation, r, and is given by T and T0 are the excited-state lifetime of the donor in presence and absence of energy transfer. r, is the actual donor-acceptor distance, R0, the critical distance for 50% energy transfer 6 25 K2n-H 6 (R = 8.78 x 10' @J cm 0) 8.78 x 10-25 is the product of fundamental constants K2, the orientation factor n, the refractive index of the medium between the interacting ions 139 0, the quantum yield of the donor in the absence of the acceptor J, the spectral overlap integral J = ;E(v)e(v)s‘“dv fF(v)dv F(v) is the luminescence intensity of the donor at frequency (cm-l) 8(v) is the molar extinction coefficient of the acceptor (M'l cm'l) Ln(III)-transition metal ion and Ln(IIIyLn(III) energy transfer were applied in distance measurements. Recently, energy transfer involving Tb(III) to Fe(III) in transferin(loo) (R0 = 27.1 A, r = 2512 H) and Eu(III) to Co(II) in galactosyltransferase(101) (R0 = 20 l, r = 18 i 3 K) were reported. In neither case are confirmatory crystallographic studies available. Horrocks gt_gt.(ll) used the inter Ln(III) ion energy transfer to measure the distances between the calcium-binding sites in thermolysin. Thermolysin has four Ca2+ binding sites, S(l) to S(H), and one Zn2+ binding active site. The exchange-inert nature of Ln(III) ions occupying site 8(1) of thermolysin makes it possible to substitute a luminescent ion (e.g., Tb(III) or Eu(III)) at this site and a different absorbing ion (e.g., Nd(III), Pr(III), Ho(III), or Er(III)) at sites 8(2) and S(H). Nonradiative energy transfer can be 1H0 monitored by measuring the effect of the presence of energy acceptor ions on the excited-state reciprocal lifetime of the donor ion. The r values estimated from the energy transfer efficiencies are in good agreement with the X- ray distance of 11.7 K between calcium sites 8(1) and S(H), except for the Tb3+ - Nd3+ results due to some contribution from a dipole-quadrupole mechanism of energy transfer. 11. Energy Transfer in Rare-Earth Metal Chelates (102) in 19u2 that B— It was discovered by Weissman diketone coordination compounds of tervalent europium, terbium, and Samarium exhibit unusual luminescence proper- ties when excited by near ultraviolet light. The compounds emit visible radiation characteristic of the rare-earth ions. Thus, intra-Hf electronic transitions, which are known to originate from levels derived from the electro- static and spin-orbit interactions among the Hf electrons within the rare-earth ions, occur whenever the excitation is carried out in the intense ligand absorption bands. These ligand bands (molar extinction coefficient ~10“ to 105) are n-n* in nature and are characteristic of the co- ordinated ligands surrounding the central chelated ion. Weissman realized that the energy was being pumped into the electronic system of the complex characterized by the w-electronic levels of the ligands and was subsequently migrating to the control chelated rare-earth ion, from 1H1 whence characteristic luminescence (line emission) of the ion occurred. He designated this process intramolecular energy transfer. A. Paths of Energy Migration A schematic energy level diagram for rare earth chelate possessing low-lying Hf electronic state is given in Figure (6O)(103). After excitation of a chelate to a vibrational level of the first excited singlet state (S0 + Sl)’ the molecule undergoes rapid internal conver- sion to the lower vibrational levels through interaction with the solvent matrix. The excited singlet state may be deactivated by combining radiatively with the ground state (80 + 81), resulting in molecular fluorescence, or the molecule may undergo radiationless intersystem cross- ing from the singlet to the triplet system. Again by internal conversion the molecule may reach the lowest triplet state, Tl' From this state, it can then combine radiatively with the ground state by means of a spin- forbidden transition (SO + T1) giving rise to a typical long-lived molecular phosphorescence. Alternatively, the molecule may undergo a radiationless transition from the triplet system to a low-lying rare-earth ion state(loua). The latter states are derived from the Hf electronic con— figuration of the coordinated trivalent rare—earth ion. After this indirect excitation by energy transfer, the Figure 60. 1H2 Rare Earth Sunglat Triplet Ion States 511% ENERGY Schematic energy level diagram for a rare earth chelate possessing low-lying Hf electronic states: ._,, radiative transitions;.~qh radiationless transitions.lo 1H3 metal ion may undergo radiative transition to a lower ion state resulting in a characteristic line emission, or it may be deactivated via radiationless processes. Direct trans- fer of energy from the excited singlet state to the low- lying rare-earth ion states has been shown to be unim- portant~ 4... 'fi c 0. ~9- \. 5 ll a) ,‘ l. o l C: ' . CD \ Q) . m l. g l 2 \‘\ L’— .\.J I’ ----l- 350 450 550 650 Wavelength (nm) Figure 66. Emission spectra of dilute solutions (3 x 10'5 M) of KEEu(C 2H13NO3) ] in ethanol at room temperature 1----), e hanol at 77°C ( ----- ), and DMSO at room temperature ( ). 158 and phosphoresces around (20,500 cm-l) which corresponds (122) to the energy of the lowest triplet state in these chelates. In Figure (67) we constructed an energy level diagram which shows the important radiative and radiation- lesswtransitionswithin the complex. Intramolecular energy transfer occurs from the ligand lowest triplet (as an 5 l or D0 of the metal (as an energy acceptor) followed by radiative transition (fluorescence) energy donor) to the 5D to the 7F. Two necessary conditions are required for sharp fluorescence from the rare-earth chelates through indirect excitation of the rare-earth ion via energy trans- fer mechanism: (1) at least one acceptor level should lie below the lowest triplet (donor) level of the isolated ligand, (2) the lifetime of triplet state should not be too short to achieve a good inversion and not too long to retain the inversion while waiting for the action to com- mence. In water, the metal complex is completely hydrolyzed to give both absorption and emission spectra characteris- tic of the free ligand. In DMSO, DMF,and acetonitrite red emission from the metal ion was observed at room tem- perature; whereas,in ethanol this emission was only ob- served at 77°K. This can be due to the difference in the stability of the metal chelate in these solvents. In highly coordinating solvents DMSO, DMF, and acetonitrile, the complex is more stable, maylxsthrough solvent coordina- tion which increases the coordination number to 8 or 9. 159 __ D. _L 50 5.7 3:32:09:— co. 335+ _ ___lI = 653| 7:: 420 .0: OquLonuO££ 7 an.“ 2 20- Arc. 11.. , 02.9.3933.» 1:00: l0“- 0 EB 335 Energy level diagram for KEEu(C12Hl3NO3)2]. Figure 67. CHAPTER VI CONCLUSION AND FUTURE WORK In the carbonyl compounds, the emission spectra showed an excited state proton transfer. No ground state proton transfer occurs in these compounds as evidenced by their absorption spectra. Both ground-state and excited-state proton transfer were observed in salicylaldehyde Schiff base derivatives with n-butylamine and valine. Depending on the solvent, intramolecular or intermolecular proton trans- fer was observed. In nonpolar solvents, e.g., 3MP, intra- molecular proton transfer occurred whereas,in polar sol- vents,intermolecular proton transfer to or from solvent molecules was observed. Such observation demonstrates the higher basicity of the azomethine nitrogen compared with the carbonyl oxygen in the ground state, and the enhanced basicity of both functional groups in the excited state. Salicylaldehyde anion interacts strongly with some monovalent cations resulting in the formation of chelated- type complexes. Salts of para-substituted phenols form contact ion-pairs and solvent-separated pairs depending on the solvent and on the presence of crown ether molecules. The spectral blue shift observed by decreasing the cationic 160 161 radius indicate stronger interactions between the ions form- ing the ion pair and may be a factor in determining which cation is a better enzyme activator. Evidence for excited- state dissociation of ion pairs was given. The formation of these ion pairs may play a key role in the function of monovalent cations as activators of some PLP-enzymes. The use of lanthanide ions as probes in biological systems is becoming interestingly important. Lanthanide ions, especially Eu(III) and Tb(III),are used extensively to probe the microenvironment at the active site and to measure the distances between the different binding sites in enzymes such as thermolysin. We have prepared a Eu(III) chelate with a Schiff base model compound. We have studied~its magnetic and optical properties. The complex is more likely to have an octahedral configuration. Our demonstration of energy transfer resulting in a red emission at 610 nm offers new opportunities in probing the active site of PLP-enzymes. Future work in this area is very important. Thus, time- resolved picosecond studies will undoubtedly help in iden- tifying excited-state species resulting from proton trans- fer. The study of sterically hindered model compounds, which are expected to have slower rates of proton transfer (higher energy barrier), is particularly interesting. More attention should be given to the study of possible con- formers that may exist in equilibrium in the ground state of a particular species. 162 The use of luminescence techniques offer a unique opportunity to study ion-pair association in the excited state. Our study opens the door for an extensive study of ion-association where one of the ions has convenient emission properties, e.g. fluorenyl anion. The study of time- resolved spectra of ion pairs is very interesting and im- portant in order to obtain rate constants of excited- state sodium ion or potassium ion transfer. Moreover, the study of the interaction of tryptophanase holoenzyme (tryptophanase-PLP) with Eu(III) is an obviously important experiment to measure distances between trypto- phenyl residues and the active site. An experiment where a tunable dye laser is used as an exciting source to moni- tor the excitation spectrum of Eu(III) emission and life- time measurements in the microsecond range of the holo- enzyme in aqueous media and in the presence of D20 provide unique information regarding the microscopic environment (water presence) at the active site. The use of other lanthanide ions such as Tb(III) should also be explored. REFERENCES 1'}; '- REFERENCES H. Sherman, et a1. Pyridoxine and related compounds, in "The Vitamins," Vol. III, Chap. 1H, W. H. Sebrell, Jr., and R. S. Harris (eds.), Academic, New York (195“). S.A. Hoch and R.D. DeMoss, J. Bacteriol., 11H, 3H1 (1973). H. Wada, Tryptophan Metabolism, Vol. I, Sekai Hoken Tsushinsha (ed.), Ltd., Osaka, Japan (196H), p. 77-92. F. C. Happold, and A. Struyvenberg, Biochem. J., Y. Morino and E. E. Snell, J. Biol. Chem., 2H2, 2800 (1967). D. 8. June, et al., J. Am. Chem. Soc., 101, 2218 (1979). D. S. June, C. H. Suelter, and J. L. Dye, Biochem., gt, 2707 (1981). R. J. Johnson and D. E. Metzler "Methods in Enzymology", D. B. McCormick and L. D. Wright (eds.), Vol. 18, pt. A, p. H33-H7l, Academic Press, New York (1970). H. C. Dunathan, Proc. Natl. Acad. Sci. U.S.A., 55, 712 (1966). R. B. Martin and F. S. Richardson, Q. Rev. Biophys., it, 181 (1979). W. De W. Horrocks, Jr., and D. R. Sudnick, Acc. Chem. Res., it, 38H (1981). D. Heinert and A. E. Martell, J. Am. Chem. Soc., éfl. 3257 (1962). F. Maggio, T. Pizzino, V. Romano and R. Zingales, Inorg. Nucl. Chem. Lett., l3, 1H9 (1978). H. E. Zaugg and A. D. Schaefer, J. Am. Chem. Soc., 91, 1857 (1965). 163 l5. 16. 17. 18. 19. 20. 21. 22. 23. 2H. 25. 26. 27. 28. 29. 30. 31. 16H N. S. Potts, J. Chem. Phys., 39: 809 (1952). Alberchet and W. T. Simpson, J. Chem. Phys., 23, 1H8O (1955). Watanabe,J. Chem. Phys., 23, 156H (195H). B. M. Wepster, Rev. Trav. Chim., it, 333 (1957), 77, “91 (1958). J. N. Murrell, The Theory of the Electronic Spectra of Organic Molecules (Methuen, London), (1963). E. A. Braude and F. Sondheimer, J. Chem. Soc., 375H (1955). E. E. Snell, P. M. Fasella, A. E. Braunstein and Rossi-Fanelli (ed.), "Chemical and Biochemical Aspects of Pyridoxal Catalysis", MacMillan, New York, NY (1963). D. Heinert and A. E. Martell, J. Am. Chem. Soc., 85, 188 (1963). R. J. Johnson and D. E. Metzler, "Methods in Enzym- ology", Vol. 18, Pt. A, ed. by D. B. McCormick and L. D. Wright, Academic Press, New York, NY (1970) pp. H33-H7l. W. Barton, P. J. Luechest and N. S. Snider, J. Am. Chem. Soc., 85, 18H2 (1962). A. Weller, Z. Physik. Chem. (Frankfurt), ll, 22H (1958). Y. Matsushima and A. E. Martell, J. Am. Chem. Soc., Q2. 1322 (1967). J. Llor and M. Cortijo, J. Chem. Soc., Perkin Trans. II, gig, 1111 (1977). S. Shaltiel and M. Cortijo, Biochem. Biophys. Res. Commun., Ht, 59H (1970). A. B. Kent, E. G. Krebs and E. H. Fischer, J. Biol. Chem., g3g, 5H9 (1958). M. Cortijo, I. Z. Steinberg and S. Shaltiel, J. Biol. Chem., gig, 933 (1971). S. Shaltiel and E. H. Fischer, Isr. J. Chem., 127 p (1967). 5, 32. 33. 3H. 35. 36. 37. 38. 39. H0. H1. H2. H3. HH. H5. H6. H7. H8. 165 M. Arrio Duport, Biochem. Biophys. Res. Commun., 55, 653 (1971). S. Veinberg, S. Shaltiel and I. Steinberg, Isr. J. Chem., 55, H21 (197H). F. D. Dwyer and D. P. Mellor in "Chelating Agents and Metal Chelates, Academic Press, New York (196H) pp. 335. E. E. Snell, Adv. Enzymol. Relat. Areas Mol. Biol., fig. 287 (19757: W. A. Newton, Y. Morino and E. E. Snell, J. Biol. Chem., 2H0, 1211 (1965). Y. Morino and E. Snell, J. Biol. Chem., 2H2, 5591 (1967). N. Bjerrum, th. danske Vidensk. Selsk., l, 9 (1926). R. M. Fuoss and C. A. Kraus, J. Am. Chem. Soc., 55, H76 (1933). T. R. Griffiths and M. C. R. Symons, Mol. Phys., 5, 90 (1950)- H. V. Carter, B. J. McClelland, and E. Warhurst, Trans. Faraday Soc., 55, H55 (1960). T. E. Hogen Esch and J. Smid, J. Am. Chem. Soc., 51, 659 (1955); §§J 307, 318 (195677 R. Waack and M. A. Doran, J. Ph 3. Chem., 51, 1H8 (1963); J. Am. Chem. Soc.,—EET‘165I‘TI963). Ralph S. Becker, "Theory and Interpretation of Fluores- cence and Phosphorescence", Wiley, New York, NY (1959), pp- 30- E. M. Kosower and P. E. Klinedinst, Jr., J. Am. Chem. Soc., 15, 3H93 (1956). J. Smid in "Ions and Ion-Pairs in Organic Reactions", M. Szware ed. Wiley, New York, NY (1972). T. L. Chan and J. Smid, J. Am. Chem. Soc. 55, H65H (1968). T. Shimomura, K. J. Tolle, J. Smid and M. Szwarc, J. Am. Chem. Soc., 55, 796 (1967). H9. 50. 51. 52. 53. SH. 55. 56. 57. 58. 59. 60. 61. 166 T. Shimomura, J. Smid and M. Szwarc,, J. Am. Chem. 800.. §9. 57H3 (1967). S. Claesson, B. Lundgren and M. Szwarc, Trans. Fara- day Soc., 55, 3053 (1970). W. S. LeNoble and A. R. Das, J. Phys. Chem., 15, 3429 (1970)- (a) M. Shinohara, J. Smid and M. Szwarc, J. Am. Chem. Soc., 59, 2175 (1968). (b) M. Shinohara, J. Smid, and M. Szwarc, Chem. Commtg London, 1962, 1232. (a) C. J. Pedersen, J. Am. Chem. Soc., 55, 7017 - (1967). l (b) C. J. Pedersen, J. Am. Chem. Soc., 25, 391 (1970). K. H. Wong, G. Konizer, and J. Smid, J. Am. Chem. Soc., 55, 666 (1970). J. J. Christensen, D. J. Eatough and R. M. Izatt, Chem. Rev., 13, 351 (197A). R. M. Izatt, R. E. Terry, B. L. Haymore, L. D. Hansen, N. Dalley, A. G. Avondet and J. J. Christensen, 5. Am. Chem. Soc., 55, 7620 (1976). (a) G. W. Liesegang, M. M. Farrow N. Purdie and E. M. Eyring, J. Am. Chem. Soc., 8, 6905 (1976). (b) G. W. Liesegang, M. M. Farrow, F. A. Vazquez, N. Purdie, and M. Eyring, J. Am. Chem. Soc., 55, 32H0 (1977). T. E. Hogen Esch and J. Smid, J. Am. Chem. Soc., 5;, H580 (1969). J. P. Pascault, Thesis, l'Universite de Lyon, France, 1970. (a) W. F. Edgell, A. T. Watts, J. Lyford and W. M. Risen, J. Am. Chem. Soc., 55, 1815 (1966). (b) w. F. Edgell, J. Lyford, R. Wright, w. M. Risen and R. Watts, J. Am. Chem. Soc., 25, 22HO (1970). J. C. Evans and G. Y-S. Lo, J. Phys. Chem., 55, 3223 (1965)- 62. 63. 6H. 65. 66. 67. 68. 69. 70. 71. 72. 73. 7M. 75. 167 (a) B. W. Maxey and A. I. Popov, J. Am. Chem. Soc., (b) B. W. Maxey and A. I. Popov, J. Am. Chem. Soc. 21. 20 (1969). (a) A. Tsatsas, W. M. Risen, J. Am. Chem. Soc., 55, 1789 (1970). (b) A. Tsatsas, R. W. Stearns, and W. M. Risen, J. Am. Chem. Soc., 55, 52U7 (1972). _ M. C. R. Symons, Pure and Appl. Chem., 55, 13 (1977). (a) M. C. R. Symons, J. Phys. Chem., ll, 172 (1967) (b) H. Sharp and M. C. R. Symons, Ions and Ion- pairs in Organic Reactions, M. Szwarc, Ed., Wiley, New York, NY (1972). (c) E. de Boer and E. Sommerdisk, Ions and Ion- pairs in Organic Reactions, M. Szwarc, Ed., Wiley, New York, NY (1972). F. C. Adam and S. I. Weissman, J. Am. Chem. Soc., 55, 1518 (1958). N. M. Atherton, and S. I. Weissman, J. Am. Chem. Soc., 55, 1330 (1961). P. J. Zandstra, and S. I. Weissman, J. Am. Chem. Soc. 55, uuo8 (1962). w. G. Williams, P. J. Pritchett and G. K. Fraenkel, Chem. Phys., g3, 558a (1970). r... . Hirota, J. Am. Chem. Soc., 55, 3603 (1968). Nakamura and Y. Deguchi, Bull. Chem. Soc. Japan, , 705 (1967). A. W. Rutter and E. Warhust, Trans. Faraday Soc., 66, 1866 (1970). I6?“ 2 E. S. Amis and J. F. Hinton, "Solvent Effect on Chemical Phenomena", Vol. 1, Academic Press, New York (1973) pp. 150:16H. E. G. Bloor and R. G. Kidd, Canad. J. Chem., 55, 3425 (1968). R. H. Erlich, E. Roach and A. I. Popov, J. Am. Chem. Soc., 55, “989 (1970). 76. 77. 78. 79. 80. 81. 82. 83. 84. 85. 86. 87. 88. 89. 90. 91. 168 R. H. Erlich and A. I. Popov, J. Am. Chem. Soc., 3, 5620 (1971). ’— M. S. Greenberg, R. L. Bender and A. I. Popov, 5. Am. Chem. Soc., 55, 2449 (1973). G. E. Maciel, J. K. Hancock, L. F. Lafferty, P. A. Mueller and W. K. Musker, Inorg. Chem., 5, 55A (1966). J. W. Akitt and A. S. Downs in "The Alkali Metals Symposium", The Chemical Society London (1967) p. 199. M. Y. Cahen. P. R. Handy, E. T. Roach and A. I. Popov, J. Chem. Phys., 15, 80 (1975). H. J. Maria and S. P. McGlynn, J. Chem. Phys., 55, 3399 (1970). T. P. Carter and G. D. Gillispie, Chem. Phys. Letters, 1;, 75 (1980). E. Nieboer, Struct. and Bond., 55, 1-47 (1975). R. Prados, L. G. Stradtherr, H. Donato, Jr. and R. B. Martin, J. Inorg. Nucl. Chem., 55, 689 (197A). R. D. Shannon, Acta Crystallogr., A32, 751 (1976). (a) R. H. Kretsinger and C. E. Nockolds, J. Biol. Chem., 2N8, 3313 (1973). (b) B. W. Matthews and L. H. Weaver, Biochemistry, ii, 1719 (197“)- K. Hermsmeyer and N. Sperelakis, Am. J. Physiol., 219, 1108 (1970). N. Sperelakis, M. F. Schneider and E. J. Harris, 5. Gen. Physiol., 55, 1563 (1967). (a) F. Abbott, D. W. Darnall and E. R. Birnbaum, Biochem. Biophys. Res. Commun., 55, 2H1 (1975). (b) F. Abbott, J. E. Gomez, E. R. Birnbaum and D. W. Darnall, Biochemistry, 55, A935 (1975). T. Forster, Disc. Farad. Soc., 51, 7 (1959). De W. Williams, Horrocks, Jr. and W. E. Collier, 55 Am. Chem. Soc., 103, 2856 (1981). 92. 93. 94. 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. 169 W. D. Horrocks, Jr., G. F. Schmidt, D. R. Sudnick, C. Kittrell and R. A. Bernhein, J. Am. Chem. Soc., .22, 2378 (1977). G. Stein and E. Wurzberg, J. Chem. Phys., 52, 208 (1975). J. L. Kropp and M. W. Windsor, J. Chem. Phys., _5, 2769 (1963). J. P. Freeman, G. H. Crosby, and K. E. Lawson, 5. Mol. Spectrosc., 55, 399 (1964). Y. Haas, G. Stein, and E. Wurzberg, J. Chem. Phys., §§, 2777 (1973). (a) J. L. Kropp and M. W. Windsor, J. Chem. Phys., £2. 1599 (1965). (b) J. L. Kropp and M. W. Windsor, J. Chem. Phys., 39. 2769 (1963). (c) J. L. KrOpp and M. W. Windsor, J. Chem. Phys., 55, 761 (1966). (a) W. De W. Horrocks, Jr., D. R. Sudnick, J. Am. Chem. Soc., 101, 334 (1979). (b) W. De W. Horrocks, Jr., D. R. Sudnick, Science Washington, D.C. T. Forster, Ann. Phys., 5, 55 (1948). C. F. Meares and J. E. Ledbetter, Biochemistry, 55, 5178 (1977). E. T. O'Keeffe, R. L. Hill and J. E. Bell, Bio- chemistry, 55, 4954 (1980). S. I. Weissman, J. Chem. Phys., 55, 214 (1942). R. E. Whan and G. A. Crosby, J. Mol. Spectry., 5, 315 (1962). (a) G. A. Crosby, R. E. Whan and R. M. Alire, 5. Chem. Phys., 55, 743 (1961). (b) G. A. Crosby and R. E. Whan, J. Chem. Phys., .55, 614 (1960). (c) G. A. Crosby and R. E. Whan, Naturwiss., 51, 276 (1960). (d) G. A. Crosby, Molecular Crystals, 5, 37 (1966). 170 104. (e) G. A. Crosby, R. E. Whan and J. J. Freeman, J. Phys. Chem., 55, 2493 (1962). — 105. M. Kleinerman, Bull. Am. Phys. Soc., 5, 265_(1964). 106. (a) M. L. Bhaumik and M. A. El-Sayed, J. Chem. Phys., 42, 787 (1965). (b) M. A. El-Sayed and M. L. Bhaumik, J. Chem. Phys., 52, 2391 (1963). (c) M. L. Bhaumik and M. A. El-Sayed, Appl. Opt. Supp. 5, 214 (1965). 107. A. E. Martell, "Metal Ions in Biological Systems", Vol. 2, H. Sigel, ed., M. Dekker, NY (1973), p. 207. 108. A. E. Braunstein, "The Enzymes", Vol. 2, P. Boyer, H. Lardy and K. Myrback (eds.), Academic Press, New York (1960). 109. W. H. Sebrell, Jr., and R. S. Harris (eds.), "The Vitamins", Vol. 2, Academic Press, New York (1968). 110. B. M. Guirard and E. E. Snell, J. Am. Chem. Soc., 19, 4745 (1954). 111. D. E. Metzler, M. Ikawa and E. E. Snell, J. Am. Chem. Soc., 15, 648_(1954). 112. M. Ikawa and E. E. Snell, J. Am. Chem. Soc., 19, 393 (1954). 113. A. Perault, B. Pullman and C. Valdemoro, Biochim., Biophys. Acta, 55, 555 (1961). 114. H. M. Dawes, J. M. Waters and T. N. Waters, Inorganic Chimica Acta, 55, 29 (1982). 115. R. S. Shekhawat, N. K. Sankhla and R. K. Metha, Polish Journal of Chemistgy, 55, 391 (1980). 116. J. H. Van Vleck, "The Theory of Electric and Magnetic Susceptibilities", Oxford Univ. Press (1932), pp. 226-261. 117. Cabera and Duperier, Compets Rendus, 55, 1640 (1929). 118. S. Matsumoto and Y. Matsushima, J. Am. Chem. Soc., 29, 5228 (1974). 119. C. V. Banks and D. W. Klingman, Anal. Chim. Acta, $2. 356 (1956). 171 120. M. L. Bhaumik, J. Mol. Spectrus. 55, 323 (1963). 121. L. J. Nugent, M. L. Bhaumik, S. George and S. M. Lee, J. Chem. Ph ., 55, 1305 (1964). . 122. M. I. Knyazhanskii and P. v. Gilyanovskii, Opt. Spektrosk, 55, 1083 (1973).