..'§;s.'.rv'L'.»n: '39": , ( G , If» I , o "‘ " rfv _ <, r‘r ‘3- Liam. -{ :A...‘ :."..-‘ .. " 1’: r v. I mew . . -~ A lib‘p‘. _ “a ‘. .21.?!" ' ~H‘—I‘ ‘.g..a ‘ ‘L ' "i ’"n ”final I. . ' 1"" n f" {'9- 1 . m'a‘v: 2‘ ~41; W“ .L .\ - ll‘I WMVM‘Q w- . . «nun w . 2:. m... ALE v‘nnv—w. ' QF‘ .I "~35?“- xmima . .U‘ V c. fiffixfi.“ . ‘- ,. u “a s n u l .9. p...‘ onu- ..:i “r 1. «r r1 x. . m t ‘c—v- £1. . 5v x b 3 ' ‘ Own-ho. -m—w W‘W . .. . ’E’-« I *u‘ v.2- . ‘ O‘I'IO. 1 .2 .. :55. Kt J '1‘. — qn-n I an —. _-) 4.. As v. noun-u A .- r. “dun-nug- “A 8014 I3 I llllHlll lllliilllililll M will LIBRARY 3 1293 00629 9345 Michigan State University This is to certify that the dissertation entitled Characterization of the Vibrational Properties of Metallochlorins, Chlorophylls and Chlorophyll - Binding Proteins by Resonance Raman Spectroscopy presented by Harold Fonda has been accepted towards fulfillment of the requirements for PHD degree in Chemistry W47. Aflm/ Major professor Date M MS U is an Affirmative Action/Equal Opportunity Institution 0-1277 1 PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. DATE DUE DATE DUE DATE DUE l ll 1L ___| .4 ——|L_l—_| | I MSU Is An Affirmative Action/Equal Opportunity Institution CHARACTERIZATION OF THE VIBRATIONAL PROPERTIES OF METALLOCHLORINS, CHLOROPHYLLS AND CHLOROPHYLL-BINDING PROTEINS BY RESONANCE RAMAN SPECTROSCOPY BY Harold Norman Fonda A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1989 @05954 ABSTRACT CHARACTERIZATION OF THE VIBRATIONAL PROPERTIES OF METALLOCHLORINS, CHLOROPHYLLS AND CHLOROPHYLL-BINDING PROTEINS BY RESONANCE RAMAN SPECTROSCOPY BY Harold Norman Fonda Three systems of increasing complexity from metallo- chlorin model compounds to Chlorophylls to chlorophyll-binding proteins have been studied. Resonance Raman and IR spectra of metallo-octaethylchlorins (MOEC) show significant.differences in the vibrational mode properties of metallochlorins and metalloporphyrins. The core-size correlation parameters for Zn, Cu and NiOEC reveal mixing of C.CIII and (3th stretching character in the high frequency modes. The isotopic frequency shifts observed for the V3 and um modes of fully and partially deuterated copper octaethylporphyrin establish that, for a delocalized vibrational mode, methine d2 deuteration produces half of the d‘ shift. CuOEC exhibits different patterns of frequency shifts upon o,fi and 1,6 methine deuteration that demonstrate the phenomenon of mode localization in metallo- chlorins. The vibrational mode characteristics determined for MOEC provide the basis for the interpretation of the resonance Raman spectra of metallochlorin t cation radicals and metal-substituted chlorophyll a. Resonance Raman spectra of the at cation radicals of CuOEC and selectively methine- deuterated CuOEC complexes show that unlike the situation in metalloporphyrins, changes in mode composition occur upon macrocycle oxidation in metallochlorins. Metal-substituted chlorophyll a complexes display altered mode compositions compared to MOEC as a result of addition of the isocyclic ring. Identification of the core-size sensitive modes in these complexes provides a reliable method to determine the coordination state of chlorophyll a in solution and in 2119. The results obtained for chlorophyll a in solution have been extended to chlorophyll-binding proteins. Resonance Raman spectra of the light-harvesting chlorophyll a/b protein complex (LHC) show four populations of chlorophyll a and two of chlorophyll b that are distinguished by the stretching frequencies of their conjugated carbonyl groups. Each of the Chlorophylls has a S—coordinate Mg atom. The chlorophyll b and two of the chlorophyll a populations result from H-bonds of different strength to the carbonyl groups. H-bonding to the conjugated carbonyls is proposed as a mechanism to account for the broadened and red-shifted electronic absorption bands of the Chlorophylls in LHC. To my parents and in memory of my grandfather, Neil W. Stuart. iv ACKNOWLEDGMENTS I would like to thank Dr. Gerald T. Babcock for his support and guidance; Dr. George E. Leroi for serving as Second Reader; Tony Oertling and Asaad Salehi for their collaboration on the chlorin and at cation radical work; Dwight Lillie and Matt Espe for maintaining the computer system; and Bob Kean and Chris Bender for comic relief. TABLE OF CONTENTS LIST OF TABLES ix LIST OF FIGURES Xi CHAPTER 1 INTRODUCTION 1. The Role of Chlorophyll in Photosynthesis 1 2. The Structure of Metalloporphyrins, Metallochlorins and Chlorophyll 5 3. The Electronic Absorption Spectra of Metalloporphyrins and Metallochlorins 9 4. Infrared and Raman Spectroscopy 15 5. Vibrational Mode Assignments for Metalloporphyrins 19 6. The Application of Resonance Raman Spectroscopy to the Study of Chlorophyll 25 7. Aim and Scope of this Work 26 CHAPTER 2 MATERIALS AND METHODS 1. Materials 27 i. Preparation of Chlorins 27 ii. Preparation of Chlorophyll a and b 28 iii. Preparation of Metal-Substituted Chlorophyll a 29 iv. Isolation of Chlorophyll-Binding Proteins 30 vi 2. Spectroscopic Techniques i. Electronic Absorption Spectroscopy ii. Infrared Spectroscopy iii. Resonance Raman Spectroscopy CHAPTER 3 VIBRATIONAL PROPERTIES OF METALLOCHLORINS 1. Introduction 2. Results 1. Electronic Absorption Spectra ii. Resonance Raman Spectra of MOEC iii. IR Spectra of MOEC iv. Effects of Methine Deuteration v. Resonance Raman Spectra of CuECI vi. Solid State Spectra 3. Discussion i. C.CIll and Cbe Modes ii. C,N Modes iii. CmI-I Modes iv. Ethyl Group Vibrations 4. Conclusions CHAPTER 4 VIBRATIONAL PROPERTIES OF METALLOCHLORIN f CATION RADICALS 1. Introduction 2. Results 3. Discussion 4. Conclusions 31 31 31 31 35 39 39 42 47 51 57 59 62 62 74 76 79 80 82 86 9O 97 CHAPTER 5 VIBRATIONAL PROPERTIES OF METALLOCHLOROPHYLLS 1. Introduction 99 2. Results 103 i. Electronic Absorption Spectra 103 ii. Resonance Raman Spectra 107 3. Discussion 110 4. Conclusions 116 CHAPTER 6 RESONANCE RAMAN SPECTROSCOPY OF CHLOROPHYLL-BINDING PROTEINS 1. Introduction 118 2. Results 127 i. Electronic Absorption Spectra 127 ii. Resonance Raman Spectra 129 3. Discussion 140 i. Selective Enhancement of Chlorophyll a and Chlorophyll b 140 ii. Mg Coordination State 144 iii. Carbonyl Stretching Region 147 iv. 28 kDa and Reaction Center Complex Proteins 151 4. Conclusions 153 CHAPTER 7 SUGGESTIONS FOR FURTHER WORK 155 LIST OF REFERENCES 162 viii Table Table Table Table Table Table Table Table Table Table LIST OF TABLES Observed and calculated frequencies (cm') of NiOEP for modes in the region from 900 to 1700 cm (from ref. 55) Electronic absorption maxima (nm) for MOEC complexes in benzene solution Electronic absorption maxima (nm) for MOEP complexes in benzene solution Vibrational frequencies (cm“) and Raman depolarization ratios of MOEC Vibrational frequencies (cm ), isotope shifts (cm ) and core-size correlation parameters for metallochlorin C5; and Cbe normal modes Vibrational frequencies (cm ), isotope shifts (cm ) and core-size correlation parameters for metalloporphyrin C4; and Cbe normal modes Vibrational frequencies (cm ) and isotope shifts (cm ) for metallochlorin C N modes Vibrational frequencies (cm') for CuOEC normal modes in the frequency region below 1350 cm Predicted and observed frequency shifts for porphyrin normal modes upon oxidation (from ref. 99) Mode characters predicted for CuOEC/CuOECP ix 21 41 42 44 67 68 75 77 92 94 Table Table Table Table Table Table Table 5.2 5.4 Vibrational frequencies (and) and isotope shifts (cm”) for the high frequency resonance Raman modes of CuOEC and CuOEC' Chlorophylls and chlorophyll-derivatives for which resonance Raman spectra have been reported Electronic absorption maxima (nm) for metal- substituted chlorophyll a in diethyl ether solution Electronic absorption maxima (nm) for chlorophyll a in various solvents Resonance Raman frequencies (cm”) and core-size correlation parameters for metal- substituted chlorophyll a Electronic absorption maxima (nm) of chlorophyll a and chlorophyll b Chlorophyll a and chlorophyll b carbonyl stretching frequencies (cm”) in LHC 95 101 105 107 109 127 147 Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure LIST OF FIGURES Structure and labelling scheme for chlorophyll a Structure and labelling scheme for a) metallo-octaethylporphyrin (MOEP) and b) trans-metallo-octaethylchlorin (MOEC) Molecular structure of a) FenOEP and b) FeHOEC as determined by X-ray crystallography (from ref. 35) Electronic absorption spectrum of copper octaethylporphyrin (CuOEP) in CHZCl2 solution Porphyrin M.O.'s comprising the Gouterman four-orbital model (from ref. 42) Electronic absorption spectrum of copper octaethylchlorin (CuOEC) in CHzClz solution Molecular processes associated with a) Infrared and b) Raman spectroscopy (from ref. 45) Atomic displacements in the v2 vibrational mode of NiOEP (from ref. 55) Diagram of the apparatus used for the collection of Raman spectra of KBr disc samples Structure and labelling scheme for metallo- chlorin model compounds. MOEC = t;an_- metallo-octaethylchlorin and MECI = trans- metallo-etiochlorin I xi 10 12 14 16 23 34 38 Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure 3.10 Electronic absorption spectra of ZnOEC, CuOEC and NiOEC in benzene solution Resonance Raman spectra of ZnOEC, CuOEC and NiOEC in CHzClz solution obtained with Soret excitation at 406.7 nm. Laser powers: 5, 7 and 20 mW, respectively Resonance Raman spectra of CuOEC in CIiIzClz solution obtained with Q, excitation at 488.0 nm and Qy excitation at 615.0 nm. Laser powers: 100 and 35 mW, respectively Resonance Raman spectra of NiOEC in CHZCI2 solution obtained with Q, excitation at 488.0 nm and Q, excitation at 615.0 nm. Laser powers: 70 and 40 mW, respectively IR spectra of ZnOEC, CuOEC and NiOEC in CCl. solution IR spectra of CuOEC-1,6-dz, CuOEC-ofi-d2 and CuOEC-d, in CCl. solution Resonance Raman spectra of CuOEC-1,6-dz, CuOEC-o,fi-dz and CuOEC-d, in CHZClz solution obtained with Qy excitation at 615.0 nm. Laser power: 40 mW Resonance Raman spectra of CuOEP, CuOEP-d2 and CuOEP-d. in CHZCIz solution obtained with Soret excitation at 406.7 nm. Laser power: 10 mW Resonance Raman spectra of CuOEP, CuOEP-dz and CuOEP-d. in CHzClz solution obtained with visible excitation at 514.5 nm. Laser power: 50 mW Resonance Raman spectra of CuECI in C11,,Cl2 solution obtained with Soret excitation at 406.7 nm, Q, excitation at 488.0 nm and Q, excitation at 615.0 nm. Laser powers: 6, 70 and 40 mW, respectively xii 40 43 48 49 50 52 53 55 56 58 Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Resonance Raman spectra of NiOEC in KBr obtained with Soret excitation at 406.7 nm, Q, excitation at 488.0 nm and Q, excitation at 615.0 nm. Laser powers: 20, 70 and 50 mW, respectively Resonance Raman spectra of CuOEC in KBr obtained with Soret excitation at 406.7 nm, Q, excitation at 488.0 nm and Q, excitation at 615.0 nm. Laser powers: 20, 75 and 50 mw, respectively Core-size correlation for the high frequency modes of MOEC Core-size correlation for the resonance Raman active (———) and IR active (- -) high frequency modes of MOEP Structure of metallo-methyloctaethylchlorin (MMeOEC) Electronic absorption spectrum of CuOEC” in CHzCl2 solution Resonance Raman spectra of CuOEC, CuOEC- 1,6-d2, CuOEC-a,fi-d2, CuOEC-d. and CuECI in CHzCl2 solution obtained with Soret excitation at 363.8 nm Resonance Raman spectra of CuOEC“, CuOEC- 1,6-d; , CuOEC-a,p-d; , CuOEC-d3“ and CuECI ' in CHZCIz solution obtained with Soret excitation at 363.8 nm Electron density contour maps for the am and a“ orbitals of a) free base porphine and b) free base chlorin (from ref. 123) Electronic absorption spectra of chlorophyll a and Zn, Cu and Ni-substituted chlorophyll a in diethyl ether solution Electronic absorption spectra of chlorophyll a in acetone, methanol and pyridine solution xiii 60 61 65 66 84 87 88 91 104 106 Figure Figure Figure Figure Figure Figure Figure Figure Figure 5.3 5.4 6.1 6.2 6.3 6.4 6.5 Resonance Raman spectra of chlorophyll a in acetone solution and Zn, Cu and Ni- substituted chlorophyll a in diethyl ether solution obtained with Soret excitation at 406.7 nm. Laser powers: 35, 20, 20 and 20 mW, respectively Resonance Raman spectra of chlorophyll a in acetone, pyridine, methanol and diethyl ether solution obtained with Soret excitation at 441.6 nm. Laser powers: 20, 15, 15 and 15 mW, respectively Proposed arrangement of the polypeptides and electron transfer components in the PS II/OEC complex (from ref. 154) Arrangement of pigments in the L and M subunits of the reaction center from Rhodopseudomogas yirigig (from ref. 158) Electronic absorption spectrum of LHC isolated from pea leaves Electronic absorption spectra of LHC, the 28 kDa chlorophyll a-binding protein and the LHC-depleted Photosystem II reaction center complex isolated from spinach Resonance Raman spectra of LHC at -125W: obtained with 413.1 and 441.6 nm excitation. Laser powers: 28 and 25 mW, respectively Resonance Raman spectra of LHC at -125%: obtained with 472.7 and 514.5 nm excitation. Laser power: 45 mW Resonance Raman spectra of chlorophyll a in acetone at -125°C obtained with 413.1 and 441.6 nm excitation. Laser powers: 19 and 22 mW, respectively xiv 108 111 121 123 128 130 131 132 133 Figure Figure Figure Figure Figure 6.8 6.10 Resonance Raman spectra of chlorophyll b in acetone at -125°C obtained with 441.6 and 472.7 nm excitation. Laser powers: 20 and 40 mW, respectively Resonance Raman spectra of LHC, the 28 kDa protein and the reaction center complex obtained with 406.7 nm excitation. Laser power: 20 mW Resonance Raman spectra of LHC, the 28 kDa protein and the reaction center complex obtained with 441.6 nm excitation. Laser power: 22 mW Resonance Raman spectra of LHC, the 28 kDa protein and the reaction center complex obtained with 476.5 nm excitation. Laser power: 25 mW Calculated excitation profiles for a 1600 cmq'mode and a 300 cm' mode of chloro- phyll a (-—-) and chlorophyll b (- -) in LHC 135 136 138 139 143 CHAPTER 1 INTRODUCTION 1. The Role of Chlorophyll in Photosynthesis In the process of photosynthesis in higher plants, light energy absorbed by chlorophyll is utilized to convert water and carbon dioxide into oxygen and stored chemical energy in the form of carbohydrate (1,2). Ingenhousz in 1779 showed that it is the green parts of plants that are involved. The structure of the green pigment chlorophyll (Figure 1.1) was determined by Willstatter and Stoll in the early 1900's: at the same time organomagnesium reagents were being developed by Grignard. This coincidence led to the (incorrect) idea.that photosynthesis occurred by the photodecomposition of carbon dioxide bound to chlorophyll. More recent work has shown that the photosynthetic process is considerably more complex. In our present understanding of photosynthesis in higher plants, there are two photosystems (called Photosystem I and Photosystem II) that operate in series to drive a sequence of electron transfer reactions (3). These reactions culminate in the splitting of oxygen from water and the formation of NADPH Figure 1.1 Structure and labelling scheme for chlorophyll a 3 (= reduced nicotinamide-adenine dinucleotide phosphate) and ATP (= adenosine triphosphate) to be used for the subsequent reduction of carbon. dioxide. The components of the two Photosystems in the electron transfer chain are arranged within protein complexes embedded in the photosynthetic membrane. At the heart of these Photosystems are chlorophyll a molecules that form the reaction centers, P-700 (P-700 refers to a pigment with an electronic absorption maximum at 700 nm) of Photosystem I (4) and P—680 of Photosystem II (5). The primary energy conversion processes in photosynthesis occur at these sites. The exact nature of the P-680 and P-700 reaction center Chlorophylls.haslbeen.difficult.to ascertain. P-680 is thought to be a ligated monomeric chlorophyll a (6) whereas P-700 may have a dimeric structure (7-14). other structures for P-700 that have been suggested are the enol form of chlorophyll a (15), 6-chloro-substituted lo-hydroxy-chlorophyll a (16) or chlorophyll a' (the C10 epimer of chlorophyll a) (17). The reaction center Chlorophylls have electronic absorption maxima in the visible region that are red-shifted compared to the 662 nm absorption maximum of monomeric chlorophyll a in acetone solution. Further' modification of the jproperties of’ the reaction center Chlorophylls is seen by comparison of their mid-point potentials. The mid-point potential, E; of chloro- phyll a in an organic solvent is in the region of +0.8 V (6) but for P-7oo/P-7oo“, a, = +0.49 v (13) and for P-680/P-680“, 4 E, = +1.1 V (19). The high redox potential of P--680+ is necessary for the oxidation of water: 2H20/02 + 41?, at pH 7.0, E:m = +0.82 V or under estimated physiological conditions at pH 5.0, E, = +0.93 v. The reaction center chlorophylls constitute less than 1% of the total number of chlorophyll molecules in the plant. The vast majority of chlorophylls are involved in light absorption and transfer of the excitation energy to the reaction centers. The major chlorophyll-binding protein is the light-harvesting chlorophyll a/b protein complex (LHC) , which serves both Photosystem I and Photosystem II (20,21) . LHC contains chlorophyll a and chlorophyll b molecules (chlorophyll b has a formyl group in the 3—position instead of a methyl group as in chlorophyll a) that absorb maximally at 677 and 652 nm, respectively. Additionally, there are separate sub-antenna complexes of chlorophyll a serving each Photosystem. The chlorophylls absorb maximally in the range 670 to 683 nm for the Photosystem II subantenna and 680 to 695 nm for the Photosystem I subantenna. This arrangement of light-harvesting and reaction center chlorophylls ensures that the light energy is distributed to both Photosystems and that energy transfer is "downhill" energetically to the reaction center. Directing the light energy absorbed by a large number of antenna chlorophylls to a single reaction center chlorophyll raises the photochemical turnover rate thereby increasing the efficiency of photosynthesis (22). 5 2. The Structure of Metalloporphyrins, Metallochlorins and Chlorophyll A metalloporphyrin is a tetrapyrrole macrocycle with an extensively delocalized w electron system. Metalloporphyrins with a range of structures play important roles in numerous and diverse biological processes (23). The structure of a typical synthetic metalloporphyrin, that of metallo-octa- ethylporphyrin (MOEP) is shown in Figure 1.2(a). Without the central metal ion, hydrogens are bonded to two of the pyrrole nitrogens and the structure is known as the free base. In the labelling scheme, C; and Cb1refer to the carbon atoms of the pyrrole rings and C5 to the methine bridge carbons. The four methine bridge carbons are labelled a, B, 1 and 6. The core-size of a metalloporphyrin is defined as the distance from the center of the macrocycle to the pyrrole nitrogen atoms (cg-N distance). The core-size depends on the size and coordination state of the central metal ion and is determined from the X-ray crystal structures of representative metallo- porphyrin complexes. A metallochlorin is a derivative of a metalloporphyrin in which a Cbe bond of one of the pyrrole rings has been reduced (24) . Figure 1.2(b) shows the structure of trans- metallo-octaethylchlorin (MOEC). Both gig- and trans- configurations of the reduced Cbe bond are possible. In nature, metallochlorins occur as the magnesium-containing \ \ \ \ N ’N\ (O) 8 \ \IM” 3 MOEP ,, \ / N N \ l \ Cg/Ca\cb/ X ) (b) MOEC Figure 1.2 Structure and labelling scheme for a) metallo- octaethylporphyrin (MOEP) and b) trans-metallo-octaethyl- chlorin (MOEC) 7 chlorophyll pigments (25) and in iron-containing leukocyte myeloperoxidase (26,27) and bacterial heme d enzymes (28-30). Several studies have been directed towards elucidating the structural, redox and ligand-binding properties of metallo- chlorins (31-34) to determine the factors that make a chlorin the preferred macrocycle in these instances. A comparative study of FenOEP and FenOEC by X-ray crystallography (35) showed that the average iron-nitrogen.distance is*very similar in the two complexes (1.996 A for FenOEP and 1.986 A for FenOEC) . However, the porphyrin macrocycle in FenOEP is planar but the chlorin macrocycle of FeHOEC is S, ruffled (Figure 1.3). The core-size of the chlorin macrocycle in the planar conformation is intrinsically larger than the corre- sponding porphyrin macrocycle. Ruffling of the chlorin macrocycle reduces the core-size (and hence the metal-nitrogen distance) compared to the planar conformation. This distortion is possible because the chlorin, being less aromatic, has increased flexibility. Similar conformational flexibility was observed for nickel tetramethylchlorin compared to nickel tetramethylporphyrin (36). The two main chlorophyll pigments in higher plants are chlorophyll a and chlorophyll b. A key structural feature of chlorophyll is the presence of a chlorin macrocycle with a trans-configuration of the hydrogens about the reduced our bond of ring IV. Fused onto the chlorin macrocycle is a fifth ring containing a keto carbonyl group. The X-ray crystal (a) Fen OEP ( b) F enOEC Figure 1.3 Molecular structure of a) FeHOEP and b) FenOEC as determined by X-ray crystallography (from ref. 35) 9 structure of ethyl chlorophyllide a dihydrate (37) (where R = Cfig in Figure 1.1) shows that the effect of the fifth ring is to reduce the amount of buckling of the macrocycle. This in turn increases the extent of conjugation of the keto carbonyl with the main 1 electron system. In the naturally- occurring' chlorophyll pigments the central metal ion is magnesium. The free base form of chlorophyll is called a pheophytin. Photosynthetic bacteria utilize bacteriochloro- phyll pigments in which both rings II and IV are reduced (25,38). Bacteriochlorophyll a has the same pattern of peripheral substituents as chlorophyll a except for replacement of the vinyl group by an acetyl group in the 2 position. Chlorophyll clland chlorophyll C, are another type of naturally-occurring chlorophyll pigment found in diatoms and brown algae (25,39,40). Chlorophylls c have the same oxidation state as that of a porphyrin. 3. The Electronic Absorption Spectra of Metalloporphyrins and Metallochlorins The electronic absorption spectrum of a typical metallo- porphyrin of D“, symmetry, copper octaethylporphyrin, in CH2C12 solution is shown in Figure 1.4. The.observed.bands correspond to in-plane 1r - 1* electronic transitions which for a D“1 metalloporphyrin are of E; symmetry. The intense transition around 400 nm is called the Soret or B band. The two weaker 10 _ B LU 0 Z < m m C U) m 4 4 I k A J 300 400 500 700 WAVELENGTH, nm Figure 1.4 Electronic absorption spectrum of copper octa- ethylporphyrin (CuOEP) in CHzCl2 solution 11 transitions in the visible region are labelled the a and 8 bands or Q00 and Q01. The 8 band is a vibronic overtone of the a band. The theoretical interpretation of the electronic absorp- tion spectra of metalloporphyrins is based on the four- orbital model of Gouterman (41,42). M.O. calculations (43) predict two HOMO's and two LUMO's as shown in Figure 1.5. Under D,h symmetry the two HOMO's, labelled b1 and b2, are of an and am symmetry, respectively. The a2u orbital is calculated to be higher in energy than the an orbital. The two LUMO's, labelled c1 and c2 are a degenerate pair of e, symmetry. This M.O. treatment identifies the visible bands with the transition a2,ll -> e, and the Soret band with a,“ -o e,. Because the transition dipoles are almost exactly equal the visible and Soret bands are predicted to be of equal intensity. In Gouterman's development of the four-orbital model, the aZu and am orbitals are assumed to be degenerate. The singly excited configurations (a2u1e81) and (alulef) have the same symmetry and undergo configuration interaction to produce the excited states that correspond to the B and Q transitions. For the 8 state, the transition dipoles are additive hence the strong Soret absorption and in the Q state the transition dipoles subtract, leading to the weakly allowed visible bands. 12 b1 (a2u) Figure 1.5 Porphyrin M.O.'s comprising the Gouterman four- orbital model (from ref. 42) 13 The nature of the central metal ion influences the energy and intensity of the B and Q transitions. The metal 1 orbital can conjugate with the 1r electron system of the porphyrin macrocycle. The an orbital, with electron density on the pyrrole nitrogens, has the appropriate symmetry for such interaction, but the a1n orbital does not. As the metal becomes more electropositive the energy of the a2n orbital is raised. This results in a red-shift of the electronic absorption spectrum and, for alkyl-substituted porphyrins, a decrease in the relative intensity of the Q band compared to the Soret. Reduction of a metalloporphyrin to a metallochlorin lowers the molecular symmetry from D,h to C2, taking into account S, ruffling, where the x,y degeneracy has been removed. Configuration interaction produces four states, two of which comprising the Soret band are accidentally degenerate and strongly allowed (44) . The other two states are Q, and Q,. The Q, transition is at lower energy and is weakly allowed whereas the Q, band is almost exactly forbidden. The Q, band of a metallochlorin is stronger and red-shifted compared to the corresponding Q band in a metalloporphyrin. These effects are clearly seen in the electronic absorption spectrum of copper octaethylchlorin in CHZClz solution (Figure 1.6) . 14 _ B LLI 0 Z < m (I: O (D 2 Qy00 1 l L i 300 400 500 600 700 WAVELENGTH, nm Figure 1.6 Electronic absorption spectrum of copper octa- ethylchlorin (CuOEC) in CHzCl2 solution 15 4. Infrared and Raman Spectroscopy Molecular"vibrations fall into 'the frequency' region between 100 and 4000 cm”. Two techniques available for the characterization of the vibrational modes of a molecule are infrared (IR) and Raman spectroscopy. The molecular processes involved in these two methods are.depicted in Figure 1.7 where |m> and |n> denote different quantum states of a vibrational mode in the ground electronic state of the molecule. The energy separation between the states |m> and |n> corresponds to the vibrational frequency um of that particular vibrational mode. In IR spectroscopy, direct absorption of a quantum of frequency um can promote the transition from state |m> to |n>. The vibrational mode is IR active if there is a change in the molecular dipole moment during the vibration. When a beam of light impinges on a sample of molecules, some is absorbed, some is transmitted and some is scattered. Most of the scattered light is unchanged in frequency (Rayleigh scattering). In Raman scattering, incident light of frequency uL is scattered by a molecule with a concomitant increase or decrease in frequency by an amount corresponding to the vibrational quantum.umv The process depicted in Figure 1.7(b) is referred to as Stokes (Raman) scattering because the transition from |m> to |n> results in a decrease in the frequency of the scattered light. The state |e> in Figure 1.7(b) is a virtual state representing a sum over all excited 16 a) IR b) Raman h3>"jT--T-- |n> |n> I an |m> |m> Figure 1.7 Molecular processes associated with a) Infrared and b) Raman spectroscopy (from ref. 45) 17 electronic states of the molecule. The Raman scattering event can be envisioned as two simultaneous transitions: absorption from |m> to |e> and emission from |e> to |n>. Anti-Stokes (Raman) scattering is associated with the transition |n> to |m> where the molecule has lost a vibrational quantum of energy and the scattered light.hasta higher frequency. At room temperature, most molecules are in the ground vibrational state so that Stokes scattering is more intense than anti- Stokes scattering. A vibrational mode is Raman active if there is a change in the ‘molecular' polarizability' during the vibration. The intensity of scattered radiation, Im, in photons per molecule per second for the Raman transition |m> to |n> is given by (45-48): I... = 19232315 v.‘ I. 2.. I (as)... I2 where u, is the frequency of the scattered radiation (for Stokes scattering, v, = uL - um), IL is the intensity of the incident radiation in photons per second and ((190),, is the path component (p,o = x,y,z) of the polarizability tensor. 18 From second-order' perturbation theory, the Kramers- Heisenberg dispersion formula shows the polarizability tensor to have the form: (02M)Inn = 1 M|e> + h ° Von - UL + 1F. um + UL + ll". where |m>, |n> and |e> correspond to the states in Figure 1.7(b), uw,is the energy of the transition from |m> to |e>, in.is the a th component of the dipole moment operator and F. is the homogeneous linewidth of the excited state |e>. The summation is carried out over all excited states |e> of the molecule. In this expression, the first part corresponds to a resonant term whereas the second part corresponds to a non-resonant term. As the incident laser frequency, v, approaches the energy of an electronic transition, u” the sum-over-states will be dominated by that single electronic state. The contribution of the first term to the scattering intensity increases as the energy separation Iv... - uLl is minimized. This resonance condition results in an enormous enhancement of the intensity of Raman scattering compared to non—resonance (or normal) Raman scattering. The key, therefore, to a resonance Raman experiment is to match the frequency of the incident laser radiation to an electronic absorption band of the molecule under study. Vibrations associated with the chromophore of the molecule will be enhanced. 19 Resonance Raman spectroscopy is particularly suitable for the study of molecules of biological interest (49-51). Selec- tive enhancement of the resonance Raman spectrum of a protein— bound chromophore is possible by excitation into an electronic absorption band of the chromophore. Additional advantages are that samples may be studied at low concentration and in aqueous solution. IR and Raman spectroscopy can be used to determine the vibrational frequencies of a molecule in its ground electronic state. For a molecule with a center of symmetry, the selection rules prohibit the vibrational modes from being simultaneously IR and Raman active. The IR active modes are antisymmetric with respect to inversion but the Raman active modes are symmetric. The vibrational intensities are usually different in the IR and Raman spectra so that both techniques are required for full characterization of the vibrational modes of a molecule. 5. Vibrational Mode Assignments for Metalloporphyrins Metalloporphyrins and hemoproteins have been studied widely by resonance Raman spectroscopy (45,52-54) . The system of vibrational mode assignments for porphyrins currently in use is based on the normal coordinate analysis of NiOEP (55,56). The choice of NiOEP for the normal coordinate 20 analysis was made for several reasons: 1) the complex is stable and experimental data are available not only for NiOEP but also for its methine-deuterated (NiOEP-d0 (57-59) and 15N-substituted (NiOEP-15m) (59) derivatives: 2) the X-ray crystal structure of NiOEP is known (60); 3) the symmetrical structure of NiOEP simplifies the symmetry classification and assignment of the observed vibrational bands: and 4) the internal vibrations of the ethyl groups, which are not conju- gated with the « electron system of the porphyrin macrocycle, do not appear in the resonance Raman spectra so that assign- ment of the porphyrin skeletal vibrations is made easier. In the normal coordinate calculation of the in-plane vibrations, the structural parameters used are based on the X-ray crystal structure of the triclinic form of NiOEP (60), which have been slightly modified to maintain Dm symmetry. The peripheral ethyl group is treated as a point mass of 15 amu. In this model, the in-plane vibrations are factorized into 35 gerade (9A1, + 9818 + 8A2, + 982,) resonance Raman active modes and 18 ungerade (13,) IR active modes. The A1, and A2, modes produce polarized (p) and anomalously polarized (ap) bands, respectively. The Bu and Ba modes produce depolarized (dp) bands. Table 1.1 lists the observed and calculated frequencies of NiOEP for modes in the region from 900 to 1700 cm"1 (55). The observed resonance Raman frequencies are for NiOEP in CHZClz (59) solution. The frequencies of the E, modes are obtained from.the IR spectrum of NiOEP in a 051 disk (58). 21 Table 1.1 Observed and calculated frequencies of NiOEP for modes in the region from 900 to 1700 cm'1 (from ref. 55) Symmetry No. Obs. Calc. Potential energy distribution (%) A1, 1!; 1602 1591 u(Cbe)60, v(Cb-Et)19 V3 1519 1517 u(C.Cm)41, u(C.Cb)35 in, 1383 1386 u(C.N)53, 6(C.-C,)21 V5 1025 1048 v(Cb-Et)38, u(C,Cb)23 315 um 1655 1656 y(C,C,,)49, u(C.Cb)17 un 1576 1587 u(Cbe)57, u(Cb-Et) 16 v12 1348* 1351 y(C,N)63, u(Cbe)13 V13 1220 1262 6(C,H)67, v(C,Cb)22 V“, 1151* 1095 u(C,Cb)31, u(C,,-Et)30 A2, um 1603 1600 u'(C,C,)67, u'(C,Cb)18 V20 1397 1409 v'(C,N)29, u'(C,,-Et)24 1121 1308 1281 6'(C,H)53, u'(C.Cb)18 V22 1121 1118 u'(C,N)37, u'(Cb-Et)26 uz3 1022* 1022 v'(C,Cb)26, u'(Cb-Et)20 82, 1223 1475* 1469 u'(C.C,)52, y'(C,C,,)21 V” 1409 1409 v'(C.Cb)47, u'(Cb-Et)26 V30 1159 1157 u'(Cb-Et)49, v'(C.N)28 V31 1019* 1016 6'(C,-C,)25, 6'(C.C,C.)23 Eu V37 1604 1634 u(C,C,)34, u'(C.C.,)24 V38 1557 1588 u(Cbe)56, u(Cb-Et)16 V39 1487 1486 u'(C.C.,)36, u'(C.Cb)17 v.0 1443 1411 y'(C,Cb)30, u'(Cb-Et)24 u,1 1389 1368 y(C.N)56, 6(C.C,)14 17,2 1224 1275 6(CmH)59, u(C,Cb)9 u,3 1127 1133 u'(Cb-Et)35, u'(C,N)33 v“, 1113 1080 u(Cb-Et)29, u(C.Cb)26 ms 993 997 u'(C.C.,)21, 6(C,C,C,)12 * not observed as a fundamental u = stretch, 6 = in-plane deformation 22 Assignments of the resonance Raman active modes are based on the observed polarization properties of the bands and support for the assignments comes from comparison with the isotopic frequency shifts for NiOEP-d, and NiOEP-”N3. The frequencies marked with an asterisk in Table 1.1 are not observed as fundamentals but are inferred from.a study of the overtone and combination resonance Raman bands (59). The potential energy distribution (P.E.D.) represents the percentage contribution of a particular type of force constant to the total energy of the vibrational mode. For example, C41 stretching is calcu- lated to comprise 60% of the energy for the V; vibrational mode. On the basis of the P.E.D., modes are classified as ORR stretching modes (112, 1111), 0,0,, stretching modes (V3, um, um), C,N stretching modes (u,, v12), CmH bending modes (V13, V21), etc. (55). The vibrational motion.is:more complicated.than the P.E.D. would suggest. A better representation of the vibrational mode is given by the Cartesian displacements of the atoms in the vibration. Figure 1.8 shows the form of the uz vibrational mode. With the vibrational mode assignments for NiOEP it is possible to interpret the resonance Raman data! for other metalloporphyrins and hemoproteins. Prior to the normal coordinate analysis, Spaulding gt _1. (61) had noticed that the frequency of the anomalously polarized mode in metallo- porphyrins corresponding to the 1603 cm'1 mode of NiOEP in CH2C12 solution (now identified as 1’19) exhibited an inverse 23 \ ' I x .‘ i. \ "‘.‘ O- I o ‘\ I, -K’ I \ ’ \ s I \ +- ’ ’ \ --\» ’ \ 6’ -» \\ ,’ x I <4 ’ “ " I a ‘~~( ‘rfl‘f' I ‘ r \‘ «tn-~- Figure 1.8 Atomic displacements in the uz vibrational mode of NiOEP (from ref. 55) 24 linear correlation with the core-size of the metalloporphyrin. A similar correlation was observed for two modes labelled u, and uh of metallotetraphenylporphyrins by Huong and Pommier (62) who expressed the relationship between the vibrational frequency and core-size by the equation: v=K(A-d) where u is the vibrational frequency in cmq, d is the core- size of the metalloporphyrin in A, A is the intercept and K is the slope. The vibrational modes of metalloporphyrins between 1450 and 1700 cmq'are collectively referred to as the core-size sensitive modes. The frequencies of the resonance Raman modes um,ug, um, um and um have been shown to exhibit a linear correlation with core-size (63-66). The slope or R value is proportional to the amount of CJI, stretching character in the P.E.D. of the vibrational mode. Expansion of the core produces a deformation of the C,CmC, bond angles while the pyrrole rings maintain their structural integrity (67). Weakening of the CJ% bonds is reflected by a lowering of the vibrational frequency. Modes with.(;C, stretching character (V3, um and um) are more sensitive to core-size changes than are the (3be stretching modes (u2 and um). Determination of the core—size sensitivity thus provides a useful means of assigning vibrational modes in related metalloporphyrin systems. 25 6. The Application of Resonance Raman Spectroscopy to the Study of Chlorophyll From the discussion of the role of chlorophyll in photo- synthesis, it is apparent that plants have adapted a single type of chromophore (chlorophyll a) to serve distinctly different functions in the reaction center and antenna complexes. The spectral and redox properties of the chloro- phylls are determined by the specific interaction with their protein environment. These interactions are possible through ligation of the central magnesium atom or through the peripheral substituents on the chlorophyll macrocycle. In principle, resonance Raman spectroscopy with its ability to probe the molecular structure of a chromophore within a protein environment provides an attractive means for the investigation of chlorophyll-protein interactions in vivo. The application of resonance Raman spectroscopy to the study of chlorophylls in yitrg and in the antenna complexes of higher plants has been pioneered by Lutz (68-76) . The reaction center chlorophylls in higher plants are more difficult to study because of their low concentration relative to the antenna chlorophylls. Improvements in the biochemical procedures for the isolation of the reaction center protein complexes (77-79) offer promise that resonance Raman spectroscopy will be successful in establishing the structures of the reaction center chlorophylls (79). 26 7. Aim and Sc0pe of this Work In the work to be presented, three systems of increasing complexity from synthetic metallo-octaethylchlorins to chlorophylls to chlorophyll-binding proteins have been studied. As described in Chapter 3, resonance Raman and IR data on MOEC and selectively methine-deuterated CuOEC have been obtained to characterize the vibrational modes of metallochlorins. TheSe results provide the basis for the interpretation of the resonance Raman spectra of metallo- chlorin n cation radicals (Chapter 4) and.metallochlorophylls (Chapter 5). In turn, it is shown in Chapter 6 how the study of the isolated chlorophylls in solution can be used to understand the resonance Raman spectra of chlorophyll-binding proteins. CHAPTER 2 MATERIALS AND METHODS 1. Materials 1. Preparation of Chlorins trans-HZOEC was prepared by the reduction of Fem(OEP)Cl (Aldrich Chemical Company) with sodium metal in iso-amyl alcohol under nitrogen (80). Chromatography on alumina (Grade I) with benzene/ether (9:1, v/v) was used to eliminate residual HZOEP (81) . CuOEC and ZnOEC were prepared from HZOEC by standard metal-insertion procedures (82) . Chromatography on silica gel with benzene was used to remove any traces of unreacted HzOEC. NiOEC was prepared by refluxing nickel acetate with HZOEC in glacial acetic acid/CHZClz (5:1, v/v) under nitrogen (83). Chromatography on magnesium oxide with hexane/benzene ( 9:1, v/v) followed by chromatography on silica gel with benzene was used for purification. Methine-deuterated HZOEC-d“ HZOEC-a,B-dz, and HZOEC- 7, 6-d2 were prepared from HZOEC by Asaad Salehi (Michigan State University). HZOEC-d, was prepared by treatment of HZOEC with 27 28 DZSO,/DZO (9:1, v/v) (84) . Reduction of the DZSO,/DZO ratio to 6:1 (v/v) afforded HZOEC-7,6-d2 (85) . Re-exchange at the 1 and 6 positions of HZOEC-d, with HZSO./HZO (6:1, v/v) produced HZOEC-oJB-d2 (85) . The proton NMR spectrum (A. Salehi, ref. 85) of tLa_n_s_-HZOEC shows two peaks at 9.7 and 8.9 ppm, corre- sponding to the a,fi and 1,6 methine hydrogens, respectively. Deuteration at the 0,6 methine positions is seen in the disappearance of the 9.7 ppm peak and at the 1,6 methine positions by disappearance of the 8.9 ppm peak. In HZOEC-d“ both signals are completely absent. Deuterium substitution is quantitative for HZOEC-d, and HZOEC-1,6-d2. HZOEC-a,fi-dz showed 90% deuteration at the (1,8 positions and complete proton recovery at the 1,6 positions. Oxidation of CuOEC-1,6-d2 with 2,3-dichloro-5,6- dicyanoquinone in benzene yielded CuOEP-dz (86), which was purified by chromatography on silica gel with CHZClz. ii. Preparation of Chlorophyll a and b Chlorophyll a and b were prepared from fresh spinach leaves according to the procedure of Omata and Murata (87). Chromatography on DEAE-Sepharose CL-6B (Pharmacia) with acetone removed carotenoids and pheophytins. Separation of chlorophyll a and b was achieved by chromatography on Sepharose CL-6B with hexane/z-propanol (20:1, v/v) to elute 29 chlorophyll a followed by hexane/Z-propanol (10:1, v/v) to elute chlorophyll b. All operations were carried out under dim light in a cold box (10°C) . iii. Preparation of Metal—Substituted Chlorophyll a Chlorophylls were extracted from spinach leaves by treatment with acetone. Dioxane and water were added to precipitate the chlorophylls which were then collected by centrifugation (88). The crude chlorophyll extract was dissolved in ether and 1 N HCl added to convert the chlorophylls to pheophytins. The ether extract was washed with water, dried over NaZSO, and then evaporated. Pheophytin a was separated from pheophytin b by the use of the Girard 'T' reagent ( ( carboxymethyl) tri-methylammonium chloride hydrazide) (89). Chromatography on alumina with CH2C12 yielded pure pheophytin a free of carotenoids (as confirmed by TLC on alumina). Nickel, copper and zinc-substituted chlorophyll a were prepared from pheophytin a following the procedure of Boucher and Katz (90) by using the metal acetate in chloroform/acetic acid under nitrogen. 30 iv. Isolation of Chlorophyll-Binding Proteins The light-harvesting chlorophyll a/b protein complex (LHC) was isolated from pea leaves by the procedure of Burke gt a1. (21). Thylakoid membranes with a chlorophyll concen- tration of 0.8 mg/ml were solubilized in 0.78% Triton X-100 and then fractionated on a 0.1 to 1.0 M sucrose gradient. The LHC fraction was identified as a deep red fluorescent band when illuminated from behind. This fraction was removed and aggregated by the addition of MgCl2 (to give a Mg”' concentration of 7 mM) so that purified LHC could.be collected by centrifugation. The chlorophyll a/b ratio and total chlorophyll concentration were determined by the procedure of Mackinney (91). LHC, the 28 kDa chlorophyll a-binding protein and the LEO-depleted Photosystem. II :reaction. center' complex from spinach. were isolated from. Photosystem II :membranes and supplied by Demetrios Ghanotakis (University of Michigan) (92). 31 2. Spectroscopic Techniques 1. Electronic Absorption Spectroscopy Electronic absorption spectra were recorded for solutions of metallochlorins, chlorophylls and chlorophyll-binding proteins on a Perkin Elmer Lambda 5 UV/vis spectrometer. Spectra were routinely recorded before and after Raman experiments to ensure that the laser beam had not induced decomposition of the sample. ii. Infrared Spectroscopy Infrared spectra were recorded on a Perkin Elmer model 1750 FTIR spectrometer with 2 cmd'resolution. Spectra in CCl, solution were obtained by using a 0.2 mm pathlength cell with KBr windows and.the pure solvent as reference. Samples pressed into KBr discs were recorded with an air reference. iii. Resonance Raman Spectroscopy Resonance Raman spectra were obtained on a computer- controlled Spex 1401 double monochromator equipped with a cooled RCA C31034 photomultiplier tube and photon-counting electronics (93). Data were collected at 1 cmd'intervals with 32 a 1 s dwell time and 5 cm"1 resolution. The measured Raman frequencies are accurate to 11 cm”. Laser powers employed are noted in the figure captions. Generally, the lowest power consistent with maintaining an acceptable signal to noise ratio was used. Excitation at 406.7 and 413.1 nm was provided by a Coherent model 90K krypton ion laser: at 441.6 nm by a Liconix model 4240 He-Cd laser: at 454.5, 457.9, 465.8, 472.7, 476.5, 488.0, 496.5, 501.7 and 514.5 nm by a Spectra Physics model 165 argon ion laser and at 615.0 nm from Rhodamine 590 in an argon ion pumped Spectra Physics model 375 dye laser. Collection of the Raman scattered radiation was achieved with either a 90° scattering or 180° backscattering geometry. The 90° scattering geometry permits ‘measurement of the depolarization ratio of the Raman bands. The depolarization ratio, p is defined as p=I‘L/I.., where IL and I" are the intensities of light scattered perpendicular and parallel, respectively, to the polarization of the incident laser beam. For polarized Raman bands, p<§flt depolarized.bands, p=$flt and for anomalously polarized bands, p>°/,. The measured depolarization ratios are accurate to 10.1. Samples were contained in a spinning cylindrical quartz cell to avoid photodecomposition. As a rule, the concentration of the sample was adjusted to give an optical absorbance of 2.0 in a 0.5 mm pathlength cuvette at the excitation wavelength. The 180° backscattering geometry was used for concentrated samples. These could be de-gassed and sealed under argon in quartz EPR 33 tubes. This method was particularly suited to highly fluo- rescent compounds (eg. ZnOEC). Attachment.of a low temperature Dewar enabled Raman spectra to be collected at temperatures down to -140%:(93). Backscattering was also used for samples pressed into KBr discs. Figure 2.1 shows a diagram of the apparatus designed and built for use with KBr disc samples. Translator __1 “P 34 KBr disc holder Collection optics .L__ Backscattering mirror \ Prisms Laser beam Figure 2.1 Diagram of the apparatus used for the collection of Raman spectra of KBr disc samples CHAPTER 3 VIBRATIONAL PROPERTIES OF METALLOCHLORINS 1. Introduction The presence of a chlorin macrocycle is a characteristic structural feature of chlorophylls. Knowledge of the vibra- tional properties of simpler metallochlorin model compounds is essential to the successful interpretation of the more complex resonance Raman spectra of chlorophyll. Resonance Raman spectroscopic studies of metalloporphyrins and hemo- proteins (45,52-54) have benefitted greatly from the avail- ability of a consistent set of vibrational mode assignments based on the normal coordinate analysis of NiOEP (55,56). In contrast, metallochlorins have received less attention and a consensus on their characteristic vibrational properties has yet to emerge. Metallo-octaethylchlorins (MOEC), (Figure 1.2(b)) with their highly symmetric structure are particularly suitable as metallochlorin model compounds. IR spectra of ZnOEC, CuOEC, NiOEC, Mg(OEC)py2 (py = pyridine) and Fem(OEC)X (where X = F, Cl, Br and I) were first reported by Ogoshi g; a;. (94) in 35 36 1975. The spectra were similar to those of the corresponding metallo-octaethylporphyrin (MOEP) complexes except for the appearance of new bands in the 1500 to 1700 cm"1 region that were observed to be metal sensitive. Resonance Raman spectra of metallochlorins were reported in 1979 by Ozaki gt__;. (95) for CuOEC, CuOEC-1,6-d2, CuOEC-”N“ NiOEC, Fem(0EC)x (x = F, C1) and Fem(OEC)Im2 (Im == imidazole) in CHZClz solution with Q, excitation at 488.0 nm. Based on the isotopic frequency shifts for the CuOEC complexes, metallochlorin mode assign- ments were proposed for some the high frequency bands by direct comparison with NiOEP. These assignments were later extended to include low frequency modes from a study of iron OEC and OEP complexes in a variety of spin, oxidation and ligation states (66,96). Andersson and co-workers reported resonance Raman spectra of FempPP (pPP = photoprotoporphyrin IX dimethyl ester) and FemDC (DC == deuterochlorin IX dimethyl ester) with Soret, Q, and Q, excitation (97) . Mode assignments were again proposed by comparison with the analogous iron protoporphyrin and deuteroporphyrin complexes. As pointed out by Boldt gt gt. (83), such an approach assumes that the form of the normal modes are unchanged. In their resonance Raman study of NiOEC, which was supported by normal coordinate calculations, the latter authors concluded that the metallochlorin normal modes were substantially altered. A number of modes in the reduced ring macrocycle appear to be localized rather than delocalized over the whole 37 macrocycle as occurs in the case of metalloporphyrins. Support for their assignments came from comparison with the resonance Raman spectra of Cu diol chlorin-d, (diol chlorin = cis—3' ,4'- dihydroxy-2,4-dimethyl-deuterochlorin IX dimethyl ester) in KBr recorded by Andersson gt g_. (98). In this Chapter, the IR and resonance Raman spectra of ZnOEC, CuOEC, NiOEC, CuECI (ECI = etiochlorin I) (Figure 3.1) and selectively methine-deuterated CuOEC complexes obtained with Soret, Q, and Q, excitation are presented. Both the IR and resonance Raman spectra are necessary to characterize fully the metallochlorin vibrational modes. By using the symmetric MOEC complexes the reliance on a single metal and the complications introduced.by other peripheral substituents are avoided. Metal substitution with Zn, Cu and Ni covers a wide range of core-sizes without introducing the ligand, spin and oxidation-state effects associated with the Fe complexes. The amounts of C4;,and CRR stretching character in the core- size sensitive modes are distinguished by the sensitivity of the vibrational frequency to metal substitution. Comparison of the spectra of CuOEC and CuECI identifies those modes with a contribution from cg; and ca; (5 = substituent) stretching and C“; bending coordinates as in the case with the analogous porphyrin species (99). The frequency shifts observed for CuOEC upon d, as well as oz,fi-dz and 7,6-d2 methine deuteration confirm the mode assignments and permit an assessment of the extent of localization for a given mode. The results obtained 38 MOEC: R|-R8=C2H5 M=Zn,CU,Ni MECIo R‘,R3,R5,R7 =CH3 M=CLI Fi2:‘?¢:"-"’6"'-"e‘Csz Figure 3.1 Structure and labelling scheme for metallochlorin model compounds. MOEC == transemetallo-octaethylchlorin and MECI = trans-metallo-etiochlorin I 39 support the concept of mode localization introduced by Boldt gt _a_;. although the mode composition for several prominent modes differs from that deduced in their analysis. 2. Results i. Electronic Absorption Spectra Figure 3.2 shows the electronic absorption spectra of ZnOEC, CuOEC and NiOEC in benzene solution. The spectra display the characteristic features of metallochlorins namely, separate Q, and Q, transitions with the peak absorbance of the Q,00 band about half that of the Soret absorption band. The absorption maxima are listed in Table 3.1 together with the ratio of the Q,00 to Soret oscillator strength, r for each complex. The oscillator strength ratios were determined by measuring the areas under the Soret and Q,00 bands for the absorption spectra plotted on a wavenumber scale. The energies of the Soret and Q, transitions follow the order Ni > Cu > Zn. Also, the ratio of the Q,00 to Soret oscillator strength decreases from N1 to Cu to Zn. 40 398 r. 615 400 390 ”J L) I: < 00 n: 614 0 U) DO < 617 ZnOEC 590 502 533 572| l 350 450 550 1 650 WAVELENGTH, nm Figure 3.2 Electronic absorption spectra of ZnOEC, CuOEC and NiOEC in benzene solution 41 Table 3.1 Electronic absorption maxima (nm) for MOEC complexes in benzene solution Soret QxOl ono Qyoz Qyo1 Qyoo 1' ZnOEC 401 502 538 572 590 617 0.08 CuOEC 400 495 530 571 591 614 0.09 NiOEC 398 491 523 571 590 615 0.13 r = nyOO/fSoret. These observations are in accord with the Gouterman four-orbital model (41,42). For alkyl-substituted complexes the a&,orbital is lower in energy than the alu orbital. As the metal becomes more electropositive, the energy of the an orbital is raised relative to the am orbital. This effect is manifested in the electronic absorption spectrmm as a red- shift of the Soret and Q bands and a decrease in the ratio of the Q00 to Soret oscillator strength. Table 3.2 summarizes electronic absorption spectral data for MOEP complexes in benzene solution. The Soret and Q bands red-shift and the oscillator strength ratio decreases in the order Ni:>Cu:>Zn. The oscillator strength ratios for the MOEC complexes are larger than the respective MOEP complexes reflecting the allowed nature of the Q,00 transition in metallochlorins. 42 Table 3.2 Electronic absorption maxima (nm) for MOEP complexes in benzene solution Soret Q01 Q00 r ZnOEP 404 533 569 0.04 CuOEP 400 527 563 0.05 NiOEP 393 518 552 0.07 r = fQOO/fSoret ii. Resonance Raman Spectra of MOEC The resonance Raman spectra of ZnOEC, CuOEC and NiOEC in CHZCl2 solution obtained with Soret excitation at 406.7 nm are shown in Figure 3.3. The vibrational frequencies for these species and the depolarization ratios for CuOEC are listed in Table 3.3. Five modes in the frequency region from 1490 to 1700 cm‘1 corresponding to the 1644, 1602, 1584, 1545 and 1507 cm'1 modes of CuOEC are observed to be metal dependent. Inspection of the polarized resonance Raman spectra of CuOEC at 406.7 nm (not shown) reveals that the 1545 cm'1 band consists of a 1547 cm'1 polarized mode and a 1543 cm"1 anoma- lously polarized mode. 43 m to ‘9. 1 1 N m .. 10 3‘3 2 v* 9 . n 7 l h 2 c ' 9 n T 1.0 I N :2 g 1 T to in ‘ 8 ts w z I Q T - d ' 0 N 9 N Q 1 o 1 7 8 m -m 1’.’ I I N '4 l 1.9 . N .. O : a M. ' n ' r~ V CuOEC o «e l“ ‘2 ‘ N T 1 6 a: E! Q T t 1 I :3 n n v w . g; '53 {G .. 9 $ I 2 1 '2 9 d i U m n 1 n _ I g ZnOEC T rs 1b 9 I n N L? I 1 ‘ 1 l5OO l7OO CNr' Figure 3.3 Resonance Raman spectra of ZnOEC, CuOEC and NiOEC in CHZClz solution obtained with Soret excitation at 406.7 nm. Laser powers: 5, 7 and 20 mW, respectively 44 Table 3.3 Vibrational frequencies (cm”) and Raman depolarization ratios of MOEC NiOEC ouoec 2nosc 406.7 498.0 615.0 IR 406.7 p 488.0 p 615.0 p IR 406.7 1R -- -- -— 1740 -- -- -- 1741 -- ~- -- -- -- 1712 -- -- -- 1708 -- 1708 1651 1652 1652 1652 1644 0.8dp 1643 0.5p 1644 0.4p 1644 1623 1624 1610 1613 1613 1611 1602 0.6p 1601 1.0ap 1602 0.5p 1602 1597 1596 1599 1592 1599 1591 1594 0.4p 1584 0.7dp 1583 0.7dp 1584 1570 1570 1554 1554 1554 1553 1547 0.3p 1546 0.6p 1548 0.4p 1547 1534 1538 -- -- -- -- 1543 1.6ap -- —- 1543 -- 1529 1517 1517 1517 -- 1507 0.2p 1506 0.3p 1507 0.4p 1504 1491 1491 -- -- 1502 1499 -- -- 1486 0.3p 1484 -- ~- -- 1483 -- -- -- -- -- -- -- -- 1465 1466 1466 1464 1466 0.7dp 1465 1.0ap 1465 0.5p 1464 1463 1466 -- -- -- 1457 -— -- -- —- -- 1457 -— —- -- 1453 -- -- -- 1452 -- 1452 -- —- -- -- —- -- -- 1443 1440 1441 1403 1404 1401 —— 1399 0.6p 1402 0.7dp 1401 0.5p -- 1399 -- -- -- —- 1397 -- -- -- 1396 -- 1394 1392 -- -- -- 1388 0.3p 1391 1392 0.4p -- 1387 -- -- 1383 -- -- -— -- -- 1384 -- 1383 -- —- -- 1378 -- -- -- 1375 -- 1375 1373 1373 1372 -- 1372 0.3p 1372 0.3p 1373 0.3p -- 1369 -— 1363 1363 -- 1365 1361 0.3p 1361 0.3p -- -- 1359 -- Table 3.3 (cont'd.) 45 NiOEC ouosc ZnOEC 406.7 488.0 615.0 IR 406.7 p 498.0 p 615.0 p IR 406.7 IR 1349 1351 -- -- 1352 0.2p 1350 0.3p -- -- 1349 ~- 1316 -- -- 1315 1318 0.6p 1318 1.4ap 1318 0.7dp 1318 1320 1320 1307 1306 1308 -- -- -- 1310 0.6p —- -- -- -— -- -- 1303 —- -- -- 1301 -- 1301 1275 1275 1276 1274 1276 0.3p 1275 0.7dp 1277 0.4p 1273 1275 1272 1266 1264 1266 1269 1266 0.3p 1265 0.8dp 1267 1262 1264 1266 1241 -- 1239 -- 1238 0.4p -- 1238 0.4p -- 1237 -- -- -- -- 1233 -— —- -- 1235 -- 1235 1218 1219 1219 1217 1216 0.6p 1212 0.6p 1215 0.4p 1211 1207 1205 1204 1202 1202 1200 -- 1199 0.3p 1198 0 3p 1197 1198 1194 -— 1180 -- 1180 -- 1182 0.5p -- 1183 -- 1183 1154* 1154* -- -- 1155* 1157* 1154* -- 1158 -- -- -- -- 1144 -- -- -- 1146 -- 1147 1142 -- 1140 -- 1141 0.3p ~- 1141 0.4p —- 1141 ~- 1125 1123 1124 1123 1129 1128 0.5p 1128 0.5p 1128 -- 1132 -- -- -- -- -- -- -— 1119 -- 1118 -- -- -- -- -- -- -- -- -- 1111 -- -- -- 1109 -- -- -- 1107 -- 1106 -- -- -- 1094 -- -- -- 1095 -- 1094 -- -- -- 1084 -- -- -- 1086 -- -- -- -- -- 1063 -- -- -- 1064 -- 1063 46 Table 3.3 (cont'd.) NiOEC CuOEC ZnOEC 406.7 488.0 615.0 IR 406.7 p 488.0 p 615.0 p IR 406.7 IR -- 1059 -- 1058 -- -- -- 1058 -- 1058 1028 1027 -- -- 1026 0.3p 1206 0.3p -- -- 1025 -- 1022 1022 1020 -- 1021 0.5p 1020 0.4p 1021 0.4p -- 1020 ~- -- -' -- 1014 -- -- -- 1015 -- 1013 -- -- 994 996 -- -- -- 988 -- 984 -- 960 961 -- -- 960 961 0.4p -- -- -- -- -- -- 957 -- -- -- 956 -- 957 932 932 -- 932 -- 932 932 932 -~ ~- 919 921 922 921 -- -- 923 918 -- 914 -- -- -- -— 909 0. 3p -- -- 406.7, 488.0 and 615.0 refer to laser wavelengths (nm) used Q, excitation, respectively. p - depolarization ratio: p = polarized, dp = depolarized, ap = anomalously polarized. * overlapped by Cit-,Cl2 solvent band for Soret, Q, and 47 Resonance Raman spectra of CuOEC (Figure 3.4) and NiOEC (Figure 3.5) in CHzCl2 solution were obtained with Q, excitation at 488.0 nm and with Q, excitation at 615.0 nm. The intense fluorescence of ZnOEC prohibited the measurement of its resonance Raman spectra with visible excitation. The frequencies are listed in Table 3.3 along with the depolarization ratios for CuOEC. The 488.0 nm excitation spectra agree with those reported by Ozaki gt gt. (95). For CuOEC, bands that are not seen with Soret excitation are observed at 1199, 960 and 932 cm'1 with Q, and Q, excitation: at 1182 cm'1 with Q, excitation only: and at 1486, 1310 and 923 cm'1 with Q, excitation only. The 1486 cm'1 band of CuOEC is metal sensitive, corresponding to the 1502 cm"1 band of NiOEC observed with Q, excitation. iii. IR Spectra of MOEC For a metallochlorin of C2 symmetry, the resonance Raman active modes are also IR active. The IR spectra of ZnOEC, CuOEC and NiOEC in CCl, solution are shown in Figure 3.6 and the vibrational frequencies are listed in'Table 3.3. The total number of bands observed by IR spectroscopy is greater than with resonance Raman spectroscopy as a result of IR activity of the internal vibrations of the ethyl groups. These modes can be identified by comparison with the IR spectrum of NiOEP reported by Kincaid gt gt. (100). The two closely spaced modes 48 E 9. l 1 ‘ m n fi 50 ‘ I XeX-GI. nm . N N N 6 7 ‘3 d : 9 . ' e ' N T 8 v 19 Q' o ‘- g = 3 a ' - m 1 1 n . * 6 I ‘ o m 9 . N l O 3 9 5 Xex-488.0nm N ' 9 ~ -~ , 9 Sn . T 9 '1':- :2 Q E .1 I I 8 $8 “3 n =g 8 I II T o N ‘ Q I q “A“ I I ‘ ' I I T l 900 H00 I300 I500 I700 CM" Figure 3.4 Resonance Raman spectra of CuOEC in CHzCl2 solution obtained with Q, excitation at 488.0 nm and Q, excitation at 615.0 nm. Laser powers: 100 and 35 mW, respectively 49 8 e 8 1 I g 1 :8 x -615.0 nm N ex 4 T 8 9 2 . = 9.‘ 9 I ' 6 l E. i 1 , v ‘1‘ - | . to m N l a N o: 1 1 N e o 8 1 v - N ' 2 1 10 "1'5 )‘ex' 488.0 nm is _ ~ 6 n v I 252 10 "' In T O 3 9 m n T B (D ' l '2 l T T l ISOO I700 Figure 3.5 Resonance Raman spectra of NiOEC in CHzCl2 solution obtained with Q, excitation at 488.0 nm and Q excitation at 615.0 nm. Laser powers: 70 and 40 mW, respect1vely 50 ZnOEC 0050. www— .350' Y 'Izoo' uoov 1000 Moo , 500 IGOO 1 I700 CM'I 3.6 IR spectra of ZnOEC, CCl‘ 1n CuOEC and NiOEC Figure solution 51 at 1547 and 1543 cm"1 of CuOEC that can be resolved in the Raman spectrum by depolarization ratio measurement are apparent in the IR spectrum; similarly the 1534 cmq'resonance Raman mode of ZnOEC is seen to consist of two modes at 1538 and 1529 cm". iv. Effects of Methine Deuteration Figure 3.7 shows the IR spectra of CuOEC-1,6-d2, CuOEC-a,B-d2 and CuOEC-d, in CCl, solution and Figure 3.8 shows the resonance Raman spectra of the methine-deuterated CuOEC complexes obtained with deexcitation at 615.0 nm. Different patterns of frequency shifts are observed for deuteration at the 1,6 and 02,3 positions. The 1644, 1602, 1584, 1507 and 1486 cmfl‘modes of CuOEC exhibit frequency shifts upon both a,fi and 7,6 deuteration that are additive for CuOEC-d,. Only the 1584 and 1507 cm'1 modes show equal frequency shifts upon a,fi and 1,6 deuteration. The 1644 and 1486 cm’1 modes are more sensitive to 1,6 deuteration whereas the 1602 cmfl‘mode shows a large shift upon a,fi deuteration. The 1543 cm"1 mode shows no shift upon a,fi deuteration but a shift of -17 cmd‘upon 1,6 and d, methine deuteration. Below 1350 cm“, many of the vibrational modes of CuOEC are sensitive to methine deuteration. However, the complicated pattern of frequency shifts observed for CuOEC-a,fi-d2, 52 l :0 In: I00. I80. Iv; low: l5: lvON_ Insu— Inn“. It». Infin— Ian. nun! Inn‘— «up! unan— Invfl loan. :00! 1.8-dz < CuOEC- '0,B'02 CuOEC - - _ CuOEC ' 64 'wfl . 0 ' xz‘oo' v‘v I400 I800 CM“1 3.7 IR spectra of CuOEC-1,6-d2, and -d2 CuOEC-a,B CCl, solution 1n Figure CuOEC-d, 53 N 9 1 3 ‘ In <3! I d CuOEC-y.8-d2 v q N 9'. . h v 9 I Iv .- Ja a ' I q 1 v N 9 3 ‘3 ‘ :- In 9 I g I 8 CUOEC'0.B‘02 Io ' 9 Q .— . m 0| ' 9 I " N 10 T v N . Q " v v . 9 CUOEC'd4 w v ~ 2* i 1 ‘ I 1 T 1 1 900 |200 I300 1500 I700 CM" Figure 3.8 Resonance Raman spectra of CuOEC-1,6-dz, CuOEC- a,fi-d2 and CuOEC-d, in CHzClz solution obtained with Qy excitation at 615.0 nm. Laser power: 40 mW 54 CuOEC-1,6-d2 and CuOEC-d, does not permit a one-to-one correlation of these modes in the various CuOEC complexes. Nevertheless, it is possible to recognize several modes that clearly show (1,3 or 1,6 localization in this frequency region. In order to interpret the frequency shifts for the methine-deuterated copper chlorin complexes, two methine- deuterated copper porphyrin compounds were prepared. In the CuOEP-d. species, all four methine carbons are deuterated; in the CuOEP-d2 complex, only two of the four bridging carbons are deuterium substituted at adjacent methine positions. These compounds allow assessment of the effect of full and partial deuteration on modes that are delocalized throughout the macrocycle. Resonance Raman spectra of CuOEP, CuOEP-dz and CuOEP-d, were recorded with Soret excitation at 406.7 nm (Figure 3.9) and with visible excitation at 514.5 nm (Figure 3.10) . The C,Cm stretching modes V10, 1119 and V3 of CuOEP show frequency shifts of -13, -22 and -10 cm'1 in CuOEP-d... For CuOEP-d2, the frequency shifts of the um and V3 modes are half of those observed for CuOEP-d," The 1119 mode shows a --14 cm'1 shift for CuOEP-d2 but only a further -8 cm’1 shift for CuOEP-d," The frequencies of the <3be stretching modes V2 and V11 are not shifted upon (12 or d. methine deuteration (101). The results for the um and V3 modes indicate that the vibrational mode is delocalized over the porphyrin macrocycle and that the mode composition is not altered by methine deuteration. The V19 mode, although delocalized, is mixed with 55 - l38l CuOEP - I59l - l38l N m '12 I CuOEP-d4 - l593 f 1 1 T 1 900 "'00 I500 I500 I700 CM" Figure 3.9 Resonance Raman spectra of CuOEP, CuOEP--d2 and CuOEP-d. in CHzCl2 solution obtained with Soret excitation at 406.7 nm. Laser power: 10 mW 56 2 .. a I v- “I, G N ' 0 ‘° T F - In '- o I v P l CuOEP o: [x co F co 0 v T o {B a g ‘— ID 1- P I CuOEP-d2 “ o o In .. N V | to In P ‘30 co ‘ o I; g to "' v- v- I '3‘ I .. ,_ o I 1?- “ so I N I . S’- -1 a v ,_ *- 2?» . 8 I . I v- V ' T .1 I 1 I I v ‘ 900 1 1 00 1300 1 500 1700 CM’1 Figure 3.10 Resonance Raman spectra of CuOEP, CuOEP-d2 and CuOEP-d, in CHzCIZ solution obtained with visible excitation at 514.5 nm. Laser power: 50 mW 57 can bending character (56) and most likely changes in mode composition upon deuteration. Taken together, these results for partially and fully methine-deuterated metallo-octa- ethylporphyrins provide a basis for the analysis of the corresponding metallo-octaethylchlorin complexes. Most impor- tantly, they show that for a delocalized mode, methine d2 deuteration can be expected to produce roughly half of the d, shift. v. Resonance Raman Spectra of CuECI For metalloporphyrins, the change in peripheral substituents that occurs upon going from CuOEP to CuEPI (Figure 3.1) affects those modes with a contribution from Cbe and CbC, stretching and CbC, bending coordinates (99) . A similar dependence of the frequency of the analogous modes on peripheral substituents is expected for the metallochlorins and provides a means by which these modes may be identified. Figure 3.11 shows the resonance Raman spectra of CuECI in CHzClz solution obtained with Soret excitation at 406.7 nm, Qx excitation at 488.0 nm and Q’ excitation at 615.0 nm. The 1584, 1547 and 1543 cm’1 modes of CuOEC show increases in frequency of +4, +4 and +3 cm”, respectively, as a result of the change in peripheral substituents on going to CuECI. In addition, the 1238, 1215 and 1198 cm‘1 modes of CuOEC also exhibit increases in frequency. The change in peripheral 58 -124I -982 .H40 -4270 -4550 )\ e x«615.0nrn ~1220 998 -I205 ‘93l Xe, - 488.0nm ‘flGOI '1600 )‘ex- 406.7nm -I644 '4000 I 900 1 I 300 1 I300 1 1500 CM" Figure 3.11 Resonance Raman spectra of CuECI in CHZClz solution obtained with Soret excitation at 406.7 nm, Qx excitation at 488.0 nm and Q, excitation at 615.0 nm. Laser powers: 6, 70 and 40 mW, respectively 4—1 I700 59 substituents lowers the frequencies of the 1372, 1361, 1350 and 1277 cm'1 modes of CuOEC by -4, -5, -3 and -‘7 cm“, respectively, and the 1021 cmd'mode appears to split into two components at 998 and 982 cm'1 in CuECI. vi. Solid State Spectra For their normal coordinate analysis of NiOEC, Boldt g; _l. (83) obtained resonance Raman spectra of NiOEC pressed into Nagxh pellets. However, differences occur in the number and frequencies of the resonance Raman modes of NiOEC in Nazso, and in CHzClz solution. In order to compare the results obtained here with those of Boldt gt, al., solid state resonance Raman spectra were recorded for NiOEC (Figure 3.12) and CuOEC (Figure 3.13) in KBr pellets. The high frequency modes of NiOEC and CuOEC are shifted to lower frequencies in KBr. Additionally, the high frequency bands of NiOEC display an.altered.patternxof relative intensities with 406.7 nm Soret excitation for the KBr sample and Boldt's NazSO. sample compared to NiOEC in CHZClz solution. Resonance Raman spectra of CuOEC in KBr on the other hand, showed the same pattern of relative intensities as in 01,012 solution with Soret, Q, and (g.excitation. The 1572 cmq'band in the NiOEC spectrum appears to be an artifact of the solid state since its relative intensity increased as the ratio of NiOEC’to KBr in the pellet was lowered. No evidence for this mode could be found in the 60 -1019 —1233 J -1275 - 1548 I “1144 Au = 615.0 nm — 1394 I ~1123 ~1307 O 6: ~ :2 to 5 v .I T ' I; CD (D Q 1- m -I v- I u) I 8 7L"=488.0 nm I} [s ‘D 53'- ' 3 m 8 lu=406.7 nm J o a $3 to ‘3 S E g I I '3 U 52 _, 1- v- N v- ' s: ' I 'T ' I ' I v I f 1 900 1100 1300 1500 1700 CM" Figure 3.12 Resonance Raman spectra of NiOEC in KBr obtained with Soret excitation at 406.7 nm, Q, excitation at 488.0 nm and Q, excitation at 615.0 nm. Laser powers: 20, 70 and 50 mW, respectively 61 1018 CD - It. s In 3; " v- - N I I P m I v A. = 615.0 nm - 1- OX 1- 8 n ' 51‘ 3', 8 - ' (‘0 a '- ow "’ 0" n [s O | v I n " N n '- 0 v- V‘ | v- -I I U 1396 - 1364 - 1539 g -1459 -1499 1582 C; 1581 1598 -1631 I" = 488.0 nm . K“ = 405.7 nm 'I :2 80 n 1‘33 . " I “ g s l I 5 3 ‘ I I .1 d ' I ' T ‘ I 1 1 900 1100 1300 1500 1700 CM'1 \ Figure 3.13 Resonance Raman spectra of CuOEC in KBr obtained with Soret excitation at 406.7 nm, Q, excitation at 488.0 nm and Q, excitation at 615.0 nm. Laser powers: 20, 75 and 50 mW, respectively 62 IR spectrum of NiOEC in KBr or for an analogous mode in the resonance Raman spectra or IR spectrum of CuOEC in KBr. 3. Discussion i. C,,CIn and Cbe Modes Upon reduction of a egg bond in a metalloporphyrin to form a metallochlorin, the molecular symmetry is lowered from D“, to C2. For C2 symmetry, the resonance Raman active and IR active vibrational modes belong to the A and B symmetry species. The resonance Raman active polarized A18 and depolarized Bu modes of D“ symmetry are expected to produce polarized modes of A symmetry in C2. The anomalously polarized A2, and depolarized le5 modes correlate with depolarized B modes. The IR activeIEg modes split into A and B. The A and 8 modes are both IR and Raman active. Under C2 symmetry, eight cg; stretching (4A + 48) and four cg; stretching (2A + 28) modes are predicted. Four of the ' QC, and two of the CbCb stretching modes are derived from E,I modes. In their initial study, Ozaki gt a;. (95) assigned the resonance Raman modes of CuOEC at 1643 (p), 1584 (dp) (see 102) and 1506 (p) cm'1 as corresponding to the v10 (81,, dp), V19 (A2,, ap) and V3 (A1,, p) modes of NiOEP, respectively. The observed polarizations of these three modes with 488.0 nm 63 excitation are consistent with the D“, to C2 symmetry correlation. The 1546 cm'1 mode was presumed to consist of a polarized and an anomalously polarized component. The ap mode that shifts to 1527 cm"1 on 1,6 deuteration was assigned to the other C4; mode. If this mode is not derived from an Eu mode it would have to correspond to v28 (82,, dp) of NiOEP calculated at 1469 cm'1 (55). The 1602 (ap) and 1546 (p) cm'1 modes of CuOEC, which.are unshifted upon 1,6 deuteration, were assigned as corresponding to u; (A1,,p) and un (Bludp). However, the anomalous polarization of the 1602 cm'1 mode is inconsistent with this mode being derived from a porphyrin.Au mode. Such a symmetry-lowering approach does not adequately account for the vibrational mode assignments for the metallo- chlorins. For these compounds our results indicate that there is mixing of C,,CIn and Cbe stretching character in the high frequency modes that is not predicted by direct comparison with the normal modes of NiOEP. This is apparent in the frequency region from 1475 to 1700 cm'1 where seven modes, corresponding to the 1644, 1602, 1584, 1547, 1543, 1507 and 1486 cm’1 modes of CuOEC, are observed to be metal, and hence core-size, dependent. 64 The relationship between vibrational frequency and core- size of a metalloporphyrin often follows the empirical expression (62): v=K(A-d) where u is the vibrational frequency in cm”, d is the center to nitrogen distance or core-size of the metalloporphyrin in A, and K (cmfl/A) and A (A) are parameters characteristic of the macrocycle. The physically useful parameter is the slope or K value, which is proportional to the amount of 0,0, stretching character in the vibrational mode. Ozaki gt gt. (66) have previously applied this kind.of analysis to metallo- chlorins; they achieved core-size variation by working with iron complexes in different spin and oxidation states. The iron chlorins, however, are susceptible to the same sorts of difficulties in such an analysis as are encountered with the corresponding iron porphyrins (63,65) and for this reason our analysis has been carried out for metals other than iron. The correlations between vibrational frequency and core-size for the high frequency modes of the Ni, Cu and Zn complexes of OEC and OEP are shown in Figures 3.14 and 3.15. The same core-sizes are assumed for the porphyrin and chlorin com- plexes, namely Ni (1.958 A) (60), Cu (2.000 A) (103) and Zn (2.047 A) (104). The core-size correlation parameters, K and A, are listed in Table 3.4 for MOEC and in Table 3.5 for MOEP. From.the chlorin K'values, the 1644, 1602, 1543, 1507 and 1486 65 1660 .. P VIBRATIONAL FREQUENCY, cm"1 1450 I I . J I . . . 4 I 1 .950 2.050 CORE-SIZE, A Figure 3.14 Core-size correlation for the high frequency modes of MOEC 66 1 660 VIBRATIONAL FREQUENCY, crn'1 1450 . . . 1.950 2.050 CORE-SIZE, A Figure 3.15 Core-size correlation for the resonance Raman active (-—-) and IR active (- -) high frequency modes of MOEP 67 mucoauaumnsm Hmuocawuom uo omcmno co vufinm xocmsvmuu .nm .000 2000 . .HUNSU OH Umozo EONH "umnsm no . has? a. e. 33 00.. won“ 2.: S I 3: 33.3.5.0.0. mm: «a: 3 8.3.0.0.. £35.06. 3.0 :3 o o? T «I 3: ~03 33.3.2.0.0. .3....n.3.0.0.. $3 23 m G...q.3.0.0.. :4. RN o 2: w- 6.. 83 53 33.2.0.0. 4510.0. 32 I m I .0.0.. .$.£.0.0.. CS 62.. n+ SI SI o 2.3 II 8.3.0.0.. .3366. ~an 33 s .06.. 3.: m3 4+ nI o 7 $3 33 33.3.0.0. $3 23 6 3.3.0.0.. 83 82 m 3:26.366. 2.0.0.. $5 ha 4+ mu T T 33 $3 33.5.0.0. $3 83 v 8.3.0.0.. 33 :3 n 8.3.0.0.. 5... 23 o 2.. NI NT 33 33 3.5.0.0.. 23 33 N 2.5.0.0.. $3 33 H 8.3.0.0.. £35.06. 3.1. Zn 6 o? 7 nu 33 $3 ucoscmamo¢ .030 .93 .oz ucoficmwmm< 92 317.53 00390 .0 :7»... £730 omono omofiz nomad: uou muasmou mwm>nacu ouocfiouooo amsuoz 4 x .:< so :4 ac moooa Hnfiuos 3&0 one Jflu :wuoHnooHHnuoa now wuouofimuan acaumaouuoo ouwonuoo can A7500 muuanu meouomw ..?auv mofiocosvouu Hoseaununfi> v.n manna 68 Table 3.5 Vibrational frequencies (cmq), isotope shifts (cm“) and core-size correlation parameters for metalloporphyrin C5; and C“% normal modes Au Au Av Kc A Mode' CuOEP d; d. subst" (cm'1/A) (A) P.E.D.‘ V10 1639 -6 -13 +3 405 6.05 y(c,c_)49, u(c.c.,)17 y, 1591 +1 0 +6 236 8.74 u(c,,c,,)6o, 0(Cb-Et)19 6,, 1584 -14 -22 -- 449 5.53 u'(c_c,)67, v'(C.Cb)18 .1, 1572 0 -3‘1 +6 203 9.73 u(c,,c,,)57, u(c.-I'-:t)16 v3 1505 -5 -1o +2 382 5.94 u(c,c_)41, u(c,c,,)3s 6,, -- -- -- -- --- ---- 6(cbc.)57, va-Et)l6' .3, 1554‘ -- -— -- 426 5.66 u(c.c_)34, u'(c.c.)24' .3, 1483‘ -- -- -- 278 7.34 u'(c.c_)36, u'(c,N)17 Mode designation and P.E.D. from the normal coordinate analysis of NiOEP (55). Au subst: frequency shift on change of peripheral substituents from CuOEP to CuEPI (99). calculated for ZnOEP, CuOEP and NiOEP using data from ref. 105. see ref. 101. P.E.D. for v37 and u38 have been interchanged (100). from ref. 100. 69 cm‘1 modes of CuOEC have substantial C,Cm stretching, but for the 1584 and 1547 cm'1 modes, contribution from the CbCI. stretching coordinate is more apparent. The 1486 cm'1 mode of CuOEC shows the highest K value. Significantly, the K values for the other chlorin C,C, stretching modes are all lower than the porphyrin K values for the Raman active C,Cm stretching modes: v10 (K=405), 019 (K=449) ,and 03 (K=382) and the IR active V38 (K=426) . The chlorin K values are close to the K value for the porphyrin IR active 039 (K=278) , which has 36% C.C.. stretching character in the calculated P.E.D. for NiOEP. 0f the two chlorin Cbe stretching modes, the K value of the 1584 cm"1 mode is virtually identical to that of ya (K=236) whereas the K value of the 1547 cm”1 mode is lower than that of 1111 (K=203) (106) . The different K values that we observe for the chlorins indicate altered mode compositions compared to the MOEP complexes: in particular, they suggest that there is mixing of C,CIn and Cbe stretching coordinates in the high frequency normal modes. To characterize further the high-frequency modes we have examined their sensitivity to methine deuteration and to a change in peripheral substituents upon going from OEC to ECI. The frequency shifts observed for CuOEC-d. and CuECI (Table 3.4) can be compared to CuOEP-d, and CuEPI (Table 3.5). In methine-deuterated CuOEP-d“ the C,,CIll stretching modes um, um and V3 exhibit frequency shifts of -13, -22 and -10 cm"1 compared to CuOEP but the €be stretching modes V; and V11 are 70 unaffected. In addition, Kincaid gt g_. observed -11 and -7 cm'1 shifts for the cg; stretching modes u” and u” from the IR spectra of NiOEP and NiOEP-d4, (100). The 1644, 1602, 1584, 1543, 1507 and 1486 cm"1 modes of CuOEC show substantial frequency shifts of -10, -13, -9, -17, -13 and -10 cm'1 upon ch methine deuteration. The magnitude of the frequency shifts indicates that these modes contain CK; stretching character. Comparison of the resonance Raman spectra of CuOEP with CuEPI showed that a change in peripheral substituents resulted in an increase in frequency of the 04x stretching modes 0, and un by +6 cm"1 (99). For CuOEC, we observe that the 1584, 1547 and 1543 cm’1 modes increase in frequency by +4, +4 and +3 cmfl'in CuECI. These modes are therefore predicted to contain (:bCb stretching character. The mode characters deduced from IL methine deuteration and the change in peripheral substit- uents (Table 3.4) are in good agreement with those predicted from the K values for MOEC. The results of the core-size correlation together with the analysis of the wavenumber shifts incurred upon methine deuteration clearly demonstrate that there is mixing of the v(C,C,) and u(Cbe) internal coordinates in the high frequency normal modes. The above results show that the mode characters observed for the high-frequency modes of CuOEC cannot be assigned by direct comparison .with the normal modes of NiOEP. The resonance Raman spectrum of CuOEC-1,6-dz reported by Ozaki gt gt. (95) showed no shift of the 1602 cm"1 mode of CuOEC upon 71 1,6 deuteration and led these authors to conclude that the 1602 cm"1 mode corresponded to the porphyrin V2 01,0,> stretching mode. We observe that the 1602 cm'1 mode shifts to 1589 cm'1 in CuOEC-d, indicating substantial C,CInn stretching character. To reconcile these deuteration effects the frequency shifts for CuOEC-1,6~dz, CuOEC-0,3-d2 and CuOEC-d. were compared. As seen in Figures 3.7 and 3.8 and Table 3.4, the high-frequency modes exhibit different shifts upon 0,3 and 1,6 deuteration. The large (~13 cm'l) frequency shift of the 1602 cm'1 mode results from deuteration at the 0,3 positions but deuteration at the 1,6 positions has little effect on the frequency of this mode. 0n the other hand, the 1543 cm'1 mode of CuOEC is shifted by ~17 cm’1 upon 1,6 deuteration and is unaffected by deuteration at the 0,3 positions. Both the 1644 and 1486 cm'1 modes of CuOEC show unequal frequency shifts upon 0,3 and 7,6 deuteration with the larger shift resulting from 1,6 deuteration. The frequency shifts upon deuteration that we observe for CuOEC support the idea of mode localization in metallochlorins as proposed by Boldt gt _l. (83) from their study of NiOEC. In Table 3.4 our results are compared with the assignments of Boldt gt gt. for NiOEC. Before proceeding to our analysis, we note that.differences occur in the number and frequencies of the resonance Raman modes of NiOEC in Nazso. and in CHzCl2 solution reported here. The 1644, 1608 and 1572 cm'1 modes of NiOEC in NaZSO, pellets are not observed in 72 either the resonance Raman or IR spectra of NiOEC in solution and appear to be artifacts of the solid state spectra of NiOEC. These modes are not considered further in this analysis. The 1648, 1614 and 1512 cm"1 modes of NiOEC in Nazso, were assigned by Boldt gt at. to C3;,stretching modes that are 1,6 localized, 0,3 localized, and delocalized, respectively. In CuOEC, we observe the analogous modes at 1644, 1602 and 1507 cm'l. From their methine d, shifts and apparent insensitivity to changes in peripheral substituents at the c; positions, these modes are assigned to 0.0, stretching vibrations in agreement with Boldt gt g_. Furthermore, the observation of equal and additive frequency shifts upon 0,3 and 1,6 deuteration shows that the 1507 cmq‘mode is delocalized over the chlorin macrocycle. Our deuteration shifts also confirm that the 1602 cm'1 mode is 0,3 localized. The 1644 cm’1 mode shows both 0,3 and 1,6 C$;,stretching character but with a greater localization in the'1,6 methines than in the 0,3 methines. This is a slight alteration from the assignment of Boldt gt gl. who represented this mode as a 50/50 in-phase linear combination of the vm and.u$h modes of NiOEP. Since we observe frequency shifts upon both 0,3 and 1,6 deuteration, a more accurate representation of this mode could be derived from a 50/20 in-phase combination of um and.u,h. The 1644 and 1507 cm'1 modes of CuOEC retain much of the character of the um and v3 modes of NiOEP. 73 The compositions deduced for the 1584 and 1486 cm"1 modes of CuOEC do not agree with those proposed by Boldt gt g1. The 1584 cm'1 mode exhibits mixed Cbcb and 0,0, stretching character as judged by the intermediate K value, the +4 cm’1 shift upon change of peripheral substituents and the ~9 cm"1 shift upon d. methine deuteration. The equal ~4 cm"1 shifts upon 0,3 and 1,6 deuteration indicate a delocalized mode. Boldt gt 3].. assigned the analogous 1590 cm'1 mode of NiOEC in Nazso. to C,C,,, stretching localized in the 0,3 methines. Disagreement is also found for the 1492 cm'1 mode of NiOEC in Nazso. which was assigned by Boldt gt g1. to C,C,, stretching localized in rings I, II and III. We observe the corresponding mode at 1486 cm'1 for CuOEC in CH2C12 solution with Q, excitation. The high K value and ~10 cm'1 shift upon d. methine deuteration indicate 0,0. stretching rather than 0.0,, stretching character. The frequency shifts upon 0,3 and 1,6 deuteration show a greater 1,6 localization for this mode. For CuOEC, we assign the 1547 cm’1 mode to (3be stretching and the 1543 cm'1 mode to a mixture of Cbe stretching and 1,6 localized C.C, stretching. The observation of these two modes in CuOEC is interesting since the normal coordinate analysis of Boldt gt g1. predicts four modes in this region at 1561, 1557, 1540 and 1532 cm". Two of the modes (calculated at 1561 and 1557 cm“) are predicted to have (2be stretching character while the calculated 1540 cm’1 mode is predicted to have C,Cb stretching character. In their resonance Raman spectra of 74 NiOEC in Nazso, only a single mode at 1546 cm'1 was observed and this mode did not shift in NiOEC-1,6~d2. This 1546 cm'1 mode of NiOEC was assigned by Boldt gt g1. to C,,Cb and 0,3 localized C,C_, stretching calculated at 1532 cm". The Cbcb stretching character seen in the 1547 and 1543 cm'1 modes supports the normal coordinate analysis although the 0,0, stretching character in the 1543 cm’1 mode has 1,6 localization and not 0,3 localization. ii. C,N Modes Under C2 symmetry, six C.N stretching modes would be expected for a metallochlorin (107). The resonance Raman spectrum of CuOEC-”N, recorded by Ozaki gt gl. (95) showed that the 1402, 1372, 1361, 1318, 1157 (see 108) and 1128 cm’1 modes of CuOEC shifted by ~1, ~6, ~5, ~5, ~9 and ~19 cm", respectively, upon ”N, substitution. These modes are therefore assigned to C,N stretching. In .Table 3.6, we have summarized the Raman active modes observed for CuOEC and NiOEC in CHZClz solution and for NiOEC in Nazso, along with the assignments proposed by Boldt gt g1. (83) from their normal coordinate analysis. In this region, solution spectra gave somewhat better resolution than pellet data as the 1367 cm'1 mode of NiOEC in Nazso. (and in our spectra in KBr, Figure 3.12) has two components in CHZCl2 solution at 1373 and 1363 cm". Similarly, for CuOEC in KBr (Figure 3.13) , we observe a single Table 3.6 Vibrational frequencies (cm”) and isotope shifts for metallochlorin C,N modes Ian") Normal coordinate analysis Au Au results for NiOECb NiOEC CuOEC 15N." subst. No. Obs. Calc. Assignment 1404 1402 -1 o 14 1402 1489 uC,Cb(I,II,III), uC,N(I,III,IV) 1373 1372 ~6 -4 -- 1363 1361 -5 -5 17 1367 1339 uC,N(I,II,III,IV), VCbe(I,II,III) 1306 1318 -5 -3 20 1308 1301 6C,H(0,3), uC,Cb(I,III) 1154 1157 -9 -3 27 1152 1147 quEt(I,II,III), VC,N 1123 1128 -19 -3 29 1124 1083 quEt(I,II,III), SC,C,N (IV) ‘ from ref. 95. b from ref. 83. 76 1364 cm’1 mode that corresponds to the 1372 and 1361 cm'1 bands of CuOEC in solution. The normal coordinate analysis of NiOEC indicates that there is mixing of the C,N character with C,Cb, Cbe and CbEt stretching coordinates. The involvement of the Cb atoms in these vibrational modes can be seen from the sensitivity of four of the modes to a change of peripheral substituents in CuECI (Table 3.6). The 1308 and 1124 cm'1 modes of NiOEC in Nazso, were assigned to modes 20 and 29, respectively, but the P.E.D.'s have no contribution from C,N stretching. In order to account for the observed 15N), frequency shifts it would perhaps be better to assign the 1318 cm"1 mode of CuOEC to mode number 19 (uC,N(I,III,IV), 6C,H(1,6)) and the 1128 cm‘1 mode to number 28 (quEt(I,II,III), vC,N(I,III)). iii . Cmfi Modes The normal coordinate calculations of Boldt gt gt. (83) indicate that the 0,3 and 1,8 hydrogen motions of NiOEC are mixed with porphyrin skeletal modes and furthermore, that the 1,6 hydrogen motions are mixed with C1,}! deformations on ring IV. Table 3.7 collects our information on the more prominent modes that have can contributions. As was found for the skeletal modes, the 0.11 modes can be divided into three classes: a) those that are localized to the 0,3 porphyrin-like 77 Table 3.7 Vibrational frequencies (cm'l) for CuOEC normal modes in the frequency region below 1350 cm'1 CuOEC 0,3-d2 7,6~d2 d. NiOEP mode # 1. 0,3 localized 1277 -- 1275 -- -- 1215 946 1215 946 um -- 1325 -~ 1322 uu -- 1178 -- 1180 u“ 2. 1,6 localized 1238 -~ 1224 ~- 1198 ~- -- 1173 1141 1140 914 914 78 part of the molecule, b) those that are 1,6 localized, and c) those that are substantially delocalized throughout the macrocycle. The 0,3 CmH deformations show clear porphyrin-like behavior. Comparison of the spectra of CuOEC and CuOEC-0,3-d2 (Figure 3.8) reveals that the mode at 1215 cmd'shifts to 946 cmd'and an intense band at 1178 cmd'appears upon deuteration. This behavior is analogous to that exhibited by the um and v“ modes of NiOEP (55,56,59). Moreover, the appearance of the 1325 cm’1 mode in the Cu0]i:C-0,3~d2 spectrum reflects that of V12 in NiOEP, which is usually not observed in the natural abundance spectrum (55). Boldt. gt, gt. predicted. an 0,3 localized mode at 1277 cm'1 and the spectra in Figure 3.8 confirm this. However, this vibration does not have a corresponding metalloporphyrin normal mode. The vibrational coordinates of the above four modes are 0,3 localized as they are unperturbed by 1,6-d2 deuteration and their frequencies in the 0,3-d.2 species persist virtually unchanged in CuOEC-d“ Some indication that the 0,3 CtH bending deformations contribute to more delocalized modes can be seen in the behavior of the cluster of modes in the 1190-1250 cmd'region, which shows complicated intensity and frequency shifts in all three of the deuterated complexes relative to CuOEC. The 1,6 CIll deformations were suggested by Boldt gt gt. to be extensively mixed with ring modes and thus to exhibit 79 complex behavior. Nonetheless, the data of Figure 3.8 provide several key insights into the characteristics of these motions. Upon 1,6-d2 deuteration the mode at 1141 cm'1 disappears and new modes at 914 and 1173 cmd'become apparent. The 1238 cm'1 mode apparently downshifts to 1224 cm”1 upon the same substitution. The behavior of the 1141 and 914 cmfl’modes indicates that these motions are strongly localized as their frequencies carry through essentially unchanged in the 0,3-d2 and d, species, respectively. A judgement as to the extent of localization of the 1173 cmq'mode is rendered difficult since this region is obscured by the strong 0,3 localized 1180 cm’1 mode in CuOEC-0,3-d2 and CuOEC-d.. Soret excitation data (not shown), however, suggest that.this mode is also 1,6 localized. iv. Ethyl Group Vibrations The IR spectrum of CuOEC shows seven modes at 1464, 1452, 1375, 1064, 1058, 1015 and 956 cm'1 that are insensitive to metal substitution, methine deuteration and a change in peripheral substituents and are therefore assigned to ethyl group vibrations. These modes are not observed in the resonance Raman spectra except for the 1464 cmd'mode that is present in nearly all of the resonance Raman spectra of the chlorin complexes that we have examined. The chlorin macrocycle does not influence the internal vibrations of the ethyl groups since the frequencies are all close to the values 80 l. (100) from the IR spectrum of NiOEP reported by Kincaid gt in an argon matrix. 4. Conclusions From our analysis of the resonance Raman and IR spectra of the MOEC complexes we have shown that the high frequency vibrational modes of metallochlorins cannot be assigned by direct comparison with the normal modes of metalloporphyrins. The normal coordinate analysis of NiOEC performed by Boldt gt gt. therefore represents the starting point for the descrip~ tion of the vibrational modes of metallochlorins. In the work presented here, the mode compositions deduced for metallo~ chlorins by metal substitution, methine deuteration and a change in the ‘peripheral substituents of OEC Ihave been compared with the assignments proposed by Boldt gt gt. The overall agreement is fairly good although several mode assignments need to be modified in the normal coordinate analysis. The agreement between our results and the normal coordinate analysis of Boldt gt gt. is especially gratifying considering that it is the first normal coordinate treatment of a metallochlorin. The wealth of information obtained from the resonance Raman spectra of metalloporphyrins and hemoproteins has been possible largely because of the accepted 81 set of mode assignments based on the normal coordinate analysis of NiOEP published by Abe gt gt. in 1978 (55). However, several other groups have also performed normal coordinate treatments for metalloporphyrins (109-116): the first, by Ogoshi gt gt. (109) appearing in 1972. Even now there is not complete agreement on metalloporphyrin normal mode assignments, especially for the IR active Eu modes. Recent years have seen new normal coordinate analyses that have attempted to include the peripheral substituents in the calculations (115,116). Similarly, futher normal coordinate treatments for metallochlorins can be expected. We hope that our results for CuOEC and the methine-deuterated CuOEC complexes will provide the basis for a new normal coordinate analysis to improve the vibrational mode assignments for metallochlorins. CHAPTER 4 VIBRATIONAL PROPERTIES OF METALLOCHLORIN fl CATION RADICALS 1. Introduction In photosynthesis, the primary charge separation event occurs at the reaction center chlorophylls P~700 of Photosystem I (4) and P~680 of Photosystem II (5). In this process, an electron is removed from the II system of the chlorophyll macrocycle resulting in the formation of a chlorophyll 1r cation radical. P~700 is thought to have a dimeric structure in the ground neutral state but in the oxidized state, P~7OO+ the unpaired electron appears to be localized, on one of the. two chlorophylls (11-14). Lutz compared the resonance Raman spectra of the bacterio- chlorophyll a cation radical in solution and in the reaction center, P~870 of Rtodopseugomonas sphaeroides (117). It was concluded that the positive charge is localized on one of the two bacteriochlorophylls of P~87O+ on the time scale of the resonance Raman effect (~107” s). Establishment of the vibra- tional properties of metallochlorin Ir cation radicals is essential if resonance Raman spectroscopy is to be used to characterize the reaction center chlorophylls. 82 83 The vibrational properties of metalloporphyrin t cation radicals have been characterized by resonance Raman spec- troscopy in this lab by T. Oertling and A. Salehi (99, 118-120). The results from these studies showed similar potential energy distributions in the normal modes for both the neutral and oxidized MOEP species. In other words, the mode compositions do not change upon oxidation of the porphyrin macrocycle. The frequencies of the cg; stretching modes were observed to increase, while those of the C5; and (aN stretching modes decreased. In collaboration with these authors, preliminary results of the investigation into the resonance Raman spectral characteristics of metallochlorin w cation radicals have been reported (121). The n cation radicals of metallo-methyloctaethylchlorin (MMeOEC) (Figure 4.1) were chosen for the initial resonance Raman study. These compounds do not have vicinal protons on the reduced ring (122) so that oxidative dehydrogenation back to the porphyrin is not possible. The resonance Raman spectra of the Cu, Co and Ni complexes of MeOEC show two modes in the 1500-1520 and 1620-1650 om’1 regions that appeared to be analogous to the C,C, stretching modes V3 and um, respectively of metalloporphyrins (121). Upon oxidation of the chlorin ring, the I’m-like mode shifted by ~10 cm'1 but the v3~like mode exhibited.a slight increase in frequency. The K values for the I’m-like mode indicated C,,Cm stretching character in both MMeOEC (K=385) and MMeOEC” (K=375). The decrease in frequency 84 MMeOEC M = C0, CU, Ni Figure 4.1 Structure of metallo-methyloctaethylchlorin (MMeOEC) 85 upon oxidation for the vm~like mode is consistent with the results of Oertling gt gt. (120) for a C5; mode of MOEP. The lower’ K 'value and increase in frequency’ upon oxidation suggested that the ug~like mode of MMeOEC consisted of mixed C.C, and Cbcb stretching character. In MMeOEC” it was proposed that the frequency increase from the Cbe character offsets the frequency decrease from the C5; character (121). The study of the I cation radicals of MMeOEC ‘was performed before the analysis of the vibrational mode characteristics of the MOEC complexes had been completed. To explore further the vibrational properties of metallochlorin 1f cation radicals, the resonance Raman spectra of the 1r cation radicals of CuOEC, CuECI (Figure 3.1) and. the 'methine~ deuterated CuOEC complexes (CuOEC-0,3-d2, Cu0EC~—y,6~d2 and CuOEC-d0 are presented in this Chapter. Comparison of the spectra of the n cation radicals with those of the neutral species shows that our initial interpretation requires modification and that unlike the situation in metallo- porphyrins, changes in mode composition occur upon oxidation of metallochlorins. 86 2. Results Metallochlorins were oxidized chemically at room temperature by using the AgClO../ CHZClz method described by Salehi gt g_. (118). The electronic absorption spectrum of CuOEC“ in CHZCl2 solution is shown in Figure 4.2. The spectrum displays a split Soret band and a disappearance of the visible bands. Resonance Raman spectra of CuOEC, CuECI, CuOEC-0,3-d2, CuOEC-7,4‘S-d2 and CuOEC-d, and their Ir cation radicals were obtained with Soret excitation by using the 363.8 nm line from a Coherent Innova 100-20 Ar“ laser. Spectra were recorded by T. Oertling on a Spex 1877 triple monochromator equipped with OMA III electronics. Figure 4.3 shows the resonance Raman spectra of the neutral copper chlorin complexes in CHZClz solution. The resonance Raman spectra of the corresponding 1 cation radicals are shown in Figure 4.4. In the high frequency region, the 1642, 1601, 1582, 1547 and 1507 cmq'modes of CuOEC are replaced by another set of modes at 1632, 1611, 1605, 1560 and 1511 011'1 in CuOEC“. 87 389 374 m o z < m n: 8 <—— X2 :0 < 478 I l l J 300 400 500 600 700 WAVELENGTH, nm Figure 4.2 Electronic absorption spectrum of CuOEC+' in CHZCl2 solution 88 five I kex = 363.8 nm a) Cu trans- OEC >._..wzm.rz_ Z Cu > Zn z Mg which is the same trend observed for the MOEC complexes. Conjugation of the keto carbonyl on ring V causes a further red-shift of the absorption bands compared to the respective MOEC complex. The oscillator strength ratios, although much higher than for 104 _ 392418 647 37 422 399 Cu 423 8 603 Z 3‘, 405 503 546 65, a: O U) m < 376 Zn 430 411 383 Mg 1 L J 350 450 550 650 750 WAVELENGTH, nm Figure 5.1 Electronic absorption spectra of chlorophyll a and Zn, Cu and Ni-substituted chlorophyll a in diethyl ether solution 105 MOEC, also follow the order Ni > Cu > Zn zMg. Two 7) bands are observed on the high energy side of the Soret band. These are forbidden in unsubstituted porphyrins but become allowed when there is a conjugated carbonyl substituent on a porphyrin or reduced porphyrin. Table 5.2 Electronic absorption maxima (nm) for metal- substituted chlorophyll a in diethyl ether solution Metal '72 711 Soret 0:01 ono Qyoi Qyoo 1' Mg 380 409 428 531 575 614 660 0.28 Zn 376 405 423 518 562 606 654 0.25 CD 399 422 503 546 603 650 0.40 N1 371 392 418 496 536 590 647 0.53 r = nyOO/fSoret The electronic absorption spectra of chlorophyll a in acetone, methanol and pyridine solution are shown in Figure 5.2 and the absorption maxima are listed in Table 5.3. The spectra of chlorophyll.a in acetone and diethyl ether solution are characteristic of 5-coordinate Mg. In pyridine solution, the Mg of chlorophyll a is 6-coordinate and under these conditions, the Qx00 band shifts to 640 nm. The Mg of chloro- phyll a is also 6-coordinate in methanol solution but there is no shift of the Qx00 band. However, the Soret band is lower in intensity and broadened. 106 _ 430 662 411 383 Acetone 432 5‘5 418 579 535 666“ m :fim t) z < m ‘6 m Methanol m 4 443 618 580 536 671 422 n 396 Pyridine 640 620 . l 1 l n l n l 350 450 550 650 750 WAVELENGTH, nm Figure 5.2 Electronic absorption spectra of chlorophyll a in acetone, methanol and pyridine solution 107 Table 5.3 Electronic absorption maxima (nm) for chlorophyll a in various solvents SCI-vent '72 "1 soret Q10]. 0:00 03701 QyOO Diethyl ether 380 409 428 531 575 614 660 Acetone 383 411 430 535 579 616 662 Methanol 390 418 432 536 580 618 666 Pyridine 396 422 443 --- 640 620 671 ii. Resonance Raman Spectra Figure 5.3 shows the resonance Raman spectra of chlorophyll a and the metal-substituted chlorophyll a com- plexes obtained with Soret excitation at 406.7 nm. The resonance Raman spectrum of chlorophyll a was recorded in frozen acetone at -125°C. Spectra of Zn, Cu and Ni chl a were recorded. in. diethyl ether solution. at room 'temperature. Vibrational frequencies are listed in Table 5.4. Twelve modes of Cu chl a exhibit a core-size dependency. The core-size correlation parameters K and A in the relation v = K (A - d) (62) for these twelve modes in M chl a are also included in Table 5.4. 108 . , . ‘ v 1 - 1 9 0 1100 1300 1500 1700 CM-1 Figure 5.3 Resonance Raman spectra of chlorophyll a in acetone solution and Zn, Cu and Ni-substituted chlorophyll a in diethyl ether solution obtained with Soret excitation at 406.7 nm. Laser powers: 35, 20, 20 and 20 mW, respectively 109 Table 5.4 Resonance Raman frequencies (cmq) and core-size correlation parameters for metal-substituted chlorophyll a Mg Zn Cu Ni K A - 2 E00 " hV 'iI‘ E01 -1111 -11" where , and the second term to resonance with the first vibrational level of the Soret excited state, |i1>. 142 In the calculation of the excitation profile, the Franck-Condon factors and are approximated by the extinction coefficient, 6, for the Soret band maximum. This simplified expression predicts an excitation profile for a vibrational mode to have two peaks, one at the 0-0 tran- sition and one at the 0-1 transition. As the vibrational frequency approaches the magnitude of the damping term the two peaks in the excitation profile coalesce. Calculated excitation profiles for a high frequency 1600 cm'1 mode and a low frequency 300 cm'1 mode of chlorophyll a and chlorophyll b are shown in Figure 6.12. The extinction coefficient and linewidth are obtained from the electronic absorption spectrum of the chlorophylls in acetone solution. The positions of the Soret maxima are for the chlorophylls in LHC. The parameters used for chlorophyll a are: e = 101.5 mM’lcm'l, I‘ = 440 cm'1 and Eoo= 22936 cm'1 (= 436 nm): and for chlorophyll b: e = 148.0 mM'lcm'l, I‘ = 640 cm'1 and E008 21053 cm'1 (= 475 nm) . Figure 6.12 shows two peaks in the excitation profiles for the 1600 cm”1 mode but only one for the 300 cm'1 mode. The 1600 cm’1 mode of chlorophyll a is enhanced with 413.1 nm or 441.6 nm excitation but not with 472.7 nm or 514.5 nm excitation. The 1600 cm’1 mode of chlorophyll b is enhanced with 441.6 or 472.7 nm excitation but not with 413.1 nm or 514.5 nm excitation. The low frequency 300 cm'1 mode of chloro- phyll a is enhanced with 441.6 nm excitation. For chlorophyll b, the low frequency 300 cm'1 mode is enhanced with 472.7 nm 143 Chl a Chl b " f‘\/\ I, \ E “ v = 1600 cm-1 (D 2 \ Ill I— \ E \ Z < \ IE. 5 \ < m EXCITATION WAVELENGTH, nm Chla >_ v=300cm'1 t 3 DJ Chlb I'- E /\ '2' ’ \ s ’ \ n: g «2 I, I2\ In 1- ‘- N ‘6‘ <- 3 / S \\ IT: I \.L EXCITATION WAVELENGTH, nm Figure 6.12 Calculated excitation profiles for a 1600 cm"1 mode and a 300 cm'1 mode of chlorophyll a (—) and chlorophyll b (- -) in LHC 144 excitation. These calculated excitation profiles provide a useful means of interpreting the pattern of chlorophyll a and chlorophyll b vibrational modes observed in LHC with changing excitation wavelength. In previous studies by Lutz on the resonance Raman spectra of chlorophylls in yiyg it was assumed that the 441.6 nm line provided selective enhancement of chlorophyll a modes and lines from 457.9 nm to 472.7 nm could be used for selec- tive enhancement of chlorophyll b modes. The results obtained here clearly show that care must be taken in asserting that a resonance Raman spectrum corresponds to only chlorophyll a or chlorophyll b. Selective enhancement for the high and low frequency modes of chlorophyll b is indeed possible at 472.7 nm. For chlorophyll a however, the low frequency modes may be selectively enhanced at 441.6 nm and the high frequency modes at 413.1 nm. ii. Mg Coordination State The electronic absorption spectra of chlorophyll a in solution are characteristic of the coordination state of the Mg atom. The most pronounced effect is the large red-shift of the Qx00 Iband of 6-coordinate chlorophyll a in. pyridine solution compared to 5-coordinate chlorophyll a in acetone solution (Figure 5.2). One possible mechanism to broaden the 145 absorption bands of the chlorophylls in LHC would be to create a mixture of 5 and 6-coordinate chlorophyll species in the protein. In Chapter 5, the core-size sensitive resonance Raman modes of chlorophyll a were identified by metal substitution and it was demonstrated that the frequencies of these modes could be used to diagnose the coordination state of the Mg atom in chlorophyll a. From these results and those of Fujiwara and Tasumi ( 142) , the resonance Raman bands of chlorophyll a at 1527-1529, 1551-1554 and 1606-1612 cm'1 indicate 5-coordination while a set of lower frequency bands at 1518-1521, 1545-1548 and 1596-1599 cm”‘ indicate 6- coordination. The two resonance Raman bands of chlorophyll a at 1551 and 1610 cm'1 in the 413.1 nm excitation spectrum of LHC are characteristic of S-coordinate Mg. The strong 1526 cm"1 carotenoid band obscures the third core-size sensitive chlorophyll a mode. There is no evidence for any 6-coordinate chlorophyll a. For chlorophyll b, core-size sensitive resonance Raman modes at 1520-1523, 1564-1566 and 1607 cm'1 were determined by Fujiwara and Tasumi (142) to be characteristic of 5-coordination. Upon 6-coordination, the 1564-1566 cm"1 mode splits into two components at 1549-1551 and 1559 cmq'and the two other core-size sensitive modes shift to lower frequencies at 1516-1519 and 1594-1596 cm". The 441.6 nm excitation 146 resonance Raman spectrum of LHC shows two chlorophyll b bands at 1568 and 1607 cm'1 indicating S-coordination. Again, the third core-size sensitive mode is obscured by the 1526 cm’1 carotenoid band. The 1552 cmd’band in the 441.6 nm excitation spectrum is a chlorophyll a vibrational mode and not an indicator of 6-coordinate chlorophyll b. 472.7 nm excitation clearly shows the 1568 cm"1 5-coordinate chlorophyll b mode with no contributions from 6-coordinate chlorophyll b (173). The observation of 5-coordination only for the chloro- phyll a and chlorophyll b molecules in LHC rules out mixtures of 5 and 6-coordinate chlorophyll species as a mechanism to broaden the absorption bands. X-ray crystal structures of the reaction center from.‘3. ;yi;idis (157,158) and a light- harvesting bacteriochlorophyll a-protein from the green photo- synthetic bacterium Prosthegochlori s w (174) show that all of the bacteriochlorophylls are 5-coordinate. The four bacteriochlorophyll b molecules in the reaction center complex and five of the seven bacteriochlorophyll a molecules in the 2. aestuarii complex are ligated by histidine residues from the protein. Resonance Raman spectroscopy can distinguish between 5 and 6-coordination for chlorophyll but cannot identify the ligands. EL. 147 iii. Carbonyl Stretching Region In agreement with the findings of Lutz, the resonance Raman spectra of chlorophyll a and chlorophyll b in LHC and as monomers in acetone solution differ mainly in the carbonyl stretching region. Chlorophyll a has a conjugated keto carbonyl at the 9-position (Figure 1.1). In chlorophyll b there are two conjugated carbonyl groups: a keto carbonyl at the 9-position and a formyl carbonyl at the 3-position. Table 6.2 lists the bands observed for LHC in the carbonyl stretching region and their assignments. Table 6.2 Chlorophyll a and chlorophyll b carbonyl stretching frequencies (cm ) in LHC 413.1 441.6 472.7 Assignment 1697 unbound keto chl a 1693 unbound keto chl b 1686 1688 unbound keto chl a 1675 1674 H-bonded keto chl a 1659 1656 H-bonded keto chl a 1635 1636 H-bonded formyl chl b 1626 1627 H-bonded formyl chl b 148 Resonance Raman spectra of LHC obtained with 413.1 nm excitation show four chlorophyll a keto carbonyl stretching modes at 1697, 1686, 1675 and 1659 on“. 441.6 nm excitation is expected to enhance both chlorophyll a and chlorophyll b carbonyl stretching modes. The frequencies of the chlorophyll a keto carbonyl stretching modes at 1688, 1674 and 1656 cm‘1 are close to those observed with 413.1 nm excitation. The two modes at 1635 and 1626 cm'1 are assigned to chlorophyll b formyl carbonyl stretching modes. The Soret band maximum of chlorophyll b in LHC occurs at 475 nm (= 21053 cmfl) so that the E01 transition for a 1630 cm‘1 mode of chlorophyll b is expected at 22683 cm’1 or 440.9 nm. Strong enhancement of the chlorophyll b formyl carbonyl stretching modes is therefore expected with 441.6 nm excitation. Only chlorophyll b modes are enhanced in the 472.7 nm excitation spectrum. Using this excitation wavelength, the two formyl stretching modes at 1636 and 1627 cm"1 are seen as well as the chlorophyll b keto carbonyl stretching mode at 1693 cm“. The variety of chlorophyll a and chlorophyll b carbonyl stretching frequencies in LHC indicates different populations of chlorophyll molecules in the protein. Four populations of chlorophyll a and two of chlorophyll b can be distinguished from the resonance Raman spectra of LHC. The interactions with the chlorophyll carbonyl groups may account for the broadened and red-shifted absorption bands in LHC. TI 149 The two modes at 1697 and 1686 cm”1 in LHC most likely correspond to stretching vibrations of the unbound keto carbonyls of chlorophyll a molecules within two different environments in the protein. The frequency of the keto car- bonyl stretching mode of chlorophyll a is highly dependent on the solvent. The resonance Raman spectra of chlorophyll a in acetone, diethyl ether and pyridine solution at -125°C show the keto carbonyl stretching frequency at 1681, 1686 and 1681 cm", respectively (Figure 5.4) . Fujiwara and Tasumi (142) and Koyama et al. (175) have reported resonance Raman spectra of monomeric chlorophyll a in polar solvents at room temperature. The keto carbonyl stretching frequency ranged from 1680 to 1702 cm'1 with the frequency increasing as the dielectric constant of the solvent decreased. The two carbonyl stretching modes at 1697 and 1686 cm'1 are in the frequency range expected for an unbound keto carbonyl of monomeric chlorophyll a. The 413.1 nm excitation spectrum of LHC shows two additional chlorophyll a keto carbonyl stretching modes at 1675 and 1659 cm". Lutz has proposed that the chlorophyll carbonyls that are not free are H-bonded to amino acid residues in the protein (76) . The resonance Raman spectrum of chlorophyll a in methanol solution (Figure 5.4) shows the H-bonded keto carbonyl stretching frequency at 1661 cm". The frequency of this mode increases to 1668 cm”1 in methanol solution at room temperature (142,175). Koyama gt; :1. observed that the H-bonded keto carbonyl stretching frequency for 150 chlorophyll a in other alcohol solutions ranged from 1670 to 1675 cm”1 and appeared to be independent of the dielectric constant of the solvent. The 1675 and 1659 cm"1 bands may therefore be attributed to chlorophyll a H-bonded keto carbonyl stretching modes with differing H-bond strengths. H-bonding to the conjugated keto carbonyl of chlorophyll a can. produce a red-shift. of the electronic absorption spectrum. Babcock and Callahan (176) determined that H-bonding to the conjugated formyl group of heme a model compounds lowered the carbonyl stretching frequency and resulted in a red—shift of the a (0%) absorption band. The magnitude of the carbonyl stretching frequency decrease and red-shift both increase as the strength of the H-bond increases. The low frequencies of the 1675 and 1659 cm'1 chlorophyll a keto carbonyl stretching modes and the red-shifted absorption bands in LHC can be explained by H-bonding. A larger red-shift is expected for the chlorophyll a population in LHC that has the 1659 cm”1 keto carbonyl stretching frequency. The same type of analysis can be applied to the chloro- phyll b carbonyl stretching frequencies in LHC. The 1693 cm‘1 mode in the 472.7 nm excitation is attributed to the unbound keto carbonyl of chlorophyll b. H—bonded formyl carbonyls are observed at 1636 and 1627 cm”; Thus there are two populations of chlorophyll b molecules in LHC with H-bonded formyls of differing strength but with their keto carbonyls free. 151 iv. 28 kDa and Reaction Center Complex Proteins Ghanotakis gt _1. have isolated an oxygen-evolving PS II reaction center core complex from spinach that contains approximately 60 chlorophylls per PS II (92). Treatment of PS II membranes with octyl glucopyranoside separates LHC from the reaction center complex of 9 polypeptides with molecular weights of 47, 43, 33, 32, 30, 28, 20, 10 and 9 kDa. Solu- bilization of the reaction center complex with dodecyl fi-D- maltoside followed by gel-filtration chromatography leaves a PS II reaction center core complex containing the 47, 43, 33, 32, 30 and 9 kDa polypeptides. The 28 kDa chlorophyll-binding protein is also isolated in this procedure. The LHC, 28 kDa protein and LHC-depleted PS II reaction center complexes isolated by Ghanotakis gt a1. were studied by resonance Raman spectroscopy. The resonance Raman spectra of LHC from spinach (Figures 6.9-6.11) are similar to those of LHC from pea (Figures 6.5 and 6.6). These spectra show the presence of both chlorophyll a and chlorophyll b in spinach LHC with 5-coordination of the Mg atoms. Two populations of chlorophyll b molecules can be distinguished by the stretching frequencies of their formyl carbonyls at 1631 and 1641 cm'1 in the 441.6 and 476.5 nm excitation spectra. The chlorophyll a carbonyl stretching region is not resolved since spectra were recorded at 4°C as compared to -125°C for LHC from pea. 152 Resonance Raman spectra of the 28 kDa and.reaction center complex proteins recorded with 406.7 and 441.6 nm excitation show only chlorophyll a and carotenoid vibrational modes. The reaction center complex preparation still contains 60 chloro- phylls per PS II. The resonance Raman spectra are dominated by the antenna chlorophylls so that it 'is not possible to observe vibrational modes of the reaction center chlorophylls or the two pheophytin a molecules in each reaction center. More importantly, the lack of any chlorophyll b vibrational modes in the 476.5 nm excitation spectra of the 28 kDa protein and reaction center complex confirms the absence of chloro- phyll b in these two complexes. Ghanotakis g; a_. suggested that the 28 kDa protein is either a separate protein intermediate between LHC and the PS II reaction center complex or a component of LHC. The 28 kDa protein has a molecular weight that is typical of LHC polypeptides but these authors were unable to isolate the 28 kDa protein from LHC. The absence of chlorophyll b indicates that the 28 kDa protein is not part of LHC. y-.f'r'1"l ‘1': ‘ . 153 4. Conclusions LHC contains three types of chromophore: chlorophyll a, chlorophyll b and carotenoids. The primary function of this protein.is to absorb light. Comparison‘with the spectra of the chlorophylls in solution shows that the electronic absorption bands of LHC are broadened and red-shifted. Broadening of the absorption bands allows light to be absorbed over a greater wavelength range thereby increasing the efficiency of light energy utilization. This is accomplished by the creation of four separate populations of chlorophyll a and two of chloro- phyll b in LHC. Different coordination states of the Mg atoms are not a factor since all of the chlorophylls are 5- coordinate. H-bonds to the conjugated keto carbonyls are responsible for creating two of the chlorophyll a populations. The local protein environments may account for the two other chlorophyll a populations. The chlorophyll b molecules are distinguished by the strength of the H-bond to their conjugated formyl carbonyls. Each.of these species.has its own distinct absorption spectrum that contributes to the overall absorption spectrum of LHC. H-bonding to the unconjugated ester carbonyls may also occur and help to orient the chlorophyll molecules in the protein but will not influence the absorption spectrum. ’7 154 Determination of the structures of the chlorophylls in LHC is essential to understanding their function. Once a photon has been absorbed by LHC, the excitation energy needs to be transferred to the reaction center. Elucidation of the mechanism of exciton migration from the antenna chlorophylls to the reaction center requires knowledge of the arrangements of the chlorophylls in the protein and the relation of LHC to the reaction center polypeptides. Help in this area is likely to come from biochemical procedures for isolating the individual chlorophyll-binding proteins and.determining their amino acid sequences. CHAPTER 7 SUGGESTIONS FOR FURTHER WORK The original goal of this research project was to explore the vibrational properties of chlorophyll and chlorophyll- binding proteins using resonance Raman spectroscopy. Resonance Raman spectra.of chlorophyll a and.chlorophyll b‘were recorded in solution. and. in ‘the light-harvesting' chlorophyll a/b protein complex. However, a major obstacle in the inter- pretation of these spectra was the lack of a set of vibrational mode assignments for chlorophyll. Attention was therefore focused on determining the vibrational charac- teristics of metallochlorins as models for the more complex chlorophyll. In contrast to the approach taken by other groups (45,66,95-98), the vibrational modes of metallochlorins could not be assigned by a direct comparison with the normal modes of metalloporphyrins. The starting point for the interpretation of the resonance Raman and IR spectra of metalloporphyrins is the normal coordinate analysis of NiOEP performed by Abe gt g1. (55,56). The set of mode assignments for NiOEP are widely accepted in resonance Raman studies of metalloporphyrins 155 n are: Mann-mi, 5‘ 156 although there is still disagreement over the form of the IR active Eh modes. The NiOEP normal coordinate analysis attributes modes ya, and was to C.CI and Cbe stretching, respectively. These assignments are supported by the work of Willems and Bocian on acetyl and formyl-substituted deuteroporphyrins (177,178) . Other groups have argued that the assignments for V37 and V3. should be reversed (65,100,179) since it is u“ that exhibits the larger core-size dependency and deuteration shift characteristic of C.C. stretching. Oertling gt g1. have shown that changing the peripheral substituents from OEP to EPI can identify those resonance Raman modes with a contribution from (3be and CbC, stretching and CbC, bending coordinates (99) . Comparison of the IR spectra of MOEP and MEPI would be useful for clarifying the E; mode characters. TheIE; modes for NiOEP have been recalculated by Abe (compare ref. 56 with 55) but the mode characters for ufl and v33 were unchanged. A possible source of error in the determination of the Eutmodes is the use of resonance Raman frequencies from NiOEP in CIIIZCl2 solution but IR frequencies from NiOEP in a CsI disk. NiOEP has two crystal structures: a planar triclinic form (60) and a buckled tetragonal form (180). We have observed resonance Raman spectra of both forms of NiOEP in KBr disc samples depending on the ratio of porphyrin to KBr in the sample. Ideally, resonance Raman and IR spectra of metalloporphyrins should be measured in the same medium. r..044..d:.t.a_e .— 4.010.443.«Alirelwflreetuou.1.. ...e.....¢.l..h..o. . .. To. :1 a. a .114. ... . 4. I. _. I ..I.. a... .4 - . . It... . acnisrdli.‘ 157 Numerous factors such as the central metal, axial ligation, peripheral substitution, macrocycle distortions, ring oxidation and reduction influence the vibrational properties of metalloporphyrins in biological systems. Many of these effects are well understood but one area that requires further investigation is to determine the conse- quences of symmetry-lowering. The mode assignments for metalloporphyrins are based on the normal coordinate analysis of the highly symmetric NiOEP molecule. Protoheme and heme a are two examples of naturally-occurring porphyrins that have conjugated peripheral substituents. The presence of conjugated substituents eg. vinyl and carbonyl groups, lowers the molecular symmetry thereby allowing the Eu modes to become resonance Raman active. Analyses have been reported for protoporphyrin IX (65,116,179) as well as formyl and acetyl- substituted deuteroporphyrins (177,178). Further studies of model compounds are needed to establish precisely how far the symmetry must be lowered to induce Raman activity of the Eu modes (and IR activity of the resonance Raman modes). Determination of excitation profiles would be useful since Shelnutt has developed expressions for calculating excitation profiles of metalloporphyrins that include the effects of symmetry-lowering (181,182). A special case of symmetry-lowering is the reduction of a Cbe bond in a metalloporphyrin to form a metallochlorin. We have seen that the high frequency modes of metallochlorins are 158 simultaneously IR and resonance Raman active and exhibit mixing of C.C, and Cth stretching character. The normal coordinate analysis of NiOEC by Boldt gt g1. (83) represented several modes as linear combinations of normal modes of NiOEP. For example, the 1648 cm'1 mode of NiOEC in NazSO. was described as a 50/50 linear in-phase combination of um and ya,“ It is not clear whether this is simply a convenient representation of the form of the metallochlorin normal mode or if it is actually derived by mixing of the metalloporphyrin v10 mode with the E“l mode V37 as a result of symmetry-lowering. There are several experiments to complete the charac- terization of the vibrational modes of metallochlorins. Our study has concentrated on the modes above 900 cm"1 which leaves the assignments for the low frequency modes unresolved. Also, a complication in the assignment of the Cu}! modes is that the normal coordinate analysis of NiOEC predicts that the a,fi and 7,6-hydrogen motions are mixed with porphyrin skeletal modes and that the 7,6-hydrogen motions are further mixed with on: deformations on ring IV. Involvement of the ring IV CbH modes can be tested by preparing CuOEC with deuterium substitution on ring IV. Results for this complex as well as the selectively methine-deuterated CuOEC complexes could then be utilized in a refinement of the normal coordinate analysis for MOEC . .13.. an. 3! .- 159 Once the mode assignments for metallochlorins have been firmly established then the features that are of interest to the study of metallochlorins in nature e.g. peripheral sub- stituents and axial ligation can be explored. Excitation profiles of the intensities and depolarization ratios of the metallochlorin Raman bands are a topic for investigation. Ozaki gt g1. (95) measured the depolarization ratios for the high frequency modes of CuOEC with 488.0 nm excitation and attempted to correlate the metallochlorin resonance Raman modes with those of NiOEP. However, the resonance Raman bands of CuOEC show a wide variation in the depolarization ratio depending on the excitation wavelength. For example, the 1602 cm'1 mode of CuOEC has a depolarization ratio of 0.6 (p) at 406.7 nm, 1.0 (ap) at 488.0 nm and 0.5 (p) at 615.0 nm. Excitation profiles and depolarization ratio studies would help in understanding the resonance Raman scattering mechanism and symmetry effects in metallochlorins. Although chlorophyll contains a chlorin macrocycle, the effects of the peripheral substituents and ring V prevent a direct correlation of the resonance Raman modes of chlorophyll with those of MOEC. Boldt gt g1. examined a series of Ni chlorins in which the structural features of chlorophyll a were added stepwise (83). However, only a single metal was utilized. Our results for metal-substituted chlorophyll a show that ring V plays a major role in determining the vibrational properties of chlorophyll. The normal modes of metallochlorins 160 and metalloporphyrins cannot be directly compared but characterization of porphyrin model compounds with an isocyclic ring would help in understanding the effects of ring V in chlorophyll. Suitable model compounds are protochlo— rophyll a, which is a porphyrin with the same peripheral substituents as chlorophyll a (183,184) and the porphyrin compounds prepared by Kenner and co-workers (185-187). The peripheral substituents also complicate the determination of the vibrational mode characters of chlorophyll. It is possible that vibrational modes of the conjugated carbonyl groups are mixed with those of the chlorophyll macrocycle. Centeno observed that several resonance Raman modes of heme a contained contributions from the conjugated formyl group by use of O“‘substitution.of the formyl (188). This substitution was achieved through preparation of the Schiff base of heme a followed by hydrolysis with Hgf‘. The same approach may be used for chlorophyll a except that the Schiff base must be prepared for pyrochorophyll a (189). Reaction of chlorophyll a with amines to form the Schiff base would instead result in the rupture of ring V (190,191). H-bonding to the conjugated keto and formyl carbonyls of chlorophyll a and. chlorophyll b, respectively, has. been proposed as a mechanism to account for the broadened and red-shifted electronic absorption bands of the chlorophylls in LHC. This conclusion is based on a comparison with the results obtained for H-bonded heme a model compounds (176). 161 To support the LHC results, a further study of H-bonding to the chlorophyll carbonyls could be carried out. Aggregation of chlorophyll in H-bonding solvents such as alcohols can be avoided by using the pheophytins or metal-substituted chlorophylls. A final suggestion concerns the manner in which future resonance Raman studies of metalloporphyrins are conducted. Resonance Raman work on metalloporphyrins and hemoproteins for the past decade have relied almost exclusively on the normal coordinate analysis of NiOEP performed by Abe gt g1. in 1978 (55) . While the mode assignments for NiOEP have benefitted the study of porphyrins, investigations of other macrocycles have been limited. Currently, the best approach is that of the Bocian group. In recent years, these workers have explored a variety of new systems including metallochlorins (83, 192, 193) , chlorophyll model compounds (83), bacteriochlorins (193), a bacteriochlorophyll a model compound (194) and metallo- porphyrin anion radicals (195). The resonance Raman spectra were supported by normal coordinate calculations for each system thus removing the reliance on the NiOEP mode assignments. It is anticipated that computer calculations will assume a much greater role in future resonance Raman studies of metalloporphyrins and other related systems. n LIST OF REFERENCES 10. 11. 12. 13. 14. 15. LIST OF REFERENCES Clayton,R.K. Photosynthesis: Physical Mechanisms and Chemical Patterns: Cambridge Univ. Press: Cambridge, MA, 1980. Forti,G. in Photosynthesis: Amesz, J., Ed.: Elseiver: Amsterdam, 1987: pp 1-20. Hill,R.: Bendall,F. Nature (London) 1960, 186, 136-137. Kok,B. Biochem. Biophys. Acta 1956, 22, 399-401. Doring,G.: Stiehl,H.H.: Witt,H.T. Z. Naturforsch. 1967, 22b, 639-644. ' Davis,M.S.: Forman,A.: Fajer,J. Proc. Natl. Acad. Sci. USA 1979, 76, 4170-4174. Norris,J.R.: Uphaus,R.A.: Crespi,H.L.: Katz,JuJ. Proc. Natl. Acad. Sci. USA 1971, 68, 625-628. Philipson,K.D.: Sato,V.L.: Saueer. Biochemistry 1972, 11, 4591-4595. Norris,J.R.; Scheer,H.: Druyan,M.E.: Katz,J.J. Proc. Natl. Acad. Sci. USA 1974, 71, 4897-4900. Shipman,L.L.: Cotton,T.M.: Norris,J.R.: Katz, J.J. Proc. Wasielewski,M.R.: Norris,J.R.: Crespi,H.L.: Harper,J. J. Am. Chem. SOC. 1981, 103, 7664-7665. Den Blanken,H.J.; Hoff,A.J. Biochim. Biophys. Acta 1983, 724, 52-61. Dikanov,S.A.; Astashkin,A.V.: Tsvetkov,Yu.D.: GOldfeld,M.G. Chem. Phys. Lett. 1983, 101, 206-210. O'Malley,P.J.: Babcock,G.T. Proc. Natl. Acad. Sci. USA 1984, 81, 1098-1101. Wasielewski,M.R.: Norris,J.R.: Shipman,L.L.: Lin,C.-P.: Svec,W.A. Proc. Natl. Acad. Sci. USA 1981, 78, 2957-2961. 162 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 163 D6rnemann,D.: Senger,H. Photochem. Photobiol. 1986, 43, 573-581. Watanabe,T.: Kobayashi,M.; Hongu,A.: Nakazato,M.: Hiyama,T.: Murata,N. FEBS Lett. 1985, 19, 252-256. Setif,P.; Mathis,P. Arch. Biochem. Biophys. 1980, 204, 477-485. Klimov,V.V.: Allakhverdiev,S.I.: Demeter,S.: Krasnovskii,A.A. Dokl. Akad. Nauk. SSSR 1979, 249, 227-230. Thornber,J.P. Ann. Rev. Plant Physiol. 1975, 26, m 127-158. F Burke,J.J.: Ditto,C.L.: Arntzen,C.J. Arch. Biochem. Biophys. 1973, 137, 252-263. Glazer,A.N. Ann. Rev. Biochem. 1983, 52, 125-157. Dolphin,D., Ed. The Porphyrins; Academic Press: New York, 1978: Vol. VII. h Scheer,H. in The Porphyrins: Dolphin,D., Ed.: Academic Press: New York, 1978: Vol. II, pp 1-44. Svec,W.A. in The Porphyrins: Dolphin,D., Ed.: Academic Press: New York, 1978: Vol. V, pp 341-399. Sibbet,S.S.; Hurst,J.K. Biochemistry 1984, 23, 3007-3013. Babcock,G.T.; Ingle,R.T.: Oertling,W.A.: Davis,J.C.: Averill,B.A.; Hulse,C.L.: Stufkens,D.J.: , Bolscher,B.G.J.M.: Wever,R. Biochem. Biophys. Acta 1985, 828, 58-66. Timkovich,R.: Cork,M.S.; Gennis,R.B.; Johnson,P.Y. J. Am. Chem. SOC. 1985, 107, 6069-6075. Chang,C.K. J. Biol. Chem. 1985, 260, 9520-9522. Chang,C.K.: Wu,W. J. Biol. Chem. 1986, 261, 8593-8596. Strauss,S.H.: Silver,M.E.: Ibers,J.A. J. Am. Chem. Soc. 1983, 105, 4108-4109. Strauss,S.H.: Thompson,R.G. J. Inorg. Biochem. 1986, 27, 173-177. Feng,D.: Ting,Y.-S.: Ryan,M.D. Inorg. Chem. 1985, 24, 612-617. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 164 Stolzenberg,A.M.: Strauss,S.H.: Holm,R.H. J. Am. Chem. Soc. 1981, 103, 4763-4778. Strauss,S.H.; Silver,M.E.: Long,K.M.: Thompson,R.G.: Hudgens,R.A.: Spartalian,K.: Ibers,J.A. J. Am. Chem. Soc. 1985, 107, 4207-4215. Gallucci,J.C.: Swepston,P.N.: Ibers,J.A. Acta Cryst. 1982, B38, 2134-2139. Chow,H.-C.: Serlin,R.: Strouse,C.E. J. Am. Chem. Soc. 1975, 97, 7230-7237. Jensen,A.; Aasmundrud,0.: Eimhjellen,K.E. Biochim. VB Biophys. Acta 1964, 88, 466-479. ' Strain,H.H.: Cope,B.T.: McDonald,G.N.: Svec,W.A.: Katz,J.J. Phytochemistry 1971, 10, 1109-1114. Dougherty,R.C.: Strain,H.H.; Svec,W.A.: Uphaus,R.A.: Katz,J.J. J. Am. Chem. SOC. 1970, 92, 2826-2833. 1T Gouterman,M. J. Chem. Phys. 1958, 30, 1139-1161. Gouterman,M. J. Molec. Spectrosc. 1961, 6, 138-163. Longuet-Higgins,H.C.: Rector,C.W.: Platt,J.R. J. Chem. Phys. 1950, 18, 1174-1181. Weiss,C. in The Porphyrins: Dolphin,D., Ed.: Academic Press: New York, 1978: Vol. III, pp 211-223. Kitagawa,T.: Ozaki,Y. Structure and Bonding 1987, 64, 71-114. Tang,J.: Albrecht,A.C. in Raman Spectroscopy: Szimanski,H.A., Ed.: Plenum Press: New York, 1970: pp 33-68. Clark,R.J.H.: Stewart,B. Structure and Bonding 1970, 36, 1-80. Rousseau,D.L.: Friedman,J.M.: Williams,P.F. in Topics in Current Physics: Weber,A., Ed.: Springer-Verlag: Berlin, 1979: Vol. 11, pp 203-252. Carey,P.R.: Salares,V.R. in Advances in Infrared and Raman Spectroscopy: Clark,R.J.H.: Hester,R.E., Eds.: Heyden: London, 1980: Vol. 7, pp 1-58. Parker,F.S. Applications of Infrared, Raman, and Resonance Raman Spectroscopy in Biochemistry; Plenum Press: New York, 1983. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 165 Spiro,T.G., Ed. Biological Applications of Raman SpectroscOpy; Wiley: New York, 1988; Vol. 3. Felton,R.H.: Yu,N.-T. in The Porphyrins: Dolphin,D.,Ed.: Academic Press: New York, 1978: Vol. III, pp 347-393. Spiro,T.G. in Iron Porphyrins: Lever,A.B.P.: Gray,H.B., Eds.: Addison-Wesley: Reading, MA, 1983: Part Two, pp 89-159. Babcock,G.T. in Biological Applications of Raman Spectroscopy: Spiro,T.G., Ed.: Wiley: New York, 1988: Vol. 3, pp 293-346. Abe,M.: Kitagawa,T.; Kyogoku,Y. J. Chem. Phys. 1978, 69, 4526-4534. Abe,M. in Spectroscopy of Biological Systems: Clark,R.J.H.: Hester,R.E., Eds.: Wiley: London, 1986; pp 347-393. Kitagawa,T.: Ogoshi,H.: Watanabe,E.; Yoshida,z. J. Phys. Chem. 1975, 79, 2629-2635. Kitagawa,T.: Abe,M.: Kyogoku,Y.: Ogoshi,H.: Watanabe,E.: Yoshida,z. J. Phys. Chem. 1976, 80, 1181-1186. Kitagawa,T.: Abe,M.: Ogoshi,H. J. Chem. Phys. 1978, 69, 4516-4525. Cullen,D.L.: Meyer,E.F.Jr. J. Am. Chem. Soc. 1974, 96, 2095-2102. Spaulding,L.D.: Chang,C.C.: Yu,N.-T.; Felton,R.H. J. Am. Chem. Soc. 1975, 97, 2517-2525. Huong,P.V.: Pommier,J.-C. C. R. Acad. Sci. Paris 1977, 0285, 519-522. Spiro,T.G.: Stong,J.D.: Stein,P. J. Am. Chem. Soc. 1979, 101, 2648-2655. Ca11ahan,P.M.: Babcock,G.T. Biochemistry 1981, 20, 952-958. Choi,S.: Spiro,T.G.: Langry,K.C.: Smith,K.M.: Budd,D.L.: La Mar,G.N. J. Am. Chem. Soc. 1982, 104, 4345-4351. Ozaki,Y.: Iriyama,K.: Ogoshi,H.: Ochiai,T.: Kitagawa,T. J. Phys. Chem. 1986, 90, 6105-6112. Teraoka,J.: Kitagawa,T. J. Phys. Chem. 1980, 84, 1928-1935. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. 80. 81. 82. 83. 166 Lutz,M. C. R. Acad. Sci. Paris 1972, 8275, 497-500. Lutz,M. and Breton,J. Biochem. Biophys. Res. Commun. 1973, 53, 413-418. Lutz,M. J. Raman Spectrosc. 1974, 2, 497-516. Lutz,M.: Kleo,J. C. R. Acad. Sci. Paris 1974, D279, 1413-1416. Lutz,M.: Kleo,J.: Gilet,R.: Henry,M.: P1us,R.: Leickman, J.P. in Proc. 2nd. International Conference on Stable Isotopes, Oak Brook, IL: Klein,E.R.: Klein,P.D., Eds.: U.S. Dept. of Commerce: Springfield, VA, 1975: pp 462-469. Lutz,M. in Lasers in Physical Chemistry and Biophysics; Joussot-Dubien,J., Ed.: Elsevier: Amsterdam, 1975: pp 451-463. Lutz,M. Biochem. Biophys. Acta 1977, 460, 408-430. Lutz,M.: Brown,J.S.: Remy,R. in Chlorophyll Organization and Energy Transfer in Photosynthesis: Wolstenholme,G.: Fitzsimons,D.W., Eds.: Excerpta Medica: Amsterdam, 1979: pp 105-125. Lutz,M. in Advances in Infrared and Raman Spectroscopy: Clark,R.J.H.: Hester,R.E., Eds.: Wiley Heyden: London, 1983: Vol. 11, pp 211-300. Nanba,O.: Satoh,K. Proc. Natl. Acad. Sci. USA 1987, 84, 109-112. Akabori,K.: Tsukamoto,H.: Tsukihara,J.: Nagatsuka,T.: Motokawa,0.: Toyoshima,Y. Biochem. Biophys. Acta 1988, 932, 345-357. Ghanotakis,D.F.: de Paula,J.C.: Demetriou,D.M.: Bowlby,N.R.: Petersen,J.: Babcock,G.T.: Yocum,C.F. Biochim. Biophys. Acta 1989, 974, 44-53. Whitlock,H.W.Jr.: Hanauer,R.: Oester,M.Y.: Bower,B.K. J. Am. Chem. Soc. 1969, 91, 7485-7489. Eisner,U.: Linstead,R.P.: Parkes,E.A.: Stephen,E. J. Chem. SOC. 1956, 1655-1661. Falk,J.E. in Porphyrins and Metalloporphyrins: Elseiver: New York, 1964: p 798. Boldt,N.J.: Donohoe,R.J.: Birge,R.R.: Bocian,D.F. J. Am. Chem. Soc. 1987, 109, 2284-2298. 84. 85. 86. 87. 88. 89. 90. 91. 92. 93. 94. 95. 96. 97. 98. 99. 100. 101. 167 Bonnett,R.: Stephenson,G.F. Proc. Chem. Soc. 1964, 291. Salehi,A.: Fonda,H.N.: Oertling,A.W.; Babcock,G.T.: Chang,C.K. J. Labeled Compd. Radiopharm. 1988, XXV, 1333-1337. Eisner,U. J. Chem. Soc. 1957, 3461-3469. Omata,T.: Murata,N. Plant 8 Cell Physiol. 1983, 24, 1093-1100. Iriyama,K.: Ogura,N.: Takamiya,A. J. Biochem. 1974, 76, 901-904. Kenner,G.W.: McCombie,S.W.: Smith,K.M. J. Chem. Soc., Perkin Trans. 1973, 1, 2517-2523. Boucher,L.J.: Katz,J.J. J. Am. Chem. Soc. 1967, 89, 4703-4708. Mackinney,G. J. Biol. Chem. 1941, 140, 315-322. Ghanotakis,D.F.: Demetriou,D.M.: Yocum,C.F. Biochim. Biophys. Acta 1987, 891, 15-21. Kean,R.T. Ph.D. Thesis: Michigan State University: East Lansing, MI, 1987. Ogoshi,H.: Watanabe,E.: Yoshida,z.: Kincaid,J.: Nakamoto, K. Inorg. Chem. 1975, 14, 1344-1350. Ozaki,Y.: Kitagawa,T.: Ogoshi,H. Inorg. Chem. 1979, 18, 1772-1776. Ozaki,Y.: Iriyama,K.: Ogoshi,H.: Ochiai,T.: Kitagawa,T. J. Phys. Chem. 1986, 90, 6113-6118. Andersson,L.A.; Loehr,T.M.: Chang,C.K.: Mauk,A.G. J. Am. Chem. Soc. 1985, 107, 182-191. Andersson,L.A.: Sotiriou,C.: Chang,C.K.: Loehr,T.M. J. Oertling,W.A.: Salehi,A.: Chang,C.K.: Babcock,G.T. J.Phys. Chem. 1989, 93, 1311-1319. Kincaid,J.R.: Urban,M.W.: Watanabe,T.: Nakamoto,K. J. Phys. Chem. 1983, 87, 3096-3101. The un mode appears as a shoulder on the broad V; band in the resonance Raman spectra of CuOEP and CuOEP-d2 with Soret excitation at 406.7 nm. With 363.8 nm excitation, the spectra of CuOEP and CuOEP-d, obtained by Oertling gt g1. (99) show an increase in the relative intensity of 102. 103. 104. 105. 106. 107. 108. 109. 110. 111. 168 uu compared to u, and no frequency shift upon d, methine deuteration. Ozaki gt g1. (95) assigned the 1602 cm1 mode of CuOEC as dp and the 1584 cm1 mode as ap. These polarizations should be reversed, i. e. 1602 (ap) and 1584 (dp). Moustakali,I.: Tulinsky,A. J. Am. Chem. Soc. 1973, 95, 6811-6815. Collins,D.M.: Hoard,J.L. J. Am. Chem. Soc. 1970, 92, 3761-3771. Kitagawa,T.: Ogoshi,H.: Watanabe,E.: Yoshida,z. J. Phys. Chem. 1975, 79, 2629-2635. The K values reported by Ozaki gt 31. (66) for the FeOEP high frequency resonance Raman modes follow the same relative order as the K values based on Ni, Cu and ZnOEP (Table 3.5) but are all higher: um (K=495), u, (K=322), V19 (K=576), yn (K=288) and v3 (K=414). Using the same porphyrin core-sizes for the FeOEC complexes and their previous mode assignments for chlorins, the K values obtained for the chlorins were: "9m" (K=472), ":12" (K=404), "um" (K=332), "uu" (K=367) and '93" (K=347). On this basis, the K values for the chlorin Cbe modes "u," and "um" are higher than the chlorin C.CIII modes "1119" and "v3". Moreover, the frequencies predicted for CuOEC and ZnOEC based on the FeOEC core-size correlation parameters of Ozaki are too low (CuOEC: 1633, 1593,1579,1538 and 1500 cm 1: ZnOEC: 1611, 1574,1564, 1521 and 1484 cm 1). The six C,N stretching modes are derived from the 9,, tin, Mm: V222 and u“ modes of NiOEP. The most recent normal coordinate treatment of the Eu modes of NiOEP by Abe (56) assigned the P.E.D. of u,5 to C3; stretching and not CJJ stretching (55). The 1157 cm'1 mode of CuOEC and 1154 cm'1 mode of NiOEC are overlapped by the 1154 cm'1 band of CHzClz. The presence of these modes is confirmed by_ the resonance Raman spectra in KBr that show an 1155 cm1mode for CuOEC and an 1153 cm1 mode for NiOEC. Ogoshi,H.: Saito,Y.: Nakamoto,K. J. Chem. Phys. 1972, 57, 4194-4202. Stein,P.; Burke,J.M.: Spiro,T.G. J. Am. Chem. Soc. 1975, 97, 2304-2305. Sunder,S.: Bernstein,H. J. Raman Spectrosc. 1976, 5, 351-371. 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124. 125. 126. 127. 169 Susi,H.: Ard,J.S. Spectrochim. Acta 1977, 33A, 561-567. Gladkov,L.L.: Gradyushko,A.T.: Shulga,A.M.: Solovyov,K.N.: Starukhin,A.S. J. Mol. Struct. 1978, 47, 463-493. Warshel,A.; Lappicirella,A. J. Am. Chem. Soc. 103, 4664-4673. 1981, Gladkov,L.L.: Solovyov,K.N. Spectrochim. Acta 1986, 42A, 1-10. Lee,H.: Kitagawa,T.: Abe,M.: Pandey,R.K.: Leung,H.-K.: Smith,K.M. J. MOl. Struct. 1986, 146, 329-347. Lutz,M.: Kleo,J. aiochim. Biophys. Acta 1979, 546, 365-369. Salehi,A.: J. Oertling,W.A.: Am. Chem. Soc. Babcock,G.T.: 1986, 108, 5630-5631. Chang,C.K. Salehi,A.: Oertling,W.A.: Inorg. Chem. 1987, 26, Babcock,G.T.; 4296-4298. Chang,C.K. Oertling,W.A.: Salehi,A.: Chung,Y.C.: Leroi,G.E.: Chang,C.K.: Babcock,G.T. J. Phys. Chem. 1987, 91, 5887-5898. Salehi,A.: Oertling,W.A.: Fonda,H.N.: Babcock,G.T.; Chang,C.K. Photochem. Photobiol. 1988, 48, 525-530. Chang,C.K. Biochemistry 1980, 19, 1971-1976. Spangler,D.: Maggiora,G.M.: Shipman,L.L.: Christoffersen,R.E. J. Am. Chem. Soc. 1977, 99, 7470-7477. Forman,A.: Renner,M.W.: Fujita,E.: Barkigia,K.M.: Evans,M.C.W.: Smith,K.M.: Fajer,J. Israel J. Chem. 1989, 29, 57-64. . x-ray crystal structures of the metalloporphyrin cation radicals Mg(C10f)TPPt (126) and Zn(c1o{)TPP“ (127) show non-planarity of the porphyrin macrocycle. This distortion is attributed to non-bonding interactions of the coordinated perchlorate ion and phenyl groups (120,128-130). Barkigia,K.M.: Spaulding,L.D.: Fajer,J. Inorg. Chem. 1983, 22, 349-351. Spaulding,L.D.: J. Am. Chem. Eller,P.G.: Soc. 1974, 96, Bertrand,J.A.; 982-987. Felton,R.H. 128. 129. 130. 131. 132. 133. 134. 135. 136. 137. 138. 139. 140. 141. 142. 143. 144. 170 Scholz,W.F.: Reed,C.A.: Lee,Y.J.: Scheidt,W.R.: Lang,G. J. Am. Chem. SOC. 1982, 104, 6791-6793. Reed,C.A.: Mashiko,T.: Bentley,S.P.: Kestner,M.E: Scheidt,W.R.: Spartalian,K.: Lang,G. J. Am. Chem. Soc. 1979, 101, 2948-2958. Masuda,H.; Taga,T.: Osaki,K.: Sugimoto,H.: Yoshida,Z.-I.: Ogoshi,H. Inorg. Chem. 1980, 19, 950-955. Strekas,T.C.: Spiro,T.G. Biochim. Biophys. Acta 1972, 263, 830-833. Spiro,T.G.: Strekas,T.C. Proc. Natl. Acad. Sci. U.S.A. 1972, 69, 2622-2626. Brunner,H.: Mayer,A.: Sussner,H. J. Mol. Biol. 1972, 70, 153-156. Fonda,H.N.; Babcock,G.T. in Progress in Photosynthesis Research: Biggins,J.,Ed.; Martinus Nijhoff: Dordrecht, The Netherlands, 1987: Vol. I, pp 449-452. Fujiwara,M.: Tasumi,M. J. Phys. Chem. 1986, 90, 5646-5650. Heald,R.L.: Callahan,P.M.: Cotton,T.M. J. Phys. Chem. 1988, 92, 4820-4824. Lutz,M.: Kleo,J. Biochem. Biophys. Res. Comm. 1976, 69, 711-717. Cotton,T.M.: Van Duyne,R.P. Biochem. Biophys. Res. Comm. 1978, 82, 424-433. Cotton,T.M.: Parks,K.D.: Van Duyne,R.P. J. Am. Chem. Soc. 1980, 102, 6399-6407. Cotton,T.M.: Van Duyne,R.P. J. Am. Chem. Soc. 1981, 103, 6020-6026. Wagner,W.-D.: Waidelich,W. Appl. Spectrosc. 1986, 40, 191-195. Fujiwara,M.: Tasumi,M. J. Phys. Chem. 1986, 90, 250-255. Shipman,L.L.: Cotton,T.M.: Norris,J.R.: Katz,J.J. J. Am. Chem. SOC. 1976, 98, 8222-8230. Lutz,M.: Hoff,A.J.: Brehamet,L. Biochim. Biophys. Acta 1982, 679, 331-341. 145. 146. 147. 148. 149. 150. 151. 152. 153. 154. 155. 156. 157. 158. 159. 160. 161. 162. 163. 171 Robert,B.: Vermeglio,A.: Lutz,M. Biochim. Biophys. Acta 1984, 766, 259-262. Robert,B.: Lutz,M. Biochim. Biophys. Acta 1985, 807, 10-23. Agalidis,I.: Lutz,M.: Reiss-Husson,F. Biochim. Biophys. Acta 1984, 766, 188-197. Robert,B.; Lutz,M. Biochemistry 1986, 25, 2303-2309. Kuhlbrandt,W. Nature (London) 1984, 307, 478-480. Li,J. Proc. Natl. Acad. Sci. USA 1985, 82, 386-390. Babcock,G.T. in Photosynthesis: Amesz,J., Ed.: Elseiver: Amsterdam, 1987: pp 125-158. Berthold,D.A.: Babcock,G.T.: Yocum,C.F. FEBS Lett. 1981, 134, 231-234. Ghanotakis,D.F.: Babcock,G.T.: Yocum,C.F. Biochim. Biophys. Acta 1984, 765, 388-398. Mathis,P.; Rutherford,A.W. in Photosynthesis: Amesz,J., Ed.: Elseiver: Amsterdam, 1987: pp 63-96. Michel,H. J. Mol. Biol. 1982, 158, 567-572. Deisenhofer,J.: Epp,0.: Miki.K.: Huber,R.: Michel,H. J. Mol. Biol. 1984, 180, 385-398. Deisenhofer,J.: Epp,0.: Miki,K.: Huber,R.: Michel,H. Nature (London) 1985, 318, 618-624. Michel,H.: Deisenhofer,J. in Progress in Photosynthesis Research: Biggins,J., Ed.: Martinus Nijhoff: Dordrecht, The Netherlands, 1987: Vol. I, pp 353-362. Allen,J.P.: Feher,G. Proc. Natl. Acad. Sci. USA 1984, 81, 4795-4799. Chang,C.-H.: Schiffer,M.: Tiede,D.: Smith,U.: Norris,J. J. Mol. Biol. 1985, 186, 201-203. Allen,J.P.: Feher,G.: Yeates,T.O.: Komiya,H.: Rees,D.C. Proc. Natl. Acad. Sci. USA 1987, 84, 6162-6166. Wright,K.A.: Boxer,S.G. Biochemistry 1981, 20, 7546-7556. Eccles,J.: Honig,B. Proc. Natl. Acad. Sci. USA 1983, 80, 4959-4962. 164. 165. 166. 167. 168. 169. 170. 171. 172. 173. 174. 175. 176. 172 Katz,J.J.: Shipman,L.L.: Cotton,T.M.: Janson,T.R. in The Porphyrins: Dolphin,D., Ed.: Academic Press: New York, 1978: Vol. V, pp 401-458. Cotton,T.M.: Trifunac,A.D.: Ballschmiter,K.: Katz,J.J. Biochim. Biophys. Acta 1974, 368, 181-198. Ward,B.: Callahan,P.M.: Young,R.: Babcock,G.T.: Chang,C.K. J. Am. Chem. Soc. 1983, 105, 634-636. Maggiora,L.L.: Maggiora,G.M. Photochem. Photobiol. 1984, 39, 847-849. Ward,B.: Chang,C.K., Young,R. J. Am. Chem. Soc. 1984, 5 106, 3943-3950. Hanson,L.K.: Chang,C.K.: Ward,B.: Callahan,P.M.: Babcock,G.T.: Head,J.D. J. Am. Chem. Soc. 1984, 106, 3950-3958. I French,C.S.: Brown,J.S.: Lawrence,M.C. Plant Physiol. 1972, 49, 421-429. Brown,J.S. Ann. Rev. Plant Physiol. 1972, 23, 73-86. Babcock,G.T.: Callahan,P.M.: Ondrias,M.R.: Salmeen,I. Biochemistry 1981, 20, 959-966. Lutz claimed to observe both 5 and 6-coordinate chlorophyll b in LHC based on the relative intensities of two low frequency bands at 312 and 304 cm1 (75). The 312 cm1 component was attributed to 5-coordinate chlorophyll b and the 304 cm1 component was attributed to 6-coordination. Lutz has proposed (76) for chlorophyll b that a single band at 300- 310 cm 1with a half-bandwidth of 25 cm1 or less at 30 K and 27 cm1 or less at 300 K indicates 6-coordination. A S-coordinate Mg is indicated by the presence of an additional band.at 310- 320 cm 1with an increase of half-bandwidth to 28- -35 cm1. However, these criteria are not reliable predictors of the Mg coordination state. For example, chlorophyll b in acetone solution at room temperature is predicted by Lutz to have 6-coordination (76) but the frequencies of the core-size sensitive modes indicate S-coordination (142). Matthews,B.W.: Fenna,R.E.: Bolognesi,M.C.: Schmid,M.F.: Olson,J.M. J. Mol. Biol. 1979, 131, 259-285. Koyama,Y.: Umemoto,Y.; Akamatsu,A.; Uehara,K.: Tanaka,M. J. Mol. Struct. 1986, 146, 273-287. Babcock,G.T.: Callahan,P.M. Biochemistry 1983, 22, 2314-2319. 177. 178. 179. 180. 181. 182. 183. 184. 185. 186. 187. 188. 189. 190. 191. 192. 193. 194. 195. 173 Willems,D.L.: Bocian,D.F. J. Am. Chem. Soc. 1984, 106, 880-890. Willems,D.L.: Bocian,D.F. J. Phys. Chem. 1985, 89, 234-239. Choi,S.: Spiro,T.G.: Langry,K.C.: Smith,K.M. J. Am. Chem. Soc. 1982, 104, 4337-4344. Meyer,E.F.Jr. Acta Cryst. 1972, 828, 2162-2167. Shelnutt,J.A. J. Chem. Phys. 1981, 74, 6644-6657. Houssier,C.: Sauer,K. Biochim. Biophys. Acta 1969, 172, 476-491. Houssier,C.: Sauer,K. Biochim. Biophys. Acta 1969, 172, 492-502. Cox,M.T.: Howarth,T.T.: Jackson,A.H.: Chem. Soc. Perkin I 1974, 512-516. Kenner,G.W. J. Cox,M.T.: Jackson,A.H.; Kenner,G.W.: McCombie,S.W.: Smith,K.M. J. Chem. Soc. Perkin I 1974, 516-527. Kenner,G.W.: McCombie,S.W.: Smith,K.M. J. Perkin I 1974, 527-530. Chem. Soc. Centeno,J. Ph.D. Thesis: Michigan State University: East Lansing, MI, 1987. Pennington,F.C.: Strain,H.H.: Svec,W.A.: Katz,J.J. J. Am. Chem. Soc. 1964, 86, 1418-1426. Pennington,F.C.: Boyd,S.D.: Horton,H.: Taylor,S.W.: Wolf,D.G.: Katz,J.J.: Strain,H.H. J. Am. Chem. Soc. 1967, 89, 3871-3875. Pennington,F.C.; Boettcher,N.B.: Katz,J.J. Chemistry 1974, 3, 204-212. Bioorganic Boldt, N.J.: Bocian,D.F. J. Phys. Chem. 1988, 92, 581-586. Donohoe,R.J.: Atamian,M.: Bocian,D.F. J. Phys. Chem. 1989, 93, 2244-2252. Donohoe,R.J.: Frank,H.A.: Bocian,D.F. Photochem. Photobiol. 1988, 48, 531-537. Atamian,M.: Donohoe,R.J.: Lindsey,J.S.: Bocian,D.F. J. Phys. Chem. 1989, 93, 2236-2243. "‘IVIIIIIEIJIIIII‘IIIIIIEF