r". . '7‘ .L‘ Y-. v 3: < : c 4‘: A ‘t‘a'a'auk‘ai . 4- " u) M?W 4".mult V ‘ u '3 "6 . __ ‘ 7‘. 1,: E: _ a v "V Hund'nmi .— H .- '7 -‘ .23.: ’0» .‘A. , 6:“: “)3"? '.-.Yf’{‘-' "‘Wmfifi ~‘ at l1: 7?; ‘ *3. 3M ‘1’? u. k #1.. , \ 1‘ ' .1 t '33”: . $11,,“ a." “uh I v‘fi‘r‘v‘pfi‘ ‘ “5:“- - 33g 4 V I" v x > b . . ‘ l A _ o 4:12;! '- ' ‘ . i ‘ " ,.I'-' 1%":- C. , A' . M ".4 k h w ‘ V93 .. _ ‘ . .- ’_ j)"; m g" , - . H x .' 3“, - O- ':: \ ‘ ~ - Z ’ I ; I . , '4‘ ‘ H. — -’ ' '5 .‘ , _ ' "H" 1‘ V - ’ f ‘ ‘ '-‘ . , ‘L ‘ I‘ - - . _ t - -‘ h. ' 55$" . ‘Ih [unhmai ‘6 - ‘ ll. ‘ L‘ ‘7 ' “‘- nfiff.‘ u. “as.“ .3 LIBRARY Michigan State University This is to certify that the dissertation entitled Vibrational, Electronic and Structural Properties of High Valent Catalytic Intermediates of Horseradish Peroxidase presented by W. Anthony Oertling has been accepted towards fulfillment of the requirements for [40 degreem CAQM/s/‘v/ M/ 7, Mm Major professor "(II-$.- T— .:_._ A - .- In , I “Ln—l 0.1277, bViSSI,J RETURNING MATERIALS: Place in book drop to LJBRARJES remove this checkout from u. your record. FINES will be charged if book is returned after the date stamped below. VIBRATIONAL, ELECTRONIC AND STRUCTURAL PROPERTIES OF HIGH VALENT CATALYTIC INTERMEDIATES OF HORSERADISH PEROXIDASE By W. Anthony Oertling A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1987 ABSTRACT VIBRATIONAL, ELECTRONIC AND STRUCTURAL PROPERTIES OF HIGH VALENT CATALYTIC INTERMEDIATES OP HORSERADISH PEROXIDASE By W. Anthony Oertling The active site of the two-electron oxidized compound I intermediate of horseradish peroxidase (HRP-I) is usually described as an oxoferryl porphyrin n cation radical. In order to assess separately the effects of each oxidation equivalent on the structural, electronic and vibrational properties of this heme complex we present a resonance Raman (RR) characterization of a series of model compounds. From comparison of the FeO stretching frequency, v(Fe0), of synthetic oxoferryl hemes and oxoferryl heme proteins the predominant determinant of the v(FeO) is identified as a trans-ligand effect, with a lesser effect coming from H-bonding of the oxo ligand. Through a separate RR study of metalloporphyrin Tr cation radicals (MP+°) the vibrations of the oxidized porphyrin macrocycle are analyzed and compared to those of the parent metalloporphyrin (MP). The frequency of the vibrational modes in the 1450-1700 cm"1 range of the MP“ are found to be an inverse linear function of porphyrin center-to-pyrrole nitrogen distance, similar to that which describes the vibrational frequencies of the parent compound. Thus, as is the case for the MP, the high frequency vibrations of the MP4” reflect porphyrin core geometry. W. Anthony Oertling In the enzymatic cycle, HRP-I is reduced by substrate to the more stable HRP—ll. These two intermediates are thought to contain similar oxoferryl structures. Based on RR measurements of the v(FeO) by others, the oxo ligand of HRP-II is thought to be H-bonded at pH 7 but not as pH 11. We present RR scattering from a flowing HRP-I sample prepared by rapid mixing. The v(FeO) is very similar to that of HRP-II at high pH. This is taken to suggest that, unlike HRP-II, there is no H-bond to the oxo ligand of HRP-I at pH 7. A mechanism, consistent with these RR observations and past kinetic studies, is proposed for the reduction of HRP-I by p—cresol. The frequencies we measure for the in—plane vibrations of HRP-I, however, are not characteristic of a MP4". Possible explanations for this result are discussed. ii ACKNOWLEDGMENTS For their various contributions to this work, my thanks is given to: the members of my family for support and encouragement; G.T. Babcock, G.E. Leroi and C.K. Chang for critical reading and guidance; Jerry Babcock, Pat Callahan and John McMahon for instruction in the measurement of Raman scattering; the entire supporting staff of the Chemistry Department, in particular Ron Haas and Deak Watters, for technical services; Dwight Lillie, Bob Kean, Harold Fonda and Ralph Thiim for computer programming. Special thanks is extended to my "partner", Asaad Salehi, for his sincerity, enthusiasm and perserverance. Thanks also goes to Carol Smith, Sharon Corner, Margy Lynch and Bill Draper for their efforts in the preparation of this manuscript. Kris Babcock and Mari’a Centeno are acknowledged for their fine cooking and Anheuser-Busch and Miller Breweries for their fine products, without which this work would have been much more difficult. iii TABLE OF CONTENTS Page LIST OF TABLES ...................................................................................... vi LIST OF FIGURES .................................................................................... viii Chapter 1 - GENERAL INTRODUCTION . .................................................... 1 Significance of the Oxoferryl and Porphyrin n Cation , Radical Structures ................... ......... ................................................. 3 Aim and Strategy of the Thesis ........ . ................................................... 5 Metalloporphyrin 1r Cation Radicals and Compound I ............................. 5 References ....................................................................................... 14 CHAPTER 2 - CHARACTERIZATION OF SIX-COORDINATE OXOFERRYL PROTOHEME BY RESONANCE RAMAN AND OPTICAL ABSORPTION SPECTROSCOPY Summary ............................ ...................................................... 16 Acknowledgments ............................................................................... 21 References 6C Notes .................................... . ...................................... 22 CHAPTER 3 - RESONANCE RAMAN SPECTROSCOPIC DETECTION OF DEMETALLATION OF METALLOPORPHYRIN 1r CATION RADICALS Summary ............................... . ............................................................ 24 Acknowledgments .............................................................. . ................ 31 References 6C Notes ............................................................................ 32 CHAPTER 4 VIBRATIONAL, ELECTRONIC AND STRUCTURAL PROPERTIES OF COBALT, COPPER AND ZINC OCTAETHYLPORPHYRIN n CATION RADICALS Summary .................................................... .... ................................... 34 Introduction ...................................................................................... 35 Experimental ........................................... . ................ . ....................... 38 Results Effects of oxidation on the high frequency porphyrin core vibrations ............................................................. 39 One-electron oxidation of CuOEP and ZnOEP ................................ 40 iv P88! Porphyrin diacid impurities . ........... ..... . ...... ............. ..... 46 One- and two-electron oxidation of CoOEP ......... ............ ...... 52 Neal. uv RR Spectra 00000000000 oooooooooooooo o oooooooooooooooooooooooooooooooooooooo 58 Visible RR spectra ..... o ................. o ooooooooooooooooooooooooooooooooooooooooo 64 Low temperature effects ............................................................. 70 Discussion ............. . ..................................... ................ 74 Structural correlations ....................... . ........................................ 75 Porphyrin core geometry ............................................................. 80 Ring buckling effects ................................................................... 84 Cobaltic octaethylporphyrin 1r cation radicals ............................... 85 Effects of one-electron, metal centered oxidation and axial ligation ..................... . ......................... . ........................ 86 Comoar‘ 20104 .............. ............. . ..................... 89 001110131)+ 23: ..... ....... .................................... 90 Conclusions .............. . ......................................................................... 92 Acknowledgments ................................................... 4. ........................... 9 4 References and Notes ............................................... .. ........................ 95 CHAPTER 5- DETAILS OF PEROXIDASE CATALYSIS SUGGESTED BY RESONANCE RAMAN MEASUREMENTS OF IRON-OXYGEN STRETCHING FREQUENCIES OF HORSERADISI-I PEROXIDASE INTERMEDIATES Introduction ...................................................................................... 1 0 1 Experimental Materials ........ . ....... . ................... . ............................................... 104 HRP-I ................... ........... . ............................ .............. 104 HRP-II ....................................................................................... 106 Model compounds ........................................................................ 106 Instrumentation ............................................. . ......................... 106 Results High frequency RR scattering from HRP intermediates .................. 110 Model compounds ......................... . .............................................. 1 1 3 Frequency shifts .. ...... ..................................... . ...................... 115 Low frequency RR scattering from HRP intermediates .................. 116 Excitation profiles ...... . ................ . ............... . ....... . ...................... 120 Discussion Prediction of HRP-I vibrations from structural correlations ......... ............................................................ 130 HRP-II ...................................................................................... 131 Photochemistry ............................ ........ ..... ............. .. ..... 131 Spin delocalization . ....... .. ........ . ........ ...... ...... 132 HRP-1" ......... ........... ........ ........... ....... . ..... 133 Kinetics and mechanism of HRP-I reduction .................................. 135 Conclusions ....................... ................................... .............. 136 Future Work ............... . ............. . ............ . ................. ........ . ..... 137 Acknowledgments ............... ......... ......... . ......... ........................ 138 References ....................................................................................... 139 2-] LIST OF TABLES Comparison of v (FeO) for Various Oxoferryl Species ..................... Resonance Raman Frequencies (cm‘l), Depolarization Ratios and Optical Absorption Maxima (nm) for Parent MOEP Compounds and their Corresponding n Cation Radicals, MOEP+ CIO4‘ .............................................................. Core Size Parameters from Crystal Structures of Metallo- porphyrin 17 Cation Radicals and Related Compounds ..................... Resonance Raman FrequenciesI andO tical Absorption Maxima for Axially Ligated CuII , N 1 and Co11 IOEP Complexes ...... ............ . ............. ........... .. ....................... Comparison of Resonance Raman Frequencies (cm 1) and of Depolarization Ratios for Cobaltic OEP" Complexes ................. vi Page 2 0 83 88 91 1-1 1-2 1-4 LIST OF FIGURES (a) Ferriprotoporphyrin IX. (b) Metallooctaethylporphyrin. Carbons a (or a) are bonded to the nitrogen atoms. Car- bons b (or B ) are bonded to the peripheral substituents. The methine, CH, positions are meso. The metal center, M, may be further ligated in the axial positions normal to the macrocycle plane ...................................................................... Optical absorption spectra of native HRP (N), HRP-I (I) and HRP-II (11). Reference 31. .................................................. The atomic orbital (AO) structure of two HOMOs of porphine. The A0 coefficients are proportional to the size of the circles; solid lines indicate positive values, dashed lines negative. The view is from the positive 2 axis. The straight dashed lines indicate the nodes of the alu ( 1!) orbital. Reference 14. .................. . ............ ..... . ............. .............. Optical absorption spectra of (a) cobaltic octaethyl- porphyrin 11 cation radicals, (b) compound I enzyme transients. 2A2u type spectra (—). 2A1u type spectra (--). Reference 16. .......................................................................... Optical absorption spectra of compound I intermediates (2Alu type). (a) native horse erythrocyte catalase (CAT) and CAT-I. Reference 26. .......... ........................................................... (b) native chloroperoxidase (CPO) and CPO-I. Reference 24. ...... Optical absorption Spectra of compound I intermediates. (a) native ferric bromoperoxidase (BPO) and BPO—I. Reference 27. ..... . ............. ........ . ................................. (b) native lignin peroxidase (MP) and LiP-I. Reference 29. ......... (c) Rapid scan spectrophotometric measurements for the reactions of intestinal peroxidase (IPO) and lactoperoxidase (LPO) with hydrogen peroxide at pH 7.1. Dashed lines repre- sent native enzyme. the numbers show time in ms from the stop flow to the end of the wavelength scan. Thus, early times represent compound I spectra, and late times compound 11. At intermediate times, mixtures were obtained. Reference 28. ....... vii Page 6 10 10 11 ll ..... 12 2—1 3-1 3-2 4-2 Page Absorption spectra of native cytochrome c peroxidase (CC?) and CcP-I. Reference 30. CcP-I contains an amino acid radical rather than a porphyrin radical. ......... .. ............................. 13 Resonance Raman s ectra of 18O=FeIV(Im)PPDME (upper trace), and 16O=FeI (Im)PPDME (lower trace). Vertical line denotes v (FeO); * denotes solvent peak. Spectra were obtained in toluene-d3 at -130 C with 15 mW (406.7 nm) incident on the sample. Inset: optical absorption spectrum of O=FeIV(Im)PPDME in toluene at -90 C. The shoulder at 619 nm is due primarily to u—oxo dimer contamination. ................... 18 Resonance Raman spectra excited at 406.7 nm. (a) CoIIOEP+‘ClO4‘ prepared with AgClO4; (b) CoIHOEP+'ZClO4‘ prepared with Fe(ClO4)3; (c) CoInOEP+'2ClO4", prepared with Fe(ClO4)3 and a trace amount of HCIO4; (d) H4OEP2+2CIO4", prepared from HZOEP + HCIO4. All samples were dissolved in CH2C12. A spinning quartz cell and 'v 20 mW laser power were used. Solvent bands are labelled with an *. ......................................... ................... 27 RR spectra of ZnOEP+'ClO4. (a) xex = 390 nm; (b) - (d) show a time course at A ex = 406.7 nm for 'v 1 ml of sample in a quartz spinning cell with a laser power of 16 mW. The time values indicate total irradiation time at the end of the scan. ........... . ............................................................................ 30 Electronic absorption spectra of (a) CuOEP+'ClO4" and (b) ZnOEP+’ClO4' (0.2 mM) in dry CHZCIz. Extinction coefficients were taken from references 29 and 4, respectively. ..... 42 RR spectra of Cu and ZnOEP and their corresponding cation radicals. (a) CuOEP; (b) CuOEP+'ClO4'; (c) ZnOEP; (d) ZnOEP+'ClO4‘, excitation at 390 nm (1.5 mJ/pulse, [ZnOEP+'l 20.2 mM) reflects vibrations of the monomer. Cchlz bands are marked with an *. CW laser power 20-35 mW. at 363.8 nm for (ch). ........................................................ 45 Resonance Raman spectra excited at 406.7 nm. (a) CoIIOEP+'ClO4"; (b) ComOEP+'2ClO4" containing small amounts of porphyrin diacid impurity; (c) H4OEP2+ZCIO4‘; (d) a mixture of H30EP+CIO4’ and H4OEP2+2CIO4‘ resulting from treatment of OEPHZ with AgClO4; viii Page (e) H40EP2+2Br-. RR bands of the solvent, dry CH2C12, are marked with an". Vertical dashed lines mark the vibrations of the diacid, H4OEP2+ZBr‘. Laser power, 20 mW. .................................. 48 Electronic absorption spectra of porphyrin free base acids. (a) H4OEP2+2CIO4"; (b) a mixture of H30EP+CIO4' and H4OEP2+2CIO4" corresponding to Fig. 4-3d. Solvent, dry CHZCIZ. ...... . ............ ...... . ................... 51 Electronic absorption spectra of CoOEP and its oxidation products. (a) CoOEPH; (b) CoIIOEP+'ClO4’(—-); (c) ComOEP”'2ClO4‘("°). Solvent, dry CHZCIZ. ....................................................... , .......... 5 4 Electronic absorption spectra of oxidation products of CoOEP. (a) Com(MeOH)20EPBr'(-); (b) ComOEPBr' (——); (c) ComOEP+'ZBr (m), the small feature at 401 nm is due to 1% H4OEP2+ZBF contamination as discussed in the text. Solvent, dry CH2C12, except for (a) which contains W 5% methanOl (M9011). 00000000000000 oooooeoeeo 0000000000000 e ooooooooo oooooeooeo oooooooooooooo 57 RR spectra excited at 363.8 nm ( ~35 mW) of CoOEP and its oxidation products. (a) CoOEP; (b) Com(MeOH)20EPClO4‘; (c) CoHOEP+'CIO4‘; (d) CoIIIOEP+’2ClO4‘; ................................................................ 60 (e) Com(MeOH)20EPBr‘; (f) ComOEPBr'; (g) CoIIIOEP+'2Br-. Solvent, dry CH2C12 except (b) and (e) which contain 'h 596 methanol. Solvent bands are marked with an *. ............................ 63 RR spectra excited at 514.5 nm (100 mW). (a) CoOEP; (b) Com(MeOH)20EPBr"; (c) ComOEPBr". CH2012 bands marked *. ............................................................. 66 ix 4-9 4-10 4-1 1 4-12 5-1 5-2 5-4 Page RR spectra of ComOEP+'2Br'. xex = (a) 620 nm, 300 mW; (b) 647.1 nm, 250 mW; (c)—(d) 676.4 nm, 100 mW. CHZCIZ bands marked *. Spike (**) in (b) is Rayleigh scattering at 676.4 nm. .......................................................... . ....................... 69 Electronic absorption 3 ectra showing the conversion of CoHOEP+'ClO4‘ to Co (H20)20EPCIO4‘ as the temperature is lowered. Solvent, "dry" CH2C12; path length 2 2mm (EPR tube). ..................................................... . ......... ....... . ........... 73 Porphyrin core vibrational mode frequencies (for Raman allowed bands, 1450-17 00 cm‘l) v_s_. porphyrin core size for the indicated MOEP complexes and their corresponding cation radicals, MOEP+’CIO4'. Mode assignments are accord- ing to reference 33a. Open symbols correspond to parent MOEP frequencies; filled symbols correspond to MOEP+°CIO4' frequencies. The regression analysis does not include the Ni (ng) points. Ct-N distances are from reference 45. ....................... 78 Electronic transition energies vs. porphyrin core size for the indicated MOEP and MOEP‘F‘ClOf complexes. Open and filled symbols correspond to absorption band maxima of MOEP and MOEP“'CIO4‘, respectively. .................................... 81 End view (a), longitudal section (b) and cross-section (c) of the eight-jet, tangential mixers. Ref. 48. ..................................... 107 Instrumental configuration used for pulsed resonance Raman measurements of flowing HRP-I samples prepared by rapid mixing. Laser irradiation at 390 nm (5 ns pulses, 1 mJ/pulse, 10 Hz) was provided by pumping Exciton LD390 laser dye in the Quanta Ray Pulsed Dye Laser (PDL-l) with the third harmonic (355 nm) from the Nd:Yag DCR-lA laser. Raman scattering from the sample (at position X) was collected with a Spex 1459 Illuminator, dispersed in a triple mono— chromator, and detected and analyzed by using the EGétG PAR 1420 diode array detector and associated OMA II electronics. RR scattering excited at 390-435 nm from HRP-II samples in a cooled (2-5 C) cylindrical quartz spin- ning cell was also measured using this instrumentation. The 10 Hz rep rate was employed on all occasions. ............................... 109 High frequency RR spectra of HRP samples at 2-10 C, 1 mJ/pulse. Sample conditions and total accumulation times: (a) spinning cell, 20 min; (b) flowing in the capillary of the rapid mixer at 0.65-0.85 ml/min, 40 min; (c) spinning cell, 22 min. ......... . ............................................... 112 RR spectra of model compounds in CHZCIZ. Conditions: 40-50 mW incident on sample in spinning cell at 25 C. Accumulation times: (a) 5.3 min; (b) 10 min. .............................. 114 X 5-6 5-7 5-8 5-9 5-1 0 Page Low frequency RR spectra of HRP samples obtained under 390 nm excitation. Total accumulation times: (a) 85 min; (b) 160 sample: 144 min, 180 sample: 78 min; (c) 52 min; ((1) 160 sample: 48 min, 180 sample: 48 min. The feature at 981 cm‘1 in (b) is due to sulfate ion in the H21802 preparation. ................................................................. 1 1 8 High frequency RR scattering from HRP-II, pH 10.8. Accumulation time for each spectrum was 10 min using 1 mJ/pulse. Excitation wavelengths were as indicated. The 1049 cm‘1 peak is from 0.2 M NO3' used as an internal standard to measure relative intensities. The 984 cm"1 feature is from 804‘. ................................................... 122 High frequency RR scattering from HRP-I, pH 7. (a) 1.9 mJ/pulse, 4 min; (b) 15 mW, 0.5 s/cm'l; sample flow rate, 0.22 m1/min; similar results were obtained flowing at 0.7 ml/min; (c) 1 mJ/pulse, 40 min. .............................. 125 Low frequency RR scattering from HRP-II, pH 10.8. Accumulation time was 10 min for (a)-(d), 20 min for (er), using 1 mJ pulses. The 983 cm“1 peak is from 0.2 M SO ‘ used as an internal intensity standard. The 1049 cm‘ feature is from NO3‘. ............................................... 127 Excitation profiles of v7 = 680 cm‘l, v (FeO) = 787 cm'1 and v 4 = 1379 cm‘l. 129 Schematic diagram of the active site of HRP catalytic intermediates, showing HRP-I, the postulated photo- product HRP-1*, the reaction of HRP-I with a phenolic substrate, and HRP-II. ............................................................... 134 xi CHAPTER I GENERAL INTRODUCTION Aerobic organisms reduce molecular oxygen (02) to water (H20). Partial reduction of dioxygen can result in the formation of hydrogen peroxide (H202) and superoxide (02"). These compounds are potentially dangerous to the organism and are removed from the cell by enzyme catalases, peroxidases, and superoxide dismutases. While catalase enzymes convert hydrogen peroxide directly to water and dioxygen, peroxidases utilize hydrogen peroxide to oxidize various organic and inorganic compounds.1 The active site of native peroxidases and catalyses most often contain a five-coordinate ferriprotoporphyrin IX prosthetic group as shown in Figure 1-1a. The proximal ligand to the iron may vary and is typically a histidine nitrogen (e.g. horseradish peroxidase, HRP),2‘4 cysteine sulfur (e.g. chloroperoxidase, CPO),4’5 or tyrosine oxygen (e.g. bovine catalase, CAT).6 Horseradish peroxidase isolated from plants is the best characterized and most readily available of the peroxidase enzymes. The general catalytic mechanism of plant peroxidases is depicted by the following scheme proposed originally for HRP:7 CH2 a u cu CH3 H3C ca=ca2 asc CH3 C'Hz EH2 CH2 CH2 l I COOH COOH b c'H3 C|H3 CH2 C‘Hz 3 2 HBC‘HZC CHZ-CHS 9:2 9‘2 CH 3 CH3 Figure 1-1 . (a) Ferriprotoporphyrin IX. (b) Metallooctaethylporphyrin. Carbons a (or a ) are bonded to the nitrogen atoms. Car- bons b (or 8 ) are bonded to the peripheral substituents. The methine, CH, positions are meso. The metal center, M, may be further ligated in the axial positions normal to the macrocycle plane. Native (Fem) + H20 + compoundI(Fe"V") + H20 compoundI(Fe"V") + AH + compound II(FeIV) + A' compound II(FeIV) + AH + native (Fem) + A' + H20 2A' + products In the first step of this sequence, the heme reacts with hydrogen peroxide to form compound I (green), which is two oxidation equivalents above the native ferric state. Compound I rapidly oxidizes the substrate in the second step, forming the more stable, second intermediate, compound 11 (red). In the final step of the reaction, compound II relaxes to regenerate the native'ferric state of the enzyme (brown). The structural, electronic and magnetic properties of peroxidase compounds I and II have been investigated for a number of years.8 Most studies have used HRP owing to the ease of isolation and purification9 as well as the stability of its compound I (HRP-I) relative to other peroxidases.1 The absorption spectra of native HRP, HRP-I and HRP-II appear in Figure 1-2. HRP-I is currently described as a spin-coupled oxoferryl porphyrin 1r cation radical complex.4 In this structure one electron has been removed from the ferric center to yield FeIV and the other electron has been removed from the porphyrin ring to yield a delocalized porphyrin 1r cation radical. The FeIV center is six-coordinate, being bound on the proximal side by histidine nitrogen and on the distal side by an oxime group, yielding a O=FeIV-N(His) configuration axial to the heme ring, reducing the 1r cation radical but leaving the oxoferryl nitrogen ligation unchanged.2‘4 Significance of the Oxoferryl and Porphrin 1r Cation Radical Structures The oxoferryl n cation radical structure is considered common to catalytic Absorbonce ' 7 0.08 0.06 0.04 0.02 0’1 l l l l" J l 4 l " O 250 300 350 400 450 500 550 6CD 650 Wavelength (nm) Figure 1-2 . Optical absorption spectra of native HRP (N), HRP-I (I) and HRP-II (II). Reference 31. intermediates of a range of heme peroxidase and catalase enzymesl‘4 and possibly to cytochrome P450.10 The 1r cation radical structure has also been implicated in the photosynthetic process of both higher plants and bacteria11 and in the catalytic cycles of nitrite and sulfite reductases.12 The oxoferryl structure has also been proposed for an intermediate in the reaction of cytochrome oxidase with dioxygen.13 Thus the significance of both the oxoferryl and the porphyrin ncation radical structures to the redox and electron transport functions of metalloporphyrins in nature is clear. Aim and Strategy of the Thesis The aim of this work is to characterize the vibrational, electronic and structural properties of the first intermediate of the catalytic cycle of HRP, HRP-I, by using resonance Raman spectroscopy in conjunction with rapid mixing techniques applied to the enzyme in vitro. The difficulties associated with this measurement, the means adopted to circumvent them, and the results obtained by the technique are discussed in Chapter 5. In addition to the work with the enzyme, we utilize systematic spectroscopic studies of various model compounds to explore separately the properties of the oxoferryl-nitrogen linkage and the porphyrin n cation radical. This work is described in Chapters 2-4. Metalloporphyrin I Cation Radicals and Compound I Chapters 2 and 3 describe a spectroscopic study of highly symmetric (D4h) metallooctaethylporphyrin n cation radicals (see Figure 1-1b). Oxidation of the porphyrin ring occurs by the removal of an electron from either of the nearly degenerate highest filled nmolecular orbitals (HOMOs). In D4h symmetry these oribtals are designated a1u(1r) and a2u(1r). Their structure is shown in Figure 1--3.14 As representative examples for discussion we choose two cobaltic HOMO'I "32“ (17)" Figure 1—3 . The atomic orbital (A0) structure of two HOMOs of porphine. The A0 coefficients are proportional to the size of the circles; solid lines indicate positive values, dashed lines negative. The view is from the positive 2 axis. The straight dashed lines indicate the nodes of the alu ( 1!) orbital. Reference 14 . octaethylporphyrin 1r cation radicals, CoIIIOEP+'ZBr‘ and CoIIIOEP+'2CIO4‘. These two compounds are thought to occupy the 2A1u and 2A2u electronic ground states, respectively. Much emphasis has been given to characterization of the spectroscopy of these (and other) metalloporphyrin 1r cation radicals (MP+') in terms of these two electronic configurations. While this approach has been successful in the first approximation, especially for interpreting the EPR and ENDOR spectra of these compounds,15116 some controversy has emerged based on NMR results.17 The original extension of the model compound studies to the enzyme intermediates is presented in Figure 1-4. Based on the similarity of the optical absorption spectra of HRP-I and CAT-I to that of ComOEP+‘ZCIO4' and CoIIIOEP+'ZBr‘, respectively, Dolphin at 21.16 proposed porphyrin 1r cation structures for the enzyme intermediates. It is further suggested that both electronic states are represented by HRP-I (2A2u) and CAT-I (2Alu)- EPR and ENDOR measurements have confirmed the if cation radical formulation for HRP-1.13119 However, the question of electronic ground state for these and other compound I structures is unclear. The results for EPR and Mossbauer studiesZI‘23 often seem to contradict classifications suggested by optical absorption spectra.20:24 More specifically, the EPR and Mossbauer measurements of compound I type intermediates often give evidence of magnetic coupling between the 3:1 oxoferryl and the S=l/2 porphyrin n cation radical spin systems. Such interaction is not expected if the porphyrin radical resides in the a1u(1r) orbital owing to the lack of spin density on the nitrogen. atoms (see Figure 1-3). However, chloroperoxidase compound I (CPO-I) and certain reconstituted HPR-I species exhibit both 2A1u type absorption spectra and magnetic coupling. Thus a contradiction arises. Furthermore, the magnetic coupling between paramagnetic metal centers, such as the oxoferrryl structure and porphyrin 1r cation radicals, is not well understood“,25 Part of this confusion is caused I I j I A I oilmOEPz’ZCI " i: 1.0 1.0 713 0.5- ‘ 0.5 7" Z Co1m1oeeg’zar" , I 1 1 l g 1 .0 ‘3 O I f 1 I l O .- 9 x x U U 1.0- ~ 1.0 0.5r “0.5 0 0 Figure 1-4. Optical absorption Spectra of (a) cobaltic octaethyl- porphyrin 11 cation radicals, (b) compound I enzyme transients. 2 A2u type spectra (——). 2A1u type spectra (- -). Reference 16. by the lack of knowledge of the effects of symmetry lowering on the spectral properties of metalloporphyrin 1r cation radicals.22 In the natural compound I systems the molecular symmetry is lowered (from D4h) by asymmetric peripheral substitution of the ring, mixed axial ligation and possibly further by porphyrin core distortions. Neglecting core distortions (i.e. doming or buckling of the porphyrin ring system) the symmetry of compound I is CS or lower. An understanding of the relationship between the different mechanisms of symmetry lowering and the electronic ground state as well as an understanding of how each influence of spectroscopy of the MP‘" are necessary to resolve these questions. It has been suggested that S4 type ruffling of the porphyrin ring may provide an overlap pathway for antiferromagnetic coupling in ferric porphyrin «cation radicals.25 Thus, with our MP‘" model work we take a structural approach which focuses on understanding the effects of the porphyrin core geometry on the vibrational and electronic properties of the MP“. It is anticipated that the resulting correlations can be used to suggest core geometries for unknown structures. Below, in Figures 1-5, 1-6 and 1-7 optical absorption spectra of various catalase and peroxidase compound I intermediates are collected. 10 Figure 1-5 . Optical absorption spectra of compound I intermediates (zAlu type). . (a) native horse erythrocyte catalase (CAT) and CAT-I. Reference 26. 2 w40- I 301— '0b 1 clllllllllill IILIILIIILIIIIIIILIIIII CPO-I 340 380 420 475 SIS 555 595 635 675 (b) native chlor0peroxidase (CPO) and CPO-I. Reference 24 . lode ( qu") 11 l _L l l l l l L l 360 400 «0 400 520 so 500 1540 600 m1.» Figure 1-6. Optical absorption spectra of compound I intermediates. (a) native ferric bromoperoxidase (BPO) and BPO-I. Reference 27. 140 1 1 1 35 .5. :: L1? 120 - :3 — 30 E: ,1 'i 100 - i: — 25 :1 A ' I ... ' l 1 , : g 00-— 5 I -« 20 7 1' : 2 " : E 60b ! 1 ‘ '5 I l m 1' 1' 4O 1 20 ’ \‘~--I~-"~+ o 400 500 600 700 k (nm) (b) native lignin peroxidase (LiP) and LIP-I. Reference 29. Absorbonce 12 Absorbonce 1.00 450 500 550 600 650 Wavelength (nm) (e) Rapid scan spectrophotometric measurements for the reactions of intestinal peroxidase (IPO) and, lactoperoxidase (LPO) with hydrogen peroxide at pH 7 .1. Dashed lines repre- sent native enzyme. the numbers show time in ms from the stop flow to the end of the wavelength scan. Thus, early times represent compound I spectra, and late times compound II. At intermediate times, mixtures were obtained. Reference 28. l3 :Euyne 150 w ———-. uzoz jg A ;‘ . \I g IOC ., \I r M 2 E V 50 a, .N‘. 250 300 350 400 450 500 550 600 650 700 wave LENGTH (mp) Figure 1-7. Absorption spectra of native cytochrome c peroxidase (CcP) and CcP-l. Reference 30. CcP-I contains an amino acid radical rather than a porphyrin radical. 10. 11. 12. REFERENCES Bolscher, B. Ph.D. Thesis. Dunford, H.B.; Stillman, J.S., Coord. Chem. Rev. 1976, _1_9_, 187-251. Frew, J.E.; Jones, P., Adv. Inorg. Bioinorg. Mech. 1984, _3_, 175-213. Hewson, W.D.; Hager, L.P. in "The Porphyrins"; Dolphin, D., Ed.; Academic Press: New York, 1979; Vol. VII, pp. 295-332. Cramer, S.P.; Dawson, J.H.; Hodgson, K.O.; Hager, L.P., J. Am. Chem. Soc. 1978,_1_0_0_, 7282-7290. Murthy, M.R.N.; Reid, R.J.; Sicignano, A.; Tanaka, N.; Rossmann, M.G., J. Mol. Biol. 1981, _1_§_2_, 465-499. (a) Chance, B., Arch. Biochem. Biophys. 1952, 4_1, 416-424; (b) George, P., Nature 1952, .122, 612-613. (a) Theorell, H.; Ehrenberg, A, Arch. Biochem. Biophys. 1952, 4_1_, 442- 461; (b)Brill, A..;S Williams, R.J.P, Biochem. J. 1961, L8, 253- 263; (c)Blumberg, W.E.; Peisach, J.; Wittenberg, B.A.; Wittenberg, J. B., J. Biol. Chem. 1968, _2_43, 1854-1862; (d)Wittenberg, B..A; Kampa, L.; Wittenberg, J. B.; Blumberg, W. E.; Peisach, J., J. Biol. Chem. 1968, _2_43 1863- 1870; (e) Peisach, J, Blumberg, W. E.; Wittenberg, B. A.; Wittenberg, J.B., J. Biol. Chem. 1968, m, 1871- 1880. (a) Keilin, D.; Hartree, E. F., Biochem. J. 1951, 4_9, 88-109; (b) Shannon, L..;M Kay,F.., LeW,J..,Y J. Biol. Chem. 1966, _2__41, 2166- 2172. Groves, J.T.; Haushalter, R.C.; Nakamura, M.; Nemo, T.E.; Evans, B.J., J. Am. Chem. Soc. 1981, 103, 2884-2886. Parson, W.W.; Ke, R. in "Photosynthesis: Energy Conversion by Plants and Bacteria"; Govindjee, Ed.; Academic: New York, 1983. Chang, C.K.; Hanson, L.K.; Richardson, P.F.; Young, R.; Fajer, J., Proc. Natl. Acad. Sci. USA 1981, _7_8, 2652-2656. 14 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 15 Blair, D.F.; Witt, S.N.; Chan, 8.1., J. Am. Chem. Soc. 1985, 1_0_7_, 7389-7399. Maggiora, G.M., J. Am. Chem. Soc. 1973, _95, 6555-6559. Fajer, J.; Davis, MS. in "The Porphyrins", Dolphin, D., Ed.; Academic Press: New York, 1979; Vol. IV, pp. 197-256. Dolphin, D.; Forman, A.; Borg, D.C.; Fajer, J.; Felton, R.H., Proc. Natl. Acad. Sci. USA 1971, Q8, 614-618. Morishima, l.; Takamaki, Y.; Shiro, Y., J. Am. Chem. Soc. 1984, 106, 7666-7672. Schulz, C.E.; Devaney, P.W.; Winkler, H.; Debrunner, P.G.; Doan, N.; Chiang, R.; Rutter, R.; Hager, L.P., FEBS Lett. 1979, 103, 102-105. Roberts, J.E.; Hoffman, B.M.; Rutter, R.; Hager, L.P., J. Biol. Chem. 1981, 256, 2118-2121. DiNello, R.K.; Dolphin, D., J. Biol. Chem. 1981, 256, 6903-6912. Rutter, R.; Valentine, M.; Henrich, M.P.; Hager, L.P.; Debrunner, P.G., Biochemistry 1983, _2_2, 4769-47 74. Schultz, C.E.; Rutter, R.; Sage, J.T.; Debrunner, P.G.; Hager, L.P., Biochemistry 1984, 2_3_, 4743-47 54. Rutter, R.; Hager, L.P., J. Biol. Chem. 1982, 257, 7958-7961. Palcic, M.M.; Rutter, R.; Araiso, T.; Hager, L.P.; Dunford, B.H., Biochem. Biophys. Res. Comm. 1980, 93, 1123-1127. Gans, P., Buisson, G.; Duee, E.; Marchon, J.-C.; Erler, B.S.; Scholz, W.F.; Reed, C.A., J. Am. Chem. Soc. 1986, 108, 1223-1234. Schonbaum, G.R.; Chance, B. in the "The Enzymes"; Boyer, P.D., Ed.; Academic Press: New York, 1976, Vol. 13, pp. 363-408. Manthey, J.A.; Hager, L.P., J. Biol. Chem. 1985, _2_§_0, 9654-9659. Kimura, S., Yamazaki, 1., Arch. Biochem. Biophys. 1979, 1_9_§_, 580-588. Renganathan, V.; Gold, N.H., Biochemistry 1986, _2_5_, 1626-1631. Yonetoni, T., J. Biol. Chem. 1965, _240, 4509-4514. Hewson, W.D.; Hager, L.P., J. Biol. Chem. 1979, 153, 3182-3186. CHAPTER 2 CHARACTERIZATION OF SIX-COORDIN ATE OXOFERRYL PROTOHEME BY RESONANCE RAMAN AND OPTICAL ABSORPTION SPECTROSCOPY. SUMMARY As a model for oxoferryl intermediates of heme enzymes, we have synthesized (l-methylimidazole) oxoferryl protoporphyrin IX dimethyl ester, O=FeIV(Im)PPDME, and characterized it by using optical absorption and resonance Raman spectroscopy. We observe an FeO stretching frequency, v(FeO), of 820 cm‘1 for O=FeIV(Im)PPDME, which is assigned by its shift to 784 cm"1 upon substitution of 16O with 180. While the optical absorption spectrum of the six-coordinate protoheme model is very similar to that of oxoferryl myoglobin, its v(FeO) is 23 cm'1 higher. Based on results from five- and six-coordinate oxoferryl porphyrin models, we attribute this higher frequency to weaker imidazole ligation and an absence of protic environmental effects. *Robert T. Kean, W. Anthony Oertling, and Gerald T. Babcock, J. Am. Chem. _S_o_c., 1987, in press. 16 17 Oxoferryl species, O=FeIV, have been postulated in the catalytic cycle of cytochrome c oxidase,1 as the oxygen donating species in cyctochrome P-450,2 and as intermediates in the reactions of catalases and peroxidases.3 Given the diverse chemistry catalyzed by these various enzymes, heme pocket modulation of the chemical reactivity of the O=FeIv unit seems likely. Resonance Raman detection of the v (FeO) in various protein species and model compounds supports this notion. In a comparison of oxoferryl peroxidase species,4 oxoferryl myoglobin,5 and five- and six-coordinate heme model compounds,617 the frequency of v(FeO) varies by '1: 85 cm‘1 (see Table 2-1). This is in strong contrast to v(Fe—Oz) which varies by only '1: 10 cm“1 in protein species and heme model compounds.8 In addition, there are distinct differences in the optical spectra of the various oxoferryl protein species.9"11 Previously reported oxoferryl model compounds have given insight into the factors affecting the v(FeO) frequency, but since they were made on non-physiologically active hemes, they do not give specific information about the optical spectra or the other Raman active vibrations of the protoheme-containing oxoferryl protein species. To address these points, we present here optical and Raman data for a six-coordinate, imidazole-ligated, oxoferryl protoheme model compound. The (l-methylimidazole) oxoferryl protoporphyrin IX dimethyl ester, O=FeIV (Im)PPDME, was prepared according to reference 12. The optical absorption spectrum138 is inset in Figure 2-1. The shoulder at 619 nm is due primarily to u-oxo dimer contamination. The peak positions compare favorably with those of oxoferryl hemoglobin,9 and oxoferryl leghemoglobin,10 ('1: 418, 545, and 575 nm) but are most similar in both wavelengths and relative intensities to oxoferryl myoglobin (m 420, 550, and 580 nm at pH 6.8).11 The spectrum of O=FeIv (Im)PPDME, like the oxoferryl globin species, is distinct from that of horseradish peroxidase compound 11 (HRP-II; m 418, 527, and 555 nm).14 In Figure 2-1 we Ao‘z406.7 nm “6 x5 5 5 3 511 “ 677 I 3 n g 619 I e E 350 V ' I 111161111011? (my I ' ' I100 z t! _2_ z e s . I g 7114 751 W '80 ~Av=36-‘ ‘ cm" 820 ..0 Wfl/V 1 ' 1 ' I r 1 . ' 1 ' 1 . I ' soo FREQUENCY (cm") 900 Figure 2-1. Resonance Raman s ectra of 18O=FeIV(Im)PPDME (upper trace), and 16O=FeI (Im)PPDME (lower trace). Vertical line denotes v (FeO); * denotes solvent peak. Spectra were obtained in toluene—d3 at -l30 C with 15 mW (406.7 nm) incident on the sample. Inset: Optical absorption spectrum of O=FeIV(Im)PPDME in toluene at -90 C. The shoulder at 619 nm is due primarily to u—oxo dimer contamination. 19 present the intermediate frequency region of the Raman spectrum13b of O=FeIv (Im)PPDME. The peak at 820 cm‘1 shifts to 784 cm'1 upon substitution of 16O by 18O, and is assigned to FeO stretching vibration. The 36 cm"1 shift is expected for an oxoferryl structure. The v(FeO) frequency observed for O=FeIv (Im)PPDME does not vary with temperature over the range -90 C to -190 C or with porphyrin ring substituents (TPP or OEP).15 This frequency is compared with the u (FeO) frequencies of other oxoferryl species in Table 2-1. The high frequency region of the O=FeIV(Im)PPDME spectrum (not shown) contains features of both HRP-II16 and oxoferryl myoblobin5 yet it is not identical to either.15 The similarity of the optical spectrum of oxoferryl myoglobin with that of O=FeIv (Im)PPDME indicates that the two are electronically very similar. The difference in the v (FeO) frequencies suggests environmental perturbations. Differences in the v (FeO) frequency between the peroxidase species and oxoferryl myoglobin have been discussed in terms of hydrogen bonding effects“:e and out-of-plane iron effects.5 It has been suggested that 11 charge donation from the protein into the porphyrin may be important in the mechanism of horseradish peroxidase,17 but it is not known whether this has any affect on the v(FeO) frequency. Comparison of these systems with the model compounds gives further insights into the variables which affect the v (FeO) frequency. A trans ligand effect is seen very clearly in the model compounds (see Table 2-1). The five—coordinate models display the highest 0 (FeO) frequency, while the strong ligand (Im), six-coordinate samples display the lowest frequency. Electron density from the sixth ligand, along the z axis (normal to the heme plane), may compete with the ferryl oxygen for o-bondingl8 and weaken the FeO bond. Alternately, if the iron is initially displaced toward the oxygen in the five-coordinate species, the presence of a strong sixth ligand may pull the 20 Table 2-1 Comparison of v(FeO) for Various Oxoferryl Species Species 11 \1(Fe0)b Ref. HRP-II pH 6.0 776 4c,e HRP-II pH 11.0 787 4b,c CcP-I 767 4d oxoferryl Mb 797 5 O=FeIVTMP 843 21 o=FeIVTPP . 852 6 o=FeIV0EP 852 6 O=FeIV(THF)TpivPP 829 7 O=FeIV(Im)TpivPP 807 7 O=FeIV(Im)PPDME 820C O=FeIV(Im)TPP 820 O=FeIV(Im)OEP 820 aAbbreviations: HRP-II, horseradish peroxidase com- . pound II; CcP-I, cytochrome c peroxidase compound 1; Mb, myoglobin; TMP, tetramesitylporphyrin; TPP, tetraphenylporphyrin; OEP, octaethylporphyrin; THF, tetrahydrofuran; Im, l-methylimidazole; TpivPP, "picket fence" porphyrin, tetra(o-pivaloylphenyl)- porphyrin; PPDME, protoporphyrin IX dimethyl ester. b Frequency in cm‘l. C This work and reference 15. 21 iron into plane causing greater 1! interaction between iron and porphyrin orbitals, and a weakening of the FeO 1! -bond.5 The difference in the v(FeO) frequency of our models (in toluene) versus the (TpivPP)7 species (in THF) appears to results from solvent effects since temperature and ring substituent effects were ruled out above. This may reflect stronger imidazole binding in the more polar and non-aromatic THF. Although hydrogen bonding of the oxo ligand to the distal histidine could be involved, the lower v(FeO) frequency of oxoferryl myoglobin may indicate stronger proximal imidazole ligation in myoglobin than in the model compounds. This seems reasonable since the imidazole is protein bound in myoglobin and may not easily move or rotate to less strongly ligating configurations available to the free solution models. Similar trans ligand effects may contribute to the difference in v(FeO) frequencies observed for HRP-II and oxoferryl myoglobin. For the respective five-coordinate ferrous enzymes, the higher u(FeH-imidazole) frequency for HRP indicates stronger imidazole ligation.19920 For the oxy complexes, v(FeH-imidazole) is also higher for HRP than for myoglobin but its \1(Fe-02)8c is lower. A similar inverse relationship between (Fen-CO) and trans ligand strength has been reported for monomeric insect hemoglobins18a and heme model compounds.18b ACKNOWLEDGMENTS We thank Dwight Lillie for data handling and graphics software. This research was supported by NIH grant GM 25480. 10. REFERENCES AND NOTES (a) Wikstrom, M., Proc. Natl. Acad. Sci. USA 1981, _7_8, 4051-4054; (b) Blair, D.F.; Witt, S.N.; Chan, 8.1., J. Am. Chem. Soc. 1985, 107, 7389-7399. Groves, J.T. in "Cytochrome P-450: Structure, Mechanism and Biochemistry"; Ortiz de Montellano, P. Ed; Plenum Press: New York, 1985; Chapter I. Hewson, W.D.; Hager, L.P. in "The Porphyrins", Vol. VII; Dolphin, D. Ed.; Academic Press: New York, 1979, pp 295-332. (a) Terner, J.; Sitter, A.J.; Reczek, C.M., Biochim. Biophys. Acta 1985, _8_28_, 73-80; (b) Hashimoto, S.; Tatsuno, Y.; Kitagawa, T., Proc. Japan Acad. 1984, §_0_(B), 345-348; (c) Sitter, A.J.; Reczek, C.M.; Terner, J., J. Biol. Chem. 1985, _2_60, 7515-7522; ((1) Hashimoto, S.; Teraoko, J.; Inubushi, T.; Yonetani, T.; Kitagawa, T., J. Biol. Chem. 1986, 261, 11110-11118; (e) Av(FeO) value of 775 cm'1 (HRP-II, pH 7, H20 outfit?) is reported which shifts to 777 cm‘1 in D20 buffer. Hashimoto, S.; Tatsuno, Y.; Kitagawa, T., Proc. Natl. Acad. Sci. USA 1986, _83, 2417-2421. Sitter, A.J.; Reczek, C.M.; Terrier, J., Biochim. Biophys. Acta 1985, 828, 229-235. (a) Bajdor, K.; Nakamoto, K., J. Am. Chem. Soc. 1984, 106, 3045-3046; (b)Proniewicz, J.M.; Bajdor, K.; Nakamoto, K., J. Phys. Chem. 1986, _9_0_, 1760-1766. Schappacher, M.; Chottard, G.; Weiss, R., J. Chem. Soc., Chem. Commun. 1986, _2_, 93-94. (a) Spiro, T.G. in "Iron Porphyrins", Part II; Lever, A.B.P.; Gray, H.B. Ed.; Addison Wesley: Reading, Mass, 1983; pp. 107-133; (D) Kean, R.T., unpublished results; (c) Van Wart, H.B.; Zimmer, J., J. Biol. Chem. 1985, _2_06, 8372-8377. Dalziel, K.; O'Brien, J.R.P., Biochem. J. 1954, 26, 648-659. Aviram, 1.; Wittenberg, B.A.; Wittenberg, J.B.; J. Biol. Chem. 1978, 253, 5685-5689. 22 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 23 George, P.; Irvine, D.H., Biochem. J. 1952, _5_2_, 511-517. LaMar, G.N.; de Ropp, J.S.; Latos-Grazynski, L.; Balch, A.L.; Johnson, R.B.; Parish, D.W.; Cheng, R., J. Am. Chem. Soc. 1983, 105, 782-787. (8) Optical absorption spectra were obtained by using a house build optical dewar mounted in a Perkin-Elmer Lambda 5. The sample was contained in an EPR tube which was cooled to -90°C by flowing cold nitrogen gas; (b) Raman spectra were obtained with a Spex 1401 scanning monochromator (with PMT detection) by using 15 mW incident power at 406.7 nm (Spectra-Physics model 164 Kr ion) in a backscattering geometry. The samples, contained in EPR tubes, were spun continuously in a dewar while the desired temperature was maintained by flowing cold nitrogen gas. Blumberg, W.E.; Peisach, J.; Wittenberg, B.A.; Wittenberg, J.B., J. Biol. Chem. 1968, 243, 1854-1862. Kean, R.T.; Babcock, G.T., manuscript in preparation. Terner, J.; Reed, D.E., Biochim Biophys. Acta 1984, 789, 80-86._ Shelnutt, J.A. ; Alden, R.G.; Ondrias, M.R., J. Biol. Chem. 1986, 261, 1720-1723. (a) Gersonde, K.; Kerr, B.A.; Yu, N-T.; Parish, D.W.; Smith, K.M., J. Biol. Chem. 1986, 261, 8678-8685; (b) Kerr, B.A.; Mackin, H.C.; Yu, N-T., Biochemistry 1983, 2_2_, 4373-437 9. Teraoko, J.; Kitagawa, T., Biochem. Biophys. Res. Comm. 1980, _9_3, 694-700. Teraoka, J.; Kitagawa, T., J. Biol. Chem. 1981, 256, 3969-3977. Hashimoto, S.; Tatsuno, Y.; Kitagawa, T. In "Proceedings of the Tenth International Conference on Raman Spectroscopy; Petiolas, W.L.; Hudson, B. Ed.; Univ. of Oregon: Eugene, 1986, pp. 1-28, 1-29. CHAPTER 3 RESONANCE RAMAN SPECTROSCOPIC DETECTION OF DEMETALLATION OF METALLOPORPHYRIN s CA'HON RADICALS. SUMMARY Soret resonance Raman (RR) spectrum of samples of several different metalloporphyrin 11 cation radicals (MP“) prepared by both chemical and electrochemical oxidation in CH2C12 reveal the presence of free base diacid salts produced by demetallation of the complexes. The large extinction coefficient of the diacid salt allows its selective enhancement by the RR technique at concentration levels not always evident in absorption spectra but sufficient to cause serious artifacts in the RR spectra of these samples, making recognition and analysis of the scattering from the MP“ impossible. Proper control of experimental conditions, particularly choice of excitation frequency, can eliminate these complications and produce RR spectra of MP“ free of contributions from the diacid salt. * W. Anthony Oertling, Asaad Salehi, Chi K. Chang, and Gerald T. Babcock, J. Phys. Chem., 1987, in press. 24 25 Resonance Raman (RR) spectroscopy has been utilized to extract structural and electronic information from naturally occurring metalloporphyrin systems and their models, and has recently been applied to synthetic metalloporphyrin n cation radicals (MP“).1"3 The preparative techniques for MP“ generation, however, rarely produce a homogeneous sample.4 Owing to the selective enhancement afforded by Soret excitation RR, it is possible to detect trace amounts (less than 396) of porphyrin free base diacid salts in samples of MP“ prepared by chemical and electrochemical oxidation in CH2C12. If unrecognized, the presence of the diacid salt may cause serious artifacts in the RR spectra of these samples that prevent accurate analysis of scattering from the MP“.1 Proper control of experimental conditions, particularly the choice of excitation frequency, can produce RR spectra of MP“ free of contributions from the diacid.3’5 One-electron oxidation of Co and Zn octaethylporphyrin (OEP) in dry CHZCIZ leads to divalent metal porphyrin 1r cations.3v6 Laser excitation at 363.8 nm selects against resonance enhancement of possible residual starting materials and produces good quality RR spectra.3 RR spectra of these MP“ samples excited at 406.7 nm, however, are often dominated by bands not present in spectra of 363.8 nm and not assignable to parent MP modes.7 The frequencies of these bands (most noteworthy are those at 1394 and 1558 cm'l) show no metal dependence. These facts suggest the presence of an impurity which absorbs strongly near 406.7 nm. Figure 3-1a and 3-1b show RR scattering excited at 406.7 nm from samples of ConOEP“ClO4‘ prepared with AgClO4 and ComOEP“2ClO4‘ prepared with Fe(ClO4)3, respectively.8 Although the RR spectra of these two species excited at 363.8 nm are very similar3, with excitation at 406.7 nm additional bands appear in the spectrum of the cobaltic OEP“ relative to the cobaltous sample. Figure 3-1 . 26 Resonance Raman spectra excited at 406.7 nm. (a) CoHOEP“ClO4' prepared with AgClO4; (b) ComOEP“2ClO4‘ prepared with Fe(ClO4)3; (c) ComOEP“2ClO4‘, prepared with Fe(ClO4)3 and a trace amount of H0104; (d) H4OEP2+2CIO4‘, prepared from H2039 + H0104. an samples were dissolved in CHZCIZ. A spinning quartz cell and 'v 20 mW laser power were used. Solvent bands are labelled with an *. 27 *MNVT I; 2CK) c)ComOEPI +HCIO4 >._._mZm:.Z_ Z<§2+ r RAMAN SHIFT (cm") 28 Figure 3-1c shows scattering from a sample of ComOEP“ZClO4‘ prepared with Fe(ClO4)3 and a trace amount of HClO4. The added intensity of the spurious bands in this spectrum) suggests that the impurity is caused by acid promoted demetallation of the MP“ and protonation of the resultant free base porphyrin. Figure 3-1d shows the RR spectrum of the diacid salt H4OEP2+2CIO4', confirming this as the interfering species. Thus, RR spectra of MP“ samples exicted at 406.7 nm may contain artifacts due to free base diacid salts present at concentrations too low to be obvious in the optical absorption spectrum yet high enough to dominate the RR scattering at 406.7 nm. This is clear from the Soret optical properties of the impurity: the Soret band sharpens dramatically and red-shifts to 406 nm (e = 430m M‘lcm'l) upon formation of the diacid from OEPH2.9 Thus, the narrowed bandwidth, the increased extinction coefficient a and the coincidence of the Soret transition energy with the laser line at 406.7 nm, explain the selective scattering from the impurity. Figure 3-2 shows scattering excited at 390 and 406.7 nm of ZnOEP“ClO4‘. The absence of bands at 1394 and 1558 cm"1 indicates no contributions from the impurity with 390 nm excitation.10 Spectra excited at 406.7 nm show bands at 1394 and 1559 cm'1 as depicted in Figure 3-2b. Unlike the ComOEP“2ClO4‘ sample, which appears to be stable to prolonged laser irradiation, the contributions from the impurity (marked by the dashed lines) increase dramatically with repeated scanning of the ZnOEP“ClO4' sample until they dominate the spectrum as shown in Figure 3-2c-d. The solvent bands at 704 and 1423 cm‘1 show the opposite trend and decrease in intensity, indicating stronger absorbance at 406 nm as the impurity concentration increases. Indeed, for this MP“ sample with the Soret band at 387 nm, the Soret absorbance at 406 nm of the protonated porphyrin was easily seen to increase in parallel with 29 Figure 3—2. RR spectra of ZnOEP“ClO4. (a) Aex = 390 nm; (b) - ((1) show a time course at Aex = 406.7 nm for 'v 1 ml of sample in a quartz spinning cell with a laser power of 16 mW. The time values indicate total irradiation time at the end of the scan. 30 ZnOEP“C1021 * e051 IIIIIIIIII 111111 ..l111lll 1- .Ilmn: m m m n n 7.11. "W h 7h m 63 %2 @3 n m0 4 \I O to d w m lllll I..I.1|. .Illlmmm 1m . 11 III 111111 8a mom- 1 l INIJ1 I I lfiom >._._mzw._.z_ Z<§H_mzwi_.z_ c)ZnOEP Z<2._._mzm._.z_ 24.24.". RAMAN SHIFT (cm") 49 are from the diacid salt impurities. Figure 4-3c shows the RR Spectrum of the interferring Species, H4OEP2+2C104'. Addition of AgClO4 to a solution of HZOEP in CH2C12 results in a mixture of mono- and diacid species which produces the RR spectrum shown in Figure 4-3d. Whereas the vertical lines identify the vibrations of the diacid, the additional peaks labelled in Figure 4-3d (e.g. those at 1505, 1582 and 1634 cm‘l) most likely arise from the monoacid species. This suggests that the mono- and diacid species have different vibrational properties, possibly owing to the lowered symmetry resulting from the inequivalence of the x and y axes in the monoacid structure. The formation of diacid in the sample of ComOEP“2CIO4‘ prepared with Fe(C104)3 probably results from aqueous acid (i.e. HC104) present in the reagent. Thus, acid demetallation of the MP followed by protonation of the resultant free base could compete with the oxidation process. The lack of artifacts in the spectrum in Figure 4-3a illustrates that this process does not occur when anhydrous AgClO4 is used to oxidize CoOEP. The spectrum in Figure 4—3d is of interest because it suggests the possibility of both monoacid and diacid contamination of MP“ samples. Furthermore, the absence of aqueous acid impurities in the AgClO4 reagent suggests that porphyrin acid formation is a true by-product of the oxidation process in certain cases, and not merely the result of "wet" oxidant. Figure 4-3e illustrates that changing the counterion for the diacid from C104” to Br‘ decreases the frequencies of the intense vibrational bands above 1300 cm“1 by approximately 3 cm’l. The above observations are all pertinent to the interpretation of the RR Spectrum of ComOEP+'ZBr‘ obtained with 406.7 nm excitation by Kim 3 31:,20 which we suggest results from H40EP2+2Br' (possibly mixed with H30EP+Br‘) impurities present in the sample. The strong enhancement of the diacid vibrational modes with 406.7 nm excitation, even when the species occurs as a minor impurity (less than 396) 50 Figure 4-4 . Electronic absorption spectra of porphyrin free base acids. (6) H40EP2+20104‘; (b) a mixture of H30EP+CIO4’ and H40EP2+2C104‘ corresponding to Fig. 4-3d. Solvent, dry CH2C12. e(mM"' cm") 51 U T 1 I l I I I I r I I I I l I T I I 400« - 300 2001 1 o) 114 051:“ 2610; I001' ' Lm o. 1 2, .38. b) H40EP 2CIOZ H30EP+ Clo; - ~200 " x5 I . ' PI00 I K“—“/—Jd>““—:====ho VrIiITIIU'UI 300 400 500 600 700 WAVELENGTH (nm) 52 in MP‘“ samples, is rationalized by its optical properties. Figure 4-4a shows the optical absorption spectrum of the H4OEP2+ZCIO4‘ sample used for the RR measurement shown in Figure 4—3c. Figure 4-4b shows the absorption spectrum of a mixture of mono- and diacids produced by the addition of AgClO4 to OEPHZ in CH2012 and corresponds to the sample used for the RR spectrum in Figure 4-3d. Contributions to the absorption spectrum from the monoacid are evidenced by the high energy shoulder (N390 nm) in the Soret and by the absorption maximum at 518 nm.36 While this sample is predominantly (>8096) monoacid, the RR spectrum (Fig. 4-3d) is dominated by the diacid vibrations, owing to the large extinction coefficient (430 mM‘lcm‘l) and close correspondence of the Soret band (406 nm) of the latter species .to the laser line (406.7 nm). The extent of demetallation and diacid formation in MOEP“ sample is dependent on a number of factors, including the oxidation method, solvent and core size of the MOEP’“. In some cases it is enhanced by laser irradiation in the Soret band region.21 By exciting well to the blue of the absorption maxima of these potential contaminants, we avoid the complications imposed by these species upon RR measurements. One- and two—electron oxidation of CoOEP. Figure 4-5 shows uv—visible absorption spectra of CoOEP and products of its oxidation in dry CH2C12 in the presence of ClO4‘. The spectrum of the starting material, CoOEP, and the product of ring centered, one-electron oxidation, CoIIOEP+'ClO4', are shown in Figure 4-5a and b, respectively. Figure 4-5c shows the spectrum of the two-electron oxidation product, CoIIIOEP+'ZClO4'. As we recently demonstrated,22 cobaltous OEP‘“ is extremely sensitive to the presence of axial ligands which coordinate more strongly than CIO4“. Such coordination promotes electron transfer from CoII to the electron—depleted porphyrin. Addition of methanol (MeOH) to CHZCIZ solutions of ConOEP+'CIO4' produces 53 .9038 iv .Emzom .A...V-vo_o~.+mmo=_oo 3 xiv-¢o_o.+mmo=oo 3V mTESS 3 duos—V95 :oSwExo m: 98 mmOoO mo 23on :23..on 8.8.533 .3 95mm. 54 2:5 Iszmqm><>> Om. CON OmN ('_w3,_ww) 9 55 a cobaltic species, CoIH(MeOH)20EPClO4‘, with an absorption spectrum similar to, but red-shifted with respect to that of the parent CoOEP species. Because water coordinates and produces a cobaltic species similar to the dimethanol adduct (see Fig. 4-10, below), rigorously anhydrous conditions are necessary to provide a homogeneous preparation of CoIIOEP+'ClO4‘. The absorption spectrum of this species (Fig. 4—5b) closely resembles that of NiOEP+'CIO4',29 and has features usually attributed to the 2A2u state (as does CoIIIOEP+'ZClO4", Fig. 4-5c). Figure 4-6 shows absorption spectra of three species obtained by Br2 oxidation of CoOEP in CHZCIZ. One-electron oxidation in this case also produces two distinct species, but both exhibit metal centered oxidation. One-electron oxidation by Brz followed by dropwise addition of MeOH produces the six-coordinate Com(MeOH)20EPBr" shown in Figure 4-6a. The absorption spectrum of this species is typical of six-coordinate cobaltic porphyrins37 and similar, although slightly red-shifted, to that of the dimethanol adduct, CoH(MeOH)20EPC104', described previously22 (see Table 4-1, below). One—electron oxidation of CoOEP by Brz in dry CHZCIZ produces a presumably five-coordinate cobaltic porphyrin which displays the spectrum shown in Figure 4-6b. For this compound, ComOEPBr', the Soret band (373 nm) blue-shifts with respect to that of CoOEP (391 nm), and the extinction coefficient ratio of the visible to the Soret band increases. The spectrum of the two-electron oxidized 1r cation radical, CoIIIOEP+'ZBr‘, is shown in Figure 4-6c. The near uv region contains three distinct bands, at 290 (not shown), 345 and 410 nm, and there are at least two bands in the visible region at m 600 and 670 nm. These features (split Soret with an intense red-shifted visible band) are considered typical of the 2A1u electronic state.9923924 The feature at 401 nm in Figure 4-6c is attributed to the diacid salt (H4OEP2+ZBr'), the presence of which 56 .8035 3552. *m a 953:8 :02; A3 ..8 3088 .SONIO En £528 .38“ 2.: E comm-.36 mm 53538350 ....muémmOv: 8H 3 26 m_ E: 2:. «a 9538 :25 2: xiv ammimmoaoo A8 xiv (ammo—=8 3v x¢-..mmm0£:o§=_oo 3 .mmOoO no 32695 53258 go 950on .8390me 3.8.5005 . E 2&5 57 ASE Ihozm4m><>> fiOON (,_u13'_wu1)9 58 explains the subtle variations in band shape and maximum of the 410 nm transition evident in different preparations of this compounds”24 We estimate the diacid salt contamination in the preparation represented by Figure 4-6c to be «196.38 Near uv RR spectra of one- and two-electron oxidation products of CoOEP. Figure 4-7 collects RR spectra produced by 363.8 nm excitation of various products of CoOEP oxidation. The effects of metal centered, one-electron oxidation accompanied by axial metal ligation are demonstrated by comparing the spectra of CoOEP and Com(MeOH)2OEPClO4‘ depicted in Figure 4-7a and b. The oxidation state marker, v 4, increases from 1379 to 1383 cm‘l, reflecting depopulation of the porphyrin 1r* orbitals caused by metal oxidation.39 There is little systematic change in frequency of the modes above 1400 cm'l, indicating that the core size of the porphyrin ring does not change significantly upon metal oxidation and addition of axial MeOH ligands.32940 The v2 mode, which occurs at 1599 cm'1 in CoOEP, is very weak in this spectrum of Com(MeOH)20EPClO4‘, but it is seen at 1596 cm‘1 in spectra excited at 406.7 nm (not shown). The remaining high frequency bands of CoOEP (Fig. 4-7a) at 1512, 1575 and 1647 cm“1 are assigned to v3, VII and v10, respectively. The weak mode at 1476 cm"1 may correspond to v23. Although this mode is not observed in Soret excited spectra of other MOEP complexes, it is reported for NiPP (PP=protoprophyrin IX) at 1482 cm'l.40 These frequencies are consistent with the CaCm stretching character of this mode. Assignments for Com(MeOH)20EPCIO4 are similar. Figure 4-7c and (I show the spectra of the cobaltous and cobaltic OEP+° complexes. Neglecting differences in relative intensity produced by the differences in Soret absorption, the RR spectra of CoIIOEP+'ClO4' and CoIHOEP+'2ClO4' are essentially identical and (above 1300 cm'l) radically different from those of the neutral ring compounds in Figure 4-7a and b. These observations provide a strong fingerprint basis for 59 Figure 4-7 . RR spectra excited at 363.8 nm ( ~35 mW) of CoOEP and its oxidation products. _ (a) CoOEP; (b) Com(MeOH)20EPClO4‘; (c) CoIIOEP+'ClO4‘; (d) ComOEP+'ZClO4‘; 60 Xe, = 363.8 nm be >._._mzw._.ZH Z<_>_.:mzm._.ZH 232$”. IIOO ISOO ISOO - I700 900 RAMAN SHIFT (cm") 64 Visible RR spectra of one- and two-electron oxidation products of CoOEP. While Soret (or B band) resonance Raman excitation of neutral metalloporphyrins enhances totally symmetric modes (A1g in D4h symmetry) via a Franck-Condon (FC) scattering mechanism, excitation in the visible absorption (Q band) of these species enhances primarily nontotally symmetric modes (Blg, 82g and Azg) via a Herzberg—Teller (HT) mechanism.41 In particular, excitation in the B or Q(0-1) band strongly enhances A2g modes of the MP,42 well known for their inverse polarization.43 Figure 4-8a shows both polarization components of the spectrum of CoOEP obtained in Q(0-1) resonance under 514.5 nm excitation. The A2g modes are easily identified in the I 1 scan at 1123, 1310, 1394 and 1598 cm'1 and are assigned as V22, V 21, V 20 and V19, respectively. Although theory predicts that the value of the depolarization ratio, p 5 IL /I.. , will be on for these modes, the measured values are usually finite for various reasons.43:44 The large p values obtained from Figure 4-8a indicate that in CHZCIZ solutions CoOEP assumes a planar configuration and D4h symmetry as does NiOEP in solution.45 In agreement with Soret RR data for these ring-neutral species, Figure 4-8b shows that little change occurs in the spectrum as the metal center is oxidized and ligated axially by MeOH; the A2g modes of Com(MeOH)20EPBr' are apparent at 1128, 1315, 1394 and 1597 cm'l. The RR spectrum of ComOEPBr“ (Fig. 4-8c), however, is different from that of other ring-neutral MOEP compounds excited at 514.5 nm, of which the previous two examples are typical. While the Agg modes in the spectra of CoOEP and Com(MeOH)20EPBr’ are absent in the spectrum of ComOEPBr", the appearance of ap modes at 1054, 1077, 1252, 1274 and 1572 cm"1 is clear from the 11 scan in Figure 4—8c. The 1572 cm"1 band is reported in the early study by Spaulding _e_t_ _a_l_.,45 and this frequency is noted as an exception to the trends presented in that work. This band may not be assignable to v19 since it is 91:30 cm'1 too 65 Figure 4-8 . RR spectra excited at 514.5 nm (100 mW). (a) CoOEP; (b) Com(Me0H)20EPBr"; (c) ComOEPBr‘. CH2012 bands marked *. t VON- 66 ewowa OEP a) C0“ b) Co"'(MeOH)2 OEP 8r“ >thmHZ_ Z<§._._mzm._.z_ Z<_2<>> 00» com com b b L L . . P i I; « mMm 6mm e (I . . / woo amo «8.3528 «In 3.0.38.8 m0¢ mhm BONVSHOSSV 74 CoIIIOEP+'2ClO4‘ will also produce Com(MeOH)2OEPCIO4‘, although much more methanol is required than in the previous case. The redox chemistry involved here is unknown, although it may be similar to the room temperature reduction of FeIn(C1')TPP+°ClO4' by imidazole (1m) which produces Fem(Im)2TPPCl‘.10b Thus, strong axial ligation of the metal center may destabilze the porphyrin Tr cation radical. Similarly, Soret RR measurements at 406.7 nm on frozen samples of CoIHOEP+'2ClO4‘ (not shown) show increased contributions from a CoHI(H20)20EPClO4' type species with repeated scanning. Subsequent warming of these samples causes conversion back to cobaltic 1r cation radicals, as evidenced by the room temperature absorption spectrum, indicating reversible reduction in the low temperature case. Reversible conversion of NiHTPP+' to NimTPP at low temperatures has also been reported; however, no axial ligation was implicated.5t46 In general, low temperature RR studies are limited by elevated fluorescence levels and by the inability to measure depolarization ratios of frozen samples. Further complications are imposed by the increased aggregation of MOEP‘" compounds which occurs upon cooling.4929t31947 We note that under anhydrous conditions at room temperature both Collosr+°Clo4' and ComOEP+'2ClO4" exhibit striking stability to relatively high power laser irradiation at 514.5 nm. Thus, in view of the above, attempts to utilize cryogenic RR techniques to stabilize these compounds is of questionable expediency and will not easily produce reliable spectra. Room temperature visible excitation RR studies of these 2A2u type cation radicals will be limited, not by sample instability, but by the weak scattering cross-section offered by the samples. DISCUSSION In this section the four orbital model of the neutral metalloporphyrin 75 electronic states is introduced briefly in order to provide a framework for our analysis of the RR and uv—visible absorption spectra of the oxidized MP species. Next, spectra of the one-electron, ring-oxidized, MHOEP+°CIO4" systems are shown to exhibit structural trends in both the vibrational frequencies and electronic transition energies which are similar to those of the parent MOEP systems. The one-electron, metal-oxidized cobaltic OEP system is then discussed in order to establish the electronic and structural effects of metal oxidation accompanied by axial ligation. Finally, .the above correlations and analysis are used to examine structural and electronic properties of the two-electron oxidized ComOEP+'2X‘ species. In the four-orbital model of Gouterman48 the optical absorption bands of the MP originate from transitions involving the nearly degenerate HOMOs (highest occupied molecular orbitals) of the porphyrin 1r system, 81u(TI') and a2u(1r), and the degenerate LUMOs (lowest unoccupied molecular orbitals), eg(1r*). Transitions to the CI (configuration interaction) mixed excited states, Q and B, produce the a and Soret bands, respectively. The red-shifts in the Soret and a band maxima produced by metal variation in the series Ni, Co, Cu, Zn, MgOEP are explained by variations in metal electronegativity producing a totally symmetric perturbation to the electronic states of the porphyrin. Red-shifts accompanying axial coordination are explained similarly, as an effective variation in electronegativity of the metal resulting in charge donation into the a2u(n) orbital of the ring system“. Structural correlations. The frequencies of many of the vibrations of the porphyrin macrocyclic are linear functions of the center-to-pyrrole nitrogen distance, d, or the core size.3:’3t45 This dependency is described by the expression v = K(A-d) where K and A are parameters characteristic of the porphyrin macrocyclic.49 For a given vibrational mode, the K value, i.e. the slope of 76 the line, is indicative of the P.E.D. (potential energy distribution), and increases with the percentage of Cacm stretching character.32t50 Correlations of vibrational frequency to core size for a series of metal substituted porphyrins can be used to establish normal mode assignments. This approach has been used to assign vibrations of metallochlorophyll 3 complexes51 and is particularly useful when combined with vibrational data from isotopically substituted porphyrins.52 In the following, we use depolarization ratios and core size correlations from RR measurements to assign vibrational modes in MOEP+' species. Similar analysis has been applied to establish RR band assignments of iron octaethylchlorin complexes.53 Table 4-1 collects RR frequencies from 1300-1700 cm'1 obtained with Soret excitation of divalent metal OEP+° complexes with C104" counterions and juxtaposes them with known frequencies of the corresponding starting materials. Depolarization ratios measured with 363.8 nm excitation and wavelengths of the Soret band maxima are included. For MOEP+'C104' complexes with core sizes of 2.00 A or less (i.e., M = Ni, Co and Cu), our Soret RR results ( )ex = 363.8 nm) agree well with those of Kim _e_t_§t_l_.20 ”ex = 406.7 nm). This is because, for oxidation using AgClO4 or electrochemical techniques, diacid impurities occur only for complexes with core sizes larger than 2.00 A (i.e., M = Zn and Mg).21 Here we use RR frequencies for NiOEP+’CIO4" taken from Kim gt. 91.20; however, we employ alternative vibrational mode assignments consistent with those we have made for other MOEP“ species. Figure 4-11 shows plots of these vibrational frequencies as a function of core size, assuming that there is no appreciable change in d upon oxidation of the porphyrin ring. The core size measurements are those used by Spaulding _e_t_ _a_1_.45 Given the assumptions inherent in the plot (see below), Figure 4-11 indicates that the high frequency RR bands of the MOEP’” in the 1400-1700 cm‘1 range are metal .Ibn 00:93.2... cm genome» coma—33.0.: Iona voaaaaonco 0.53 3.04.2592. am < .23 a .«o '05:: unfit 6030393 I: 9an sum: $0.530- 0503 {moose—.05:— 5 sea“ 30:0... cemuantoaoaono . ON OUBULUHOG . SM" UUCOLUMUG‘ 77 A is awn Hon umm ~mm I 5" man man in is ooc pan ~mn ~mm «when I I an.m ppm Ah.ovp~m~ Am.cv~mw~ Am.ovuvw~ ummu code ch.wv cm.m Anovvmav Am.ovm~m~ Am.ovmnw~ Ap.ov>vw~ mmmu a «9.0 can Ac.ovocw_ Av.ovo~m~ Av.ovcsm~ «Nan puma Amm.pv mm.m Amwuvneu Am.ov~mm~ Av.°v~mm~ an.ovamm~ «cum 3 8.» 2" 8.882 8.882 3832 82 noun". Aoa.mv mm.a Apuuvpmn Am.ovmmm~ Am.°vmmm~ Aw.ovvpm~ when a I I om.n can Av.oV>pv. “v.9vbmvn Av.ovncm~ -m~ one Amm.mv mm.m vfimamvnmn Av.cvnm¢~ An.ov~on~ Am.oV-m~ mung a .- an.ovuva_ an.ovowm~ an.ovnmn~ cpnu zvo an.ovm>m~ An.cvmpn~ 0A«.ovm>n~ «and a 3; ExTBC . +265 1.596 .188 emcee a. name; .322 < a tween A5050 0 .wc~o.+.—m9‘ .3825: .8300 : ucmvcoaaouhou :93 van nuance—loo mug econ... you AIS algal! comanbouaz 1.03.5 van 0036: comacntuaaoa .Aullov gmocosaoun cal-E 00:93:3— eHIQ gnaw“. 78 I650 - VIBRATIONAL FREQUENCY u(cm") Figure 4-11 . |600 I550 l500 5 11+. .NiiDZd) Ni Co Cu Zn 3 l J l J l l— L 4 l j l J I l.940 |.960 I300 2.000 2.020 2.040 2.060 d Ct-N(Z\) Porphyrin core vibrational mode frequencies (for Raman allowed bands, 1450-17 00 cm'1)13_. porphyrin core size for the indicated MOEP complexes and their corresponding cation radicals, MOEP+'CIO4'. Mode assignments are accord- ing to reference 33a. Open symbols correspond to parent MOEP frequencies; filled symbols correspond to MOEP+'C104' frequencies. The regression analysis does not include the Ni (02d) points. Ct-N distances are from reference 45. 79 dependent and correlate to core size in the same manner as the analogous bands of the neutral parent species. The close agreement between the K values obtained for the neutral MOEP and the MOEP“ for each mode suggests that little change occurs in the normal mode composition upon oxidation of the porphyrin ring.54 Together with the relative agreement of the depolarization ratio measurements for modes of the MOEP and MOEP‘", these core size correlations provide support for our band assignments in the MOEP“ spectra. Further confirmation will be provided by RR measurements on MOEP“ incorporated with isotopic substitutions, which are in progress. Our results thus far suggest that, in general, normal mode calculations for the parent MP will remain useful for interpreting the vibrations of the MP+° species. Additional evidence of the correlation of vibrational frequencies of the porphyrin core stretching modes above 1400 cm'1 to core size is provided by IR studies of MOEP“ compounds.12 If we use complexes which presumably exhibit D4h solution geometries,55,56 i.e. CoInOEP+'2ClO4', FeIIIOEP+'2C104’ and FemOEP+'2CF3803', we can correlate their diagnostic IR band frequencies12 to the core sizes provided by Spaulding _et 11:45 Although the number of points is too few to establish a reliable correlation, we can estimate K and A from these data to be 420 cm’I/A and 5.67 A, respectively. They compare remarkably well to K and A values calculated for V38 from IR measurements of MOEP complexes obtained at 15K in argon matrices.52 While there is some controversy as to the normal mode composition of v33,32’33952 these IR frequencies yield K and A values of 421.3 cm'l/A and 5.69 A, and imply predominantly CaCm stretching character for this mode. For CoOEP, v33 is reportedi’2 at 1565 cm‘l, whereas the diagnostic IR band of ComOEP“'2ClO4' is reported at 1554 cm'l. The ~10 cm‘1 differences in these frequencies is in good agreement with the frequency decreases in CaCm stretching modes resulting from porphyrin 80 ring oxidation observed in our Raman work. This, and the agreement between K and A values, suggest that the diagnostic band of MOEP“ complexes corresponds to v 33 of the parent M OEP. In addition to RR frequencies, Table 4-1 also lists Soret maxima for MOEP and MOEP“ compounds. These values, along with a band energies for the parent MOEP, are plotted as a function of a core size in Figure 4-12. Absorption data for MgOEP (408 nm Soret, 580 nm a band) and MgOEP+'ClO4' (392 nm Soret) in CHZCIZ have been added. Our spectrum of the MgOEP+'ClO4" resembles the previous measurement by Dolphin _e_t_ al.4 The Soret maximum for MgOEP+°ClO4‘ used in Figure 4-12 represents the lowest energy transition in the near uv envelope of the absorption spectrum. The qualitative similarity of the trends for all transition energies is striking and can be explained by the dominance of the a2u(1r)+ eg(1r*) character of the transitions in determining the core size dependencies of these values. For the neutral MOEP the aguUT) orbital energy may be considered to vary linearly with core size while the eg(TT*) and a1u(1T) orbitals remain relatively constant.48’57‘59 Thus, the a2u( 1' ) +eg(" *) character of the excited state determines the core size dependency of the and Soret band energies. Calculations by Edwards and Zerner23 predict that, in the case of the 2A2u cobaltic porphin radical, the dominant transition in the Soret region has 3096 agu(")+eg(1T *) character, while for the 2A111 cobaltic species the least energetic transition in the Soret envelope has 70% a2u( 1r)->eg(1r"') character. Thus the correlations in Figure 4-12 suggests that the a2u( 1r)+eg(1r*) character of these transitions is responsible for the core size dependency of the Soret band energies in both the MOEP and MOEP‘" compounds. Porphyrin core geometry. The key assumption made in establishing the parameters K and A in the relation v = K(A-d) for the MOEP‘" series is the 81 27 O O Q 26 _\ I E 25 ). 0 A 0: n LIJ 2 Lu l9 - Z - 9 l8 *- [_— <75 L 0 Z o (X <12 I7 *- E L- Ni Co Cu Zn Mg '6 1 I L L 4 I 1 l 4 J L I l l.960 l.980 2.000 2.020 2.040 2.060 0 Ct-N (A) Figure 4-12 . Electronic transition energies vs. porphyrin core size for the indicated MOEP and MOEP-F C104" complexes. Open and filled symbols correspond to absorption band maxima of MOEP and MOEP+‘CIO4", respectively. 82 invariability of the core size (d) upon oxidation of the parent MOEP. Table 4-2 summarizes the relevant parameters which describe the four X-ray crystal structures of metalloporphyrin 1r cation radicals determined to date, and compares them to the structures of the analogous parent compounds. The table also compares perchlorate ligated ferric TPP and OEP as a control to establish possible steric effects limited to TPP complexes. In each case the agreement between the structures of MP and MP“ is good in terms of the Ct-N distances and the metal displacements from the plane of the pyrrole nitrogens. However, the porphyrin core structures of MgTPP’" and ZnTPP‘“ display larger deviations from planarity than do the model parent structures. This is not considered to represent a significant general structural deviation between MP and MP4” and is explained rather as a result of steric interaction of the bulky C104' counter ions and the mesophenyl groups resulting from crystal packing forces.l3 Two other examples support this idea: 1) While crystalline CuTPP+'SbC16‘ exhibits "unusually large ruffling of the porphyrin core", the CHZCIZ solution structure is planar;55 2) Crystalline Fem(ClO4')TPP exhibits core ruffling almost as large as the perchlorate ligated MTPP‘“ examples, whereas Fem(C104")OEP has a planar core (see Table 4-2). Thus, we conclude that CH2012 solution structures of the type MOEP+'CIO4' do not exhibit significant ruffling of the porphyrin core, and the lines described by Figure 4-11 and 4-12 are useful references, uncomplicated by the effects of macrocyclic distortions. The success of the correlation attests to this and suggests that in general we can assume that in solution the Ct-N distance of the MP‘“ and probably other aspects of core geometry are comparable to those of the parent MP. Thus, it is assumed that no significant changes in geometric structure of. the porphyrin ring accompany the abstraction of an electron from the a2u(1r) or a1u(1r) orbital of the metalloporphyrin in solution. .Amv .muu oonv uu aucmaumv 02a umam;xo ca luau msmuhzauon Lumaoof .o:~o> bag“ nag“ «nag on flaw: oucduumv Ziao on» ondna o~u>oouool 0:0 loam vouoaanmv am so 0:» Mn .OOGCUOMU Zia-Dv .uousuuauun ovw~ag u x auaommu>a loam ouczo>o ca nanououaou nwsbu .000:¢auwv ZIUE IOLM UUCfldOMU Zlao OHIAQUHCO 0H LUVHO Cm Uvflfiflflfl um USMC) Imfih a .amuhgauoauhvwzhnnuuou «human n.o« m.o mam." mm aaaxwozuvuuuoa vo.on m~.o sea." no emoxwozovuuuou gauged . mam." mm mHopm guano ~mm.~ em unease suzaaeeeag o e u .+ . - . a e a o «no a. a wee N am aeauxomuvuuuou sueeaz g o nee N no .+eaa Aiozovuuaou v_.oa _m.o _oo.« we aesx xv on gnomes. m.o “a.“ no m~onm amaa ~00 on u - HH~ eHeeuag e i .+ i a“. e_.ou an.o eeo.~ Hm easapasaceu e.ou pen.e eeo.~ m2 .+aaexwo~oveu a np~.o emo.~ ow amp.o~=.uz e.o» ne.o amo.~ a” .+maaxwo~ovuz fi¢ Adv Adv a(u EM Afiu ozeeau .eum a: ofieesm :Teo 2:50 ..em a: OLOU Ziuo ziuo ouoo .+ .nvczoaloo vououvm _vco uflcumvam c0300 I ambasaLaOAAqux .«o nonaaoauam 13280 5.5 aLOuOfluuum ouwm 0.30 .NIQ 0H£GH 84 Ring buckling effects. Distortions of the porphyrin core are known to weaken the bonding in the macrocyclic, thus lowering the frequencies of stretching vibrations of the core and causing negative deviations from the relation v = K(A-d).67 The classic example is shown in Figure 4-11 by the vibrational frequencies of the ng form of NiOEP,45 which displays deviations from planarity in the macrocyclic of :I: 0.5 A.“ Spiro gt _a_l_.67 conclude that the effects of core distortions on vibrational frequencies, although significant, are minor in comparison to changes in Ct-N distance, demonstrating that core size is the main determinant of the vibrational frequencies of the porphyrin macrocycle. Our results indicate that the relationship of core size to vibrational frequency is similar for the oxidized porphyrin ring. It is expected that the effects of core distortions, such as S4 ruffling, will be manifest also in the vibrational frequencies of the MP‘“, producing negative deviations from the correlations of Figure 4-11. This completes our structural analysis of the one-electron, ring—oxidized MHOEP+'CIO4' system. Before continuing, it is useful to summarize our conclusions thus far. If one assumes macrocyclic planarity of the CH2C12 solution structures of the MOEP series considered here, the close correspondence of the K values in the expression describing the porphyrin core vibrational frequencies v = K(A-d) for both the MOEP and MOEP“ implies: 1) the corrections of our mode assignments, 2) the essential invariance in the P.E.D. of the normal modes of the porphyrin ring upon oxidation, and 3) the absence of core ruffling in oxidized structures of type MIIOEP+'CIO4‘ (for M = Ni, Co, Cu and Zn). The key assumption that the porphyrin core geometry (especially Ct-N distance) remains intact upon oxidation is based on comparison of crystal structures of MP45’60‘66 and MP+'13’55956 species and has already been suggested.13’55 The success of the correlation of vibrational frequencies to 85 core size presented here suggests that the assumption is valid for the MOEP+'CIO4' species. Since they seem independent of 2A2” y_s_. 2A1u electronic state character, these linear correlations establish a reference point useful to structurally based vibrational analysis of all MP’" species. As an example, we predict structures for the two-electron oxidized cobaltic systems, CoIIIOEP+'2X'. It is anticipated that a similar approach can be applied to both ferric and oxoferryl porphyrin 1r cation radical models and ultimately to compound I type intermediates of heme enzymes. Cobaltic octaethylporphyrin I cation radicals. The CoIIIOEP+'2X' system is particularly interesting owing to the variety of spectrally distinct two-electron oxidation products possible as a function of counterions X'. Characterization of the effects of different counterions in terms of the extent of axial ligation and the relationship of axial ligation to the electronic ground and excited states is essential to understanding the rich variety of absorption spectra displayed by this system upon variation of X".69 In order to compare the electronic spectra of ComOEP+'2X’ to the parent system, ConOEP, it is necessary to determine how oxidation of the metal and addition of axial ligands influence the electronic states of the OEP“ system. For example, the agreement between the Soret maxima of CoHOEP (391 nm) and its two-electron oxidation product ComOEP+'ZClO4’ (393 nm), would be misleading without acknowledging the counterbalancing effects present. Recognition of the blue-shifted Soret maximum of the cobaltous species CoHOEP+'C104' (376 nm) isolates the effect of ring oxidation upon the Soret absorption maximum, and reveals a blue-shifted band of diminished intensity compared to the parent CoOEP. Axial ligation by the C104“ ion is unlikely in the cobaltous OEP+' because of the presence of an electron in the metal dzz orbital. Thus, we consider this and the other MHOEP+'C104' complexes discussed here to contain four-coordinate metal 86 centers, devoid of the effects of metal oxidation and axial ligation. Empirically we see that oxidation of the ring causes a 1000 cm'1 blue shift in the Soret maximum with reSpect to the parent CoOEP. This is compensated for in CoInOEP+’2ClO4' by a 1150 cm-1 red-shift produced by oxidation at the metal center accompanied by weak axial ligation by the C104‘ ions. Compared to the cobaltous case noted above, weak ligation of the cobaltic center by both ClO4" ions in the latter complex is more likely because the d6 configuration of CoIII leaves the dzg orbital empty. Thus, we consider CoIIIOEP+'2ClO4' to represent a six-coordinate complex, albeit with weak axial ligands. This example illustrates the importance of recognizing the combined effects of metal oxidation and axial ligation on the electronic states of the OEP’" systems in order to interpret the absorption spectra correctly. Effects of one-electron, metal centered oxidation and axial ligation. In order to assess these effects on the electronic transitions of the OEP’“ system, a good starting point is to consider the effects of metal oxidation and ligation in the parent OEP system. For this analysis we utilize the ring-neutral Com(MeOH)20EPX' as an example of a weakly ligated, six-coordinate cobaltic complex. Comparison of the spectra of this species to those of CoOEP will establish the effects of ligation and oxidation of the metal. These effects will be compared to the effects of axial ligation alone on Cu and NiOEP. Shelnutt _e_t_ 11.53 have shown that the red-shifts in the a and Soret bands of Cu and Ni porphyrins upon axial ligation are accompanied by expansion of the porphyrin ring as revealed by decreases in RR frequencies of the core size sensitive modes. The core expansion accompanying axial ligation in Ni porphyrins, confirmed by similar RR measurements by Kim gt a_l.69, is much greater than in Cu porphyrins and involves a change in spin state and subsequent occupancy of the dx2_y2 orbital of the Ni.70 These spectroscopic results are explained 87 by a relative increase in the porphyrin 8211M) orbital energy level caused by the interaction of the metal atom with its axial ligands, and are thus in agreement with the model for axial ligation proposed by Gouterman.48 Table 4-3 collects RR frequencies and absorption maxima for Cu, Ni and CoOEP and their axially ligated systems. We emphasize that with these examples we consider the effects of ligation alone for Cu and N iOEP, and ligation accompanied by metal oxidation for CoOEP. Inspection of Table 4-3 reveals that the red-shifts in the optical absorptions of Com(L)20EPX‘ (where L = methanol or l-methylimidazole) with respect to those of CoOEP are similar to those encountered upon ligation of CuOEP or NiOEP. Indeed, the magnitude of the Soret absorption red-shift for the cobalt system (1180 cm'l) is between that of the copper (970 cm'l) and nickel (1800 cm’l) systems. However, the RR data clearly show that axial ligation and autoxidation of CoOEP is not accompanied by core expansion, as there is no decrease in the vibrational frequencies of Com(L)20EPX" compared to CoOEP. The slight increases in v 4 frequencies reflect a decrease in the eg( 1r*) orbital occupancy upon oxidation of CoII to CoIII and are not related to core size. (The Cu and NiOEP systems show the opposite behavior and exhibit decreases in the v4 frequencies). RR wavenumbers reflect the structure of the ground state of the scattering species. Thus, red-shifts in the absorption spectra without concomitant decreases in RR frequencies suggest stabilization of the excited state, i.e. a decrease in the cg” *) rather than an increase in the a2u(1r) orbital energies upon axial ligation and metal oxidation of these species. These considerations indicate that the spectral red-shifts observed upon metal oxidation and ligation of CoOEP are not simply comparable to those caused by ligation alone of Cu and NiOEP. Whether these conclusions apply to ligation alone of cobaltous porphyrins under anaerobic conditions cannot be determined from this analysis. 88 bad -v mew amm awe Ham wfiw man .55 «Bow .ozoaoaeifiosl n :: uca 652..an u Ea 6523.53 n :3 “com: mco33>oenn< .33. do: $53330 ..8 3:259: m9 0 mmm mmm cmm Ham Hmm cum «mm .55 coma wwwfi new“ bvwa mHmH comm mama bmmfi So mmmw mama mam“ mam“ puma “cm“ mono—oEoo anon—co v5 =5 .==o 8:35 382 ..8 2.68: 5598.2 3830 v.3 memo—.2595 .353— anaconda mumfi wmmfi mama pmmfi mung mama mum“ $3 Nam" cmmfi mama Nana mom~ an: fig: Ami—=3 Egg «ivm:ban. dc 98 can moocoeouom a wfimfi mama mama «flag ~m¢~ mHmH oncmfi mama a: awn“ Nana awn“ mung cpmfi mama m>m~ m>m~ v: . mam. cocoaomom m txmmONAECEoO Liam—02:32 :aoo -eoaamomEOoz :zoo amozoo oomofioazz ofiommoamez ammotzavzo wmcaxofimmoqfiv 89 COmOEP+'2ClO4‘. Having established the combined effects of metal oxidation and weak axial ligation on the electronic transitions and RR vibrations of ConOEP, we can consider like effects in the analogous 1r cation radical system, CoIIOEP+'. Comparison of Soret maxima and RR frequencies (Fig. 4-7c and d) of CoIIOEP+'ClO4' to those of ComOEP+'ZClO4‘ reveals Optical red-shifts without porphyrin cation radical core expansion in the latter compound. The spectral differences between these two species are analogous to those observed when comparing CoHOEP to Com(MeOH)20EPC104'. In the former ( 1r cation) case the Soret maximum shifts 1150 cm‘l, or from 376 to 393 nm; while in the latter (ring-neutral) case, the shift is from 391 to 409 nm or 1180 cm‘l. In both cases, RR results imply no change in the porphyrin core size. The similarity of these examples suggests that in both instances we are comparing a four-coordinate cobaltous species to a six-coordinate cobaltic one and that ring planarity and D4}, symmetry are maintained throughout. Furthermore, the absorption spectra imply that the behavior of the Soret absorption in these TI’ cations is little changed from that of the ring-neutral analogues. As discussed above, a key assumption in our analysis concerns the effects of ring oxidation & s3 on macrocycle geometry. Based on X-ray studies of crystaline compounds,13 we assume that in general the core geometries of porphyrin 1r cations, MPJ", are not significantly different from those of the parent compound MP. By "parent" we specifically refer to a complex as nearly identical as possible to the MP‘“ in all aspects other than macrocyclic oxidation state. That is, the oxidation state and coordination geometry of the central metal must be the same. For example, the "parent" compound for ConOEP+'C104" is CoHOEP, and the "parent" for CoIH(CIO4-)20EP+° is Com(MeOH)20EPClO4'. (The cobaltic Ti cation radical complex is now written in a manner as to imply axial ligation by the perchlorate anions. Attempts 90 to exchange the perchlorate ligands for methanol result in reduction of the macrocycle and formation of the cobaltic dimethanol ring-neutral complex.) Thus, our analysis suggests that the core sizes of all four of these complexes are the same, approximately 1.97 A. COmOEP“ZBr'. To date, EPR and ENDOR measurements are the most reliable criteria to establish the MP“ electronic ground state (2A2u or 2A1“).8 EPR measurements of ComOEP“2ClO4' and ComOEP“ZBr' are consistent with the 2A2u and 2A1u states, respectively.“ Perhaps the most interesting aspect of the correlations presented in Figure 4-11 and 4-12 is the apparent insensitivity of the RR frequencies and Soret transition energies of the MIIOBP“CIO4' complexes to radical electronic state. This implies that we can use these correlations for structural analysis of OEP“ compounds of either radical type, and we can extend our discussion to include ComOEP“ZBr".71 Aside from differences in relative intensities, the RR spectra of ComOEP“2ClO4‘ and ComOEP“2Br' are similar (see Fig. 4-7d and g; Fig. 4-9). Table 4-4 compares the RR frequencies of these two complexes. The vibrational frequencies of the latter (presumably a 2A1u complex), however, are 6-7 cm‘1 lower than those of the former (presumably a 2A2u complex), with the as yet unexplained exception of v10. Because the core size correlations seem independent of radical electronic state, and because the frequency differences between these two complexes are small, we speculate that they can be attributed to differences in porphyrin core geometry caused by the different axial ligation of these two species. Decreases in frequency of the macrocyclic vibrations (above 1400 cm‘l) may result from either expansion or buckling of the porphrin core. It is reasonable to assume that ComOEP“ZBr’ is a six-coordinate complex with a planar porphyrin core. The v3, v11, v2 and v10 frequencies (Table 4-4) can then be used, along with R and A values 91 Table H Comparison of Resonance Raman Frequencies (cm-1) and of Depolarization Ratios for Cobaltic OEP“ Complexes COmOEP+'ZClO4" ComOEP+'2Br‘ v4 1361 (0.2) ? 03 1505 (0.5) 1498 (0.3) v11 1605 (0.7) 1598 (0.5) 92 1617 (0.5) 1611 (0.4) 910 1642 (0.7) 1649 (0.7) 92 from Table 4-1, to predict a core size of 1.99 :i: 0.02 A for this complex. This is to be compared to 1.974 :t 0.008 A predicted similarly for the core size ComOEP“ZClO4". However, the possibility that the vibrational frequencies of CoInOEP+’ZBr' are lowered from those of ComOEP“2ClO4‘ as a result of porphyrin core distortions, rather than core expansion, cannot be ruled out by this analysis.72 CONCLUSIONS While the interpretation of the spectra described in this paper essentially ignores radical designation as 2A2u or 2A1“, other researchers have emphasized the species symmetry of MP“ compounds. Characterization of the MP“ as 2A2u or 2A1”, as revealed by spin density profiles, has been based primarily on MO calculations,47a:73'75 as well as EPR and ENDOR studies.498’249473973t75'77 These studies have established many useful generalizations concerning radical type. For instance, the porphyrin ring substituents have a profound influence, and structures of the type MTAP“ (where A is an alkyl or aryl group in the meso position) exhibit 2A2u characteristics,8 while MHP“ (where HP=hydroporphyrin) tend to be classified as 2A1u radicals.73,76t77 Structures of the type MOAP“ (octaalkylporphyrin where the Cb substituents are not necessarily identical) may exhibit either state depending primarily on the metal center and the axial ligands.8’24t73 Examples of the latter type include the high valent catalytic intermediates of heme peroxidases and catalases. The compound I structures of these enzymes are usually decribed as oxoferryl protOporphyrin IX 1; cation radicals, O=Fe1V(X‘)PP“.2 The electronic state is considered to be mediated by the identity of the sixth ligand X', and MO calculations predict that ligation by tyrosine or thiolates promotes the 2A1u configuration (e.g. CAT-I and CPO-I, 93 respectively)73’76 while imidazole ligation generally favors a predominantly 2A2" state (e.g. HRP-0.73,” EPR and ENDOR studies, however, are interpreted to suggest a 2A2u configuration for both HRP—I and CPO-~I.78 The EPR measurement of CPO-I (chloroperoxidase compound I) was judged inconsistent with a 2A1u configuration owing to the relatively strong antiferromagnetic coupling (between the 8:1 oxoferryl structure and the S=l/2 1f cation radical) suggested by the spectra. The authors reason that such coupling is more likely in the 2A2u configuration where relatively large spin densities are placed on the pyrrole nitrogens.73 Distortions of the porphyrin ring may provide an orbital overlap pathway for the strong I J |=250 cm'l, antiferromagnetic coupling displayed by FeHI(Cl’)TPP“SbC16'.55’56 While such magnetic coupling was previously thought to occur only in 2A2u radicals,8 more recent interpretations76 suggest magnetic coupling between the paramagnetic metal center and porphyrin 1r cation radical can occur in either radical type, with the radical type influencing the magnitude of the interaction.553 While the weak coupling, |J|=2 cm‘l, displayed by HRP-I78 has been attributed to possibly a planar configuration of the porphyrin ring,56 the stronger, IJ [=37 cm'l, coupling for CPO-I78 may reflect a more buckled configuration (other explanations are also possible11176’78979). Recent work clearly indicates that evaluation of porphyrin core geometry must be addressed in order to interpret results from magnetic based techniques. The similarity of the vibrational frequencies in the 1400-1700 cm"1 range of the pr0posed 2A2u and 2A1u 1r cation radicals ComOEP“2ClO4' and ComOEP“2Br', respectively, suggest that these frequencies are insensitive to radical type. There are, however, small wavenumber differences which can be attributed to relatively minor differences in porphyrin core geometry of these two compounds. Because these geometric differences are as yet 94 speculative, other interpretations are plausible. Although these vibrations of the porphyrin macrocyclic may well be insensitive to radical electronic state, the abnormally high values of the v (Cbe) modes provide a clear basis to identify ring centered oxidations, particularly in MOEP“ species.22 Both the frequency and relative intensity of the v4 mode (1340-1370 cm'l) may reflect 2A2u y_s_. 2A1u character. More information concerning electronic states will be provided by RR excitation profiles. Our intent in this work is to establish an approach to evaluate core geometry of oxoferryl porphyrin 1r cation radicals in enzymes in order to complement the NMR, EPR, and ENDOR results for these sytems. ACKNOWLEDGMENTS T.O. thanks Dwight Lillie. and Bob Kean for their computer programming and Harold Fonda for discussion. This work was supported by NIH grants GM-25480 (G.T.B.) and GM-36520 (C.K.C.), and NSF grant CHE-86-10421 (G.E.L.). 10. ll. 12. REFERENCES AND NOTES (a) Frew, J.E., Jones, P. in "Advances in Inorganic and Bioinorganic Mechanism"; Academic Press: New York, 1984; Vol. 3, pp. 175-215; (b) Hewson, W.D.; Hager, L.P. in "The Porphyrins"; Dolphin D., Ed.; Academic Press: New York, 1979; Vol. 4, pp. 295-332. (a) Norris, J.R., Sheer, R.; Katz, J.J. in "The Porphyrins"; Dolphin, D., Ed.; Academic Press; New York, 1979; Vol. 4, pp. 159-195; (b) Lubitz, W.; Lendziar, F.; Plota, M.; Mobins, K.; Trinkle, E., Springer Series in Chem. Phys. 1985, 42, 164-173. Fuhrhop, J.H., Struct. Bonding 1974, _1_8_, 1-67. Dolphin, D.; Mulijinani, Z.; Rousseau, K.; Borg, D.C.; Fajer, J.; Felton, R.H., Ann. N.Y. Acad. Sci. 1973, 206, 177-200. Wolberg, A.; Manassen, J., J. Am. Chem. Soc. 1970, 9_2_, 2982-2991. Phillippi, M.A.; Goff, H.M., J. Am. Chem. Soc. 1982, 104, 6026-6034. Carnieri, N.; Harriman, A., Inorg. Chim. Acta 1982, §_2_, 103-107. (a) Fajer, J.; Davis, M.S. in "The Porphyrins"; Dolphin D., Ed.; Academic Press: New York, 1979, Vol. 4, pp. 197-256, (b) Fajer, J.; Borg, D.C.; Forman, A.; Felton, R.H.; Vegh, L.; Dolphin, D., Ann. N.Y. Acad. Sci. 1973, 399, 349-364. (a) Browett, W.R.; Stillman, M.J., Inorg. Chim. Acta 1981, 49, 69-77; (b) Browett, W.R.; Stillman, M.J., Biochim. Biophys. Acta 1981, 660, 1-7. (a) Godziela, C.M.; Goff, H.M., J. Am. Chem. Soc. 1986, _1_0_8_, 2237-2243; (b) Goff, B.M.; Phillippi, M.A., J. Am. Chem. Soc. 1933, 13, 7567-7571. Morishima, 1.; Takamuki, Y.; Shiro, Y., J. Am. Chem. Soc. 1984, 106, 7666-7672. Shimomura, E.T.; Phillippi, M.A.; Goff, H.M., J. Am. Chem. Soc. 1983, 103, 6778-6780. 95 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 96 (a) Barkigia, K.M.; Spaulding, L.D.; Fajer, J., Inorg. Chem. 1983, 22, 349-351; (b)Spaulding, L.D.; Eller, P.G.; Bertrand, J.A.; Felton, R.H., J. Am. Chem. Soc. 1974, g, 982-987. Nadezhdin, A.D.; Dunford, H.B., Photochem. Photobiol. 1979, _2_9_, 899-903. (a) Teraoka, J.; Ogura, T.; Kitagawa, T., J. Am. Chem. Soc. 1982, 104, 7354-7356; (b) Van Wort, H.B.; Zimmer, J., J. Am. Chem. Soc. 1985, 107, 3379-3381. (a) Oertling, W.A.; Babcock, G.T., J. Am. Chem. Soc. 1985, 107, 6406-6407; (b) Ogura, T.; Kitagawa, T., submitted. (a) Cotton, T.M.; Parks, K.D.; Van Duyne, R.P., J. Am. Chem. Soc. 1980, _1_(2, 6399-6407; (b)Lutz, M. in "Advances in Infrared and Raman Spectroscopy"; Clark, R.J.H., Ed.; Wiley, New York 1984, Vol. XI, pp. 211-300. Yamaguchi, H., Nakano, M.; Itoh, K., Chem. Lett. 1982, 1397-1400. (a) Chottard, G.; Battioni, P.; Battioni, J.-P.; Lange, M.; Mansuy, D., Inorg. Chem. 1981, _2_0, 1718-1722; (b) Aramaki, S.; Hamaguchi, H.; Tasumi, M., Chem. Phys. Lett. 1983, g, 555-559. Kim, D.; Miller, L.A.; Rakhit, G.; Sprio, T.G., J. Phys. Chem. 1986, 20, 3320-3325. Oertling, W.A.; Salehi, A.; Chang, C.K.; Babcock, G.T., J. Phys. Chem., submitted. Salehi, A.; Oertling, W.A.; Babcock, G.T.; Chang, C.K., J. Am. Chem. Soc. 1986, 108, 5630-5631. Edwards, W.D.; Zerner, M.C., Can. J. Chem. 1985, 6_3_, 1763-1772. Dolphin, D.; Forman, D.C.; Borg, J.; Felton, R.H., Proc. Natl. Acad. Sci. USA 1971,68, 614-618. Wang, C.B.; Chang, C.K., Synthesis 1979, 548-549. Falk, J.E. in "Porphyrins and Metalloporphyrins"; Elsevier: New York, 1964, 798. . Alternative structural assignments for these compounds, particularly Com(MeOH)20EPBr‘, are possible. A mixed diaxial species of the type (Br‘)Com(MeOH)OEP was prOposed by Johnson, A.W.; Kay, I.T., J. Chem. Soc. 1960, 2979-2983 and may better explain the similar, yet distinct absorption maxima of Com(MeOH)20EPClO4' and Com(MeOH)20EPBr‘ (see Table 4-1 of the Discussion section). Further addition of MeOH to the latter comfiound produces a sample with identical absorption maxima to that of CoI (MeOH)20EPClO4'. For this work, however, a distinction between dimethanol ligation YE.- mixed axial ligation of these complexes is not necessary. See also: Sugimoto, H.; Ueda, N.; Mori, M., Bull. Chem. Soc. Jpn. 1981, 5_4, 3425-3432. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 97 Under certain conditions, however, molecular oxygen (02) was observed to promote oxidation of metalloporphyrins. For instance, in the presence of excess (“01 mM) tertbutyl ammonium perchlorate, oxidation of CoOEP to ComOEP“2ClO4' was observed, presumably via the formation of HClO4 in CH2012 solution, as discussed by Fukuzumi, S.; Ishikawa, K.; Tanaka, T., Chem. Lett. 1986, 1-4. Upon laser irradiation in Soret resonance at 406.7 nm or 390 nm, CHzClg solutions of CuOEP, ZnOEP, CoOEP and ConOEP“ClO4' in the presence of 02 were observed to undergo one-electron oxidation requiring various exposure times for each species. These effects were not observed with 363.8 nm excitation. Similar photochemical oxidation of CoTPP has been discussed by Gasyna, Z.; Browett, W.R.; Stillman, M.J., Inorg. Chem. 1984, 2_3_, 382-384; and by Yamamoto, K.; Hoshino, M.; Kohno, M.; Ohya-Nishiguchi, H., Bull. Chem. Soc. Jpn. 1986, _5_9, 351-354. Both MgOEP and MgOEP“Br' were difficult to work with for RR experiments; many spectra of these- compounds were collected at the outset of our study, but they are not heavily relied on in this work. Fuhrhop, J.H.; Mauzerall, D., J. Am. Chem. Soc. 1969, 91, 4174-4181. Konishi, S.; Hoshino, M.; Imamura, M., J. Am. Chem. Soc., 1982, 104, 2057-2059. Fuhrhop, J.H.; Wasser, P.; Riesner, D.; Mauzerall, D., J. Am. Chem. Soc. 1972, 9_4_, 7996-8001. Spiro, T.G. in "Iron Porphyrins'" Lever, A.B.P.; Gray, H.B., Eds., Addison-Wesley: Reading, M.A. 1983; Part 2, pp. 89-159. (a) Abe, M.; Kitagawa, T.; Kyogoku, Y., J. Chem. Phys. 1973, 99, 4526-4534; (b)Gladkov, L.L.; Solovyov, K.N., Spectrochim. Acta 1986, Vol. 42A, 1-100 ’Boldt, N.J.; Donohoe, R.J.; Birge, R.R. and Bocian, D.F., J. Am. Chem. Soc. 1987, in press. (a) Brunner, H.; Mayer, A.; Sussner, H., J. Mol. Biol. 1972, 10, 153; (b) Yamamoto, T.; Palmer, G.; Gill, D.; Salmeen, I.T.; Rimai, L., J. Biol. Chem. 1973,2481 5211-5213. Corwin, A.H.; Chivvis, A.B.; Poor, R.W.; Whitten, D.G.; Baker, W.E., J. Am. Chem. Soc. 1968, _9_0_, 6577-6583. (a) Whitten, D.G.; Baker, B.W.; Corwin, A.H., J. Org. Chem. 1963, 3, 2363-2368; (b) Tsutsui, M.; Valapoldi, R.A.; Hoffman, L.; Suzuki, K.; Ferrari, A., J. Am. Chem. Soc. 1969, 9_1, 3337-3341. We emphasize that the presence of trace amounts of diacid salts in preparations of MP“ in CH2C12 is not always evident from absorption spectra. For example no diacid contributions to the absorption spectrum in Figure 4-5c of CohIOEP“2ClO4' are apparent, yet RR excitation at 406.7 nm reveals their presence as shown by Figure 4-3b. Likewise, RR measurements at 406.7 nm of ComOEP“2Br' preparations invariably reflect the presence of porphyrin acid impurities. Earlier, the blue-shifted 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 98 Soret absorption of meta110porphyrin dications allowed the detection of porphyrin diacid impurities from uv-visible absorption spectra of these samples (see ref. 47a). It was not until the application of the RR technique, however, that the presence of the diacid was recognized in metalloporphyrin 1r cation radicals (ref. 21). Spiro, T.G.; Strekas, T.C., J. Am. Chem. Soc. 1974, 96, 338-345. Choi, S.; Sprio, T.G.; Langry, K.C.; Smith, K.M.; Budd, D.L.; LeMar, G.N., J. Am. Chem. Soc. 1982, 104, 4345-4351. (a) Stewart, B.; Clark, R.J. H., Struc. Bonding (Berlin) 1979, 3__6, 1- 80; (b) Rousseau, D. L.; Friedman, J. M.; Williams, P. F, Topics Curr. Phys. 1978, _1_1_, 202- 252. Shelnutt, J.A., J. Chem. Phys. 1981, 74, 6644-6657. Spiro, T.G.; Strekas, T.C., Proc. Natl. Acad. Sci. USA 1972, _6_9_, 2622-2626. Verma, A.L.; Mendelsohn, R.; Bernstein, H.J., J. Chem. Phys. 1974, 61, 383-390. Spaulding, L.D.; Chang, C.C.; Yu, N.-T.; Felton, R.H., J. Am. Chem. Soc. 1975, _9_7_, 2517-2524. Johnson, E.C.; Niem, T.; Dolphin, D., Can. J. Chem. 1978, 56, 1381-1388. (a) Fajer, J.; Borg, D. C.; Forman, A.; Dolphin, D.; Felton, R. H., J. Am. Chem. Soc. 1970, _9__2, 3451- 3459; (b)Mengersen, C.; Subrammanian, J.; Fuhrhop, J. H, Mol. Phys. 1976, 3_2, 893- 897. (a) Gouterman, M., J. Chem. Phys. 1959, 30, 1139-1161; (b) Gouterman, M., J. Mol. Spec. 1961, 6, 138-163. Huong, P.V.; Pommier, J.C., C.R. Acad. Sci., Ser. C. 1977, 285, 519-522. (a) Warshel, A., Rev. Biophys. Bioeng. 1977, 6, 273; (b) Callahan, P.M.; Babcock, G.T., Biochemistry 1981, _2_0, 952-958. (a) Fu31wara, M.; Tasumi, M. J. Phys. Chem. 1986, _9_0, 5646- 5650; (b) Fonda, H. N.; Babcock, G. T., Proc. 7th Int]. Cong. Photosyn. 1986, in press. Kincaid, J.R.; Urban, M.W.; Watanabe, T.; Nakamoto, K., J. Phys. Chem. 1986, 90, 5646-5650. Ozaki, Y.; Iriyama, K.; Ogoshi, H.; Ochiai, T.; Kitagawa, T., J. Phys. Chem. 1986, 10, 6105-6112. The decrease in K value upon oxidation for v10 is not considered +significant owing to the uncertainty in the v10 measurement of ZnOEP+ "C104 by Soret Excitation RR. Visible excitation of this compound at 647.1 nm, which should more clearly show the vm band, was not possible owing to intense fluorescence emission centered around 675 nm. Thus it is possible that the relatively high 910 value measured for ZnOEP“ CIO4‘ anomalously 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 99 lowers the K value (for 910) of the MOEP“ series. (a) Scholz, W.F.; Reed, C.A.; Lee, Y.J.; Scheidt, W.R.; Lang, G., J. Am. Chem. Soc. 1982, iii, 6791-6793; (b) Buisson, G.; Deronzier, A.; Duée, E.; Gans, P.; Marchon, J.-C.; Regnard, J.-R., J. Am. Chem. Soc. 1982, 10_4_, 6793-6796. Gans, P.; Buisson, G.; Duee, E.; Marchon, J.-C.; Erler, B.S.; Scholz, W.F.; Reed, C.A., J. Am. Chem. Soc. 1986, 108, 1223-1234. (a) Kitagawa, T.; Ogoshi, H.; Watanabe, E. Yoshida, Z., J. Phys. Chem. 1975, _7_9_, 2629-2635; (b) Shelnutt, J.A. ; Ondrias, M.R., Inorg. Chem. 1994, _23, 1175-1177. (a) Shelnutt, J.A.; Straub, K.D.; Rentzepis, P.M.; Gouterman, M.; Davidson, E.R., Biochemistry 1984, _2_3, 3946-3954; (b) Shelnutt, J.A.; Alston, K.; Ho, J.-Y.; Yu, N.-T.; Yamamoto, T.; Rifkind, J.M., Biochemistry 1986, _2_5, 620-628. the model presented here is an oversimplification. Similar trends in the eg(1r *) orbital energy levels (LUMO) are revealed by reduction potentials for MOEP complexes. Comparative measurements of the first oxidation potential reflect trends in the HOMO energies. The results show that both the HOMO and the' LUMO rise in energy in the series Ni Cu, Zn, MgOEP; however, the HOMO levels increase more, thus producing the spectral red-shift. See ref. 29 and Fuhrhop, J.-H.; Kadish, K.M.; Davis, D.G, J. Am. Chem. Soc. 1973, §_5_, 5140-5147. Timkovich, R.; Tulinsky, A., J. Am. Chem. Soc. 1969, 11, 4430-4432. Collins, D.M.; Hoard, J.L., J. Am. Chem. Soc. 1968, 92, 3761-3771. Mitra, S. in "Iron Porphyrins"; Lever, A.B.P.; Gray, H.B., Eds.; Addison-Wesley: Reading, Mass, 1982, Part II, pp. 3-42. Scheidt, W.R.; Cohen, I.A.; Kastner, M.R., Biochemistry 1979, 18, 3546-3552. Fleischer, E.B.; Miller, C.; Webb, L., J. Am. Chem. Soc. 1964, _8_6_, 2342-2343. Masuda, H.; Taga, T.; Osaki, K.; Sugimoto, H.; Yoshida, A.-I.; Ogoshi, H., Inorg. Chem. 1980, 950-955. Reed, C.A.; Mashiko, T.; Bentley, S.P.; Kastner, M.E.; Sheidt, W.R.; Spartalian, K.; Lang, G., J. Am. Chem. Soc. 1979, 101, 2948-2958. Spiro, T.G.; Strong, J.D.; Stein, P., J. Am. Chem. Soc. 1979, 101, 2648-2655. Meyer, E.F., Acta Cryst. 1972, 828, 2162-2167. Setsume, J.; Ikeda, M.; Kishimoto, Y.; Kitao, T., J. Am. Chem. Soc. 1986, 108, 1309-1311. 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. 100 Kim, D.; Su, Y.O.; Spiro, T.G., Inorg. Chem. 1986, 2_5, 3988-3993. Actually the question of radical ground state for these compounds is not completely clear-cut. Of the C104' complexes used in the correlations, only MgOEP“ and ZnOEP“ have been firmly assigned a ground electronic state (2A1u) based on EPR studies (ref. 8). Coupling between the paramagnetic metal centers and the porphyrin radical of CoIIOEP+'ClO4" and CuOEP“ render these com lexes EPR silent. Although the uv-visible absorption spectra of Cu, Co and NiOEP“ imply 2A2u character for their ClO4‘ complexes, the electronic state of CuOEP“ClO4‘ is assigned as 2A1u based on NMR spectra (ref. 10a). To our knowledge, the EPR spectrum of NiOEP“ClO4‘ (ref. 29) has never been discussed in terms of 2A2u or 2A1u ground state symmetry. Based on RR frequencies, the precurser compound, ComOEPBr', is predicted to have a contracted core, 911.957 :t 0.009 A. We also note that this one-electron oxidized compound displays certain Raman absorption spectroscopic features analogous to a structure known to possess a buckled porphyrin core, namely the u-nitrido dimer, (FeOEP)2N. See: (a) Hoffman, J.A.; Bocian, D.F., J. Phys. Chem. 1984, _8_8_, 1472-1479; (b) Scheidt, W.R.; Summerville, D.A.; Cohen, I.A., J. Am. Chem. Soc. 1976, fl, 6623-6628; for spectra and crystal structure of u-nitrido dimers, respectively. However, it is not yet clear if the geometry of this precurser (not "parent" as defined above) is relevant to the structure of the two-electron oxidized product, CoIIIOEP+‘ZBr‘. Hanson, L.K.; Chang, C.K.; Davis, M.S.; Fajer, J., J. Am. Chem. Soc. 1981, 103, 663-670. Loew, G.H.; Herman, 2.8., J. Am. Chem. Soc. 1980, 102, 6174-6175. Fujita, 1.; Hanson, L.K.; Walker, F.A.; Fajer, J., J. Am. Chem. Soc. 1983, 105, 3296-3300. Fujita, E.; Chang, C.K.; Fajer, J., J. Am. Chem. Soc. 1985, 107, 7665-7669. Chang, C.K., Hanson, L.K.; Richardson, P.F.; Young, R.; Fajer, J., Proc. Natl. Acad. Sci. USA 1981,_'_7_8_, 2652-2656. (a) Roberts, J.E.; Hoffman, B.M.; Rutter, R.; Hager, L.P., J. Biol. Chem. 1981, _2_5_6_, 2118-2121; (b) Rutter, R.; Hager, L.P., J. Biol. Chem. 1982, _2_5_7_. 7958-7961; (c) Schultz, C.E.; Rutter, R.; Sage, S.T.; Debrunner, P.G.; Hager, L.P., Biochemistry 1964, 2_3_, 4743-4754. Sontum, S.P.; Case, D.A., J. Am. Chem. Soc. 1985, 107, 4013-4015. CHAPTER 5 DETAILS OF PEROXIDASE CATALYSIS SUGGESTED BY RESONANCE RAMAN MEASUREMENTS OF IRON-OXYGEN STRETCHING FREQUENCIES OF HORSERADISH PEROXIDASE INTERMEDIATES INTRODUCTION Horseradish Peroxidase (HRP) uses hydrogen peroxide to catalyze the oxidation of indole acetic acid (a growth hormone) in plant roots. The generally accepted reaction pathway, as pr0posed independently by Chance1 and George,2 occurs in four steps: 1. HRP+ H202 + HRP-I + H20 2. HRP-I + AH + HRP-II + A' 3. HRP-II + AH + HRP + A‘ + H20 4. 2A' + products Other pathways, however, may have physiological significance.3 The prosthetic group of the enzyme in the native state (HRP) consists of a ferric protOporphyrin IX ligated on the proximal side of the heme plane by histidine nitrogen. In the first intermediate state, compound I (HRP-I), the heme most likely retains the histidine imidazole (Im) ligand and undergoes two-electron oxidation, forming 101 102 an oxoferryl protoporphyrin IX 17 cation radical, O=FeIV(Im)PP“. The second intermediate, compound II (HRP-II), is formed via the oxidation of the substrate (AH) which leaves the low-spin (S=l) oxoferryl-imidazole linkage intact but reduces the porphyrin macrocyclic back to the neutral state.4'6 The intermediates of the HRP reaction are of interest because of the possible general application by nature of similar structures in the catalysis of not only other peroxidases, but also catalases, oxygenases and oxidases.7 The oxoferryl structure of HRP-I and HRP-II is well characterized by a variety of physical techniques. Three unpaired electrons (S = 3/2) are indicated by magnetic susceptibility measurements of HRP-I,8 consistent with a low-spin (S=1) FeIV ferromagnetically coupled to a S=1/2 radical center. Massbauer measurements of HRP-I and HRP-II are compatible with similar FeIv configurations for both intermediates.9 170 electron nuclear double resonance (ENDOR) Spectroscopy of HRP-I prepared with H21702 demonstrates the incorporation of a single 170 atom in the first intermediate.10 Proton nuclear magnetic resonance (NMR) studies of the second intermediate and its synthetic models provide further evidence for the oxoferryl structure in HRP-II.11 Although somewhat controversial,12:13 recent evidence from extended X-ray absorption fine structure (EXAFS) spectroscopy presents detailed structures for HRP-I and HRP-II with relatively short (1.65 A) FeO bond lengths and contracted (1.99 A) heme center-to-pyrrole nitrogen distances.14115 Recent resonance Raman (RR) measurements of the FeO stretching frequency, v(FeO), of HRP-1116917 and model oxoferryl porphyrinsls“21 confirm the double-bonded character and short bond length of the oxoferryl structure. Furthermore, the frequencies of the porphyrin core vibrational modes are consistent with the 1.99 A core size for HRP-II“:23 and for the synthetic oxoferryl complexes“):24 We present here RR evidence that these structural features are maintained 103 in HRP-I as well. RR studies of HRP-II also reveal a pH dependency for v(FeO) which suggests H-bonding of the oxo ligand to a protonated distal histidine with a pKa of 8.6.25126 We report a v (FeO) for the first intermediate at pH 7 which is very similar to that of HRP-II at pH 11 in the absence of H-bonding. This suggests that the oxoferryl of HRP-I is not H-bonded at pH 7. The significance of these structural features to the kinetics and mechanism of peroxidase catalysis is discussed. RR studies of the first intermediate are complicated by its reactivity and photolability.27 Past studies”,29 were interpreted to suggest that cryogenic techniques may be insufficient to stabilize HRP—I to laser irradiation in the region of the Soret absorption (300-450 nm). Because of this we have adopted an alternative approach, using pulsed, near-uv RR excitation of flowing samples of HRP-I generated by rapid mixing.30 Similar approaches have been applied to reaction intermediates of cytochrome c oxidase, and both pulsed31132 and continuous wave (CW) laser excitation33 have been used. The high frequency (1200-1700 cm‘l) spectrum of HRP-I we obtained earlier with pulsed RR excitation30 has been independently confirmed by Ogura and Kitagawa34 with CW excitation. We have also carried out an extensive RR study of porphyrin 11 cation radicals (P“) and other oxidized forms of Co, Cu and Zn octaethylporphyrin (OEP).35937 We present here preliminary results from an additional RR study of synthetic ferric porphyrin cation radicals.38 Based on these recent results from the P“ model compounds and further characterization of the RR scattering from HRP-II presented here, we have reanalyzed the high frequency RR spectrum obtained for HRP-l. We find no vibrational frequencies characteristic of a metalloporphyrin 11 cation radical (MP“). In view of the evidence in favor of the P“ formulation for HRP-I from uv-visible absorption,39 electron paramagnetic resonance (EPR),40941 ENDOR,42 and NMR43144 Spectroscopies, we offer possible explanations for the RR results. 104 EXPERIMENTAL Materials. Horseradish peroxidase was purchased from Sigma (Type VI) and used without further purification. Sodium phosphate and sodium carbonate buffers (50 mM) were used at pH 7 and 10.8, respectively, for all enzyme and hydrogen peroxide solutions. 30% H21602 was purchased from Mallinkrodt. H21302 was made from 1302 (98%, Cambridge Isotope Laboratories) using glucose oxidase (SIGMA) according to the procedure of Asada and Badger.45 HRP and peroxide solutions were quantified photometrically by using extinction coefficients at 403 7117146 and 240 nm47 of 103 mM'lcm‘1 and 43.6 M‘lcm"1, respectively. — HRP-I. Flowing HRP-I samples (pH 7) for RR measurements were prepared by rapid mixing of equal volumes of cooled solutions of 0.1 mM HRP and 1 mM hydrogen peroxide. These were loaded in 10.0 ml syringes and driven through two eight-jet tangential mixers in series“8 with a Sage 355 syringe pump. After mixing, the HRP-I sample passed through a 0.5 mm i.d. capillary. Raman scattering from HRP-I was measured both with pulsed excitation at 390 nm and 420 nm, and with CW excitation at 406.7 nm. Enzyme solutions were collected during the measurement and subsequently allowed to relax back to the native state overnight, whereupon they were pre-filtered (0.45 11111 pore size, Gelman) and concentrated (the enzyme concentration is halved during the mixing experiment) in a Diaflow ultrafiltration cell (Amicon, PM 10) and used again. HRP-I samples prepared from the recycled solutions had a much longer lifetime owing to the removal of potential substrates present in the preparations.49 Initial RR spectra of HRP-I were obtained using 0.65 - 0.85 ml/min flow rates which produced dead times of 3.1 - 2.4 s between mixing and laser excitation. As a spatial beam width of 1.5 mm incident on the 0.5 mm i.d. capillary was used, these flow rates insured that 3.6 - 4.7 scattering 105 volumes (one scattering volume is 0.2 111) passed through the capillary between laser pulses. After preliminary characterization of the spectrum under these conditions, we determined that a 0.22 ml/min flow rate (corresponding to a 9.3 3 dead time) produced identical results for the recycled preparations. At this slower flow rate, 1.2 scattering volumes passed between pulses, corresponding to the minimum necessary to- insure that each pulse was incident on a fresh sample aliquot. This was used to increase accumulation time for the low frequency spectra which suffered from intense background scattering from the capillary walls. Under the conditions described above for our system we estimate that 1.0 mJ pulses at 390 nm provide 3» 35 photons absorbed/molecule, for those molecules from which Raman scattering is collected, regardless of flow rate. For the CW experiment at 406.7 nm, we estimate 22 and 69 photons/molecule absorbed by the sample at flow rates of 0.70 and 0.22 ml/min, respectively, for a 15 mW laser power, and 0.5 mm beam waist. Thus, for pulsed excitation, photon flux is independent of flow rate above a threshold which insures that each pulse is incident on a fresh sample aliquot. Photon flux then depends directly (and exclusively) on pulse energy if spatial beam width and capillary i.d. (which determine scattering volume) remain constant. On the other hand, for CW excitation, photon flux is inversely proportional to flow rate and directly proportional to laser power for a system in which the laser beam waist and capillary i.d. are equal (i.e. all photons are incident on the sample and all sample molecules are illuminated). The formation of HRP-I was confirmed for each rapid mixing experiment by visual detection of the flowing green solution in the capillary prior to the onset of the RR measurement. After laser irradiation a sample aliquot was collected and checked by a uv-visible absorption measurement during each run. For the recycled samples these spectra were essentially that of HRP-I. For 106 the less stable fresh samples these spectra reflected a mixture of HRP-I and HRP-II, confirmed by monitoring the return to the native state. The total time from mixing to the absorption spectrum was typically '1. 3 minutes or longer. The formation of HRP-I prior to the RR measurement was further confirmed by an independent optical absorption experiment in which the rapid mixer with a section of rectangular tubing (0.3 mm path, Vitrodynamics) in place of the capillary was used. HRP-II. HRP-II samples were prepared in the cooled cylindrical quartz cell which was spun during RR measurements. Addition of stochiometric amounts of p-cresol (Aldrich) and 1.5 x stochiometric quantities of hydrogen peroxide to solutions of native HRP at pH 7 resulted in formation HRP-II. RR spectra of these samples were collected with OMA detection for 2.5 min after which absorption spectra were monitored and the experiment was repeated with a fresh sample. HRP-II samples at pH 10.8 were prepared by adding 1.5 equivalents of peroxide to solutions of the native enzyme. RR spectra of these more stable solutions were collected for 10 min, and absorption spectra were monitored before and after laser irradiation. Model compounds. Cl'FemOEP was prepared from OEPHZ (Porphyrin Products) by standard methods.50 Cl‘FemOEP“CIO4‘ was prepared by A. Salehi by stirring Fe(ClO4)3 in a CH2C12 solution of Cl'FemOEP. Sample integrity was monitored by optical absorption before and after RR measurements. CH2C12 was freshly distilled from CaHz. Instrumentation. Laser pulses (5 ns duration, 10 Hz) from 390-435 nm were provided by a DCR-IA based Quanta Ray System by using three laser dyes (Exciton): LD390 (in methanol), DPS (in dimethylformamide) and Stilbene 420 (in methanol). CW emission at 363.8 nm came from a Coherent Innova 90-5 Argon ion laser. Raman scattering from these sources was measured with a 107 A n V Lama.) P.C. diam. . 8 mm. diam. 1*: . 05 mm. Tr" View in direction of arrow X .A ll. 8 Section B-B 1_ 4 1 cm. Figure 5-1 . End view (a), longitudal section (b) and cross-section (c) of the eight-jet, tangential mixers. Ref. 48. 108 8:08:80 :a :o CESEEC was 3a.. :9. mm 3 9E. .:o:3:oE=.=m:_ £5 38: 8.5908 cm? was :8 wE: 15% 5.3.6 Bot—05:8 A0 mtmv @280 a :m 33:28 :42: Eon.“ E: 33723 a: 8:88 @2338 mm .moEobooE = <20 88888 use .6698 8...; oboe S.,: 5.: 080m of wcmm: an 033.3: “3:: 030320 28 ...oquoEo ..o:oE 23.5 a :_ vomhoammc £333.35 3: xoam a :33 @3828 $3 Cm 5333 :3 29:8 9: E0...“ mieguaom :aEwm dong tmzm._.ZH Z<_>_._._mzw._.ZH Z<§._._mzm._.ZH Z<_>_tmszZH 232$; e) 406.7 nm f) 404 nm soo‘aoo l200 I000 400 RAMAN SHIFT (cm") 129 e E C E 3 E {3- n o s! g g g HRP-I pI'I I08 I I 5.1". V4 ‘ I379 CITITI E C >_ o .___ a (I) I Z LIJ I— Z Z 4 2 VIFeO) - 787 cm" <1 0: LIJ Z ’— <1 _I LIJ 0: 117-680 cm”l 23 2'4' ' ' '2'5' ' ' T2'6 EXCITATION FREQUENCY (cm"/|OOO) Figure 5-9 - Excitation profiles of W; = 680 cm'l, v (FeO) = 787 cm"1 and V4 = 1379 cm‘l. 130 DISCUflON Prediction of HRP-l vibrations from structural correlations. The similarity between the high frequency RR spectra of HRP-l and HRP-II in Figure 5-3b and c is unexpected. The vibrational frequencies between 1450 and 1700 cm"1 are a sensitive probe of the structure of both the neutral metalloporphyrin52 and the metalloporphyrin 1r cation radical.36 Studies of model compounds suggest that oxidation of the macrocycle does not significantly influence porphyrin core geometry, particularly if the oxidation state and axial ligation of the metal center remain unchanged.36168 This is in agreement with EXAFS measurements of HRP-I and HRP-II by Penner-Hahn 91 31.15 Thus, we expect the porphyrin core sizes of HRP-I and HRP-II to be similar. Despite the structural similarities between MP and MP‘“ the electronic states of the cation are clearly distinct from the netural form and our recent studies of metallooctaethylporphyrin 1r cation radicals (MOEP+' where M = Zn, Cu, Co, and Ni) demonstrate that vibrational frequencies and relative intensities in the RR spectra of the MOEP“ species are substantially changed from those of the parent MOEP compounds.36 Thus we expect significant differences in the RR spectra of HRP-I y_s_. HRP-II, analogous to those we observed in the MOEP“ 1g. MOEP spectra. Based on our recently presented36 structural correlations and assuming a core size of 1.99 A, we expect v3, 911 and v 2 to occur at approximately 1501, 1590 and 1606 cm'l, respectively, for HRP-I. Here we have assumed that the frequencies of the v(Cbe) modes, v11 and 92, are 10 cm“1 less for PP“ compared to OEP‘" compounds as they are in the ring-neutral complexes.69 Thus, the V3 value of 1508 cm"1 for HRP-I (Fig. 5-3b), unchanged from that of HRP-ll (Fig. 5-3c), and the absence of features in the 1590 to 1610 cm'1 range, suggest that the species producing the spectrum 131 in Figure 5-3b is not a porphyrin 1r cation radical. Since the evidence for a porphyrin radical in HRP-l, particularly from ENDOR results,42 is strong, we consider here possible explanations for the RR result. HRP—II. The trivial explanation is that our HRP-I spectrum is produced by bona-fide HRP-II contaminants in the sample, produced either by photo-reduction or reduction by impurities acting as substrates. However, although similar, the RR spectra of HRP-I and HRP-II at pH 7 are clearly distinct, most notably in terms of the behavior of the oxoferryl stretching mode (Fig. 5—5). Furthermore, optical absorption spectra following each RR measurement of HRP-l solutions (particularly for the recycled samples) were like that of HRP-I. Thus, reduction to HRP-II (which would be irreversible) is not the cause of the similar RR spectra of the two intermediates (Fig. 5-3b and c). Photochemistry. Both frozen and solution samples of HRP-I are known to undergo photoreaction when exposed to light in the Soret band region.27 The photoproduct, called Y, has an absorption spectrum like that of HRP-II (i.e. characteristic of an oxoferryl, ring-neutral heme), and gives an EPR signal of a free radical located on a protein residue.70 Although a different interpretation has been offered,71 photo-induced electron transfer from a nearby amino acid to the porphyrin cation radical could leave a structure analogous to cytochrome c peroxidase compound 1.4453 This species would have an oxoferryl, ring-neutral heme with a protein free radical and would account to the EPR and absorption spectrum of the photoprod'uct.71 If this were the case, the RR spectrum of Y should resemble that of HRP-II. Thus, our spectrum (Fig. 5—3b) might represent scattering from Y despite the measures taken to prevent the photoreaction. Under our typical experimental conditions (1 mJ/laser pulse) we estimate a maximum of 35 photons/molecule was absorbed by the scattering sample. With a quantum yield of 0.003,'71 this would correspond 132 to less than 1196 photoconversion possible per 5 nanosecond laser pulse. Thus, a maximum of 1196 impurity could be reflected in the RR spectrum only if Y is formed in less than 5 ns. We estimate that a 17% impurity with an extinction coefficient at 420 nm of 100 mM‘lcm'1 would be required to produce scattering ”ex = 390 nm) at intensitites comparable to that of the green cation radical, HRP-I, with an extinction coefficient at 400 nm of 50 mM‘lcm'l. Thus, an 1196 photoconversion could feasibly cause artifacts in the HRP-I RR spectrum. The close similarity of the RR excitation profiles of HRP-l at pH 7 and HRP-ll at pH 10.8 are consistent with possible photoconversion as they suggest scattering species with similar electronic absorption spectra. On the other hand, similar experiments with CW excitation at 406.7 nm at a wide range of flow rates (from 0.2-20 ml/min) carried out both in our lab (Fig. 5-7b) and elsewhere,34 as well as pulsed measurements at laser energies ranging from 0.5-2.0 mJ/pulse, all produced similar results. These facts argue against not only artifacts due to photoconversion to Y but also possible scattering from excited states of HRP-I. Spin delocalization. Using an approach similar to ours, but with CW excitation at 406.7 nm and high flow rates (5-20 ml/min.) Ogura and Kitagawa34 recently obtained a high frequency RR spectrum of HRP-I, although with a very low signal-to-noise ratio. Their result is similar to our high frequency spectrum30 (see also Fig. 5-3b) and is also atypical of a MP‘“. As these authors point out, extensive spin delocalization of the porphyrin radical onto the axial oxoferryl-imidazole system might account for the absence of vibrations characteristic of the oxidized porphyrin in the RR spectrum of HRP-I. Delocalization of the porphyrin cation radical spin onto an axial pyridine ligand has been described for PyZnTPP‘” (TPP = tetraphenylporphyrin).72 Perhaps more significant, the spin systems of the oxoferryl (8:1) and porphyrin cation radical (S=1/2) may interact strongly in compound I structures.73174 While 133 these factors may well influence vibrational frequency shifts between oxidized and neutral porphyrin complexes, this may not be sufficient to explain the essentially identical RR spectra of HRP-II (pH 10.8) and HRP-I (pH 7) presented in Figures 5-6c-e and 5-7, respectively. ENDOR measurements clearly show that the unparied electron in the radical site of HRP-I is hyperfine coupled to both 1H and 14N nuclei.42 These could come from either the methine and b-carbon substituents and the pyrrole nitrogen of the porphyrin or from the a- and B -protons and imidazole nitrogen of the axial histidine. Although deuterium substitution ENDOR experiments which would demonstrate that the radical site is indeed the porphyrin rather than the histidine are mentioned by Roberts e_t_ £1.42, they are not presented. HRP-1". Given the possibility of a product of photo-induced electron transfer within the heme pocket, we must address the relationship of the photoproduct to the green form of HRP-I and to the catalytic sequence. We will call this putative photoproduct HRP-1* to acknowledge the possibility that it may be distinct from the product Y proposed earlier,70,71 and to emphasize that it is formally equivalent to HRP-I in oxidation state. Figure 5-10 depicts HRP-1* in rapid equilibrium with HRP-l. In the dark, formation of HRP-1* is negligible, but under high illumination HRP-1* is favored, being reached presumably through an excited state of HRP—l. HRP-1‘I lacks both the prophyrin radical and the H-bonded oxoferryl. Devoid of these two features, HRP-1* is an unreactive structure and must return to the green form, HRP-l, in order for peroxidase catalysis to continue. The heme of HRP-1* is thus identical to that of HRP-II at high pH. This accounts for the similarity of the spectra in Figures 5-6 and 5-7. In this interpretation, HRP-I and HRP-1* have metal centers with identical valences and axial ligands and thus differ only in the porphyrin oxidation state. Our RR studies of (CH3OH)2ComOEPClO4‘ and (C104‘)ZComOEP+' suggest 134 I»... HRP-I Figure 5-10. HRP-II Schematic diagram of the active site of HRP catalytic intermediates, showing HRP-I, the postulated photo- product HRP-I“, the reaction of HRP-l with a phenolic substrate, and HRP-II. 135 similar six-coordinate cobaltic solution structures with 1.97 A core sizes for both complexes.36 By analogy to these and other MOEP“, MOEP pairs, we conclude that HRP-l and HRP-1* possess similar geometries. Thus, should our HRP-I RR result actually be due to an HRP-1* photoproduct, we prOpose that the structural information contained in the RR spectrum is still valid to the green P+' form of HRP-l. Therefore, given either possible explanation (i.e. photochemistry or spin delocalization) for the similarity of the high frequency RR scattering of HRP-l and HRP-ll, the results presented in Figures 5-3 and 5—5 suggest that the core sizes of these two species are equal and that, at pH 7, the oxoferryl of HRP-I is not H-bonded as is the oxoferryl of HRP-II. In the absence of the H-bond at high pH,25,26 HRP-II is unreactive. The association between this H-bond and the reactivity of HRP-II suggests that the oxoferryl unit requires the H-bond for activation, presumably in order to facilitate the formation of leaving H20 as depicted in reaction 3, above. Thus, the absence of the H-bond in HRP-l helps explain the initial reduction of the porphyrin radical cation, rather than the ferryl structure, to form the second intermediates. Kinetics and mechanism of HRP-l reduction. A host of kinetic studies of the reaction of HRP-l with a variety of substrates suggest the involvement of a distal amino acid residue exhibiting a pKa of 5.1,5 identified possibly as histidine 42 (His-42).6b For the substrates p—cresol and L-tyrosine, the reaction proceeds by a base catalyzed mechanism.6 Figure 5-10 proposes a mechanism for the reaction of p—cresol with HRP-l at pH 7. The substrate is oriented by H-bonding to distal His-42. This is consistent with studies which show that the acid form of p—cresol will bind to the cyanide complex of HRP at a distinct but nearby binding site at pH 7.75 As an electron is transfered from the substrate to the porphyrin cation radical, a proton is transferred along the H-bond to His-42. This homolytic cleavage of the H-A bond (or H-O bond of p-cresol) 136 results in the formation of the free radical of the substrate76 along with HRP-ll. The oxoferryl of HRP-ll is H-bonded to the protonated His-42 residue.25v26 The reaction of HRP-II with the substrate is influenced by an acid group with pKa 8.6.5 Thus, the participating acid group is most likely His-42 in both cases as was speculated earlier.77 The pKa is shifted by the positive charge of the I porphyrin cation in HRP-l. This mechanism is consistent with Hammett and Okamoto-Brown plots constructed from rate constants for the reduction of HRP-l by a series of substituted phenols. These were taken to suggest a mechanism wherein the neutral substrate donates an electron to HRP-l and simultaneously loses a proton.78 CONCLUSIONS. Using RR spectroscopy applied to a flowing sample of HRP-I (pH 7) prepared by rapid mixing, we have characterized a distinct intermediate formed by the reaction of HRP with hydrogen peroxide. This compound is two (oxidation equivalents above the native state and decays through a species with an absorption spectrum characteristic of HRP-II. The RR scattering in the 1300-1700 cm'1 range obtained from this intermediate is not characteristic of a metalloporphyrin 17 cation radical and is suggestive of an oxoferryl heme with the same geometric and electronic structure as that of HRP-II (particularly _the high pH form). We have considered two possible explanations for our result: (i) photo-induced electron transfer to the porphyrin radical cation of HRP-l occuring in less than 5 ns and resulting in a new species, HRP-1*, which produces the RR spectrum, or (ii) extensive delocalization of the porphyrin radical of HRP-I onto the axial ligands. The low frequency RR scattering in the 600-1000 cm'1 range from this species reveals a v(FeO) at 791 cm‘l, typical of the non-hydrogen bonded form 137 of HRP-ll. This suggests that the oxo ligand of HRP-I is not hydrogen bonded at pH 7. This proposal is consistent with the kinetics of the reaction of HRP-I with p-cresol. Because the oxoferryl structure requires the hydrogen bond for activation,25v26 this result explains why the porphyrin radical is the site of reduction of HRP-I. We represent a mechanism for HRP-I reduction by p-cresol in which a proton from the substrate binds to His 42 and forms a hydrogen bond to the oxo ligand of HRP-ll. FUTURE WORK At the present time it is not possible to address conclusively the question of photoconversion v_s. radical delocalization to explain the high frequency RR result for HRP-l. In addition to the work presented here, we have utilized cryogenic techniques, both in aqueous (unpublished results) and antifreeze solvents,30 to obtain RR spectra of HRP-I similar to those we obtained with both pulsed, 390 nm and CW, 406.7 nm excitation. Variations of the rapid mixing approach have included the following: (i) off-resonance CW excitation at 363.8 nm which failed to produce detectable Raman scattering, and (ii) comparative CW measurements at 406.7 nm at different flow rates (we used 0.22 and 0.7 ml/min; Ogura and Kitagawa34 report flow rates of 5-20 ml/min), all of which produced similar results. Felton e_t_ 53.58 also describe results similar to these obtained under 457.9 nm excitation for HRP-I samples at temperatures of 0 to -30 C. In conclusion we must assume that if these results do not represent Raman scattering from the green MP‘“ form of HRP-I, such scattering is prohibitively difficult to obtain under CW or nanosecond pulsed laser excitation. On the other hand, in collaboration with A. Salehi, RR studies of the model compound O=FeIVOEP+’79 are in progress and will greatly facilitate further analysis. In particular, methods to produce 180=FeIVOEP+' are being developed. 138 Time-resolved Optical absorption measurements of flowing HRP-I samples illuminated by pulsed laser irradiation are planned. These experiments will be carried out in collaboration with Dr. Dewey Holten and coworkers and will conclusively address the question of the involvement of HRP-I photochemistry in these RR results. Regardless of these complications, the pH dependence of the v(FeO) observed for our putative HRP-I species should be determined. Low frequency data should be obtained from samples at pH 4, 5 and 6 in order to estimate the pKa of the proposed histidine group involved in the reaction. Also, the experiment should be repeated in D20 buffer at pH 7 to establish more firmly the absence of hydrogen bonding of the oxo ligand. Finally, RR studies of intermediates involved in the catalysis of other peroxidases (notably mye10peroxidase, lactoperoxidase and heme bromoperoxidase) have already begun. This work will be carried out in the MSU Shared Laser Laboratory with Drs. R. Wever and J. Manthey. ACKNOWLEDGMENTS We thank Dr. T. Kitagawa for manuscripts prior to publication and Drs. R. Wever and J. Manthey for discussion and critical reading. T.O. thanks A. Salehi for technical assistance. This work was supported by NIH grant GM-25480. 10. 11. 12. 13. 14. REFERENCES Chance, B., Arch. Biochem. Biophys. 1952, 4_1, 416-424. George, P., Nature 1952, 169, 612-613. Smith, A.M.; Morrison, W.L.; Milham, P.J., Biochemistry 1982, _2_l_, 4414-4419. Frew, J.E.; Jones, P. ' in "Advances in Inorganic and Bioinorganic Mechanism"; Academic Press: New York, 1984; Vol. 3, pp. 175—213. Hewson, W.D.; Hager, L.P. in "The Porphyrins", Dolphin, D., Ed.; Academic Press: New York, 1979, Vol. 7, pp. 295-332. (a) Dunford, H.B.; Stillman, J.S., Coord. Chem. Rev. 1976, _1_9, 187-251; (D) Dunford, H.B., Adv. Inorg. Biochem. 1982, _4, 41-68. Naqui, A.; Chance, B., Ann. Rev. Biochem. 1986, _5_5_, 137-166. Theorell, H.; Ehrenberg, A., Arch. Biochem. Biophys. 1952, 4_1, 442-461. Moss, T.H.; Ehrenberg, A.; Bearden, A.J., Biochemistry 1969, 8, 4159-4162. Roberts, J.E.; Hoffman, B.M., J. Am. Chem. Soc. 1981, 103, 7654-7656. LaMar, G.N.; de Ropp, J.S.; Latos-Grazynski, L.; Balch, A.L.; Johnson, R.B.; Smith, K.M.; Parish, D.W.; Cheng, R.-J., J. Am. Chem. Soc. 1983, 105, 782-787. Chance, B.; Powers, L.; Ching, Y.; Poulos, T.; Schonbaum, G.R.; Yamazaki, 1.; Paul, K.G., Arch. Biochem. Biophys. 1984, 235, 596-611. (a) Chance, M.; Powers, L.; Kumar, C.; Chance, B., Biochemistry 1986, 23, 1259—1265; (b) Chance, M.; Powers, L.; Poulos, T.; Chance, B., Biochemistry 1986, .2_5_, 1266-1270. Penner-Hahn, J.E.; McMurry, T.J.; Davis, I.M.; Balch, A.L.; Groves, J.T.; Dawson, J.H.; Hodgson, K.O., J. Biol. Chem. 1983, 258, 12761-12764. 139 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 140 Penner-Hahn, J.E.; Smith-Eble, K.; McMurry, T.J.; Groves, J.T.; Dawson, J.H.; Hodgson, K.O., J. Am. Chem. Soc. 1987, in press. Terner, J.; Sitter, A.J.; Reczek, C.M., Biochim. Biophys. Acta 1985, 828, 73-70. Hashimoto, S.; Tatsuno, Y.; Kitagawa, T., Proc. Japan Acad. 1984, 6_0, Ser. B.; 345—348. Bajdor, K.; Nakamoto, K., J. Am. Chem. Soc. 1984, 1_(_)§, 3045-3046. Proniewicz, L.M.; Bajdor, K.; Nakamoto, K., J. Phys. Chem. 1986, _9_0_, 1760-1766. Chappacher, M.; Chottard, G.; Weiss, R., J. Chem. Soc., Chem. Commun. 1988, _2, 93-94. Kean, R.T.; Oertling, W.A.; Babcock, G.T., J. Am. Chem. Soc. 1987, in press. Terner, J.; Reed, D.E., Biochim. Biophys. Acta 1984, 789, 80-86. Sitter, A.J.; Reczek, C.M.; Terner, J., Biochim. Biophys. Acta 1985, 828, 229-235. Kean, R.T.; Babcock, G.T., manuscript in preparation. Sitter, A.J.; Reczek, C.M.; Terner, J., J. Biol. Chem. 1985, 260, 7515-7522. Hashimoto, S.; Tatsuno, Y.; Kitagawa, T., Proc. Natl. Acad. Sci. USA 1986,83, 2417-2421. Stillman, J.S.; Stillman, M.J.; Dunford, H.B., Biochemistry 1975, _1__4, 3183-3188. Teraoka, J.; Ogura, T.; Kitagawa, T., J. Am. Chem. Soc. 1982, 104, 7354-7356. Van Wart, H.B.; Zimmer, J., J. Am. Chem. Soc. 1985, 107, 3379-3381. Oertling, W.A.; Babcock, G.T., J. Am. chem. Soc. 1985, 107, 6406-6407. Babcock, G.T.; Jean, J.M.; Johnston, L.H.; Palmer, G.; Woodruff, W.H., J. Am. Chem. Soc. 1984, 106, 8305-8306. Babcock, G.T.; Jean, J.M.; Johnston, L.H.; Woodruff, W.H.; Palmer, G., J. Inorg. Biochem. 1985, _2_3_, 243-251. Ogura, T.; Yoshikawa, S.; Kitagawa, T., Biochim. Biophys. Acta 1985, _8_3_2_, 220-223. Ogura, T.; Kitagawa, T., submitted for publication. Salehi, A.; Oertling, W.A.; Babcock, G.T.; Chang, C.K., J. Am. Chem. Soc. 1986, _1_Q_8_, 5630-5631. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 141 Oertling, W. A.; Salehi, A.; Chung, Y.; Leroi, G. E.; Chang, C. K; Babcock, G..,T J. Am. Chem. Soc. 1987, submitted for publication. Oertling, W.A., Salehi, A.; Chang, C.K.; Babcock, G.T., J. Phys. Chem. 1987, in press. Salehi, A.; Oertling, W.A., unpublished results. Dolphin, D.; Forman, A.; Borg, D.C.; Fajer, J.; Felton, R.H., Proc. Natl. Acad. Sci. USA 1971, 6_8, 614-618. Aasa, R.; Vanngard, T.; Dunford, H.B., Biochim. Biophys. Acta 1975, 391, 259-264. Schulz, C.E.; Devaney, P.W.; Winkler, H.; Debrunner, P.G.; Doan, N.; Chiang, R.; Rutter, R.; Hager, L.P., FEBS Lett. 1979, 103, 102-105. Roberts, J.E.; Hoffman, B.M.; Rutter, R.; Hager, L.P., J. Biol. Chem. 1981, 256, 2118-2121. La Mar, G.N.; de ROpp, J.S., J. Am. Chem. Soc. 1980, 102, 395-397. La Mar, G.N.; de Ropp, J.S.; Smith, K.; Langry, K.C., J. Biol. Chem. 1981, _2_5_8, 237-243. Asada, K.; Badger, M.R., Plant 6c Cell Physiol. 1984, 25, 1169-1179. Schonbaum, G.R.; Lo, S., J. Biol. Chem. 1972, 247, 3353-3360. Worthington Biochemical Corporation, "Worthington Enzyme Manual", 1972, p. 41. Gibson, Q.H.; Milnes, L., Biochem. J. 1964, _9_1_, 161-171. Chance, B., Arch. Biochem. 1949, _2_2_, 224-252. Adler, A.; Kampas, F.; Kim, J., J. Inorg. Nucl. Chem. 1970, 12, 3443-3445. Spaulding, L.D.; Chang, C.C.; Yu, N.-T.; Felton, R.H., J. Am. Chem. Soc. 1975, _9_7, 2517-2525. Spiro, T.G. in "Iron Porphyrins", Lever, A.P.B.; Gray, H.B., Eds.; Addison-Wesley: Reading, MA, 1983; Part 2, pp. 89-159. Abe, M.; Kitagawa, T.; Kyogoku, Y., J. Chem. Phys. 1978, _6_9_, 4526-4534. Choi, S.; Spiro, T.G.; Langry, K.C.; Smith, K.M.; Budd, D.L.; La Mar, G.N., Desbois, A.; Mazza, G.; Stetzkowski, F.; Lutz, M., Biochim. Biophys. Acta 1984, 785, 161-176. Spiro, T.G.; Strekas, T.C., J. Am. Chem. Soc. 1974, g, 338-345. 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 142 Yamaguchi, H.; Nakano, M.; Itoh, K., Chem. Lett. Jpn. 1982, 1397-1400. Felton, R.H.; Romans, A.Y.; Yu., N.-T.; Schonbaum, C.R., Biochim. Biophys. Acta 1976, 434, 82-89. Rakhit, G.; Spiro, T.G., Biochem. Biophys. Res. Comm. 1976, _7_1_, 803-808. Boersma, A.D.; Goff, H.M., Inorg. Chem. 1984, _2_3_, 1671-1676. Hanson, L.K.; Chang, C.K.; Davis, M.S.; Fauer, J., J. Am. Chem. Soc. 1981, 103, 663-670. Kim, D.; Miller, L.A.; Rakhit, G.; Spiro, T.G., J. Phys. Chem. 1986, fl, 3320-3325. Maggiora, G.M., J. Am. Chem. Soc. 1973, 95, 6555—6559. Kashigawa, H. in "Biomolecules: Electronic Aspects", Nagata gt _a_l_., Eds.; Elsevier: New York, 1985; pp. 31-50. Choi, S.; Spiro, T.G.; Langry, K.C.; Smith, K.M., J. Am. Chem. Soc. 1982, 194, 4337-4344. La Mar, G.N.; de Ropp, J.S.; Smith, K.; Langry, K.C., J. Am. Chem. Soc. 1983, 105, 4576-4580. Bangcharoenpaurpong, K.T.; Schomacher, K.T., Champion, P.M., J. Am. Chem. Soc. 1984, 106, 5688-5698. (a) Spaulding, L. D., Eller, P. G.; Bertrand, J. A; Felton, R. H, J. Am. Chem. Soc. 1974, _9_6, 982- 987; (b) Gans, P.; Buisson, G.; Duee, E.; Marchon, J. -.;C Erler, B..;S Scholz, W.F.; Reed, C.A., J. Am. Chem. Soc. 1986,108, 1223-1234. Callahan, P.M.; Babcock, G.T., Biochemistry 1981, _2_0_, 952-958. Chu, M.; Dunford, H.B.; Job, D., Biochem. BiOphys. Res. Commun. 1977, E, 159-164. Nadezhdin, A.D.; Dunford, H.B., Photochem. Photobiol., _2_9_, 899-903. Fujita, 1.; Hanson, L.K.; Walker, F.A.; Fajer, J., J. Am. Chem. Soc. 1983, LIE, 3296-3300. Sontum, S.P.; Case, D.A., J. Am. Chem. Soc. 1985, M, 4013-4015. Schultz, C.E.; Rutter, R.; Sage, J.T.; Debrunner, P.G.; Hager, L.P., Biochemistry 1984, 23, 4734-4754. Critchlow, J.E.; Dunford, H.B., J. Biol. Chem. 1972, 247, 3714-3725. Shiga, T.; Imaizumi, K., Arch. Biochem. Biophys. 1975, 167, 469-479. Dunford, H.B.; Cotton, M.L., J. Biol. Chem. 1975, 250, 2920-2932. 143 78. Job, D.; Dunford, H.B., Eur. J. Biochem. 1976, 66, 607-614. 79. Chang, C.K.; Kou, M.-S., J. Am. Chem. Soc. 1979, 101, 3414-3415.