PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. F—__——_—_—.—_———'—_=___———__ DATE DUE DATE DUE DATE DUE ll MSU In An Affirmative Acrion/Equel Opportunity Institution PHOTOPHYSICS AND PHOTOCHEMISTRY OF MULTIPLY BONDED METAL-METAL DIMERS by [-1 y Chang A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1988 ABSTRACT PHOTOPHYSICS AND PHOTOCHEMISTRY OF MULTIPLY BONDED METAL-METAL DIMERS b y I-Jy Chang The excited state dynamics and excited state chemistry of multiply bonded metal-metal complexes (M-LM) have been investigated. The lowest energy excited state, 1(56*), of dimolybdenum tetrahalo(tetrakis)phosphine (M02X4P4) has been studied by using emission and time-resolved picosecond laser spectroscopy. Emission quantum yields and lifetimes are significantly affected by the electronic configuration of the individual molecule and are only slightly affected by the eclipsed-to-staggcred conformational changes. Molecules with D2d symmetry possess high emission quantum yields and long lifetimes, complexes with DZh symmetry possess lower emission quantum yields and shorter lifetimes. Transient absorption spectra of M02X4P4 complexes display decays corresponding to prompt relaxation of the 1(66*) excited state to ground state. In addition, complexes possessing phosphines with sterically bulky constituents (e.g. monodentate or bidentate phenyl phosphines) exhibit a long-lived nonemissive excited state. It is believed that this nonemissive excited state is due to the dissociation of the phosphine ligands within a solvent cage. A new class of multiply bonded dimolybdenum phosphate dimers, M02(HPO4)44' and M02(HPO4)43', has been prepared and spectrosc0pically and electrochemically characterized. The electronic absorption spectrum of M02(HPO4)44' in acidic solution is typical MAM species with the 62 -* 56* transition occurring at 516 nm. Cyclic voltammograms of phosphoric acid solutions of M02(HPO4)44' are characterized by two reversible waves at -0.67 and -0.25 V vs. SCE that have been assigned to the M02(HPO4)43’/4' and M02- (HPO4)42'/3' couples, respectively. Oxidation of M02(HPO4)44' produces the mixed-valence dimer M02(HPO4)43‘, which exhibits an intense near-infrared absorption band which is assigned to the 5 -> 5* transition. Irradiation (x > 355 nm) of phosphoric acid solutions of the MA-M dimer leads to facile production of M02(HPO4)42' and hydrogen. The photochemical reaction mechanism is consistent with sequential one-electron oxidation of the M02 core [i.e., M02(II,II) -> M02(II,III) -9 M02(III,III)]. Electronic absorption spectra of the M02(HPO4)4n' (n = 2, 3, 4) dimers in the ultraviolet spectral region, wavelength-dependent quantum yield measurements, and photochemical studies of phosphate dimers under N20 atmospheres confirm the assignment of a n 11* photoactive state of M02 phosphato complexes. Chemistry from the 1(55") excited state of the M02(HPO4)44' complex is circumvented by proton quenching. To this end, a novel photoreceptor, Moz(02P(OC6H5)2)4, was prepared to promote the low energy multielectron reactions. This diphenyl phosphate dimolybdenum species exhibits 52 -9 55* absorption at 515 nm in CH2C12 solution. Cyclic voltammograms show two reversible waves at +0.06 and +1.00 V vs. SCE which correspond to Mo2(02P(oc6H5)2)4+/0 and Mo2(02P(oc6H5)2)42+/+ couples, respectively. One electron oxidation of M02(02P(OC6H5)2)4 yields mixed-valence Moz(02P(OC6H5)2)4+ species which is a direct analogue to the M02(HPO4)43'. In spite of the similarity of M02(02P(OC6H5)2)4 to M02(HPO4)44’, M02(02P(0C6H5)2)4 emits directly from 1(655*) excited state in nonaqueous solutions. The preserved 1(66*) lifetime is long enough to permit direct reactions of the excited state in nonaqueous solution. Visible irradiation (X > 530 nm) of solutions containing M02(02P(OC6H5)2)4 and ClCHzCHzCl results in the facile oxidation of the metal core and the production of ethylene. To My Parents, Shiow-Ing and Jong-Way ACKNOWLEDGMENTS I would like to thank Dr. Daniel G. Nocera for his encouragement, guidance and friendship during the course of this study. His scientific insight has opened my eye and directed me into a new field. I also like to express my gratitude to the present members of the group for their friendship, encouragement and many helpful discussions. More importantly, they have created a great working atmosphere to make my graduate life enjoyable. I must express my thanks to Dr. Jay R. Winkler at Brookhaven National Laboratory. I was fortunate to have the opportunity to work with him. From his guidance I learned how to use the picosecond laser instrument. Moreover, I learned how to review results in different way from him. Finally, I like to thank my parents and family, without their support and love I could not finish this study. vi TABLE OF CONTENTS Page LIST OF TABLES ......................................................................................... x LIST OF FIGURES ........................................................................................ xii CHAPTER I INTRODUCTION .......................................................... 1 CHAPTER II EXPERIMENTAL ........................................................ 3 2 A. Syntheses ................................................................................ 3 2 1. General Procedures ................................................ 32 2. Syntheses of Dimolybdenum Complexes ................................................................... 3 2 B. Solvent Purification ............................................................ 41 1. Solvents Used for Syntheses ................................ 41 2. High Purity Solvents ............................................... 42 C Instrumentation and Methods ....................................... 43 1. Absorption Spectroscopy ...................................... 43 2. Steady-State Luminescence Spectroscopy 43 3. Time-Resolved Laser Spectroscopy ................ 44 4. Magnetic Measurements ....................................... 44 5. Resonance Raman ................................................... 44 6. Electrochemical Measurements ........................ 4 5 7. Photochemistry ........................................................ 45 8. Gas Analysis .............................................................. 51 vii D. Crystal Structure Determinations ................................. 51 1. General Procedures .............................................. 5 1 2. M02(02P(OC6H5)2)4'2THF ................................... 5 1 3 . M02(02P(0C6H5)2)4BF4 ...................................... 5 6 CHAPTER 111 SPECI'ROSCOPIC STUDIES OF M02X4(PBu3)4 AND M02Cl4(LL)2 COMPLEXES ................................................................. 5 9 A. Background ............................................................................. 5 9 B. Results and Discussion ....................................................... 6 5 1. Absorption Spectroscopy ...................................... 6 5 2. Emission Spectroscopy .......................................... 6 9 3. Transient Absorption Spectroscopy ................ 9 2 CHAPTER IV PHOTOCHEMISTRY OF NOVEL DIMOLYBDENUM PHOSPHATE COMPLEXES IN ACIDIC SOLUTION ................ 1 15 A. Background ............................................................................. 115 B. Results and Discussion ....................................................... 117 1. Magnetic and Spectroscopic Characterization of M02(HPO4)4“‘, (n = 2, 3, 4) Species ................................................. 1 17 2. Oxidation-Reduction Chemistry ......................... 133 3. Photochemistry ......................................................... l3 8 CHAPTER V PHOTOCHEMISTRY OF A NOVEL DIMOLYBDENUM DIPHENYL PHOSPHATE IN NONAQUEOUS SOLUTION .............................. 161 A. Background ............................................................................ 161 B. Results and Discussion ....................................................... 1 6 7 viii 1. Structures of M02(02P(0C6H5)2)4'2THF and M02(02P(OC6H5)2)4BF4 ............................. 167 2. Oxidation-Reduction Chemistry ......................... 1 8 0 3. Magnetic and Spectroscopic Characterization of M02(02P(OC6H5)2)4 and Moz(02P(OC6H5)2)4BF4 ............................. 1 85 4. Photochemistry ........................................................ 193 CHAPTER VI FINAL REMARKS ..................................................... 209 REFERENCES ......................................................................................... 21 3 ix LIST OF TABLES Table Page 1 State Orderings in d4-d4 Complexes: [Re2C18]2' ................ 27 2 Quantum Yield Data for Ferrioxalate Photoreaction ............. 49 3 Torsional Angles for Structurally Characterized M02X4(PBu3)4 and M02C14(LL)2 Complexes ....................... 64 4 Magnetic Susceptibility of K3M02(HPO4)4 ............................... 1 19 5 Electronic Absorption Maxima of Dimolybdenum Ions in Aqueous Solution ................................................................ 146 6 Crystallographic Data of M02(02P(OC6H5)2)4'2THF and M02(02P(OC6H5)2)4BF4 ......................................................... 168 7 Selected Atomic Positional Coordinates of M02(02P(OC6H5 )2)4'2THF .............................................................. l 7 4 8 Selected Atomic Positional Coordinates of M02(02P(0C6H5)2)4BF4 .................................................................. l 7 5 10 11 12 13 14 Selected Bond Distances of Moz(02P(OC6H5)2)4°2THF 176 Selected Bond Distances of M02(02P(OC6H5)2)4BF4 ........ 177 Selected Bond Angles of M02(02P(OC6H5)2)4°2THF ......... 178 Selected Bond Angles of M02(02P(OC6H5)2)4BF4 ............ 179 The Formal Reduction Potentials of the M02(II,III)/(II,II) and M02(III,III)/(II,III) Couples of M02 Sulfato, Phosphato, and Diphenyl Phosphate Complexes ................................................... 183 Quantum Yield Data for the Photoreaction of M02(02P(0C6H5)2)4 and CICI‘I2CH2CI in Various Nonaqueous Solutions ...................................................... 211 xi Figure LIST OF FIGURES Latimer diagram for a transition metal complex M. The relationship among the 0, 0 transition energy (BO-0) and the ground and excited state redox potentials is described by equations (1) and (2) ............. Schematic illustration of a water-splitting cycle involving photosensitizer, electron relay, and redox catalysts ................................................................................ Water-splitting photochemistry of a bifunctional catalyst involving modified semiconductor particles. CB, conduction band; EF, energy of Fermi level; R, electron relay (e.g. MV2+); and S, sensitizer (e.g. Ru(bpy)32+) (ref 30) .......................................................... Reaction cycle for the conversion of isopropanol to acetone and hydrogen by using Pt2(POP)44' as a photocatalyst .......................................................................... Photoinduced bimolecular reaction of Ir2(p -pz)2(COD)2 with 1,2-dichloroethane ....................... xii Page ..... 3 ..... 6 ..... 8 ..... l3 ..... 16 10 ll Bimolecular photoreaction pathways from a long- lived triplet excited state of an inorganic complex M. kr, radiative decay rate; km, nonradiative decay rate; and kisc, intersystem crossing rate ........................................................................................ Proposed bimolecular photoreaction pathways from a long-lived singlet excited state of an inorganic v . . . complex M . The intersystem crossrng rate, kisc’ IS 18 negligible for a large singlet-triplet energy gap .................. 21 Relative energies of the d-derived molecular orbitals in Dooh M2 and in a D4,, M2L3 complex. The ground state of the d4-d4 M2L3 complex is 1A1g(02n452) .................................................................................. 24 Experimental apparatus for typical photolysis ...................... 47 Experimental apparatus for photoreaction quantum yield measurements .......................................................................... 5 2 (a) 02d structure of quadruply bonded chlorophosphino dimolybdenum M02X4L4; (b) Newman projection of M02X4(LL)2 complexes where X = Cl, Br, I; L and LL = mono- and bi-dentate phosphine ligands, respectively ................................................... 62 xiii 12 13 14 15 16 17 18 Electronic absorption spectra of M02X4(PBu3)4 where X = Cl (——-), Br (---), and I (- - -) ................................... Energies of the lowest electronic states of M02C14(PMe3)4 and M02C14(dmpe)2 (ref 79) ................. Emission of M02C14(LL)2 complexes in CHZCIZ solution at room temperature: (a) M02C14(dmpm)2; (b) M02C14(dppe)2; and (c) M02C14(dmpe)2 ...................... Luminescence of M02C14(LL)2 complexes at 77 K in a CH2C12/toluene glass: (a) M02C14(dppm)2; (b) M02C14(dmpm)2; (c) M02C14(dppe)2; and (d) M02C14(dmpe)2 ............................................................................... Absorption and emission spectra of 62 :2 65* transitions of staggered (a) M02C14(dppe)2 and (b) M02C14(dmpe)2 complexes ......................................................... Absorption and emission spectra of 62 t‘-’ 65* transitions of eclipsed (a) M02C14(dppm)2 and (b) M02C14(dmpm)2 complexes ........................................................ Relative energies of the d-derived molecular orbitals in 02d (i.e. M02C14(PBu3)4) and 02h (i.e. M02C14(dppm)2) symmetries ............................................ xiv 66 7 0 7 3 7 7 82 87 90 19 20 21 22 Transient difference spectra of (a) MozBr4(PBu3)4 in CH2C12 and (b) M0214(PBu3)4 in CH2C12. The spectra were recorded 50 ps and 1 ns, respectively, after the 532-nm pulse of a Nd:YAG laser (FWHM = 25 ps) ................................................................................... 94 Transient difference spectra of the following M02C14(LL)2 complexes: (a) M02C14(dPPm)2. recorded 75 ps after a 355-nm laser excitation pulse; (b) M02C14(dmpm)2, recorded 25 ps after a 532-nm laser excitation pulse; (c) M02C14(dppe)2, recorded 100 ps after a 532-nm laser excitation pulse; and (d) M02C14(dmpe)2, recorded 100 ps after a 355-nm laser excitation pulse ....................................... 98 Transient kinetics for dimethylphosphine complexes: (a) M02C14(dmpm)2 recorded at 600 nm (excitation by 2nd harmonic of a NszAG, FWHM = 25 ps) (b) M02Cl4(dmpe)2, recorded at 450 nm (excitation by 3rd harmonic of a NszAG, FWHM = 25 ps) ...................... 105 Transient kinetics for M02Cl4(dppm)2 in CH2C12 solution. The decay kinetics were recorded at (a) 400 nm and (b) 620 nm (excitation by 3rd harmonic of a NszAG, FWHM = 25 ps) ..................................... 109 XV 23 24 25 26 27 28 29 Transient kinetics of M02C14(dppe)2 in CH2C12 solution. The decay kinetics were recorded at 420 nm (excitation by 2nd harmonic of a NszAG, FWHM = 25 ps) .................................................................................... 1 1 1 Temperature dependence of the corrected magnetic susceptibility of K3M02(HPO4)4 .................................................. 121 X-band (9.460 GHz) EPR spectrum of a frozen solution of K3M02(HPO4)4 in 7.5 M H3PO4 at 5 K ............... 123 Electronic absorption spectrum of M02(HPO4)44' ion in 2 M H3PO4 at room temperature ................................... 126 Electronic absorption spectrum of K3M02(HPO4)4 in 2 M H3PO4 at 25 0C .................................................................... 128 Near-infrared absorption band of K3M02(HPO4)4 in a frozen phosphoric acid solution at 77 K ........................... 131 Cyclic voltammogram of a 2.5 mM solution of the pyridinium salt of M02(HPO4)44' in 2 M H3PO4. The scan rate was 2 mV s'1 ........................................................... 134 xvi 30 31 32 33 34 Electronic absorption spectral changes during irradiation (at > 335 nm) of M02(HPO4)44' in 2 M D3PO4. Due to spectral congestion in the visibile spectral region, the 45-min trace is not illustrated between 350 and 800 nm. The absorbance sensitivity in the visible spectral region is twice that of the infrared spectral region ........................... 139 Absorption changes resulting from irradiating (x > 335 nm) 2 M D3PO4 solutions of M02(HPO4)43‘. The visible absorbance scale is 5 times greater than the near-infrared absorbance scale ............................................ 141 Electronic absorption spectral changes during irradiation (x > 335 nm) of M02(HPO4)42' in 2 M D3PO4. No absorption bands appear in the near-infrared spectral region during the photolysis reaction ............................................................................. 1 4 4 Terminating electronic absorption spectrum of photolyzed (X > 335 nm) phosphoric acid solutions of M02(HPO4)4“' (n = 2, 3,4) ....................................................... 149 Action spectrum of the photolysis reaction: M02(HPO4)43' + H+ _. M02(HPO4)42' + 1/2 H2. Quantum yields were measured by monitoring the 5 -» 5* absorption band of M02(HPO4)43’ ......................... 153 xvii 35 36 37 38 39 40 Ultraviolet absorption bands of M02(HPO4)44‘ (---), M02(HPO4)43' (—), and M02(HPO4)42' (~ - -) in 2 M H3P04 at 25 0C ...................................................................... 155 Photoreaction pathway of "M0208" complexes in acidic solutions, where "M0208" corresponds to M02(aq)x4+, M02(SO4)44', and M02(HPO4)4n' .................... 162 4 Proton quenching of M—M excited states .............................. 165 Structure and labeling scheme for the quadruply bonded complex M02(02P(OC6H5)2)4°2THF. Atoms are represented by their 50% probability ellipsoids. Due to structural congestion, the two tetrahydrofuran molecule are not illustrated ........................................................... 170 Structure and labeling scheme for the mixed-valence complex M02(02P(OC6H5)2)4BF4. Atoms are represented by their 40% probability ellipsoids ............................................................... 172 Cyclic voltammograms of (a) M02(02P(OC6H5)2)4 and (b) M02(02P(OC6H5)2)4BF4 in CH2C12 solution at 23 °C. NBu4PF6 was used as the supporting electrolyte, scan rate = 20 mV s'1 .............................................. 181 xviii 41 42 43 44 45 46 X-band (9.434 GHz) EPR spectrum of a frozen solution of M02(02P(OC6H5)2)4BF4 in CH2C12 solution at 17 K .................................................................................... 186 Electronic absorption and emission spectra of M02(02P(0C6H5)2)4 in CH2C12 solution at 25 0C ................. 189 Electronic absorption spectrum of M02(02P(OC6H5)2)4BF4 in CH2C12 solution at 25 0C ......... 191 Transient absorption kinetics of Moz(02P(OC6H5)2)4 in THF solution. The decay kinetics were recorded at 460 nm and excitation was with a 532-nm Nd:YAG laser pulse (FWHM = 25 ps) .......................................... 194 Electronic absorption spectral changes during irradiation ()‘exc > 530 nm) of M02(02P(OC6H5)2)4 in CICHZCHZCI solution at 25 °C. The visible absorbance scale is 2 times greater than the near-infrared absorbance scale .................................................... 197 FABMS spectrum of a photolyzed ClCHzCHZCl solution of M02(02P(OC6H5)2)4 ................................................... 199 xix 47 48 Electronic absorption spectrum of the product of the reaction between M02(02P(OC6H5)2)4 and 0.5 equivalent of C6H51C12 in CH2C12 solution at room temperature .......................................................................................... 201 Electronic absorption spectral changes during irradiation ()‘exc > 530 nm) of M02(02P(OC6H5)2)4 in CH2C12 solution at 25 °C ............................................................. 206 XX CHAPTER I INTRODUCTION A central theme of inorganic chemistry is the activation of small molecules by transition metal complexes. The main difficulty of small molecule activation chemistry lies in the large kinetic and/ or thermodynamic barriers which typically confront these activation reactions. To this end, inorganic photochemistry can play an important role in the activation of small molecules, because the energy of the electronically excited transition metal complex can be utilized to surmount these reaction barriers. Owing to the fact that small molecule activation reactions involve oxidation-reduction chemistry, the development of the photoredox reaction pathways of transition metal complexes is crucial to the ultimate design of photochemical activation schemes. 1 Since Adamson's initial studies in which the electron transfer reactivity of the LMCT excited state of Co(NH3)5I was demonstrated, the energetics and mechanisms of photoinduced electron transfer of a large number of transition metal complexes have been delineated.2'5 A feature of paramount importance is that the metal complex in its excited state is a considerably stronger oxidant and 1 2 reductant than in its ground state. The reduction potentials of an excited state can be estimated from its spectroscopic energy (0-0 transition defined as E0_0 (M/M*)) and the oxidation and reduction potentials of the ground state species as follows, E*,cd (M*/M') = Bored (M/M') + E 0_0 (M/M*) (1) 13’“ox (M+/M*) = 500x (M+/M) - E 0_0 (M/M*) (2) The Latimer diagram shown in Figure 1 illustrates the simple thermodynamic relations described by equations (1) and (2). This enhanced redox reactivity of electronically excited complexes has been exploited in the experimental testing of electron transfer 6-11 theories, the elucidation of photoinduced electron transfer in 12'” and the activation of small molecules for biological systems, the design of solar energy conversion schemes.l6'21 In the case of the latter, because most small molecule activation processes involve two or more electrons, present photocatalytic schemes have relied on coupling the one-electron chemistry of electronically excited complexes to drive the overall multielectron process. This photochemistry is exemplified in the water-splitting reactions of mononuclear polypyridyl complexes. The tris(bipyridyl)ruthenium (11) ion, Ru(bpy)32+, provides the cornerstone example of this photochemistry. Electronically excited Ru(bpy)32+ ion, [Ru(bpy)32+]*, produced by visible irradiation can readily transfer one electron to acceptors, such as methylviologen (i.e. MV2"’).22’23 The reduced viologen reacts with protons in the presence of colloidal platinum to Figure 1. Latimer diagram for a transition metal complex M. The relationship among the 0, 0 transition energy (130-0) and the ground and excited state redox potentials is described by equations (1) and (2). B 0-0 Figure 1 5 produce hydrogen.24 This chemistry is described in the following scheme, Rutbpm2+ ——’E- Ruibpy>32+* (3) Ru(bpy)32+* + MV2+ —. Ru(bpy)32 + MV+ (4) MV+ + H+ ——PL- MV2+ + 1/2H2 (5) Scheme I The platinum catalyst couples the one electron chemistry of MV+ to the two-electron hydrogen production chemistry by effectively storing the reducing equivalents of the viologen. The photogenerated oxidant Ru(bpy)33+ produced in Scheme I is a strong oxidant. It has been shown that Ru(bpy)33+ can react with water or hydroxide in the presence of a Rqu catalyst under suitable conditions to produce oxygen.25 In principle, the hydrogen and oxygen photochemistry of the Ru(bpy)32+ system can be combined to construct the photocatalytic water-splitting cycle schematically represented in Figure 2.26'28 This photocatalytic chemistry has been realized to a limited extent with bifunctional photosystems employing colloidal 29'33 A schematic illustration of one such semiconductors. bifunctional photosystem is displayed in Figure 3.30 An electronically excited sensitizer donates an electron to the conduction band of the colloidal Ti02 particle. Protons are reduced to hydrogen at platinum dispersed on the surface of the semiconductor particle. The oxidized sensitizer takes an electron from the valence band of the Ti02/Ru02 particle and ensuing oxidation of water is mediated Figure 2. Schematic illustration of a water-splitting cycle involving photosensitizer, electron relay, and redox catalysts. N oSwE fbfibmt e s\ r M, w L E D ME<3 >mummzm m0 Sac/m HZDSQAmHm; m0 .257: Figure 3. Water-splitting photochemistry of a bifunctional catalyst involving modified semiconductor particles. CB, conduction band; EF, energy of Fermi level; R, electron relay (e.g. MV2+); and S, sensitizer (e.g. Ru(bpy)32+) (ref 30). V 1 8 9 em 0: . n\ \ -mo + E m: \ \\ \. x r A+ENIV om mm +m WW mismd s\\\\\\\\ \\ can a ooooooooooooo 10 by RuOZ. The efficiency of the cycles such as those shown in Figures 2 and 3 are extremely low for several reasons. The reaction yield for H20 oxidation by RuOZ is negligibly small. A series of studies by 2034 and others35 have shown that a more Sutin and coworkers, efficient system for water oxidation involves cobalt(II) complexes as the water oxidation catalysts. The multistep mechanism which accounts for 02 production is shown in Scheme II34b where the cobalt catalyst and Ru(bpy)33+ is represented by [C02+] and [Ru(III)]. [Co2+] + 2H20 :2 Co(OH)2 + 211+ (6) Co(OH)2 + [Ru(III)] 2 Co(OH)2+ + [Ru(II)] (7) Co(OH)2+ + [Ru(III)] -> Co02+ + [Ru(II)] + H20 (8) Co02+ + H20 -+ [Co2+] + H202 (9) Coo2+ + [Co2+] .. CoOCo4+ -> "C0203"s (10) 2[Ru(III)] + H202 _. 2[Ru(II)] + 02 + 2H+ (11) 4Ru(bpy)33+ + 21120 .. 4Ru(bpy)32+ + 411+ + 02 (12) Scheme 11 Despite the success of replacing Ru02 with other catalysts, overall low efficiencies are retained owing to the limited diffusion of reactants, efficient back reaction of primary photoproducts, and difficulties with channelling reactivity selectively along the desired routes in these complicated multifunctional schemes. Even if these 11 problems are overcome, there remains one major problem confronting these systems: Ru(bpy)32+ multielectron chemistry is limited to water-splitting and is not applicable to other small molecule (e.g. N2, C02, CO) activation reactions due principally to the lack of suitable catalysts, such as Pt and RuOZ, to couple the single electron transfer steps to the desired multielectron transformation. An alternative approach to the ultimate design of multielectron photochemical schemes is to develop photochemical excited state reaction pathways in which an electronically excited complex can participate directly in multielectron processes. With such systems, the problems inherent with coupling one electron steps to drive overall multielectron processes are circumvented. A promising approach to photoinduced multielectron reactivity relies on coupling formal oxidation state changes of metal centers into a polynuclear core to drive the multielectron transformation. An issue of paramount importance to this approach is the necessity to maintain the structural integrity of the polynuclear core upon light absorption. For instance, a representative metal-metal bonded polynuclear system is dimanganese decacarbonyl Mn2(CO)10. From molecular orbital theory, the 12 electrons of this d7-d7 dimer reside in predominantly metal orbitals of TI and 6 symmetries. The remaining two electrons populate the metal-metal O bonding orbital formed from the overlap of the 4d22 orbitals. The lowest energy transition in Mn2(CO)10 arises from the promotion of an electron from the O bonding orbital to 0* antibonding orbital. Consequently, photoexcitation of these complexes significantly weakens the metal- metal bond, and photofragmentation to give two manganese 12 pentacarbonyl free radicals is the dominant excited state reaction pathway.36’37 As is typical of most singly bonded metal-metal systems,38'41 the photon is consumed for bond breaking and thus the polynuclear complex in its excited state is not preserved for redox chemistry. This problem of photofragmentation can be avoided by a variety of approaches. The structural integrity of metal-metal cores of dinuclear species can be preserved by bridging the metal-metal core with bidentate ligands. This approach has been most successfully exploited with d8-d8 complexes where bridging ligands such as pyrophosphite in Pt2(POP)44' (POP : (HOZP)ZO),42 pyrazolyl in Ir2(p-pz)2(COD)2 (pz = pyrazolyl, COD = cyclooctadiene)43 and diisocyanoalkanes in Rh2((CN)CnH2n(NC))444 chelate a dinuclear metal center possessing no formal metal-metal bond. Simple molecular orbital arguments suggest that the respective ground and excited state configurations of these binuclear d8-d8 complexes are represented as (d22)2(d22"‘)2 and (dzz)2(d22*)l(pz)l. Promotion of an electron from the metal-metal antibonding framework to the bonding orbital results in an increased metal-metal interaction in the 45-47 excited state. Electronic absorption and emission spectra and ”'50 of the ground and excited states of these Raman spectroscopy dimers substantiate this prediction. With the molecular structure preserved in the excited state, a rich photooxidation chemistry of d8-d8 complexes has been observed.51’54 For instance, Pt2(POP)44' photocatalytically converts isopropyl alcohol into acetone and hydrogen.51 The reaction pathway for this transformation is shown in Figure 4. In this 13 Figure 4. Reaction cycle for the conversion of isopropanol to acetone and hydrogen by using Pt2(POP)44' as a photocatalyst. 14 Pt2(POP)44‘ Cf“ CH3CCH3 ll Pt2(POP)4H4' CH3CCH3 on CH CCH Pt2(POP)4H4‘ ill 3 Pt2(Pop)44-* Pt2(POP)4H24‘ Pt2(POP)44‘ “V H2 Figure 4 relic me: (Cl equ The (11 llll of ch st ex ex ex co COi F161 3101 Con ltdu 15 reaction scheme, electronically excited Pt2(POP)44' abstracts the methine hydrogen of isopropanol to give Pt2(POP)4H4' and (CH3)2COH radical. Subsequent reaction of this radical with another equivalent of Pt2(POP)44' results in the production of Pt2(POP)4H4'. The Pt(II)Pt(III) mixed-valent Pt2(POP)4H4' intermediate undergoes disproportionation to give Pt2(POP)44' and Pt2(POP)4H24’ which has recently been observed.51b Reductive elimination of H2 from Pt2(POP)4H24' completes the photocatalytic cycle. The overall two- electron reaction is comprised of a photoinduced one-electron step followed by an ensuing radical reaction. A similar reaction pathway, observed in the reaction of Ir2(p -pz)2(COD)2 with 1,2- dichloroethane,52 is shown in Figure 5. In this scheme, the ensuing thermal reaction of the photoinduced radical is the terminating step of the reduction of 1,2-dichloroethane to ethylene. Careful examination of the above systems reveal a general characteristic of all photosystems studied to date : a triplet excited state yields photoproducts of triplet spin.2'5 More generally, excitation of a transition metal complex usually produces a singlet excited state which intersystem crosses to the lowest energy triplet excited state. Relatively small energy gaps between the singlet and corresponding triplet state in conjunction with strong spin orbital coupling results in extremely efficient conversion of the singlet state to the triplet state. The triplet excited state reacts to exclusively yield primary photoproducts arising from electron transfer (8 = l) or atom abstraction (s = 1). This chemistry is summarized in Figure 6. Conversely, multielectron processes such as oxidative-addition, reductive-elimination, and atom transfer inevitably yield singlet 16 Figure 5. Photoinduced bimolecular reaction of Ir2(p-pz)2(COD)2 with 1,2-dichloroethane. 17 h V ‘ 06’ngch r ...... Egg/Ecx‘cc l/ l I 0’ re ClCHzCH2Cl _ 'CH2CH2C l + Cl fl / 4171 P 1‘13 OVA/1 47 “Nb + C2H4 Figure 5 fl ‘‘‘‘‘ Ir—Ir~. l/ \“' ClCH2CH2C l .l 18 Figure 6. Bimolecular photoreaction pathways from a long-lived triplet excited state of an inorganic complex M. kr, radiative decay rate; km, nonradiative decay rate; and kisc, intersystem crossing rate Bimolecular photoreaction pathways from long-lived triplet excited state of an inorganic complexes M. kr, radiative decay rate; km, non- radiative decay rate; kiso intersystem crossing rate. l9 1(M*) ll \ kisc kabs kct HX hv i kl' knr kr km A/D M-H + X- L atom abstraction M+/' + A'lD+ electron transfer (l M Figure 6 20 products (Figure 7). Accordingly, spin considerations suggest that successful multielectron photoreagents will derive their reactivity from singlet electronic excited states. The successful development of singlet excited state chemistry, however, will require that intersystem crossing from the singlet state to triplet state be inhibited. In addition a problem of equal significance in the development of singlet excited state chemistry is the short lifetimes of the photoactive state. Unlike tn'plet states which possess lifetimes from microseconds to milliseconds owing to the spin-forbiddeness of the relaxation process back to ground state, singlet state decay (fluorescence) is usually only a few nanoseconds or even shorter. These intrinsically short lifetimes do not permit the singlet excited state to easily participate in bimolecular reactions. It is undoubtedly these two features of singlet excited states that has inhibited their reactivity. Thus, a critical issue to this thesis is whether inorganic photosystems possessing lowest energy excited states of pure singlet character exist with lifetime long enough to permit bimolecular reactivity. This question is addressed by beginning with multiply bonded metal-metal (M-n-M) dimers. Since their discovery in 1964, numerous M-A-M complexes, comprised primarily of rhenium, chromium, molybdenum and tungsten metal cores with a variegated array of ligating systems, have been discovered primarily due to the 55 efforts of Cotton and coworkers. The prevalence of quadruply bonded metal complexes has stimulated numerous theoretica156'63 164-78 and experimenta investigations directed toward elucidating their electronic structures. 21 Figure 7. Proposed bimolecular photoreaction pathways from a long- lived singlet excited state of an inorganic complex M'. The intersystem crossing rate, kisc’ is negligible for a large singlet-triplet energy gap. 22 1(M'*> ll km kox kisc hv HX BO kr knr 1(X-M'-H) oxidative-addition 1(M'=O) + B oxygen transfer Ml Figure 7 23 The general molecular orbital diagram depicted in Figure 865 has evolved from these studies. The metal-metal bonding in L4M ML 4 quadruply bonded dimers is most easily derived from a MM core (Dock symmetry) in which the z axis of a right-handed Cartesian coordinate system is chosen to lie along the metal-metal axis. Linear combinations of the dz29 (dxz, dyz), and (dx2-y2,dxy) atomic orbitals on each metal produce bonding and antibonding molecular orbitals of 0, TI, and 5 symmetries, respectively. The n and 5 molecular orbitals are each doubly degenerate in aDooh point group. From the nodal theorem and atomic orbital overlap considerations, the molecular orbital energy ordering of the diatomic MM core shown in Figure 8 is obtained. The equidistant positioning of eight ligands in aL4MML4 complex perturbs the bonding of the diatomic metal core by lowering the symmetry of the system to D 4),, thus, splitting the 5 orbital degeneracy. By definition of the coordinate system in Figure 8, the ligands will lie along the x and y axes and their orbitals will overlap with dx2-y2 as well as s, px, py (which are not shown in Figure 8) to form its four metal-ligand bonds. The linear combination of the ligand and metal dx2-y2 orbitals produces 5(blg) and 5*(b2u) molecular orbitals which are essentially metal-ligand O antibonding in character. This antibonding interaction causes the energies of these orbitals to increase dramatically in energy. The energies of the O, n, and 5(dxy) molecular orbitals, for the most part, are unperturbed as the ligands approach the metal core and the L4MML4 molecular orbital diagram for D4}, symmetry shown in Figure 8 is obtained. Additionally, the characteristic properties of a diamagnetic ground 24 Figure 8. Relative energies of the d-derived molecular orbitals in Dooh M2 and in a D4;1 M2L8 complex. The ground state of the d4-d4 M2L8 complex is lAlg (0211452). 25 __________ -— 821,03“) 0* .— ........ b2u(dx2-y2) ’o::—— blg(dx2-y2) ‘ e egmi‘) 11* """ . > x, x i: __. b1u(5*> :: 5* ’ .’ m bzgas) 5 " ...... _ll_ _.l.L_ eu(fl) n ....... _______________ Ji— alg(c) O ________ y y {y {Y L /L / / / I L I I- / / x/ / / x/ l X/L '5” i Figure 8 26 state, eclipsed ligand geometry (which is necessary for dxy overlap), and exceptionally short metal-metal bond exhibited by quadruply bonded complexes are also explained by this bonding scheme. According to this simple molecular orbital picture, the lowest energy transition arises from the promotion of an electron from the 5 to 5* molecular orbital. Polarized electronic absorption spectroscopic investigations of various quadruply bonded metal complexes verify that the lowest energy absorption corresponds to the 52 -> 55* transition thereby confirming molecular orbital predictions,65.68a,o9 Though conceptually quite appealing, the molecular orbital model does not provide an accurate quantitative picture of the electronic structure of quadruply bonded metal dimers, especially in terms of 5 bonding. The poor overlap of the dxy orbitals involved in 5 bonding suggests that a valence bond description of the electrons in these orbitals would be more appropriate. Such an approach was used in an ab initio calculation of the RezClgz' electronic structure.58 Some of the results of which are outlined in Table 1. In the valence bond model, the 1A1 g ground state has one electron in each dxy orbital. At slightly higher energy (0.4 eV) is the 3A2u excited state, which still has one electron in each dxy orbital, though now triplet paired. This state correlates with the 3(55*) excited state of the molecular orbital model. Lying much higher in energy are two ionic singlet excited states that arise from antisymmetric (lAzu, 1.8 eV) and symmetric (lAlg, 2.2 eV) combinations of configurations with both dxy electrons on one metal center. These two states correlate with the 1(55*) and l(5"'2) 27 Table 1. State Orderings in d4-d4 Complexes: [Re2C18]2' Ea/eV States 2.6 3A2.,(n n *) 2.6 1Egmfi‘“) 2.2 1A1g(5*5*)31[(xyA)(XyA) + (xyaxyan 1.3 1A2u(55*): lltxyAXxyA) - (xyBxyBH 1.3 3Eg(n5*) 0.4 3A2u(66*): 3[(xyA)(xyB)l 0.0 1A1g(55)3 11(XYA)(XYB)1 a Reference 58. 28 molecular orbital configurations, respectively. An important feature of this model is that it predicts a large 1A2u 4 3A2u energy gap and a relatively small separation between the lAlg ground state and the 3A2u level. This prediction is consistent with energy gaps estimated from magnetic susceptibility measurements of MAM complexes.79 Within this manifold of four states, the only spin and dipole allowed transition from the lAlg ground state is to the 1A2u excited state. This effectively corresponds to a metal-metal charge transfer (MMCT) transition and correlates with 1(52) -> 1(55*) excitation. On the basis of this valence bond model, the large 1(55"')-3(55"') energy gap should inhibit intersystem crossing and consequently the 55* excited state produced upon excitation should be of pure singlet character. This prediction has been verified by several experimental studies. The good 0-0 overlap between the absorption and emission spectra of M-Z-i-M dimers at low temperature shows that the long- lived excited states (I ~ 100 ns)8o’82 of quadruply bonded metal dimers directly correspond to the 1(55*) excited state.65’81 Moreover, the long lived excited state is not quenched by typical triplet energy transfer acceptor molecules,83 thereby further verifying the singlet character of the 55*emissive state Besides these unique long-lived singlet excited states, the quadruply bonded metal dimers possess many other features that presage their utility as polynuclear multielectron photoreagents. First, the longevity of the 55 * excited state should permit the electronically excited metal complex to participate in bimolecular reactions. Secondly, MAM complexes should be capable of maintaining their structural integrity upon irradiation. Theoretical 29 estimates predict the overall energetic contribution of the 5 bond to 58 the metal-metal bond energy to be less than 10%. Irradiation of the 52 .. 55* transition will, therefore, only slightly weaken the metal bond and photoinduced metal-metal bond cleavage reactions will be excluded in these complexes. Third, the multiple metal-metal bond may act as an electron source or sink in multielectron transfer reactions while maintaining a strong metal-metal interaction. Walton and coworkers have prepared a series of L4MML4 triply bonded metal complexes with O 2 n 4 5 25 * 2 ground state electronic 84'86 In all cases investigated, these compounds configurations. exhibit facile one electron oxidation. The absence of the 5 bond in 02114525’“2 triply bonded complexes does not greatly perturb the L4MML 4 structural unit as evidenced by the crystal structures of Re2Cl4(PhPMe2)4n+ (n = 0, 1 ,2) complexes which show only slightly shortened Re-Re bond distances (r(Re-Re) of Re2Cl4(PhPMe2)4n+, n = o, 1, 2 is 2.241(1)A, 2.218(1)A and 2.215(2)A, respectively).87 The ability of the multiple metal-metal bond to undergo only minor structural reorganization upon changes in 5 and 5* occupancies may greatly facilitate the photoredox chemistry of quadruply bonded metal complexes. Fourth, the metal-metal core is coordinatively unsaturated and substrates may readily add to the redox-active metal center.88'90 Finally, the lowest energy 52 -* 55* transition falls in the visible spectral region, and low energy photochemistry is therefore possible. It is rather surprising that the photochemistry of these systems has virtually been ignored in view of the attractive photochemical properties of quadruply bonded metal complexes. Most 30 photochemical studies to date have been confined to high energy reaction pathways. The Re2C182' 91 and M02Cl4(PBu3)492 dimers exhibit solvent assisted metal-metal bond cleavage chemistry upon UV irradiation. More generally, however, the metal—metal bond is retained and UV irradiation of M-4-M dimers in acidic solutions results in photooxidation as shown by the following equations,93'95 M02(SO4)44' + H+ -—-}l]-)—- M02(SO4)43- + 1/2 H2 (13) M02(aq)4+ + 2H+ _11_1_{_. M02(aq)(}J-OH)24+ + H2 (14) M02C184' + W “V 0‘ A... M02C13H3' (15a) M02C13H3‘ + W + Cl' —- M02C193' + H2 (15b) Photochemistry of M-4—M dimers promoted by visible light has only recently been observed. Dichloromethane or acetonitrile solutions of Re2C182' can be photooxiditized (xexc > 600 nm) to the confacial bioctahedral, RezClg', in the presence of electron acceptors TCNE (tetracyanoethylene) or DDQ (2,3-dichloro-5,6-dicyano-1,4-benzo- quinone) and chloride ion.96 However, the photochemistry basically 2 entails simple one-electron quenching of the Re2C13 ' excited state to yield Re2C19' followed by trapping by chloride ion. In view of the potentially rich multielectron photochemistry of MA-M dimers, studies aimed at the systematic development of 1(55*) excited state chemistry were undertaken. A prerequisite for elucidating the photochemistry of MAM dimers is a clear 31 understanding of their excited state dynamics. Accordingly, time- resolved picosecond laser spectroscopic studies of M02X4L4 and M02Cl4(LL)2 (where X = Cl, Br, I; L and LL = mono- and bi-dentate phosphine ligand) dimers were performed at Brookhaven National Laboratory. Transient absorption and kinetics results of these studies will presented in Chapter 111, along with a discussion of the excited state properties. Chapter IV describes our initial efforts to develop multielectron transformations by quenching reactive hydrides from MAM dimers. Investigations of the spectroscopy, electrochemistry and the photoinduced two-electron transfer chemistry of M02(HPO4)4"' (n = 2, 3, 4) in acidic solution are presented. An important result of this work was the discovery that protons efficiently quench 1(55*) excited states of MAM dimers thereby circumventing multielectron reactivity. Accordingly, Chapter V describes studies focussed on the photochemistry of quadruply bonded M02(II,II) diphenyl phosphate in non-aqueous solutions. The first observations of multielectron chemistry of M-4- dimers promoted by visible light are presented. Specifically, two- electron photoreductions of organic substrates are effected upon irradiation of nonaqueous solutions containing M02(OZP(OC6H5)2)4 with wavelengths greater than 500 nm. CHAPTER II EXPERIMENTAL A. Syntheses 1. General Procedures. Synthesis of all complexes were performed by using standard Schlenk-line techniques. All chemicals were reagent grade and were used as received unless otherwise noted. Solvents were rigorously deoxygenated and dried prior to use. For previously prepared compounds, electronic absorption spectroscopy, cyclic voltammetry and NMR (when previously reported) were employed for characterization. The details for characterization of new compounds are presented with the discussion on the preparative method. 2. Syntheses of Dimolybdenum Complexes. a. M02(02CCH3)4.97 Molybdenum acetate, the starting compound for most quadruply bonded dimolybdenum species, was prepared by refluxing 10 g of Mo(CO)6 (37.9 mmole) (Aldrich Chemical Co.) in 250 ml of o-dichlorobenzene which contained 30 ml of CH3COOH and 10 ml of (CH3CO)20. The solution was refluxed for 5 32 33 h. The needle-shaped, yellow crystals which formed with cooling on the sides of the round bottom flask were filtered and washed by EtzO three times, dried, and stored under vacuum. Molybdenum acetate is fairly air-stable but will decompose if exposed to air longer than one week. b. K4M02C18.98 Potassium chloride (6 g, 80 mmole) was added to 200 ml of 12 M HCl(aq) with stirring. The solution immediately changed from yellow to red upon the addition of 5 g of M02(02CCH3)4 (11.7 mmole). After 1 h of stirring, a bright red precipitate was collected, washed with EtOH (0 0C) and EtzO, and then dried under vacuum. Similar to M02(02CCH3)4, K4M02C18 is fairly air-stable. c. (NH4)5M02C19°H20.99 The ammonium salt of M02C184‘ was prepared by the same procedures as used for the preparation of K4M02C18, except 4 g of NH4C1 (76 mmole) was used instead of KCl. A reddish-purple solid precipitated after 1 h of stirring. The solid was collected, washed three times with EtZO and dried under vacuum. d. CS3M02Br3.100 A 100 ml 48% HBr(aq) solution containing 1 g of M02(OZCCH3)4 (2.3 mmole) was heated to 60 °C for l h, after which 2.5 g of CsBr in 50 ml of 48% HBr(aq) was added with stirring. The yellow-brown precipitate was collected by suction filtration, washed in succession with EtOH and EtZO, and dried under vacuum. The solid is air-stable. 34 101,81 - e. M02C14(PMe3)4. Dimolybdenum chlorophos- phine was prepared by either of two methods: (1) 0.5 g of (NH4)5M02C19°H20 (0.807 mmole) was added to 50 ml of a pre- mixed MeOH solution which contained 0.32 ml (3.23 mmole) of PMe3 (Strem chemicals). The final solution was stirred for 3 days, during which the color of the solution slowly changed from red to purple. The purple solid, which formed on the sides of the flask over this period, was collected by suction filtration, washed with 5 ml of H20 to dissolve unreacted (NH4)5M02C19'H20, and finally washed three times with 10 ml aliquots of MeOH. The compound was dried and stored under vacuum. (2) The second method employed Moz(02CCH3)4 as a starting material. To 40 ml of THF, 0.5 g of M02(02CCH3)4 (1.17 mmole) was dissolved. Under an argon counterflow, 0.59 ml (4.65 mmole) of Me3SiCl (Aldrich Chemical Co.) and 0.47 ml of PMe3 (4.74 mmole) were added separately by syringe. The blue solution was refluxed for 10 h. The blue residue collected from removing solvent from the reacted solution was column chromatographed on Florasil by using CH2C12 as the eluting solvent. The first blue band was collected, and the solvent was removed under vacuum to produce a blue solid. Compound produced by each of these methods exhibit identical properties. Absorption show strong absorption bands at 583 and 325 nm (E = 3110, 3350 M'1 cm'l, respectively) and a weak band at 440 nm (E = 200 M'1 cm'l). M02C14(PMe3)4 is air-sensitive and decomposes in l h when solid compound is exposed to air. Solutions of the dimer decompose within minutes. 35 f. M02C14(PBu3)4.102 This dimer is most easily prepared by Method 1 of Section A.2.e : 0.5 g of (NH4)5M02C19°H20 (0.807 mmole) was added to 40 ml of a pre-mixed MeOH solution containing 0.81 ml of PBu3 (3.25 mmole) (Aldrich Chemical Co.). The purple solution was refluxed for 3 h to yield purple crystals which were collected by suction filtration, and washed with 5 ml of H20 to remove unreacted (NH4)5M02C19'H20. The purified solid was subsequently washed with MeOH and EtzO, dried, and stored under vacuum. M02C14(PBu3)4 decomposes in air within few minutes. In CH2C12 solution, M02C14(PBu3)4 exhibits three strong absorption bands at 588, 330 and 287 nm (e: 3150, 3820, 3050 M-1 cm-l, respectiveIY). and a weak band at 454 nm (E = 260 M'1 cm'l). g. MozBr4(PBu3)4.100 A 30 ml MeOH solution containing 0.5 g of CS3M02Br8 (0.407 mmole) and 0.405 ml of PBu3 (1.63 mmole) was stirred for 3 h. The deep blue crystals, which formed during stirring, were collected, washed with H20, EtOH and Et20 and then dried and stored under vacuum. This air-sensitive complex exhibits two strong absorption bands at 600 and 352 nm and two weak bands at 485 and 460 nm (e = 3520, 5130, 210, and 162 M-1 cm'l, respectively). h. M0214(PBu3)4. Procedures analogous to those described for the synthesis of M02C14(PMe3)4 by Method 2 were employed. A 40 ml THF solution charged with 0.5 g of M02(02CCH3)4 (1.17 mmole), 0.67 ml (4.7 mmole) of Me3SiI (Aldrich Chemical Co.) and 1.17 ml of PBu3 (4.7 mmole) was refluxed for 16 h. 36 The residue, collected by pumping the reacted solution to dryness, was purified with chromatographic procedures identical to those previously described (Section A.2.e). This air-sensitive compound strongly absorbs at 590 nm. i. M02C14(dppm)2.103 0.77 g (2.0 mmole) of dppm (dppm = Bis(diphenylphosphino)methane, Aldrich Chemical Co.) was added to 40 ml of MeOH which contained 0.5 g of (NH4)5M02C19°H20 (0.807 mmole) to yield a blue-green solution. This solution was refluxed for 3 h. The blue crystals, formed during refluxing, were collected by suction filtration, and consecutively washed with H20, EtOH and EtzO. The blue solid is air-stable but in solution, it is air- sensitive, decomposing within an hour after exposure to air. Deoxygenated CH2C12 solutions of M02C14(dppm)2 exhibit two strong absorption bands at 634 nm (5 = 2490 M-1 cm-l) and 325 nm (e = 5600 M'1 cm'l) and a weak band at 462 nm (E = 900 M'1 cm'l). j. M02Cl4(dmpm)2.107 0.5 g of (NH4)5Mo2c19-H20 (0.807 mmole) and 0.26 ml (1.63 mmole) of dmpm (dmpm = bis(dimethylphosphino)methane, Strem Chemicals) were reacted in 40 ml of MeOH solution following a procedure analogous to that used for the preparation of M02C14(dppm)2. M02C14(dmpm)2 is an air- sensitive compound, and in CH2C12 solutions, exhibits two absorption bands at 604 nm (e = 1730 M'1 cm'l) and 426 nm (e = 270 M'1 cm'l). 1H NMR: 1.51(s), 3.08(p) (1H,}, = 5.1 Hz). 37 k. M02Cl4(dppe)2.103:105 This dimer is prepared by refluxing a 40 ml THF solution charged with 0.5 g of M02(OZCCH3)4 (0.17 mmole), 0.59 ml of Me3SiCl (4.65 mmole) and 0.96 g (2.4 mmole) of dppe (dppe = bis(diphenylphosphino)ethane, Aldrich Chemical Co.) for a minimum of 12 h. The solution changed from deep red to blue within 15 min, and then to mauve after 5 h of refluxing. The mauve precipitate was collected by suction filtration, and washed consecutively with H20, EtOH and EtZO. M02C14(dppc)2 is a very air-stable compound in both solid and solution forms. In CHzClz solutions, the compound exhibits four electronic absorption bands at 762, 548, 469, and 345 nm (E = 1030, 210, 780, and 8050 M '1 cm'l , respectively) I. M02Cl4(dmpe)2.106 This complex is prepared according to the previously described procedure for the synthesis of Mo2C14(PMe3)4 by Method 2. A 40 ml THF solution which contained 0.5 g of M02(02CCH3)4 (1.17 mmole), 0.59 ml of Me3SiCl (4.65 mmole) and 0.4 ml (2.4 mmole) of dmpe (dmpe = bis(dimethyl- phosphino)ethane, Aldrich Chemical Co.) was refluxed for 8 h. The volume of the final reddish-brown solution was slowly reduced to about 10 ml over 2 h by vacuum distillation. The reddish-brown crystals which formed during solvent evaporation were collected by suction filtration under a heavy argon flow. M02C14(dmpe)2 is an air-sensitive compound which decomposes within an hour in air. The absorption spectrum of this dimer exhibits three weak bands at 803, 565, and 445 nm and a strong absorption at 345 nm (E = 210, 190, 300, and 3200 M'1 cm'l, respectively). 38 m. M02(HPO4)44' Anion. The M02(II,II) phosphate complex was synthesized by ligand exchange reactions. Solutions of the dimer were prepared by three methods: (1) K4Mo2C18 (0.1 g) was added to 10 ml of a 1 M H3PO4 solution and stirred under purified argon for 4 h; (2) 0.1 g of K4M02(SO4)4107 was added with stirring to 2 M H3PO4 and allowed to react for 2 h; and (3) 1 ml of concentrated H3PO4 was added to 10 mM solution of M024+ 108 aquo dimer blanketed with argon. For each of these methods, a pink solution was obtained with a characteristic absorption band at 516 nm. Aqueous solutions of the dimer are extremely sensitive to oxygen decomposing within minutes to dark brown uncharacterized solutions. Even in the absence of oxygen, the quadruply bonded dimer slowly reacts over days to produce blue-gray solutions which exhibit a near-infrared absorption profile diagnostic of M02(HPO4)43' (vide infra). Solutions of the quadruple-bond dimer prepared by each of the three methods are contaminated inevitably with M02(HPO4)43', and accordingly, addition of K2HPO4 to any of the above solution yields solid K4M02(HPO4)4 contaminated with K3M02(HPO4)4. However, addition of zinc amalgam to these solutions smoothly converts M02(HPO4)43' to the quadruply bonded M02(HPO4)44' species, as evidenced by the disappearance of the near-infrared absorption band of M02(HPO4)43'. In fact, stirring 2 M H3PO4 solutions of K3M02(HPO4)4 over zinc amalgam provides a convenient route to the preparation of pure M02(HPO4)44' solutions. 11. K3M02(HPO4)4. The potassium salt of M02(HPO4)43' was prepared from solutions of M02(HPO4)44'. Addition of Pt foil (1 39 cm2) to 25 ml of 1 M H3PO4 containing M02(HPO4)44’ (0.4 mmole) quantitatively produces M02(HPO4)43' in 2 h. Dropwise addition of a saturated solution of KZHPO4 leads to the immediate precipitation of a fine blue-gray powder. The product was filtered under vacuum and washed with three 5 m1 portions of 1 M H3PO4 at 0 oc. Aliquots for all washings were delivered by cannula. The solid was dried and stored under vacuum. Alternatively, a more facile route to the mixed-valence dimer involved the addition of 0.505 g of K4M02C13 (0.800 mmole) to 15 ml of degassed 1 M H3PO4. After the solution was stirred at room temperature for 1 h under argon, 0.108 g of K28203 (0.400 mmole) was added to the solution via a sample charger. The red solution quickly turned blue, and a blue-gray powder precipitated from the solution. The solid was collected by using Schlenk-line techniques. Anal. Calcd (Found) for K3M02H4P4016: Mo, 27.68 (27.60). Solids and solutions of the mixed-valence dimer are sensitive to oxygen, but the compound is indefinitely stable under inert atmospheres. o. Py3M02(HPO4)4Cl.109 A solution of 2.5 m1 2 M H3PO4, charged with 0.1 g of K4M02C13 (0.16 mmole), was slowly mixed with 0.3 g of pyridinium chloride (2.62 mmole) dissolved in 2.5 ml 2 M H3PO4. The mixture was allowed to stand in air and after 2-3 days, deep purple crystals had formed. A more facile synthesis of M02(III,III) phosphate involved the use of K28208 (0.135 g, 0.5 mmole) as an oxidant which was added with stirring to 10 ml of a 1 M H3PO4 solution containing 0.316 g of K4M02C18 (0.5 mmole). The solution turned from red to colorless to violet within minutes, and a 40 violet solid precipitated immediately after the addition of 0.202 g of CsCl (1.2 mmole) in 10 ml of 1 M H3PO4. The solid was collected by suction filtration under argon. p. M02(02P(OC6H5)2)4. The M02(II,II) diphenyl phosphate was prepared by a simple ligand substitution reaction. The anion (C6H50)2P02' was generated by stirring 3.25 g of (C6H50)2P(O)OH (l3 mmole) with 0.702 g of NaOMe (13 mole) in 60 ml of MeOH. After 10 min, 0.5 g of (NH4)5M02C19°H20 (0.807 mmole) was added to the above solution which was then heated to reflux for 3 h. A pink precipitate, formed during refluxing, was collected by suction filtration and washed with three 15 ml portions of MeOH. Aliquots for all washings were delivered by cannula. The solid was dried and stored under vacuum. Yield: 76%. Suitable crystals of M02(02P(OC6H5)2)4-2THF for X-ray structural determination were obtained by layering a tetrahydrofuran solution of M02(02P(OC6H5)2)4 with cyclohexane in a Schlenk tube. Anal. Calcd (Found) for M02C43H4016P4'2C2H30: Mo; 14.4 (14.87); P, 9.29 (9.17); C, 50.46 (49.93); H, 4.24 (4.15). This newly synthesized dimolybdenum diphenyl phosphate complex is extremely air- sensitive. It decomposes within seconds upon exposure to air. This dimer is not soluble in most organic solvents, but it exhibits fairly good solubilities in halogenated hydrocarbons and THF. A CHZCIZ solution of M02(02P(OC6H5)2)4 exhibits a moderate intensity band at 515 nm , which is very similar to M02(HPO4)44' in acidic solution. 41 q. M02(02P(OC6H5)2)4BF4. The mixed-valent M02(II,III) diphenyl phosphate complex can be prepared by direct oxidation of the M02(II,II) complex. 1.001 g (0.85 mmole) of M02(02P(OC6H5)2)4 and 0.099 g (0.85 mmole) of NOBF4 (Aldrich Chemical Co.) were rigorously stirred in 30 ml of CH2C12. After 45 min the solution changed from pink to green. The volume of solution was reduced to 10 ml by vacuum distillation, whereupon 50 ml of cyclohexane was added. A green precipitate formed immediately. The solid was collected by suction filtration. Anal. Calcd (Found) for M02C43H40016P4BF4 : C, 45.20 (41.23); H, 3.16 (3.15); P, 9.71 (9.33); F, 5.96 (6.13). Yield 85%. X-ray quality crystals were obtained by layering a dichloromethane solution of M02(OZP(OC6H5)2)4BF4 with cyclohexane in a Schlenk tube. The Mo2(II,III) diphenyl phosphate complex is not as air-sensitive as the corresponding M02(II,II) dimer. Solid exposed to air is stable but not longer than 30 minutes. However, solutions of M02(02P(OC6H5)2)4- BF4 is very air-sensitive; both the solid and solution are thermally unstable, decomposing when heated higher than 80 0C. B. Solvent Purification l. Solvents Used for Syntheses. Hexane, cyclohexane and diethyl ether were refluxed over Na for no less than 6 h and freshly distilled prior to use. MeOH was also refluxed with Na. The amount of Na added to MeOH was 20% more than the amount required to react with the water contained in a freshly opened bottle (which can be calculated from the water content shown on the label of the bottle). 42 Benzene, toluene and tetrahydrofuran were refluxed with sodium and small amount of benzophenone to form the blue or purple ketyl radical. Halogenated hydrocarbons such as CH2C12, ClCHzCHZCl, and CHzBrz were refluxed over P205 for at least 6 h prior to use. Acetonitrile and acetone was purified by refluxing over CaH2 and anhydrous CaSO4, respectively. 2. High Purity Solvents. All solvents used for electrochemical, spectroscopic and photochemical experiments were purchased from Burdick and Jackson Laboratories (spectroscopic grade). Purified solvents were stored in glass containers which consisted of a 1 liter flask equipped with a Kontes high-vacuum valve. Dichloromethane, 1,2-dichloroethane and isopropanol were degassed by seven freeze-pump-thaw cycles and then vacuum distilled onto activated (heated to 250 0C under dynamic vacuum (10"6 torr) for 12 h) Linde 4 A molecular sieve contained in the storage flask. Hexane, cyclohexane, and diethyl ether were also degassed by seven freeze-pump-thaw cycles and vacuum distilled onto sodium-potassium alloy. Tetrahydrofuran, benzene, and toluene, again after seven freeze-pump-thaw cycles were distilled into a flask containing sodium-potassium alloy with a small amount of benzophenone. The purple ketyl form of benzophenone formed over a two-day period. Acetonitrile was degassed by seven freeze-pump- thaw cycles and vacuum distilled onto activated Linde 3 A molecular sieve. 43 C. Instrumentation and Methods 1. Absorption Spectroscopy. Absorption spectra were recorded with Cary 17D or Cary 2300 spectrophotometers. Extinction coefficients were determined by using high-vacuum cells consisting of a l-cm quartz cuvette and a 10-ml side arm. These two chambers were separated by two Kontes high-vacuum quick-release teflon stopcocks. For measurements of molar absorptivity coefficients in nonaqueous solution, weighed samples were placed in the cuvette and isolated by the Kontes valve. The appropriate high purity solvent was transfered to the 10-ml side arm by bulb-to-bulb vacuum distillation, and three subsequent freeze-pump-thaw cycles were performed before mixing with the sample. After the absorption spectra was recorded, the solution were pipetted out of the cell and the volume was measured. For measurements in aqueous solutions, known volumes of phosphoric acid were pipetted into the side arm and subject to five freeze-pump-thaw cycles before mixing with weighed samples of solid K3M02(HPO4)4 contained in the cuvette cell. For Moz(HPO4)44', known concentrations of M02(HPO4)44' were prepared by zinc amalgam reduction of M02(HPO4)43' solutions. Extinction coefficients were calculated from Beer-Lambert plots composed of at least seven points. 2. Steady-State Luminescence Spectroscopy. Emission spectra were recorded on an emission spectrometer constructed at Michigan State University. A detail description of the instrument is described in the doctoral thesis of Mark D. Newsham. Emission 44 quantum yields, were measured by using Ru(bpy)3(ClO4)2 as the quantum yield standard ($6 = 0.042 in H20 at 23 0C). The emission quantum yield was calculated with the following equation,110 x=\x)l x [Tle/l'lrzl x [Dx/Drl (16) where x and r designate the unknown and standard solutions, respectively, n is the average refractive index of the solution, D is the integrated area under the corrected emission spectrum, and AOc) is the absorbance/cm of the solution at the exciting wavelength X. 3. Time-Resolved Laser Spectroscopy. Picosecond laser spectroscopic experiments were undertaken in Brookhaven National Laboratory. The instrumentation has previously been described.111 4. Magnetic Measurements. EPR spectra were recorded by using a Bruker ER 200D X-band spectrometer equipped with an Oxford ESR-9 liquid-helium cryostat. Magnetic fields were measured with a Bruker ER035M gaussmeter, and frequencies were measured with a Hewlett-Packard 5245L frequency counter. Magnetic susceptibility measurements were made with a SHE 800 series variable-temperature SQUID magnetometer. 5. Resonance Raman. Raman spectra were recorded with a Spex 1401 double monochromator and associated Ramalog electronics. A Spectra Physics 165 argon ion laser was the excitation source, and incident powers were 20-40 mW. All spectra were 45 collected at a 900 scattering geometry from solid samples at room temperature. 6. Electrochemical Measurements. The oxidation-reduction potentials of dimolybdenum complexes were determined by cyclic voltammetry with a Princeton Applied Research (PAR) Model 173 potentiostat, Model 175 programmer, and a Model 179 digital coulometer. The output of the digital coulometer was fed directly into a Houston Instrument Model 2000 X-Y recorder. Cyclic voltammetry measurements of aqueous solutions were performed at room temperature by using a conventional H-cell design and a three- electrode system consisting of a polished glassy carbon working electrode (A = 0.08 cmz), a Pt wire auxiliary electrode, and a saturated calomel reference electrode. For nonaqueous solutions, a three-electrode system was again used with a standard H-cell configuration. The working electrode was a Pt button, the auxiliary electrode was a Pt gauze, and a Ag wire served as an adequate reference potential. Ferrocene was used as an internal standard. Potentials were related to the SCE reference scale by using a ferrocenium-ferrocene couple of 0.307 V vs. SCE”:2 7. Photochemistry. Sample irradiations were executed by using a Hanovia 1000-W Hg/Xe high-pressure lamp. The source for ultraviolet light at 254 nm was an Oriel 14W Hg/Ar low-pressure lamp. The beam was collimated and passed through a 10-cm circulating water filter. Photolysis experiments were performed in two-arm evacuable cells equipped with Kontes quick-release teflon 46 valves. Sample temperatures were thermostated at 15.0 i 0.5 0C in all photoreactions. The photolysis experimental set-up is shown in Figure 9. Photolyses were accomplished by using Schott color glass high-energy cutoff filters to eliminate the unwanted higher energy light. For quantum yield measurements, excitation wavelengths were isolated by using interference filters with half-band widths of less than 10 nm at the given mercury line ( e.g. 254, 313, 367, 405, 436, 546, 577 nm ). All interference filters were purchased from Oriel Corporation. Quantum yields were determined with the following expression or = [np/tl /lIo(M<1-10'AM/tl (17) where np is the number of product molecules formed during photolysis, 10(X) is the intensity of exciting light at wavelength 7t, A)‘ is the absorbance of reactants at the exciting wavelength X, and t is the time of irradiation.113 100.), determined by using the ferrioxalate actinometer, [Fe(CzO4)3]3' —h—V—. [Fe(CZO4)2]2' + 2c02 (18) follows directly from, Iom=npea+ / til-10‘“) x «>1. x tl (19) where AX is the absorbance of ferrioxalate at exciting wavelength X, (bx is the actinometer's quantum yield at X (listed in Table 2), t is 47 Figure 9. Experimental apparatus for typical photolysis. 48 a oSmE :8 29:3 3853 ~33: 53 533 l \ macs—zoto/t/ \ 5:: 5.8 3:25 3.22:6 om:o__aoou \ 53 co~m3.\.\ mace—:85 49 Table 2. Quantum Yield Dataa for Ferrioxalate Photoreaction. 7t Inm [K3Fe(C204)3]/M (D173 2+ 577 0.15 0.013 546 0.15 0.15 436 0.15 1.01 405 0.006 1.14 367 0.006 1.21 313 0.006 1.24 254 0.006 1.25 a Reference 1 13. 50 2+ the time for irradiation, and “Fe2+ is the number of Fe ions formed. 2+ ion is most easily accomplished by complexing the 2+ Analysis of Fe ion with 1,10-phenanthroline. The complexed Fe species absorbs strongly at 510 nm (E = 1.11><10'4 M'1 cm'l) in water. “Fe2+ is calculated directly from, nF.2+ =l6.023x1020xV1xV3xAFe2HIIV2 x EFe2+l (20) where V1 is the volume of the actinometer (ml), V2 is the volume of the aliquot withdrawn for analysis (ml), V3 is the final volume to which aliquot V2 is diluted (ml), Apez+ is the absorbance of ferrous phenanthroline at 510 nm, and EFe2+ is its molar extinction 114 coefficient. Quantum yields of multiply bonded molybdenum dimers were computed by reexpressing eq. (17) as follows, ox =[AA xVx 6.023XI020]/[E XIOX (1-10'A7k) x t] (21) where AA is the change of absorbance at the monitoring wavelength (1438 nm for Mo2(HPO4)43' and 516 nm for M02(02P(OC6H5)2)4), V is the volume of irradiated solution in ml, 8 is the molar extinction coefficient of the monitored absorption band (180 and 156 M'1 cm'1 for M02(HPO4)43' and M02(02P(OC6H5)2)4, respectively), A)‘ is the absorbance of reactants at excitation wavelength X, and t is the time employed for photolysis. In quantum yield measurements, the excitation procedures for standard and unknown have to be identical. Therefore, the special photolysis train, shown in 51 Figure 10, was designed to ensure that the required reproducibility was attained. 8. Gas Analysis. Gas produced from photoreactions was qualitatively identified by gas chromatographic analyses. A Hewlett- Packard 5710A gas chromatograph equipped with a 100/200 carbosieve S-II column and FID detector was employed for CC measurements. Quantitative determination of the gas evolved was made by Toepler pumping the gas above reacted solutions through two traps (liquid-nitrogen traps for hydrogen measurements, and pentane/N2(1) traps for ethylene measurements) into a calibrated volume and manometrically measuring the pressure. D. Crystal Structure Determinations 1. General Procedures. Crystal structure determinations were performed by Dr. Donald L. Ward of the X-ray Crystal Structure Facility at Michigan State University. Preliminary examination and data collection were performed with Mo K01 radiation (A = 0.71073A) on a Nicolet P3F diffractometer. All calculations were performed on a VAX 11/750 computer by using SDP/VAX.114 2. M02(02P(0C6H5)2)4°2THF. a. Data Collection. A dark purple prismatic crystal of tetrakis(diphenyl phosphate)dimolybdenum, M02(OZP(OC6H5)2)4, having approximate dimensions of 0.40 x 0.40 x 0.68 mm, was 52 Figure 10. Experimental apparatus for photoreaction quantum yield measurements. 53 2 255 noses _ao:ao .oxoa_ 553 an r 1111 . . _ .tt _ 530; 3 _ :oo . _ 29:3 (1\ n _ b p 5:: 55555:: l0. 0 ‘I "0‘. 342;. 114. c5225 53 553 3:22:85 N 5.5 533 act—2:35 biawcfih 32.53 55:3 .1 953 -1 m: 525. :58 5:25 52.55 54 mounted in a glass capillary with its long axis roughly parallel to the (p axis of the goiniometer. Cell constants and an orientation matrix for data collection were obtained from least-squares refinement, by using the setting angles of 17 reflections in the range 15 < 26 <20°. The monoclinic cell parameters and calculated volume are: a = 13.359(3), b = 29.64l(15), c = 14.719(3)A, [3 = 93.43(2)°, V = 5818(6)A3. For Z = 4 the calculated density is 1.522 g/cm3. As a check on crystal quality, omega scans of several intense reflections were measured; the width at half-height was 0.260 with a take-off angle of 6.00, indicating good crystal quality. From the systematic absences of OkO k = 2n (22) h01 h+l = 2n (23) and from subsequent least-squares refinement, the space group was determined to be P21/n. The data were collected at a temperature of 23(1) 0C by using the u) scan technique. The scan rate was 4°/min (in U) ). Data were collected to a maximum 28 of 45°. The ratio of peak counting time to background counting time was 1:1. The diameter of the incident beam collimator was 1.5 mm and the crystal to detector distance was 19 cm. b. Data Reduction. A total of 7973 reflections were collected, of which 7597 were unique and not systematically absent. As a check on crystal and electronic stability three representative 55 reflections were measured every 93 reflections. The slope of the least-squares line through a plot of intensity versus time was -325(6) counts/h which corresponds to a total loss in intensity of 68.5%. A nonlinear decay correction was applied. The correction factors on F ranged from 1.0 to 2.71 with an average value of 1.41. Lorentz and polarization corrections were applied to the data. The linear absorption coefficient is 5.964 cm'1 for Mo K01 radiation. A numerical absorption correction was made. Relative transmission coefficients ranged from 0.801 to 0.863 with an average value of 0.837. A secondary extinction correction was applied.116 Intensities of equivalent reflections were averaged. The agreement factors for the averaging of the 752 observed and accepted reflections was 2.5% based on intensity and 2.2% based on F0. c. Structure Solution and Refinement. The structure was solved by using the direct methods which revealed the positions of the Mo, 0, and P atoms. The remaining atoms were located in succeeding difference Fourier syntheses. Hydrogen atoms were included in the refinement but restrained to ride on the atom to which they are bonded. The structure was refined in full-matrix least-squares where the function minimized was £w(|Fol - IFCI)2 and the weight w is defined as 1.0 for all observed reflections. Scattering factors were taken from Cromer and Waber.117 Fe; 118 Anomalous dispersion effects were included in the values for Af' and Af" were those of Cromer.119 Only the 2363 reflections having intensities greater than 3.0 times their standard deviation were used in the refinements. The final cycle of refinement, which 56 included 351 variable parameters, have not yet converged (largest parameter shift was 2.23 times is esd) with unweighted and weighted agreement factors of, R1 = 2 11:0 - Fcl / 2 F0 = 0.091 (24) R2 = [(2 w(|Fo| - chhz/z wF02)]2 = 0.103 (25) The standard deviation of an observation of unit weight was 10.72. The highest peak in the final difference Fourier had a height of 1.30 F120 of 0.15; the minimum e/A3 with an estimated error based on 0 negative peak had a height of -1.03 e/A3 with an estimated error based on OF of 0.15. 3. M02(02P(0C6H5)2)4BF4. a. Data Collection. A dark green prismatic crystal of tetrakis(diphenyl phosphate)dimolybdenum tetrafluoroborate, M02- (02P(OC6H5)2)4BF4, having approximate dimensions of 0.40 x 0.40 x 0.8 mm, was mounted in a glass capillary with its long axis roughly parallel to the (9 axis of the goiniometer. Cell constants and an orientation matrix for data collection were obtained from least-squares refinement, by using the setting angles of 20 reflections in the range 20 < 26 <24°. The triclinic cell parameters and calculated volume are: a = 10.917(8), b = 11.793(4), c = l2.430(4)A, or = 6345(2) p = 7038(4), y = 70.64(5)° V = 5818(6)A3. For 2 = 1 the calculated density is 1.608 g/cm3. As a check on crystal quality, omega scans of several intense reflections were measured; the width at half-height was 0.260 with a take-off 57 angle of 6.00, indicating good crystal quality. From subsequent least- squares refinement, the space group was determined to be PT. The data were collected at a temperature of 29.0 0C by using the 6/26 scan technique. The scan rate was 4°/min (in 26). Data were collected to a maximum 26 of 55°. The ratio of peak counting time to background counting time was 1:1. The diameter of the incident beam collimator was 1.5 mm and the crystal to detector distance was 19 cm. b. Data Reduction. A total of 12192 reflections were collected, of which 6096 were unique and not systematically absent. As a check on crystal and electronic stability three representative reflections were measured every 93 reflections. The slope of the least-squares line through a plot of intensity versus time was —26(5) counts/h which corresponds to a total loss in intensity of 1%. A linear decay correction was applied. The correction factors on F ranged from 1.000 to 1.005 with an average value of 1.002. Lorentz and polarization corrections were applied to the data. The linear absorption coefficient is 6.625 cm'1 for Mo K01 radiation. A numerical absorption correction was made. Relative transmission coefficients ranged from 0.779 to 0.828 with an average value of 0.808. A secondary extinction correction was applied.116 Intensities of equivalent reflections were averaged. The agreement factors for the averaging of the 12192 observed and accepted reflections was 1.2% based on intensity and 1.0% based on F0. 58 c. Structure Solution and Refinement. The structure was solved by using the Patterson heavy-atom method which revealed the positions of the Mo and O atoms. The remaining atoms were located in succeeding difference Fourier syntheses. Hydrogen atoms were refined isotropically. The structure was refined in full- matrix least-squares where the function minimized was Zw(|Fol - IFcl)2 and the weight w is defined as 1.0 for all observed reflections. Scattering factors were taken from Cromer and Waber.11‘7 Anomalous dispersion effects were included in PC;118 the values for Af' and Af" were those of Cromer.119 Only the 5107 reflections having intensities greater than 3.0 times their standard deviation were used in the refinements. The final cycle of refinement, which included 436 variable parameters, has not yet converged (largest parameter shift was 3.31 times is esd) with unweighted and weighted agreement factors described by equations (24) and (25) with R1 = 0.027 and R2 = 0.027. The standard deviation of an observation of unit weight was 0.654. CHAPTER III SPECTROSCOPIC STUDIES OF M02X4(PBu3)4 AND M02C14(LL)2 COMPLEXES A. Background A prerequisite to the elucidation of the photochemistry of quadruply bonded metal dimers is a clear understanding of their excited state properties. As mentioned in Chapter I, the spectroscopic properties of quadruply bonded dimers have been intensely investigated64'78 since Cotton and coworkers first identified this class of complexes in 1964. A general conclusion of the absorption studies is that the lowest electronic transition, which typically falls in the visible region, is 52 -’ 55*.65 In contrast, the emissive properties of quadruply bonded metal dimer complexes vary significantly and a consistent description of 55* luminescence has not yet been achieved. Promotion of an electron from the 5 bonding orbital to the 5* antibonding orbital results in an excited state with an electronic configuration of 02n455* thereby reducing the overall bond order from 4 to 3. Simple group theory shows that the d22 orbital is totally 59 60 symmetric and the linear combinations of dxz and dyz orbitals are also totally symmetric; therefore, for complexes with an electronic configuration of 02n455*, there is no energy barrier for rotation about the metal-metal bond. Thus, the eclipsed geometry (D 4h) of the quadruply bonded complex in the ground state is lowered to D4d or D4 by torsional rotation of the ML4 fragments in the 1(55*) excited state.80 This free rotation provides an effective nonradiative decay pathway from the excited state thereby resulting in low emission quantum yields and short emission lifetimes. The torsional model accounts for the lack of a mirror image between the absorption and emission spectra of many quadruply bonded complexes. Presumably, sterically small ligands of M2X8 or M2X4L4 complexes permit free rotation about the metal-metal bond. The fact that the apparent origins in low temperature emission spectra for such complexes do not overlap results from distortions along two coordinates (valg(M-M) and the torsional mode); this commonly is 121 referred to as a Duschinsky effect. On the other hand, large bulky ligands of quadruply bonded metal dimers such as the phosphines in M02C14(PMe3)4 and M02Cl4(PBu3)4 are interlocked in a Dz d geometry and create a large steric barrier to rotation about the metal-metal bond. In these cases, the 02d symmetry is preserved in the 1(55*) excited state and a mirror image of absorption and emission profiles is observed.80’81 The above observations imply that the excited state dynamics of quadruply bonded metal dimers are closely related to the molecular structures in the ground and the 1(55*) excited state. Our interest in developing the chemistry of electronically excited 61 M-4—M dimers has prompted us to assess the generality of the D4}, 4* 04d torsional model, and more generally to define the factors governing the nonradiative processes of MAM dimers. These studies have become possible with Cotton and coworkers synthesis of a series of complexes with a general formula M02X4(LL)2, where X is a halide, and LL is a bridging bidentate phosphine (1.106’107’122'124 In these complexes, steric restrictions of the ligan LL ligands introduce various angles of internal rotation away from the perfectly eclipsed conformation. The rotation is defined by a torsional angle, X, shown in Figure 11. The torsional angles of a series of M02X4(LL)2 complexes as determined by X-ray crystallography are shown in Table 3. When it is equal to zero, there is no distortion and the molecule is in an eclipsed geometry which maximizes the dxy overlap. The overlap of dxy orbitals decreases with increasing x, and becomes nonbonding when a completely staggered conformation X = 45° is achieved.12421 In this chapter, we report and discuss the absorption, emission, and transient absorption spectroscopic studies of a series of M02X4(LL)2 complexes possessing various torsional angles. All the lifetime measurements and transient absorption spectra displayed in this chapter were collected at the picosecond laser facility, under the administration and directorship of Dr. Jay Richmond Winkler, at Brookhaven National Laboratory. 62 Figure 11. (a) 02d structure of quadruply bonded chlorophosphino dimolybdenum M02X4L4; (b) Newman projection of M02X4(LL)2 complexes where X = Cl, Br, I; L and LL = mono- and bi-dentate phosphine ligands, respectively. 63 : 255E 64 Table 3. Torsional angles of structurally characterized M02X4(PBu3)4 and M02C14(LL)2 complexes. MoiMo complexes XIde g ref. M02C14(PBu3)4 0 a MozBr4(PBu3)4 0 a M0214(PBu3)4 0 a M02C14(dppm)2 0 1 22 M02C14(dmpm)2 0 107 M02C14(dppe)2 30.5 123 M02C14(dmpe)2 4 0 106 a X-ray crystal structure determinations of M02X 4(PBu3 )4 complexes, have not been performed. However, the assumed eclipsed geometry for these dimers is reasonable by analogy to the data provided for M02X4(PMe3)4 complexes (ref. 78c). 65 B. Results and Discussion 1. Absorption Spectroscopy. The electronic absorption of M02X4(PR3)4 (R = alkyl) complexes has been studied in detail by 68,78,79 Gray and Trogler and their coworkers. Spectra of M02X4(PBu3)4 complexes, shown in Figure 12, exhibit a distinct band at ~600 nm possessing an large molar extinction coefficient (E~3x103 M'1 cm’l) attributable to the 52 -> 55* transition. The shift of this transition to lower energy relative to M02C184‘ (515 nm) and "M0203" complexes (510-520 nm) may be explained within a simple valence bond framework. As discussed by Trogler et al.,68 good donor ligands (such as phosphine) place significant electron density on the metal centers, thereby expanding the valence shell and increasing the 5 overlap. Increased dxy orbital overlap is manifested in the 1(55*) excited state acquiring more covalent character and, as a result, the spin pairing term contribution to the overall transition energy is reduced. Therefore the 1(55*) excited state is shifted to lower energy. The molar extinction coefficients for the 52 -’ 55* transition of quadruply bonded metal complexes, in general, are small for a transition which is fully allowed. However, 125 from the work of Mulliken, it is known that the oscillator strength of a one-electron N -’ V transition is related to the square of the orbital overlap. The relatively low intensity of this transition is therefore a consequence of the small orbital overlap (~01) of the dxy orbitals that comprise the 5 bond. To this end, the unusually large 1(55*) intensities of the M02X4(PR3)4 complexes can be attributed to the enhanced overlap of the 5 bond in phosphine ligated 66 Figure 12. Electronic absorption spectra of M02X4(PBu3)4 where X = Cl (—), Br (---), and I (- - -). 67 S 85E :52 com oov com EIONVHHOSHV 68 complexes. However as recently pointed out by Hopkins et al.,78 the intensities are much too large to be exclusively accommodated by increased orbital overlap resulting from secondary donor effects by the phosphine ligands. This has led these authors to suggest that the strong 52 -* 55* absorption of M02X4(PBu3)4 complexes is due to the stealing of intensity from the lowest energy ligand-to-metal charge transfer (LMCT) transition which occurs at ~330 nm (£~104 M’1 cm'l) in the M02X4(PBu3)4 complexes.78d The absorption spectra of eclipsed M02C14(LL)2 complexes, M02C14(dppm)2 and M02C14(dmpm)2, are similar to that of M02C14(PBu3)4. The 52 -+ 55* transition and the molar extinction coefficients are 634 nm (e = 2490 M’1 cm'l) and 604 nm (e = 1730 M'1 cm'l) for M02C14(dppm)2 and M02C14(dmpm)2, respectively. In contrast, the staggered M02C14(LL)2 complexes exhibit 1(52 -r 55*) absorptions of much lower intensity and energy. The 1(52 -’ 55*) transition of M02C14(dppe)2 maximizes at 762 nm (E = 1030 M'1 cm'l) whereas M02Cl4(dmpe)2 maximizes at 803 nm (E = 210 M"1 cm'l). The lower energy of the 1(52 -’ 55*) transition is due to the decreased dxy orbital overlap and consequently reduced energy splitting of the 5 and 5* levels. The energetic differences between the eclipsed 1(55*) and staggered 1(55"')excited states can readily be attributed to the dependence of the one- and two-electron interactions of the dxy orbital overlaps. It has been shown previously that the 5 interaction decreases as the distortion angle increases. A 40° distortion angle, such as that found in M02C14(dmpe)2, reduces the 5 interaction to 0.17 times that found at 0°.123 Inasmuch as the 5 overlap of the eclipsed molecules are 69 on the order of S ~ 0.1,65 the energetic splitting of the one-electron 5 and 5* levels of the M02C14(dmpe)2 is clearly quite small. As shown in Figure 13,79 the energetic splitting between the covalent l(52) ground state and 3(55*) excited state decreases with increasing distortion angle to a nearly degenerate pair (for x = 45°). In contrast, the energy of the ionic 1(55*) state is dominated by an electron repulsion term that is roughly invariant over the range of overlaps (0.01< S < 0.1). Thus, even near the zero-overlap limit (400 torsional angle), M02C14(dmpe)2 exhibits a 1(52 -» 55*) transition of 12500 -1 cm owing to the significant two-electron spin pairing term. The poor overlap of the dxy orbitals in the torsionally distorted molecules will be manifested in extremely small oscillator strengths. Hence, the question is not why the molar absorption of these complexes is small, but rather by what mechanism do these transitions acquire their intensity. Similar to the M02X4(PBu3)4 complexes, intensity stealing from low lying LMCT transitions most reasonably accounts for the high oscillator strengths of the 1(52 -’ 55*) transitions of these staggered M02C14(LL)2 complexes. 2. Emission Spectroscopy. The M02X4(PBu3)4 complexes provide the benchmark for the interpretation of the luminescence of quadruply bonded phosphine complexes. Solids and solutions of the M02X4(PBu3)4 complexes exhibit intense emission at room and low temperatures. The luminescence spectra mirror the absorption spectra and indeed these complexes were the models for which the torsional model was initially proposed.80 On the basis of this model, M02Cl4(LL)2 complexes in which the bidentate phosphine ligands 70 Figure 13. Energies of the lowest electronic states of M02C14(PMe3)4 and M02C14(dmpe)2 (ref 79). 71 2 oswE l o_w=< 5:59 642-264 NAoQEBEUNo—z vfimQEQVEUNoE t 0 N0 r oo~n :E: .5350 Aowwvm :8: 622 now ~ New I_u10/1(319U3 72 strap the metal-metal bond in an eclipsed conformation would be expected to exhibit strong emission directly from 1(55*) excited state. On the other hand, staggered quadruply bonded M02C14(LL)2 dimers would display weak emissions. But this is not the case. The staggered M02C14(dppe)2 and M02Cl4(dmpe)2 complexes emit with appreciable intensities in CH2C12 solution at room temperature whereas M02C14(dmpm)2 emits only weakly, and M02C14(dppm)2 is non-luminescent in CH2C12 solution at 25 °C. The emission spectra of the M02C14(dppe)2, M02C14(dmpe)2, and M02C14(dmpm)2 complexes are show in Figure 14. All four M02C14(LL)2 complexes emit in low temperature glasses (Figure 15). The intensities of the emission parallel that observed at high temperatures. Namely, the staggered complexes luminesce much more intensely than the eclipsed dimers. Typical of binuclear metal-metal emission, the luminescence bands sharpen considerably upon cooling.126 Because the complexes do not exhibit vibrational fine structure even at low temperature, the excited state dynamics can not be analyzed by conventional spectroscopic techniques. However, Stokes shifts of the complexes do provide some information. The 1(52 -> 55*) absorption and emission bands of M02Cl4(dppe)2 and M02C14(dmpe)2 are shown in Figure 16. It is obvious that both complexes exhibit large Stokes shifts. It should be noted that these Stokes shifts represent a lower limit because the photomultiplier tube of our emission instrument is attenuated significantly for A > 1000 nm. From the valence bond model of quadruply bonded metal dimers, as discussed previously, the 52 ground state of M02Cl4(dmpe)2 and M02C14(dppe)2 is nearly energetically 73 Figure 14. Emission of M02C14(LL)2 complexes in CH2C12 solution at room temperature: (a) M02C14(dmpm)2; (b) M02C14(dppe)2; and (c) M02C14(dmpe)2. 74 ' 600 700 800 A/nm Figure 14 75 (b) A/nm Figure 14 76 800 9 00 l 000 Figure 14 77 Figure 15. Luminescence of M02C14(LL)2 complexes at 77 K in a CH2C12/toluene glass: (a) M02C14(dppm)2; (b) M02C14(dmpm)2; (c) M02C14(dppe)2; and (d) M02C14(dmpe)2. 78 500 700 800 A/nm Figure 15 (a) Iem 79 (b) Figure 15 80 (C) Iem l l 1 l 800 900 IOOO A/nm Figure 15 81 Iem (d) 800 A/nm Figure 15 82 Figure 16. Absorption and emission spectra of 52 2 55* transitions of staggered (a) M02C14(dppe)2 and (b) M02C14(dmpe)2 complexes. 83 |em 000—. 000 EC\K 2 25E 000 00h 3 J . _ 4 O 1 00m L..UJ3 L-W/3 1000.. 84 'em 000? 000 2 2:3". 5:2 00m 005 30 j 85 degenerate to the 3(55*) excited state.79 Therefore, the emission clearly does not originate from 3(55*), and we assign the emission to the 1(55*) excited state. To explain the Stokes shifts of the emission spectra of M02C14(dmpe)2 and M02C14(dppe)2 complexes, we adopt the explanation for the large red-shift from absorption to emission for the Re2C182’ dimer.80 Namely, emission is observed from an equilibrium series of complexes possessing torsional angles between a maximized dxy orbital overlap (x = 0°) and the configuration with a steric minimum (X = 45°). For the M02Cl4(dppe)2 and M02Cl4(dmpe)2 complexes the equilibrium geometry is with X: 30.50 and X = 40°. Rotation to smaller angles will increase the metal- metal distance owing to steric constraints of the LL backbone. Hence rotation away from the equilibrium torsional angles will result in emission from a spectrum of complexes with metal-metal distances greater than that occurring from simple population of the 5 * antibonding orbital near equilibrium geometry. This will result in luminescence from significantly distorted M02C14(dppe)2 and M02C14(dmpe)2 excited states, and hence large Stokes shifts will be observed. This explanation is supported by the emission properties of M02Cl4(dppe)2 dimer in poly(methy1 methacrylate) (PMM) films at room temperature. The solution spectrum of the MozCl4(dppe)2 complex is frozen by incorporation into the rigid matrix of the PMM film. The emission maximum of M02C14(dppe)2 shifts from 964 nm in CH2C12 to 936 nm in the PMM film. The decrease in the Stokes shifts upon incorporation into PMM indicates that emission is from a complex geometrically more similar to the ground state structure. 86 This result is significant because it shows that the distorted M02Cl4(LL)2 dimers emit directly from staggered l(55*) excited states. Emission of M02C14(dppm)2 and M02C14(dmpm)2 complexes exhibit smaller Stokes shifts (Figure 17) than that for complexes with staggered structures. This indicates that emission arises from an excited state complex structurally similar to the ground state. This is not unreasonable because the methylene bridged LL complexes can maintain the geometry of the M-4-M complex in the excited state better than that possessing bidentate phosphines with an ethylene backbone. The important point is that the ethylene complexes are distinguished from the methylene complexes by significantly lower emission intensities of the latter (~102 less). Of equal significance is that the emission of the eclipsed M02 C14 (d p p m ) 2 and M o 2Cl4 ( d m p m) 2 complexes is 104 less than the eclipsed M02C14(PBu3)4 complexes ((156 = 0.013).80 This discrepancy can not be explained by the simple D 4h 41-) 04d torsional model which predicts nearly equivalent intensities for the two classes of eclipsed complexes. Some insight into the issue of emission intensities is provided upon closer examination of the molecular and electronic structures of the eclipsed and staggered complexes. The electronic structure of MA- M dimers is significantly affected by the symmetry of the molecule. As shown in Figure 18, the (dxz, dyz) orbitals are no longer energetically degenerate when the symmetry of the chlorophosphino dimolybdenum complexes changes from 02d (e.g. M02X4(PBu3)4) to D2}, (e.g. M02C14(dmpm)2), whereas the dxy 5-bonding and 5- 87 Figure 17. Absorption and emission spectra of 52 :2 55* transitions of eclipsed (a) M02C14(dppm)2 and (b) M02C14(dmpm)2 complexes. 88 C 05E EC\K "— 000_. 000 000 005 000 com _ _ _ _ q H. o 1 00°F 1 83B Iem l 0000 L-uJO HIV 1 000v 3 89 lem 000 5— 0.53m EEK 005 000 com 30 11000 110005 1.0005 L.Luo L w/3 90 Figure 18. Relative energies of the d-derived molecular orbitals in D2d (i.e. M02Cl4(PBu3)4) and D2}: (i.e. M02C14(dppm)2) symmetries. 91 b2(0*) -— ----------------------- —— b2u(o*) b2(M'L 0*) __ _____________________ __ b1u(M-L 0*) a1(M-L 0*) — —-‘- ag(M-L 0*) ........... —_ b3g(n*) e(n*) -::Z::: ..... ........ b2g("*) a1(5*) —— ------------------------ __ au( 5*) 132(5) 1L ........................ _‘l_ blgw) 0 ll ------------------- —'L b2u(TI) e(n) _ b3u(n) 31(0) —“— ------------------------ ——l-L ag(0) A L X L L z-éMo—Mo z—>Mo\§ Mox x/m «W L X L L \_/ (02d) (02h) Figure 18 92 antibonding interactions are not greatly perturbed. The energy gaps between 1'! and 5 states are smaller for complexes with D2 h symmetry and hence contributions (i.e. decreased energy gap) of lower lying states proximate to the excited state will increase the nonradiative decay rate, and hence result in lower emission yields for 02h complexes (i.e. M02C14(dmpm)2). Staggered M02C14(LL)2 complexes possess symmetries between the D2d and 02 h structures and therefore, to a first approximation, exhibit higher luminescence intensities than D2 h molecules but lower than those of 02d complexes. A more quantitative assignment of the nonradiative processes of these molecules will only be achieved when good theoretical calculations of the energy levels of these MA— complexes are in place. 3. Transient Absorption Spectroscopy. Kinetic studies of M02X4(PBu3)4 and M02C14(LL)2 complexes were investigated by time-resolved transient absorption spectroscopy. Owing to the relatively weak luminescence intensity of M02C14(dmpm)2 and related complexes, transient species important to nonradiative decay can remain undetected with time-resolved luminescence spectroscopy. This issue is best illustrated by picosecond absorption measurements of M02C14(PBu3)4. Time-resolved emission spectro- scopy identifies a single excited state species with a lifetime of 16 ns.80 Transient absorption spectroscopy, however, reveals two kinetic decays of lifetimes 9 and 90 us.82 The transient species with the 9 ns decay is consistent with the emission lifetime; thus, this species has been assigned to the 1(55*) eclipsed excited state. The 93 nature of the long-lived 90-ns nonemissive transient species is not yet clear, but three possibilities for the transient are: (l) the low- lying 3(55") state (~5000 cm'l); (ii) an electronic excited state of M02C14(PBu3)4 that lies only slightly lower in energy than 1(55*) state but carries very little oscillator strength to the ground state; and (iii) a highly distorted chemical intermediate of M02C14(PBu3)4. The decay has been attributed to most likely arise from a distorted chemical intermediate.82 Because the transient spectroscopy of M02C14(PBu3)4 is well defined, studies were initiated by recording the spectra of its halide analogues, MozBr4(PBu3)4 and M0214(PBu3)4. Figure 19 displays the transient absorption spectra of M02C14(PBu3)4 (X = Br, I) at 50 ps and 1 ns, respectively, after excitation with a ~25 ps pulse of 532- nm laser light. The spectra, concordant with that of M02Cl4(PBu3)4, exhibit a moderately intense absorption maximum at 440 nm, an absorption rising into the near UV, and a bleaching of the 1(55"') transition at ~600 nm. That the moderately intense absorption maxima of M02C14(PBu3)4, MozBr4(PBu3)4, and M0214(PBu3)4 are energetically similar (440, 460, and 440 nm for the chloro, bromo, and iodo, respectively) distinguishes this transition as a metal localized transition or an ligand-to-metal charge transfer (LMCT). Because the energy of the transition is invariant across the halide series, an LMCT involving the phosphine ligands is the preferred assignment.82 The extinction coefficient of this band is ~2000 M'l cm'l. This value clearly indicates a spin allowed transition but can not distinguish between the LMCT and metal-localized excitations. For the latter, assuming D 4h selection rules, five dipole-allowed one- 94 Figure 19. Transient difference spectra of (a) MozBr4(PBu3)4 in CH2C12 and (b) M0214(PBu3)4 in CH2C12. The spectra were recorded 50 ps and 1 ns, respectively, after the 532-nm pulse of a Nd:YAG laser (FWHM = 25 ps). 95 2 2.5m. EE\K 555 com coo ; _ _ e 6 o 6 o e 6 0 Adv 6 66% $6 5 0' HDNVHHOSHV V 4111 ((1611 'AG N0- 96 2 25E E:\K com 000 00v .5 _ _ ee 5 1 e e e 6 ++ 1 a. a e e r. 1 £0 + A. _ _ _ 0.0 «.0 HDNVHHOSHV V 97 electron metal-localized excitations can arise from a 1(55*) excited state: 5 .. 5*; n .. 11*;11 -» 5'(dx2_y2); 5* .. 11*; and n -» 5. These transitions will produce, respectively, 1A1 g(5*2), 1A1 g(11355*rt"‘), 1l.=.g(n355*5'), lr~:€.;(5r1"‘), and 1Eg(n5*) states. The level arising 1 from the 440 nm absorption must lie ~39000 cm“ above the ground state, and calculations suggest that58 of the five metal localized transitions, the 5* -’ TI* excitation is the most plausible assignment for this feature. However, the data of Mo2X4(PBu3)4 series do not provide an unequivocal assignment to either 5* -+ 11* or LMCT. The kinetics of M02C14(PBu3)4 excited state decay can readily be ascertained from these transient absorption spectra. For the chloro, bromo, and iodo complexes, the ~440 nm absorption and 1(55*) bleaching signals decay biexponentially. The shorter lifetime component in each case is nearly identical to the measured luminescence lifetimes (I = 9, 2, 6 us for chloro, bromo, and iodo M02X4(PBu3)4 complexes, respectively). Because the luminescence spectrum of each complex is a mirror image of the 1(52 4 55*) transition, the short-lived transient is assign to the eclipsed 1(55*) excited state. The longer lived species (I = 90, 34, 108 ns for chloro, bromo, and iodo M02X4(PBu3)4 complexes, respectively) is due to one of the three possibilities mentioned previously. In an effort to provide further insight into the nature of this longer lived non- luminescent intermediate, the transient absorption spectra of M02C14(LL)2 complexes were investigated The transient absorption spectra of the M02C14(LL)2 complexes is similar to that of the M02Cl4(PBu3)4 dimer. Spectra, collected 25- 100 ps after the excitation pulse (shown in Figure 20), exhibit a 98 Figure 20. Transient difference spectra of the following M02C14(LL)2 complexes: (a) M02C14(dppm)2, recorded 75 ps after a 355-nm laser excitation pulse; (6) M02C14(dmpm)2, recorded 25 ps after a 532-nm laser excitation pulse; (c) M02C14(dppe)2, recorded 100 ps after a 532-nm laser excitation pulse; and (d) M02C14(dmpe)2, recorded 100 ps after a 355-nm laser excitation pulse. 99 EEK 8655 occ can ace ++_ — _ : 665 + +e+++eefi HVV 8.5 s HONV 8110 no.0 100 ocw EG\K 000 cm 2.5a 00v Ev _ 170- l 0 0 170 HDNVEIHOSHV V 101 cm 2:5 e52 000 00v 60 a Z + 00.0 8.0 HDNVHHOSEIV V 102 000 :52 000 om 8wa 00v 60 _ 00.0 no.0 00.0 HONVEIHOSHV V 103 prominent bleaching signal corresponding to the 1(55*) transition. It should be noted that the bleaching of M02C14(dmpe)2 lies well into the red and is weak owing to low response of the photomultiplier tube. However the leading edge of the bleaching signal is clearly distinguishable in Figure 20(d). The spectra of the four complexes are dominated by pronounced absorption in the 430-450 nm; the spectra of M02Cl4(dmpe)2 displays an additional absorption at 525 nm. The extinction coefficient of the ~440 nm bands of M02C14- (dppm)2, Mo2Cl4(dmpm)2, and M02C14(dppe)2 are 1000-3000 M'1 cm’l. On this basis, the intensity of the 525 nm band of M02C14(dmpe)2 identifies it as the corresponding transition, and thus the 435-nm absorption of the M02C14(dmpe)2 complex is unique. This is further suggested by the fact that the band halfwidth of the 435 nm is much smaller than that of the other observed absorptions in M02Cl4(LL)2 complexes. The obvious question becomes what transition is responsible for the respective 430, 450, 470, and 525 nm absorptions of the M02C14(dppm)2, M02C14(dmpm)2, M02C14(dppe)2, and MozCl4(dmpe)2 complexes. The energy and intensities of these absorptions suggest them to be analogous to those observed in the spectra of the M02X4(PBu3)4 complexes. We have previously suggested that the ~440-nm absorption for the electronically excited M02X4(PBu3)4 complexes is attributed to either a 5* -' 11* transition or a charge-transfer transition involving the phosphine ligands. Consideration of the electronic structure of MAM complexes imply that the red-shift in the absorption along the series M02C14(dppm)2, M02Cl4(dmpm)2, M02C14(dppe)2, and M02C14(dmpe)2 is inconsistent with a 5* -’ 11* transition. The n and 104 n * orbitals of the M-4—M complexes are cylindrically symmetric about the metal-metal axis and hence the energy of these orbitals is invariant to rotation about the metal-metal bond. The 5-5 * splitting however will decrease dramatically with rotation owing to stabilization of the 5 * orbital and destabilization of the 5 orbital with a decreasing 5 interaction. To this end, the 5 * -> n * transition should blue-shift with increasing 7:. On the other hand, charge transfer transitions to the 5* orbital will red-shift as 71 increases. Therefore, the most likely assignment of the transition leading to transient absorption in M02Cl4(LL)2 and M02C14(PBu3)4 complexes is not metal-localized but ligand(phosphine)-to-metal(5*)~charge- transfer. The assignment of the 435-nm band in the spectrum of M02C14(dmpe)2 has not been elucidated at this time. Although the transient absorption spectra of the four M02C14(LL)2 complexes suggest that the complexes are similar, analysis of the lifetime decay of the transient show that the diphenylphosphine complexes (i.e. M02C14(dppm)2 and M02C14- (dppe)2) are distinctly different from their dimethylphosphine counterparts (i.e. M02C14(dmpm)2 and M02Cl4(dmpe)2). Figure 21 presents the decay kinetics of the transient species of the M02C14- (dmpm)2 and M02C14(dmpe)2 complexes. In both cases the transients decay by monoexponential kinetics with lifetimes of 43 and 960 ps for the M02Cl4(dmpm)2 and Mo2C14(dmpe)2, respectively. The decay kinetics of M02C14(dmpm)2 were obtained from the 1(55*) bleaching signal and thus the 43 ps represents the 1(55*) lifetime of the excited state. Because the bleaching signal of the M02Cl4(dmpe)2 complex is not easily observed, the direct 105 Figure 21. Transient kinetics for dimethylphosphine complexes: (a) M02C14(dmpm)2 recorded at 600 nm (excitation by 2nd harmonic of a Nd:YAG, FWHM = 25 ps) (b) M02C14(dmpe)2, recorded at 450 nm (excitation by 3rd harmonic of a Nd:YAG, FWHM = 25 ps). 106 a 655E m:\~ 04 n0 0.0 _ A _ a _ l'l'lll'llllIlllllll'll'ITTTIITTIl'll' 70- 0.0 'EDNVHHOSEIV V 107 a 2:3... m:\u 30 0.0 70 EIDNVEIHOSEIV V 108 analysis of 1(55*) is not feasible. However, it is reasonable that the prompt decay of the transient absorption is also the lifetime of the 1(5 5 I") excited state. These excited state properties of the bidentate methylphosphine complexes are parallel to those of the monoexponential 1(55*) decay of the unbridged methylphosphine dimer, Mo2Cl4(PMe3)4. Analogously, the phenyl substituted bridged phosphine complexes, M02C14(dppm)2 and Mo2C14(dppe)2, exhibit decay kinetics similar to unbridged phosphine complexes possessing bulky substituents. The transient absorption at 400 nm and 1(55*) bleaching at 620 nm decay by the same kinetics indicating that the transient spectrum arises from a single species (exemplary decays for M02C14(dppm)2 are shown in Figure 22). These data are fit extremely well with two decay lifetimes of 130 ps and 2.1 ns for M02C14(dppm)2. As shown in Figure 23, a similar biexponential decay process was observed from the 1(55*) excited state of M02C14(dppe)2 (lifetimes of 410 ps and 2.4 ns). By analogy to M02Cl4(PBu3)4, the shorter lifetime is assigned to 1(55"') excited state and the longer lifetime corresponds to a non-luminescent transient intermediate. Although the data does not allow unequivocal assignment of the long-lived intermediate, it is interesting that the lifetime decay kinetics of M02X4(PR3)4 and M02Cl4(LL)2 are related to the cone angle and basicity of the phosphine ligands. For M02X4(PMe3)4 and the bidentate analogues Mo2Cl4(dmpm)2 and M02C14(dmpe)2, the coordinated phosphines possess the smallest cone angle (PMe3= 118°, dmpe: 107°) and 127 strongest Lewis basicity. These complexes decay directly from the 1(55*) and their excited state dynamics are not mediated by any 109 Figure 22. Transient kinetics for M02C14(dppm)2 in CH2C12 solution. The decay kinetics were recorded at (a) 400 nm and (b) 620 nm (excitation by 3rd harmonic of a Nd:YAG, FWHM = 25 ps). 110 T -$--11_._.. T T (b) 0 ------------_-------.+.---- P p n a a d 560 . ) . a ( . . . P . . r4 . . lhw 0 . _ r 2 _ . . _ . . .01: . .Q . . t h n - r \M u - libiu.lflrl '0' p b p u d u 4 2 0 0 0. O 0 0 0 m02 staggered M02Cl4(LL)2 (D2) > eclipsed M02C14(LL)2 (D 2h) suggests that it is the presence of states proximate to the 1(5 5 *) singlet that are important to nonradiative decay and not the torsional angle about the metal-metal bond. A single excited state following excitation is observed for halophosphine M-4-M complexes when the substituents on the phosphines are sterically small. However, with increasing steric bulk, the presence of two transient intermediates are observed, one of which is attributable to the 1(55*) excited state and the other to a long-lived chemical intermediate. Although a definitive assignment of this intermediate is not possible, the data are consistent with formation of a coordinatively unsaturated transient produced by simple excited state dissociation of phosphine within a solvent cage. CHAPTER IV PHOTOCHEMISTRY OF NOVEL DIMOLYBDENUM PHOSPHATE COMPLEXES IN ACIDIC SOLUTION A. Background Conversion of a M-n-M dimer with a metal-metal bond order of 4 to a species of bond order of 3 conceptually represents the simplest multielectron transformation available to a MAM system. The overall reaction constitutes a two-electron oxidation of the metal- metal core. Multielectron studies were initiated by exploring the oxidation photochemistry of M-4-M dimers in acidic solution with the goal of promoting the following reaction sequence Mfr-M __fll’_... MA—M“ (26) MA-M* + HB —- 114-11w + 13' (27) M—3—M + HB —- M-3-M2+ + H2 + 13' (28) Scheme III 115 116 In this scheme, electronically excited MA-M reacts with HB in a two- electron oxidation step to produce a hydride intermediate and Bronsted base B'. Trapping the hydride with another equivalent of HB gives oxidized dimer M-3-M and hydrogen. It should be noted that the hydride production reaction is fundamentally important to excited state multielectron chemistry because it represents one of the simplest multielectron reaction pathways available to the excited quadruply bonded dimer. Practically, the reaction is important because metal hydrides are crucial intermediates in small molecule activation schemes such as those for water splitting chemistry. The hydrogen donor is not limited to a Brensted acid but in principle can involve any small molecule reactant. Thus, the reaction chemistry described in Scheme III is significant in regard to our efforts in designing new excited state reactivity patterns. Examination of Scheme III reveals that hydrogen production relies on a quadruple bond (=2 triple bond conversion. Although, conceptually quite appealing, photoreaction to triply bonded species is an exceptional reaction pathway and the excited state reaction chemistry of M-A-M dimers has been limited primarily to single- electron photoreactions. For instance, the photochemistry of M02(SO4)44’ is exemplary of typical MAM species: under relatively strong oxidizing conditions (Xexc > 254 nm), irradiation of Mo2(SO4)44' in H2804 is confined to the production of the one electron product M02(SO4)43'.93 Presumably, the chemical instability of the triply bonded sulfate dimer precludes multielectron photochemistry. Accordingly, we set out to synthetically design M-n M complexes which could undergo triple/quadruple 117 photoconversions. In this context, the discovery of the triply bonded 109 phosphate complex M02(HPO4)42' stimulated our interest because the M02 phosphate system could possibly provide insight into the mechanism of multielectron photoredox chemistry of M-n-M dimers in acidic solution with the successful preparation of quadruply bonded molybdenum phosphate dimer. The existence of the quadruply bonded M02(II,II) sulfate complex suggested that the M02(II,II) phosphate could be prepared in view of the electronic and structural similarities of HPO42' and 8042’. This chapter describes the spectroscopy and electrochemistry of M02(HPO4)44‘ and M02(HPO4)43' dimers as well as the photoinduced two-electron oxidation of M02(HPO4)44’ in acidic solution. B. Results and Discussion 1. Magnetic and Spectroscopic Characterization of M02(HPO4)4“' (n = 2, 3, 4) Species. A central result to emerge from exhaustive spectroscopic studies of multiply bonded metal- metal dimers is a consistent description of the electronic structure of these complexes in terms of a 0, n, and 5 molecular orbital framework.55“"65 One consequence of this model is the prediction that dimers of bond orders 4 (0211452) and 3 (02114) will possess a diamagnetic ground state whereas the mixed-valence dimer of bond order 3.5 (0211451) will be paramagnetic. Investigation of the magnetic properties of M02(HPO4)43' and M02(HPO4)44' were of interest because these studies in connection with a previous susceptibility measurement on the pyridinium salt of M02(HPO4)42' 118 which was shown to be diamagnetic,129 would provide magnetic data on the first homologous M-n-M series possessing a 021145" (11 = 0, 1, 2) ground state configuration. The predicted paramagnetic behavior of M02(HPO4)43' is confirmed by magnetic susceptibility measurements of solids from 5 to 300K at a field of 5500 G. Correction for the diamagnetic contribution to magnetic moment was provided by the room temperature magnetic susceptibility of the pyridinium salt of M02(HPO4)42’. The susceptibility data is shown in Table 4. The temperature dependence of the magnetic susceptibility follows Curie law behavior (Figure 24) and the magnetic moment of K3M02(HPO4)4 is 1.58 BM. In accordance with susceptibility results, EPR spectra of the pyridinium salt of frozen solution of K3Mo2(HPO4)4 at 5 K (Figure 25) show a paramagnetic species in an axial environment, g1 = 1.894, g" = 1.886. As described in detailed studies of the EPR spectra of M02(SO4)43',130 and other M-z'LM dimers, 1 31 the EPR signal is consistent with a species in which the unpaired electron is coupled between two equivalent molybdenum nuclei. Reduction of M02(HPO4)43‘ by one electron will yield a quadruply bonded complex which should exhibit rigorous diamagnetism. Unfortunately, the accurate determination of the magnetic susceptibility of M02(HPO4)44' is precluded by our inability to cleanly isolate salts of the quadruple bond complex from the mixed-valence dimer. Indeed, EPR spectra of solid samples of K4M02(HPO4)4 at 5 K show signals identical to that of K3M02(HPO4)4. That the intensity of the signal varies with sample preparation indicates that the observed paramagnetism arises from 119 Table 4. Magnetic Susceptibility of K3M02(HPO4)4. Tl K XMa/emu mol'1 1/XM/emu'lmol 5 0.066219 15.1014 6 0.053052 18.8494 8 0.038362 26.0675 10 0.030215 33.0961 12 0.024878 40.1962 14 0.021185 47.2032 15 0.019661 50.8621 16 0.018452 54.1947 18 0.016427 60.8754 20 0.014713 67.9671 22 0.013333 75.0188 25 0.011682 85.6018 27 0.010850 92.1149 30 0.009801 102.0346 35 0.008420 118.7719 40 0.007454 134,1562 45 0.006638 150.6478 50 0.006053 165.2073 55 0.005517 181.2744 60 0.005124 195.1715 65 0.004741 210.9260 70 0.004455 224.4669 75 0.004171 239.7794 80 0.003725 268.4924 85 0.003725 295.6743 90 0.003561 280.8279 95 0.003382 295.6743 100 0.003241 308.5087 110 0.003004 332.8562 120 0.002763 361.9910 129 0.002595 385.4307 140 0.002545 393.0045 Table 4 (cont'd.). 120 148 0.002321 430.9230 163 0.002301 434.5937 166 0.002071 482.8352 184 0.002086 479.2715 191 0.001923 520.0479 205 0.001913 522.6572 221 0.001723 580.4167 223 0.001754 570.0604 254 0.001565 638.8144 255 0.001612 620.3859 279 0.001436 696.5729 281 0.001470 679.7634 310 0.001415 706.5141 a Data were corrected from the diamagnetic contribution by using the magnetic susceptibility of Py3M02(HPO4)4Cl at room temperature (XM = -100x10'6 emu mol'l).129 121 Figure 24. Temperature dependence of the corrected magnetic susceptibility of K3M02(HPO4)4. 122 00¢ 00m 3 265E 00m 00v WX/I 000 000 123 Figure 25. X-band (9.460 GHz) EPR spectrum of a frozen solution of K3M02(HP04)4 in 7.5 M H3PO4 at 5 K. 124 4} L L ! L 1 l l 3300 3400 3500 H/G . 1 I 3600 3700 Figure 25 125 M02(HPO4)43' impurities and no evidence for a distinct EPR signal attributable to M02(HPO4)44’ can be discerned. Thus, to the best of our knowledge, the magnetic properties of M02 phosphate series are accommodated by the electronic structure model of M-n-M (D 4h) dimers. Solids and solutions of M02(HPO4)44’ are pink in color. Figure 26 displays the electronic absorption spectrum of the dimer in 2 M H3PO4. The lowest energy absorption band (A max = 516 nm, E = 196 M'1 cm’l) is comparable in energy and intensity to that of M02(SO4)44‘ and by analogy is assigned to the 52 -’ 55* transition. As expected for a dipole allowed transition, the band sharpens upon cooling solutions to 77 K, but the integrated intensity of the band remains constant; the absorption profile remains vibrationally featureless at low temperature. In addition to the 52 -’ 5 5 * transition, visible and ultraviolet absorption systems in M02(HPO4)44' have direct analogues in M02(SO4)44' spectra. The absorption (x = 408 nm, e = 27 M'1 cm'l) immediately to higher energy of the 52 -’ 55* transition is especially noteworthy; a similar positioned band in M02(SO4)44' has tentatively been assigned to the n -' 5* transition.129 The electronic absorption spectrum of M02(HPO4)43' (Figure 27) is dominated by a prominent absorption profile in the near- infrared (71max = 1438 nm; e = 180 M'1 cm'l) and two weak visible bands at 595 nm (6 =10 M‘1 cm'l) and 420 nm (E = 8 M’1 cm’l). Similar to the quadruply bonded dimers, the spectra of M02(II,III) sulfate and phosphate are nearly identical. The results of previous spectroscopic studies of M02(SO4)43' are consistent with the 126 Figure 26. Electronic absorption spectrum of M02(HPO4)44' ion in 2 M H3PO4 at room temperature. 127 000 005 E:\& com _ em 2.55 00m _ 00V . 00. OON ,_ulo '_w /3 128 Figure 27. Electronic absorption spectrum of K3M02(HPO4)4 in 2 M H3PO4 at 25 °C. 129 00— 000 000 R 8:5 5:2 000— 00: 00N— 000.— 000 00¢ , T I d E a 5 fl l_tu:1 l_|11|/3 130 assignment of the near-infrared absorption band to the 5 -' 5* (2B1u ‘- 2B23) transition.93’132 This significant red shift of the 5 -t 5* transition of mixed-valence M02 dimers is explained by the absence of two-electron term contributions to the transition energy for 0211451 configured ground state species. Similar to M02(SO4)43' ion, the 5 -> 5* absorption band of M02(HPO4)43' exhibits a vibrational progression in solution at room temperature. As shown in Figure 28, the vibrational peaks sharpen considerably upon cooling the phosphoric acid solution to 77 K. The 334 cm'1 spacing is consistent with a progression in the symmetric metal-metal stretching vibration. With regard to the visible absorption profile of M02(HPO4)43‘, the analogous absorption bands at 417 and 595 nm in M02(SO4)43' have been assigned by Hopkins et al. to the n -* 5* and n -r 5 transitions, respectively.129 These assignments, which resulted from a comparative analysis of M02(II,II) and M02(II,III) sulfate and M02 (III,III) phosphate spectra, are predicated on the assumption that bridging so42' and HPO42' ligands will perturb the electronic structure of a M02 core in a similar fashion. In this context, the absorption spectra of the homologous M02 phosphate series reported herein confirm this contention and accordingly support the n -* 5 and n -' 5* assignments for the visible absorption bands of Mo2(HPO4)43‘ and M02(SO4)43'. Raman spectra of solid samples of M02(HPO4)4“' (n = 2, 3, 4) at room temperature were recorded with 4880 A excitation light. The excitation frequency falls within the contour of the metal localized absorption profiles and significant enhancement of Raman peaks 131 Figure 28. Near-infrared absorption band of K3M02(HPO4)4 in a frozen phosphoric acid solution at 77 K. 132 000. 3 655E Eexa 00m. 005. _ _ 000. J 00v. _ 00m. _ 00: . aouoqlosqv 133 associated with metal-metal vibrations can be expected.133 Prominent bands are observed at 345, 352, and 356 cm’1 in Raman spectra of M02(II,II), M02(II,III), and M02(III,III) phosphate, respectively; this latter frequency is in excellent agreement with a previous study of M02(HPO4)42' in which a band at 358 cm'1 was attributed to valg(Mo-Mo).129 Assignment of the observed bands in the M02 phosphate series to valg(Mo-Mo) is consistent with their frequencies which will fall squarely within the 350-400 cm‘1 range which characterizes metal-metal vibrations in M-fl-M dimers and is also supported by the presence of energetically similar peaks in M02 sulfate dimers.134 Given the valg(Mo-Mo) assignment, a striking result of M02(HPO4)4n' (n = 2, 3, 4) Raman spectra is the observed increase in metal-metal stretching frequency with decreasing bond order. Precedent for this puzzling trend has heretofore existed as an anomaly in the Raman spectra of M02(II,II) (valg(Mo—Mo) = 371 cm'l) and M02(II,III) (valg(Mo-Mo) = 373, 386 cm'l) sulfate. An explanation for this apparent inconsistency is the ability of the bridging ligand to modulate the metal-metal frequency,135 although, the origin of this effect has not clearly been established. Nevertheless, the Raman results of the phosphate and sulfate series clearly demonstrate the ability of bridging ligands to vitiate predictions using simple bond order arguments. 2. Oxidation-Reduction Chemistry. The cyclic voltammo- gram of M02(HPO4)42' in 1 M H3PO4, shown in Figure 29, exhibits two oxidation processes. Reversible electrochemical behavior is suggested by linear plots of the cathodic and anodic currents vs. 134 Figure 29. Cyclic voltammogram of a 2.5 mM solution of the pyridinium salt of M02(HPO4)44’ in 2 M H3PO4. The scan rate was 2 mV s'l. 135 am 85E 58 .m> >\m m0- 0. HI 136 121/2 (scan rate, 12: 5-100 mV sec'l), and values of 0.98 i 0.04 for ratios of the anodic and cathodic peaks currents.112 The identities of these two electrode processes are revealed by electrochemical experiments on solutions of the quadruply bonded and mixed-valent phosphate dimers. Cyclic voltammograms identical to the one displayed in Figure 29 are observed upon scanning th electrode potential from the appropriate resting potentials of solutions of M02(II,II) and M02 (11,111) phosphates. These results lead us to assign the waves at -0.25 V and -0.67 V vs. SCE to the M02(HPO4)42' B” and M02(HPO4)43'/4' couples, respectively. The syntheses of M02 phosphate dimers and their reaction chemistry are easily understood in the context of the above electrochemcial results. Our observation that solids of M02(HPO4)44' precipitated from acidic solutions are inevitably contaminated with M02(HPO4)43’ is consistent with the negative potential of the M02(HPO4)43‘/4' couple with respect to the standard hydrogen electrode. Indeed M02(HPO4)44' in 2 M H3PO4 solution slowly converts to M02(HPO4)43’ as evidenced by the disappearance of the 1(52 -+ 5 5*) absorption band at 516 nm and the concomitant growth of the 2(5 -* 5*) absorption of Moz-(HPO4)43'. Closer analysis of the reaction reveals the quantitative conversion of quadruple bond dimer to mixed-valence species (I ”2 = 27.5 h) and chromatographic analysis identifies hydrogen as a reaction product. Toepler pumping the gas above completely reacted solutions reveals that 0.39 moles of 112/mole of M02(HPO4)44' is evolved thereby establishing the overall reaction stoichiometry as 137 M02(HPO4)44' + 11* -» M02(HPO4)43‘ + 1/2 H2 (29) Reduction of protons to hydrogen radicals is a highly energetic process [E(H+/H) = 2.6 V vs. NHE],136 and hence, direct reaction of the quadruple bond dimer with H+ to M02(HPO4)42' produce the mixed-valence dimer is not thermodynamically feasible. We propose that M02(HPO4)44’ reacts in acidic solution to M02(HPO4)42' which then reacts in a comproportionation reaction to produce mixed- valence dimer according to the following, M02(HPO4)44' + 211* -» M02(HPO4)42‘ + H2 (30) Mo2(HPO4)44' + M02(HPO4)42' _. 2M02(HPO4)43‘ (31) Consistent with reactions (30) and (31) is our observation that addition of a phosphoric acid solution of M02(HPO4)42‘ to one of M02(HPO4)44' leads to the immediate and quantitative production of Moz-(HPO4)43’. In view of previous studies of M-n—M dimers which have demonstrated large rate constants for electron exchange 83 reactions involving 5 orbitals, the facility of reaction (27) is not surprising since the comproportionation simply involves the transfer of an electron between the 5 orbitals of the M02 phosphato dimers. We therefore believe that the slow thermal conversion of M02(HPO4)44' to M02(HPO4)43' in phosphoric acid solutions, in the context of the above scheme, is a consequence of large kinetic barriers associated with reaction (30). 138 3. Photochemistry. In striking contrast to the slow thermal chemistry, reaction of M02(HPO4)44’ in acidic solution is markedly accelerated by ultraviolet irradiation. Spectral changes for the room temperature irradiation (A > 335 nm) of M02(HPO4)44' in 2 M D 3P0 4 are shown in Figure 30. An initial decrease in the intensity of the 52 -* 55* absorption band of M02(HPO4)44' is accompanied by an increase in the 5 4 5* absorption of M02(HPO4)43‘. Ensuing reaction of the mixed-valence dimer is revealed by the disappearance of the near-infrared 5 -’ 5* absorption. A series of weak changes in the visible region is observed with continued irradiation and the photolysis reaction terminates with an absorption profile distinguished by a strong absorption band at 385 nm and a weaker band at 685 nm. The absence of an isobestic point during the photolysis reaction, and the growth and decay of the characteristic near-infrared band of M02(HPO4)43' are consistent with a multistep photooxidation pathway with subsequent reaction of the M02 primary photoproduct. Further analysis of the M02(HPO4)44' photoreaction was pursued with investigations of the photochemistry of M02(II,III) and M02(III,III) phosphate dimers. Irradiation (A > 335 nm) of M02- (HPO4)43' in 2M D3PO4 produces the spectral changes illustrated in Figure 31. The intensity of the near-infrared band of M02(HPO4)43' monotonically decreases and bands at 420 and 540 nm appear during the initial stages of irradiation. With continued irradiation, the 540-nm band disappears and absorptions grow in at 385 and 690 nm. This latter spectrum is identical to the terminating spectrum of M02(HPO4)44' photolysis. The quantum yield for the 139 Figure 30. Electronic absorption spectral changes during irradiation (x > 335 nm) of M02(HPO4)44’ in 2 M D3PO4. Due to spectral congestion in the visibile spectral region, the 45-min trace is not illustrated between 350 and 800 nm. The absorbance sensitivity in the visible spectral region is twice that of the infrared spectral region. 140 on 25E :52 8: 8m. 82 8: Qua” - q 00 .50. me 00¢ 141 Figure 31. Absorption changes resulting from irradiating (A > 335 nm) 2 M D3PO4 solutions of M02(HPO4)43'. The visible absorbance scale is 5 times greater than the near-infrared absorbance scale. 142 a 65mm :22 82 8: 82 25 ._1 =8 :8 , 143 disappearance of Mo2(HPO4)43' is 0.046 (Xexc = 313 nm) whereas no reaction is observed for solutions of the mixed-valence dimer stored in the dark at room temperature. The appearance of the 540-nm band during the early stages of Mo2(HPO4)43' photolysis is consistent with the formation of M02(HPO4)42’ as a primary photoproduct. Support for this contention is provided by the photochemistry of M02(HPO4)42'. Figure 32 displays the absorption changes accompanying irradiation (A > 335 nm) of 2 M D3PO4 solutions of M02(HPO4)42‘. Smooth conversion of M02(HPO4)42' to the photoproduct, characterized by the 385- and 685-nm absorptions, is observed and in contrast to M02(II,II) and Mo2(II,III) photochemistry, isobestic points at 500 and 608 nm attest to a stoichiometric photoreaction. While the identity of this ubiquitous photoproduct has not yet been revealed, a clue to its identity is provided by comparison to the absorption spectra of other dimolybdenum oxo and hydroxo cores.l37'146 Table 5 displays absorption maxima and molar absorptivity coefficients of polynuclear molybdenum complexes. Examination of these data reveals that the absorption spectrum of the primary photoproduct of M02(HPO4)42' photolysis (Amax(€) : 385 nm (1800 M'1 cm'l); 690 nm (350 M'1 cm'1)) closely matches that of the M02(III,III)-u -dihydroxo complex.131 These results suggest that irradiation of M02(HPO4)42' triple bond dimer promotes ligand substitution of the phosphate to yield the p-hydroxo species. The slight wavelength discrepencies between the primary photoproduct and the M02(III,III)(u-OH)2(aq) complex may be due to the differences of the ancillary ligands (H20 vs. HPO42') on the M0204- 144 Figure 32. Electronic absorption spectral changes during irradiation (X > 335 nm) of M02(HPO4)42' in 2 M D3PO4. No absorption bands appear in the near-infrared spectral region during the photolysis reaction. Relative Absorbonce 145 U141, 2.5 l5 1. Ch 411 ° 4 l l I Oh I 400 500 600 700 A/nm Figure 32 146 S; 85084 .8556 .220me 5.5.5 +ee6~6mfioéomcz 2: $859. 48:85 5:49:09: 52 858m .3553 .8548 :3: 4er3560.392 ”2 $059. .0025 36562 E 69:; .6025 E +merN662 e 35:. .8095 5506692 52 can; .Scmcn 5.: 4.30892 5 8034 .8505: 4.242592 2: 8084 .825: 430892 f: 8.35% A2025 :.: +4359: .5.— mEE< 053 505338 «Bacon EBB—om 58:06. 5 5:3 65535565 .5 «8:82 :3553< 3.55055 .m 033. 147 .383 5:. p 285555 5 08 35508 550538 .832 a El «1; 3; NV— :1 S; 8208.... .853; >.> $889839: 858m .883“ $20~qumoz 8252 .855“ 82on888088382 88;? .8388 2.2.2 320388838382 838% .8va83. 88882 $83298er3va: 838% .853. .8382: 2.5.5 +n=8~5mfioéomoz A6353 n Bash. 148 01-024+ core. This hypothesis can be tested by investigating the substitution chemistry of Moz(}.1-OH)2(aq)4+ in phosphate containing solutions. Interestingly continued ultraviolet photolysis of the proposed Moz(u-OH)24+ phosphate photoproduct yields a species with the absorption spectrum shown in Figure 33. Detection of hydrogen above reacted solutions by gas chromatography suggests that the M02(p-OH)24+ core is photooxidized. This observation is confirmed by the fact that addition of K28208 to phosphoric acid 4- solutions of M02C13 yields a product with an absorption spectrum identical to the one shown in Figure 33. These results are consistent with photooxidation M02(}J-OH)24+ phosphate produce a high valent M02 oxo or hydroxo core (e.g. M02042+). The above photochemical results of M02(HPO4)4n‘ (n = 2, 3, 4,) dimers substantiate a sequential oxidation pathway. The relatively complicated series of spectral changes induced by irradiation of M02(HPO4)44' solutions are explained by a reaction scheme in which the final photoproduct (A max = 385 and 685 nm) is generated directly from M02(HPO4)42', which in turn is a primary photoproduct of M02(HPO4)43' which is the primary photoproduct of M02(HPO4)44'. Gas chromatography reveals a concomitant production of hydrogen with each of these discrete dimer photo- processes. A reaction sequence accommodating these observations is shown by equation (32). As previously noted, ensuing M02(HPO4)42' photo-oxidation leads to the generation of a presently undetermined higher valent molybdenum species and hydrogen. Thus the overall photochemical reaction corresponds to a multielectron process which involves the exchange of at least two electrons. 149 Figure 33. Terminating electronic absorption spectrum of photolyzed (7x > 335 nm) phosphoric acid solutions of M02(HPO4)4“’ (n = 2, 3, 4). 150 com com coo as: 8 2..me com oov . com _ _ _ _ _ _ ooueqlosqV 151 + H 4_ th. > 335 nm) M02(HPO4)4 \ : l/2H2 H 3_ S lit/(7x > 335 nm) 2_ M02(HPO4)4 \; M02(HPO4)4 (32) l/2H2 In an effort to ascertain the nature of the intermediates leading to hydrogen production, we photolyzed solutions of M02(HPO4)44' and M02(HPO4)43' under atmospheres of N20, which is an effective trap of highly energetic radical intermediates. Ultraviolet irradiation (xcxc > 335 nm) of M02(HPO4)44' and M02(HPO4)43' in N20 saturated 2 M H3PO4 solutions yields the one-electron photo- oxidized M02 phosphato complex and N2. These results are consistent with the generation of H atoms as primary photoproduct of M02 phosphato dimers because nitrogen, which is generated at the expense of hydrogen formation, is a product expected from a reaction 0.147 of hydrated hydrogen atoms with N2 Annihilation of two hydrogen atoms or subsequent oxidation of the starting complex by a hydrogen atom to yield H' followed by a facile proton trapping reaction will result in hydrogen production. Insight into the nature of the photoactive state of the M02 phosphate systems is provided by the wavelength dependence of M02(HPO4)4“' (n = 2, 3, 4) photoprocesses. The photochemistry of each of the M02 phosphate system is inhibited as the irradiation 152 wavelength is increased. Illustrative of this reactivity trend is the photooxidation reaction M02(HPO4)43' which proceeds promptly with ultraviolet irradiation but is suppressed upon shifting the excitation wavelength into the visible spectral region; Figure 34 summarizes the wavelength dependence of the reaction quantum yield. No photoreaction can be detected when M02(HPO4)43' solutions are irradiated at wavelengths longer than 475 nm. Between 300 and 400 nm, the quantum yield increases monotonically with decreasing irradiation wavelengths. And for wavelengths less than 300 nm, the quantum yield asymptotically approaches a limiting value of 0.05. Similarly, facile conversion of M02(HPO4)44' is observed for excitation wavelengths less than 367 nm, whereas solutions of M02(HPO4)44' do not photoreact with irradiations into the 1( 52 -> 65*) absorption band. These results imply that the photoreactivity of the M02 phosphato dimers is not associated with the lowest energy metal localized transition of the M02 core but rather with high energy excited states that lie in the ultraviolet spectral region. In an effort to identify these high energy photoactive excited states, the ultraviolet absorption spectra of the M02 phosphato complexes in 2 M H3PO4 was recorded. As illustrated in Figure 35, a prominent absorption band dominates the ultraviolet spectral region of the three phosphato dimers (X max, M02(HPO4)44' = 206 nm; xmax, M02(HPO4)43' = 210 nm; xmax, M02(HPO4)42' = 220 nm). The relative insensitivity of the band maximum to the oxidation state of the M02 core suggests that the transition is metal localized. Theoretical calculations predict that the allowed Tl -> n * transition of 153 Figure 34. Action spectrum of the photolysis reaction: M02(HPO4)43' + H+ -* M02(HPO4)42’ + 1/2 H2. Quantum yields were measured by monitoring the 6 -9 5* absorption band of M02(HPO4)43'. 154 com 554 00¢ 8m 2:5 00m OON . 0.. (p 60' 155 Figure 35. Ultraviolet absorption bands of M02(HPO4)44' (---), M02(HPO4)43' (—), and Moz(HPO4)42‘ (- - -) in 2 M H3PO4 at 25 °C. 156 .0. ........ ..... 00'- 00000000 aaaaaaaa 0000000 ...... 0000000000000000 O ....... oooooooo 0000000 ooooooo ...... ....... to. 00¢ ‘ ‘ ‘ \ “ ‘ “" ‘ “‘ ‘ “ “l‘ ‘ QQCNQHOmDaN 400 300 X / nm Figure 35 200 157 MA-M dimers should be observed in ultraviolet energy region,148 149 and polarized absorption spectroscopic studies support this contention. Accordingly, we assign the 206 nm absorption of M02(HPO4)44' to the TI -* 11* transition. The modest red-shift of the band across the M02(II,II), M02(II,III), and M02(III,III) series is consistent with a concomitant decrease in the metal-metal interaction. An additional absorption band is observed at 237 nm of M02(III,III) phosphate. That this absorption is not present at lower energies in the spectra of M02(II,III) and M02(II,II) phosphato dimers precludes its assignment to a transition corresponding to oxidation of the metal core, such as charge-transfer-to-solvent (CTTS), because absorption bands attributable to transitions of this type would exhibit a marked red-shift with reduction of the metal core. Conversely, the absence of this band in M02(II,II) and M02(II,III) phosphate spectra is presumably due to its blue-shift into the hard-ultraviolet thereby suggesting that the transition exhibits significant LMCT character. Thus, although the TH!“ and LMCT excited states of the M02(II,III) dimer are potentially photoreactive, our inability to detect charge transfer excited states in the ZOO-250 nm region for the M02(II,II) and M02(II,III) phosphato complexes implies that the photoreactivity of these species is derived exclusively from TI -’ TI* excitation. From molecular orbital point of view, this assignment is consistent with the observed chemistry because the metal centered TI * antibonding orbital exhibits significant electron density away from the metal-metal core along the z direction. Thus promotion of one electron from TI bonding to n * antibonding orbitals increases the probability of electron transfer 158 from the metal core to substrates occupying the open axial coordination sites. It is at these axial coordination sites which protons are most likely to reside. The smooth conversion of M02(II,II) phosphate to M02(III,III) phosphate dimer with the concomitant production of hydrogen clearly demonstrates the capacity of the M-ILM core to engender multielectron photochemical transformations by coupling the oxidation-reduction chemistry of the two metal centers along a controlled reaction pathway. It is evident from the reduction potentials of the M02 phosphato complexes that the overall reaction is thermodynamically favored and hence the stabilities of the M02(HPO4)44' and M02(HPO4)43' ions in H3PO4 must necessarily arise form large kinetic barriers associated with oxidation of the M02 cores. Our results demonstrate that these barriers are easily surmounted along photochemical reaction pathways. The 11 -’ 11* assignment of the photoreactive state of M02 phosphato complexes conforms with the general reactivity pattern which has developed for M-n-M dimers in acidic solution. Namely, low-energy metal- localized excited states are not responsible for the photooxidation chemistry of M-n-M dimers in protic environments. We believe that the proposed photochemical mechanism is not specific to M02 phosphato complexes but may be extended to accommodate the photochemistry of other M0208 complexes in acidic solution.94 For instance, the electronic absorption spectra of M02(SO4)43’ and M02(SO4)44' are completely analogous to the spectra of the corresponding M02 phosphato complexes in that an intense absorption band whose energy is independent of the oxidation state 159 of the M02 core (X max = 239 and 235 nm of M02(SO4)43' and M02(SO4)44', respectively) dominates the ultraviolet region. These data support assignment of the band to a 1'! -t 1'!" transition, and it is reasonable to propose that the photoreactivity of the M02(SO4)44' system, similar to the phosphato series, originates from TI -* 11* excitation. The two-electron photochemistry of the M02 phosphate system, as opposed to the one-electron transformations of the sulfate system, reflects the increased susceptibility of the M02 core toward oxidation when ligated by phosphate. This difference between the oxidation-reduction properties of M02 phosphato and sulfato complexes may derive from the fact that the ligands of the former complex possess a dissociable proton. In this connection, phosphoric acid and its ions exhibit enhanced proton ionization upon coordination to a metal; and ionization is enhanced further when the 150,151 ions chelate a metal center. In view of the relatively high charge of the M02 core and the bridging coordination geometry of phosphate, significantly reduced interactions of a proton with the 3' may be expected. Weak or complete dissociation of 2.. oxygens of P04 a proton from one or more of the HPO4 ligands will increase the overall negative charge of the complex and facilitate oxidation of the M02 core. If this model is correct then the reported electrochemistry and photochemistry should exhibit a large pH dependence. Unfortunately, the insolubility and instability of M02 phosphate dimers in H3PO4 solution at concentrations less than 2 M precludes pH dependence studies. The ability of the phosphate ion to significantly shift the oxidation potential of metal cores to more 160 positive values may be responsible for the unique photooxidation chemistry of the dimolybdenum phosphate system in acidic solution. CHAPTER V PHOTOCHEMISTRY OF A NOVEL DIMOLYBDENUM DIPHENYL PHOSPHATE IN NONAQUEOUS SOLUTION A. Background The multielectron photochemistry of quadruply bonded M02(HPO4)44' is derived clearly from sequential one-electron oxidation steps. The desired two-electron oxidative-addition pathway described in Scheme III is not observed. However, the photochemistry of the M02(HPO4)4n' system is similar to other MA-M systems in acidic solution and establishes a general reactivity pattern: M-n-M dimers do not exhibit appreciable lifetimes in acidic solution and consequently photoreactivity is confined to high energy, pseudo unimolecular one—electron processes. For the specific case of the M02(HP04)44' system, the 66* excited state is short lived and the photoreactivity is derived from the one-electron reduction of H+ by the highly energetic TI TI * state (Figure 36). The issue of paramount importance to multielectron reactivity is by what mechanism the 66* excited state is deactivated in acidic solution. 161 162 Figure 36. Photoreaction pathway of "M0208" complexes in acidic solutions, where "M0208" corresponds to M02(aq)x4+, M02(SO4)44’, and M02(HPO4)44-. Energy/av 6.0T- 2_4,_ 163 + H 1 a: A1g( nn ) 1/2 H2 + M—“iLZM 1 at: _ A2u(6 6 ) 3 at: —) A2u(55 > 1 2 ’— Alg(5 ) Figure 36 164 X-ray crystallography has identified stable hydride M—M products from the reaction of protons with the quadruple bond core.152 In these complexes, the metal in the M-M core has been formally oxidized by one electron with subsequent two-electron reduction of H+ to H'. Alternatively, a resonant structure for an M—M core is one where the core is simply protonated, /"\ __ MEM — M (a) (b) (33) \ [Ill/a We believe that for most M-‘l-M systems the major contributing structure is (b). Indeed, we have shown the 66* excited states are efficiently quenched by protons and a quenching mechanism consistent with the above discussion is shown in Figure 37. In this scheme protons quench the 66* excited state to give a "hydride-like" intermediate which expels the proton from the metal core to yield ground state M-4-M dimer and H4". Therefore, the photon energy is wasted in the proton quenching process and reactivity derived from the 66* excited state is circumvented. These results suggest to us that if 66* multielectron photochemistry is to be observed, the proton quenching of 66* excited state and the energy consuming back proton transfer step must be inhibited. Owing to our knowledge of M02 phosphate systems coupled with the demonstrated photoredox chemistry of 165 Figure 37. Proton quenching of MAM excited states. 166 “V M-4-M* Figure 37 167 M02(HPO4)44', we initiated studies on the synthesis of M02 phosphates which could be introduced into aprotic environments. To this end, we have prepared the quadruply bonded tetrakis(diphenyl phosphate)dimolybdenum M02(02P(OC6H5)2)4 complex which exhibits good solubility in nonaqueous solvents. The compound reacts with oxidants such as NOBF4 to yield the one-electron oxidized mixed-valence species M02(02P(OC6H5)2)4B F4. Both the quadruple-bond and mixed- valence species have been characterized by X-ray crystallography, electronic spectroscopic, and electrochemical methods. In contrast to the high energy photochemistry of M02(HPO4)44' in acidic solution, the photochemistry of M02(02P(OC6H5)2)4 in halogenated hydrocarbon solvent is promoted by visible irradiation. The chemistry and photochemistry of this new complex is presented in this chapter. B. Results and Discussion 1. Structures of M02(02P(OC6H5)2)4-2THF and M02(02P(OC6H5)2)4BF4. The molecular structures of M02(02P(OC6H5)2)4-2THF and M02(02P(OC6H5)2)4BF4 are typical structures of multiply bonded dimolybdenum complexes.55 Molecules of the quadruply bonded dimer occupy general positions of the space group P21/n whereas the mixed-valent species is situated on inversion sites of the P1 space group. Crystallographic data for the two complexes are listed in Table 6. Molecular structures of the two dimers and atom numbering schemes are 168 Table 6. Crystallographic Data of Moz(02P(OC6H5)2)4°2TI-IF and M02(02P(OC6H5 )2)4BF4 M02(02P(OC6H5)2)4'2THF M02(02P(OC6H5)2)4BF4 space group P21/n P1 cryst syst monoclinic triclinic a, 13.359(3) 10.917(8) b, A 29.641(15) 11.793(4) c, A 14.719(3) 12.430(4) (2, deg 63.45(2) 6, deg 93.43 70.38(4) y, deg 70.64(5) V, A3 5818(6) 1316.8(12) d, g/gm 1.522 1.608 Z 4 1 radiation Mo Kc Mo KG abs coeff(}.1(Mo 1(a)), cm-1 5.964 6.625 26 limits, deg 45 55 no. of reflens collcd 7973 12192 no. of unique reflens 7597 6096 no. of unique reflens used with F0 >30 2363 5107 R(F) .091 .027 Rw(F) .103 .027 GOF 10.7 .654 169 illustrated in Figures 38 and 39. The positional parameters for heavy atoms (e.g. Mo, P, and O) are listed in Tables 7 and 8 for M02(02P(OC6H5)2)4'2THF and M02(02P(OC6H5)2)4BF4, respectively. Tables 9 and 10 list selected bond distances; and the selected bond angles are listed in Tables 11 and 12 for quadruply—bonded and mixed-valent species, respectively. Even though the molecular structures of both compounds appear similar, closer examination reveals some significant differences between the two molecules. The most noticeable difference is the Mo-Mo bond distance which increases from 2.154 A in M02(02P(OC6H5)2)4-2THF to 2.191 A in M02(02P(OC6H5)2)4BF4. This increase of 0.04 A resulting from the oxidation of the dimer by one electron is comparable to that observed for the M02 sulfates which exhibit a lengthening of the Mo-Mo bond by 0.05 A upon oxidation (2.111 A and 2.164 A for quadruple bond and mixed- valence complexes, respectively)13°. One—electron oxidation of the quadruple bond reduces the formal 6 bond order from 1 to 1/2. Besides a decreased bond order, the increased formal charge on each metal (from +2 to +2.5) causes a constriction of the valence orbitals of the metal atoms and corresponding decreased orbital overlap. Because of the longer metal-metal bond distance in the mixed— valent species, there is less steric strain of phosphate ligands bridging the one-electron oxidized metal core. The five membered M02(02P) ring possesses trans O-Mo-O angles close to 180° in the mixed-valent species as compared to the more strained angles in the quadruple-bond compound (e.g. 1710 and 1670 respectively). This is manifested in the mixed-valent species to possess a higher 170 Figure 38. Structure and labeling scheme for the quadruply bonded complex M02(02P(OC6H5)2)4'2THF. Atoms are represented by their 50% probability ellipsoids. Due to structural congestion, the two tetrahydrofuran molecule are not illustrated. 171 Figure 38 172 Figure 39. Structure and labeling scheme for the mixed-valence complex M02(02P(OC6H5)2)4BF4. Atoms are represented by their 40% probability ellipsoids. Figure 39 174 Table 7. Selected Atomic Positional Coordinates of M02(02P(0C6H5)2)4°2THF atom x y z Mo(l) 0.82506 0.11940 0.26235 Mo(2) 0.66364 0.11702 0.25292 P(l) 0.74433 0.10992 0.45245 P(2) 0.73685 0.21593 0.28329 P(3) 0.74340 0.13519 0.06147 P(4) 0.75817 0.02147 0.24665 0(1) 0.64447 0.10769 0.39247 0(2) 0.83149 0.10697 0.40465 0(3) 0.83597 0.19041 0.28522 0(4) 0.64912 0.18690 0.27072 0(5) 0.65419 0.13151 0.10958 0(6) 0.84402 0.13110 0.11602 0(7) 0.84850 0.05124 0.23834 0(8) 0.66053 0.04395 0.23834 0(9) 0.73567 0.06964 0.52455 0(10) 0.75655 0.15712 0.50616 0(11) 0.72432 0.24283 0.37444 0(12) 0.74407 0.25134 0.20039 0(13) 0.74688 0.09620 -0.01629 0(14) 0.73748 0.18258 0.01391 0(15) 0.76189 -0.00184 0.34847 0(16) 0.75853 -0.02398 0.18216 Table 8. Selected Atomic Positional Coordinates of 175 M02(02P(0C6H5 )2)4BF4 atom x y z Mo(l) 0.00738 -0.09327 ~0.00349 P(1) 0.25799 0.01683 -0.l6557 P(2) 0.13565 -0.13277 0.19945 0(1) 0.19858 -0.10096 -0.11524 0(2) 0.08782 ~0.20395 0.15108 0(3) -0.l7874 -0.11616 0.10870 0(4) -0.07255 -0.01050 -0.15827 0(5) 0.26835 0.07531 -0.30863 0(6) 0.40088 -0.01817 -0.l4397 O(7) 0.10424 -0.20480 0.34273 0(8) 0.28892 -0.14036 0.15318 l 7 6 Table 9. Selected Bond Distances of M02(02P(0C6H5)2)4°2THF Bond Distances/A Bond Distances/A Mo(l) -Mo(2) 2.154 Mo(l) -O(2) 2.123 Mo(2) -O(1) 2.103 -O(3) 2.135 -O(4) 2.098 -O(6) 2.211 -O(5) 2.149 -O(7) 2.078 -O(8) 2.205 P(1)-O(1) 1.556 P(2) -O(3) 1.524 -0(2) 1.399 -O(4) 1.457 -O(9) 1.606 -O(12) 1.617 P(3) -0(5) 1.426 P(4) -O(7) 1.506 -O(6) 1.529 -O(8) 1.482 -O(13) 1.629 -O(15) 1.648 -O(14) 1.569 -O(16) 1.648 177 Table 10. Selected Bond Distances for Moz(02P(0C6H5)2)4BF4 Bond Distances / A Mo(l) -MO(1) 2.191 Mo(l) -O(1) 2.082 -O(2) 2.073 -O(3) 2.067 -O(4) 2.074 P(1)-0(1) 1.513 -O(3) 1.509 -O(5) 1.572 -O(6) 1.559 P(2) -O(2) 1.514 -O(4) 1.515 -0(7) 1.563 -O(8) 1.562 1 7 8 Table 11. Selected Bond Angles of M02(02P(0C6H5)2)4-2THF Bond Angle Angles/o Bond Angle Angles/o Mo(l)-Mo(2) -O(1) 97.0 Mo(2)-Mo(l) -O(2) 92.2 -O(4) 93.4 -0(3) 95.8 -O(5) 93.2 -O(6) 96.6 -O(8) 93.0 -O(7) 96.8 0(2)-MOO) -O(3) 91.0 0(1)-Mo(2) -O(4) 89.9 -O(6) 171.1 -O(5) 168.9 -O(7) 90.1 -O(8) 92.9 0(3)—Mo(1) -0(6) 89.3 0(4)-Mo(2) -O(5) 85.7 -O(7) 167.3 -O(8) 172.9 0(6)-Mo(l) -O(7) 87.7 0(5)-Mo(2) -O(8) 90.8 0(1)-P(1) 00) 115.0 -O(9) 104.5 0(2)-P(1) -O(9) 112.4 -O(10) 111.9 -O(10) 103.8 0(3)—P(2) -0(4) 113.6 -O(11) 111.8 0(4)-P(2) -O(11) 106.3 -O(12) 104.2 -O(12) 111.8 0(5)-P(3) -0(6) 117.9 -O(13) 110.8 0(6)-P(3) -0(13) 104.6 -O(14) 105.6 -O(14) 108.9 0(7)-P(4) -O(8) 114.7 -O(15) 109.8 0(8)-P(4) -O(15) 111.3 -O(16) 113.7 -0(16) 106.0 Mo(l) -0(2)-P(1) 120.0 Mo(l) -O(1)-P(1) 113.5 -O(3)-P(2) 115.8 -O(4)-P(2) 121.3 -O(6)-P(3) 112.1 -O(5)-P(3) 120.1 -O(7)-P(4) 115.2 -O(8)-P(4) 113.3 179 Table 12. Selected Bond Angles of M02(02P(0C6H5)2)4BF4 Bond Angle Angles/o MO(1)-MO(1) -O(1) 94.6 -O(2) 94.8 -O(3) 94.9 -O(4) 94.5 0(1)-M00) -O(2) 89.7 -0(3) 170.5 -O(4) 90.1 0(2)-M00) -O(3) 88.6 -O(4) 170.8 0(3)-Mo(l) -O(4) 90.1 Mo(l) -O(1)-P(1) 117.6 -0(2)-P(2) 116.9 -O(3)-P(1) 118.2 -0(4)-P(2) 117.2 180 symmetry structure, which includes a center of inversion, despite identical virtual symmetries (D 4 h) for both complexes. 2. Oxidation-Reduction Chemistry. The cyclic voltammogram of M02(02P(0C6H5)2)4 in CH2C12 is shown in Figure 40a. Two reversible oxidation processes are observed upon anodically scanning solutions containing the quadruple bond complex. Reversible electrochemical behavior is indicated by linear plots of the cathodic and anodic currents vs. 121/2 (scan rate, 12: 20- 200 mV seC‘l), and values of 0.99 :1: 0.02 for ratios of the anodic and 112 cathodic peak currents. Figure 40b illustrates the cyclic voltammogram of the mixed-valence species M02(02P - (0C6H5)2)4BF4. The two complexes exhibit identical E1/2 values (+0.067 and +0.997 V vs. SCE), with the quadruple bond having two reversible oxidations, whereas the mixed-valence complex possesses one reversible oxidation and one reversible reduction. These results clearly show that the wave at 0.067 V vs. SCE corresponds to the M02(02P(0C6H5)2)4°/+ couple. Owing to the reversibility of the wave at 0.997 V vs. SCE at all scan rates, we assign this wave to the M02(02P(0C6H5)2)4+/2+ couple. Table 13 collects the reduction potentials of the "M0203" complexes M02(SO4)44’,13° Moz(HP04)44', and M02(02P- (0C6H5)2)4. Inspection of these data obviously reveal that the ligands perturb the redox properties of the metal core in a different manner. Not surprisingly, proton association with the ligands of the M02 sulfato and diphenyl phosphate complexes is unlikely and the M02(II,III)/M02(II,II) reduction potentials are similar. It is 181 Figure 40. Cyclic voltammograms of (a) M02(02P(0C6H5)2)4 and (b) M02(02P(0C6H5)2)4BF4 in CH2C12 solution at 23 °C. NBu4PF6 was used as the supporting electrolyte, scan rate = 20 mV s'l. 182 (a) (b) O . 5 EN vs. SCE Figure 40 183 Table 13. The Formal Reduction Potentials of the M02(II,III)/(II,II) and M02(III,III)/(II,III) Couples of M02 Sulfato, Phosphato, and Diphenyl Phosphate Complexes Complexes E1/2/V vs. SCE E1 [ZN vs. SCE M02(II,III)/(II,II) M02(III,III)/(II,III) Sulfato +0.25 a Phosphato -0.67 -0.24 Diphenyl phosphate +0.06 +1.00 a Not available. 184 noteworthy that the M02(III,III)/M02(II,III) couple for the sulfato complex lies in the background of solvent (in this case H20) and heretofore has not been determined. 0n the basis of the energy separation between the M02(II,III)/M02(II,II) and M02(III,III)/M02(II,III) couples of the diphenyl phosphate complex, we predict the latter couple of the sulfato complex to lie at 1.19 V vs. SCE. As originally inferred in Chapter IV, the very negative potentials of M02(HP04)44’ oxidation processes is attributed to the dissociation of a proton from one or more of the HP042' ligands. The electronic and structural similarities of the HP042' and 029mm- ligands are virtually identical and hence the only logical explanation for the ~l.0 V shift to negative potentials for M02(HP04)44' is that the negative charge of the M02(HP04)44' is increased by proton dissociation, thereby facilitating oxidation of the M02 core. Similar to the M02(HP04)44' system, it is easy to understand the syntheses of the molybdenum diphenyl phosphates and their reactions on the basis of the oxidation-reduction potentials. Although oxidation of the quadruple bond complex by N0BF4 proceeds smoothly to the pure mixed-valence compound Moz(02P(0C6H5)2)4BF4, oxidation to M02(02P(0C6H5)2)42+ by using NOBF4 can not be achieved. From the redox potentials, it is understandable that the N0+INO couple (estimated to be ~0.85 to 1.0 V vs. SCE in nonaqueous solution)153 is insufficient to bring about the second oxidation. Even stronger oxidants, such as magic green (tris(4-bromophenyl)aminium hexachloroantimonate (E1 ,2: 1.04 V vs. SCE) or dichloroiodobenzene (estimated 0.9 to 1.2 V vs. SCE) with just enough potential to generate the dication M02(02P(0C6H5)2)42+, 1 8 5 yield only the mixed-valence M02(02P(0C6H 5)2)4+ product. This may be a result of significant kinetic barriers for generating the dication. Unfortunately, solutions of the M02(02P(0C6H5)2)4BF4 dimer decompose upon heating, and hence the oxidations must be run at or below room temperature. An alternate method to the synthesis of the dication species is by using bulk-electrolysis. The dication complex can be generated in solution at electrodes whose potentials are as high as 1.7 V vs. SCE. Unfortunately, this dication species failed to recrystallize. Although small concentrations of impurities (mostly higher-valent molybdenum complexes) from the bulk-electrolysis obscure the visible absorption region (from our experience on the M02 phosphate system, we believe that the dication should exhibit weak absorption in visible region), the absence of near-infrared absorption suggests that the dication M o 2( 0 2 P ( 0 C 6 H 5 )2 )4 2 + species is generated in solution and that electrochemical preparation of this triply bonded species is plausible. 3. Magnetic and Spectroscopic Characterization of M02(02P(0C6H5)2)4 and M02(02P(0C6H5)2)4BF4. Typical of M-‘LM complexes, M02(II,II) diphenyl phosphate has a 0211462 ground state electronic configuration and will possess a diamagnetic ground state. Oxidation by one electron to yield the paramagnetic mixed-valence complex whose EPR spectrum recorded on frozen solutions at 5 K is shown in Figure 41. An axial doublet signal with g1 = 1.909 and g" = 1.863 is observed at low temperatures and is consistent with a species in which the unpaired electron is coupled between two equivalent molybdenum nuclei. 186 Figure 41. X—band (9.434 GHz) EPR spectrum of a frozen solution of M02(02P(0C6H5)2)4BF4 in CHZCIZ solution at 17 K. 187 3400 3500 3600 3700 3800 H / G Figure 41 188 The absorption profiles of the M02 diphenyl phosphates in the visible and near infrared are directly analogous to the phosphates. Figure 42 illustrates the electron absorption spectrum of the M02(02P(0C6H5)2)4 complex in CH2C12 solution. The lowest energy absorption band ()‘max = 515 nm, E = 156 M'1 cm'l) is comparable in energy and intensity to that of M02(HP04)44'. The next higher energy absorption band at 404 nm (E = 38 M’1 cm'l) is assigned to the TI -’ 6* transition.129 The electronic absorption spectrum of M02(02P(0C6H5)2)4BF4, shown in Figure 43, is dominated by a prominent absorption profile in the near-infrared (Xmax = 1469 nm (2:142 M'l cm'1)) and a weak visible band at 600 nm (e: 27 M-1 cm'l). From the previous spectroscopic studies on M02(SO4)43', we are able to assign the near-infrared absorption band to the 6 4 6* (23m .— 213 2g) transition.93’132 As with the M02 phosphates, the large red-shift of the 6 -’ 6* transition in mixed-valence M02 dimer is explained by the absence of two electron spin pairing term contributions of the 0211461 configured ground state species. Similar to M02(S04)43‘ and M02(HP04)43‘ ions, the 6 -' 6* absorption band of M02(02P(0C6H5)2)4BF4 exhibits a vibrational progression in solution at room temperature. The 308 cm'1 energy spacing is consistent with the symmetric metal-metal stretching vibration. The lower energy of the progression as compared with the M02(S04)43' and M02(HP04)43' (352 and 334 cm’l, respectively) is probably due to the longer M-M bond distance of this mixed-valent M02 diphenyl phosphate. The 600-nm band in M02(02P— (0C6H5)2)4BF4 is a direct analogue to the lower energy absorption in M02(II,III) sulfato and phosphato complexes and is by analogy, 189 Figure 42. Electronic absorption and emission spectra of M02(02P(OC6H5)2)4 in CH2C12 solution at 25 0C. 190 Ne 2.88 8:2 com 82 . o8 . oom cow 0 " // 2 3. I 2 z 2 I 2 I 2 ill/ 2 m , oo. 19 ll‘ OON l__uJ:) L__|/\|/3 191 Figure 43. Electronic absorption spectrum of M02(02P(0C6H5)2)4BF4 in CHZCIZ solution at 25 OC. 192 2:2 CORP 000.. com? 003. cc? comp CO: COS. Q. 2:32 com com com com com owl ON? _ _ _ Nx 1’ _ _ _ _ om 193 assigned to the 11 4 6 transition. The lack of higher energy absorptions in the visible spectral region could be explained by the strong ultraviolet absorption from the 1'! -’ 11 * transition of the aromatic ring of the ligands (A max = 260 nm) which tails into the visible and presumably shields the weak absorption of the other metal-localized transitions. In spite of the similarities of the absorption properties of M02(02P(0C6H5)2)4 to M02(HP04)44', there is one major spectroscopic difference between these two compounds. The new M02(II,II) diphenyl phosphate species emits in nonaqueous solution as shown in Figure 42. This observation supports our original hypothesis that M0208 quadruple-bond dimers do not emit in aqueous solution because of proton quenching of the 66* state. Because the emission quantum yield for this complex is quite low, the lifetime of the excited state was measured by transient absorption spectroscopy. The transient absorption band at 460 nm decays exponentially after the 532-nm laser excitation pulse (FWHM = 25 ps). By using a monoexponential fitting method, the decay lifetime is 68 ns (Figure 44). This excited state lifetime is long enough to permit bimolecular reactivity. 4. Photochemistry. In striking contrast to the high energy photoreactions of M02(HP04)4“' in acidic solution, M02(02P(0C6H5)2)4 reacts with halogenated hydrocarbon solvents upon low energy irradiation. This low energy photoreaction directly originates from the long-lived 1(66*) excited state. Spectral changes for the photoreaction of M02(02P(0C6H5)2)4 in 1,2-dichloroethane 194 Figure 44. Transient absorption kinetics of M02(02P(0C6H5)2)4 in THF solution. The decay kinetics were recorded at 460 nm and excitation was with a 532 nm Nd:YAG laser pulse (FWHM = 25 ps). 195 2:22 82 3. 2:22 — [(31.13 . . nm .. 5 p51. cod 2 o' EIDNVHHOSHV V 196 upon irradiation into 62 -t 66* absorption (xexc - 530 nm) are shown in Figure 45. An initial decrease in the intensity of the 62 -’ 6 6 "' absorption band is accompanied by a concomitant increase in the near-infrared region. It is noteworthy that the near-infrared absorption arising from the photoreaction of M02(02P(0C6H5)2)4 in C1CH2CH2C1 is not identical to the absorption of M02(02P(0C6H5)2)4+BF4'. The band maximum of the photolyzed product is red-shifted (1470 nm for M02(02P(0C6H5)2)4+BF4' and 1501 nm for photoproduct) and more intense. Because Cl' is the most possible anion generated in dichloroethane solution, the oxidized photoproduct is very likely to have chloride in the axial coordination position. This assumption has been verified by Fast Atom Bombardment Mass Spectroscopy (FABMS). The highest molecular weight cluster found in photolyzed CICHZCHzCl solutions of M02(02P(0C6H5)2)4 displays a parent ion peak at 1225 amu (Figure 46) which is consistent with the molecular weight of M02(02P(0C6H5)2)4Cl (mol wt = 1223.33). We attribute the red- shifted and increased absorption intensity of the 6 -* 6* transition of the photolyzed product to coordination of chloride in the axial coordination site. This hypothesis is confirmed with the independent synthesis of the M02(02P(0C6H5)2)4Cl. The absorption spectrum of CH2C12 solutions of M02(02P(0C6H5)2)4 recorded after addition of 0.5 equivalents of the oxidant C6H51C12 exhibits a strong absorption band at 1494 nm (5 ~ 362 M‘1 cm'l) as shown in Figure 47. The intensity and energy of this band closely match that of the final near-infrared photolysis spectrum ()‘max = 1501 nm, 8 ~ 391 M'1 cm'l). The photochemically and chemically produced spectra also 197 Figure 45. Electronic absorption spectral changes during irradiation ()‘exc > 530 nm) of M02(02P(0C6H5)2)4 in ClCH2CH2Cl solution at 25 0C. The visible absorbance scale is 2 times greater than the near- infrared absorbance scale. 198 a. 23E EC\K comp 092— 009. OONF Go: 000* com com cox. com com 00?. 4 1 _1 fl 2 _ _ _ a _ _ _ SE 6 x T mx {/77 7% .. w 1W2» 8 /, o _ / H . 3,, n O 2.2 a 5:. o 2,, , // 199 Figure 46. FABMS spectrum of a photolyzed ClCHzCHZCl solution of M02(02P(OC6H5)2)4. 200 e. 2.6.6 582.2 omNF ovmr. ONNP ooww omww t P _.g3?cwfiagcwgfi23:532-5e2mgét wegégig t ‘1 4 ($2 BONVGNHBV 201 Figure 47. Electronic absorption spectrum of the product of the reaction between M02(02P(0C6H5)2)4 and 0.5 equivalent of C6H51C12 in CH2C12 solution at room temperature. 2.2. 65mm 55 K oovfi coma ocofi cow coo oov 202 . _ _ _ 2 _ 2 _ _ 2 2 _ HDNVHHOSHV 203 are in close agreement in the visible except that the 450-nm band is much more pronounced in the latter. Additionally, the 605-nm band present in the spectrum of photolyzed solutions is not as pronounced for solutions containing chemically oxidized product. Interestingly the M02(02P-(0C6H5)2)4BF4 species does exhibit a band at 605 nm and more intense absorption to the ultraviolet. These results suggest that photolyzed solutions consist of a mixture of M02(02P(0C6H5)2)4C1 and M02(02P(0C6H5)2)4+ owing to the establishment of an equilibrium between coordinated chloride and free mixed-valence species, M02(02P(0C6H5)2)4C1 3 M02(02P(0C6H5)2)4+ + Cl‘ (34) Since the absorption of M02(02P(0C6H5)2)4Cl is much stronger than M02(02P(0C6H5)2)4BF4 in the near infrared region, the final spectrum of photolyzed solutions in the near infrared region will be dominated by the appearance of M02(02P(0C6H5)2)4Cl with minor blue-shifts of the band maximum. Additionally, the visible absorption features of photolyzed solutions match those obtained from the sum of the absorption of M02(02P(0C6H5)2)4Cl and M02(02P(0C6H5)2)4+BF4'. The FABMS data in conjunction with chemical oxidation results suggest that the initial photoreaction step is M02(02P(0C6H5)2)4 + ClCH2CH2Cl .__Lv_. -CH2CH2C1 + [M02(02P(0C6H5)2)4+ a M02(02P(0C6H5)2)4Cl] (35) 204 The photogenerated chloroethane radical is a very reactive species, and thermal reaction with unreacted quadruple bond species is thermodynamically favorable, to yield the chloro-substituted product and ethylene, 'CH2CH2C1 + M02(02P(OC6H5)2)4 4 M02(02P(0C6H5)2)4C1 + CH2CH2 (36) As discussed above, the equilibrium described by reaction (34) is quickly established. Consistent with this proposed mechanism is our measurement of ethylene above reacted solutions. Gas chromatographic analysis reveals ethylene as the only detectable organic product, and manometric measurements determine 0.51 i 0.03 equivalents is produced. The slightly higher measurement of ethylene (according to reaction (35) and (36) only 0.50 equivalents of ethylene should be produced) is indicative of the secondary reaction, 'CH2CH2C1 + M02(02P(OC6H5)2)4C1 -* M02(02P(0C6H5)2)4C12 + CH2CH2 (37) which should become important as the concentration of M02(02P(0C6H5)2)4 is depleted. 0ur observation of the loss of the initial isosbestic point at long photolysis times in Figure 45 is consistent with reaction (37) becoming a significant pathway as the quadruple bond photoreacts. 205 The M02(02P(0C6H5)2)4 dimer also reacts with other halogenated reactants. The spectral changes of M02(02P(0C6H5)2)4 in CH2C12 during visible excitation (A exc > 530 nm) are shown in Figure 48. An absorption profile exhibiting a moderately intense band at 542 nm and a strong band at 420 nm combined with absorption in near infrared spectral region at 1494 nm. Although the final spectrum in the near infrared is very similar to the one- I electron oxidized species with Cl', the visible spectrum clearly differs from that obtained for ClCHZCHzCl photochemistry. These observations suggest the following photoreaction scheme, M02(02P(0C6H5)2)4 + CH2C12 JV... M02(02P(OC6H5)2)4C1 + °CH2C1 (38) 'CH2C1 + 'CH2C1 -’ C1CH2CH2C1 (39) The generated 1,2-dichloroethane can then photoreact with quadruply bonded species as described by reactions (35) and (36). However that the visible absorption spectrum is not consistent with exclusive production of M02 ( 0 2 P ( 0 C 6 H 5 )2 )4 C1 or M02(02P(0C6H5)2)4+ clearly identifies a more complicated photochemical scheme. The chloromethane radical is extremely reactive and it is reasonable to assume that the following reactions are competitive with dimerization, 'CH2C1 + M02(02P(OC6H5)2)4 -’ M02(02P(OC6H5)2)4CH2C1 (40) 'CH2C1 + M02(02P(OC6H5)2)4C1 4 CH2C1M02(02P(0C6H5)2)4C1 (41) 206 Figure 48. Electronic absorption spectral changes during irradiation ()‘exc > 530 nm) of M02(02P(0C6H5)2)4 in CH2C12 solution at 25 °C. 207 we 2&2 552K . com: ooc2 OOVH OONH coofi oc occ oov H u I 1‘ 11m 15 4 1 \ ooueqiosqv ' 208 The M02(III,III) organometallic products would definitely exhibit intense absorptions in the visible spectral region. However, the inorganic products of CH2C12 photolysis have not yet been isolated and thus it is impossible to unequivocally identify the photoreaction pathway from the data available at this time. CHAPTER VI FINAL REMARKS The studies described herein demonstrate that M-A-M complexes are good multielectron photoreagents. The redox activity of the individual metal centers of the binuclear metal core can be coupled to drive overall multielectron photoprocesses. This principle is demonstrated by the photochemistry of M02(HP04)4n' (n = 2, 3, 4) complexes in acidic solution in which the sequential one-electron chemistry of the metal centers provides the two-electron reduction of protons to hydrogen. An important discovery from these studies is that the 1(66"‘) excited state is efficiently quenched by proton transfer. Therefore, in acidic solution, the 1(66*) excited state is not preserved for bimolecular reaction. Our approach to circumvent this quenching process by employing synthetic strategies to design M-4-M complexes with similar structures but readily soluble in nonaqueous solutions was a successful one. The newly synthesized M02(02P(0C6H5)2)4 complex luminesces from the 1(66*) excited state in nonaqueous solutions at room temperature with an appreciably long lifetime of 68 ns. This long-lived singlet excited state dominates the photoreactivity of this molecule and in contrast 209 210 to M02(HP04)44', M02(02P(0C6H5)2)4 exhibits low energy two- electron photochemistry. This system is significant because it represents the first example of a multielectron process originating from the 1(66*) excited state. Although the M02(II,II) diphenyl phosphate photochemistry constitutes a two-electron process driven by a single electron, the reaction is not a concerted multielectron process of the 1(6 6 1") excited state. Molecular models of M02(02P(0C6H5)2)4 reveal that the metal-metal core, important for multielectron reactivity is shielded by the diphenyl phosphate ligands. These results suggest that M02(02P(0C6H5)2)4 photochemistry is confined to attack by substrate at the axial coordination sites of the metal-metal core. Quantum yield data support this analysis. The photoreaction quantum yield of M02(02P(0C6H5)2)4 with 1,2-dichloroethane decreases dramatically with increasing ligating ability of solvent. These data are displayed in Table 14. The quantum yield decreases from 0.040 to 0.001 by changing the solvent from 1,2-dichloroethane to acetonitrile. Computer generated molecular models of the M02(02P(0C6H5)2)4°2THF crystal structure shows that the two THF molecules occupying the axial sites insulate the metal-metal core from substrate. Consequently solvent coordination is reflected in the reduction of the photoreaction quantum yield. Therefore, substrates directly attack only one metal instead of the binuclear metal-metal core. Hence the advantages of using M-4-M complexes are attenuated by axial attack and one-electron photoredox chemistry is observed. This analysis suggests that M-A-M multielectron photo- chemistry will only be observed for substrates which can approach 211 Table 14. Quantum Yield Data for Photoreaction of M02(02P(0C6H5)2)4 and ClCHZCHZCl in Various Nonaqueous Solutions Solvent [C1CH2CH2Cl]/M (Pa ClCHzCHzCl 13(neat) 0.040 C6H6 9 0.031 THF 9 0.014 CH3CN 9 0.001 a Measured by using ferrioxalate actinometer. xexc = 546 nm. 212 the metal core in an equatorial position. Thus the steric congestion of the equatorial planes of the MA-M dimers must be reduced. The best structures would have two bridging ligands to prevent the free rotation of the 1(66 a") excited state and four small monodentate ligands to minimize steric crowding. The bridging ligands could be in either trans or cis positions. In summary, quadruply bonded MAM dimers are promising multielectron photoreagents. The photochemistry described herein represents the first examples of multielectron 1(66*) excited state chemistry. The singlet character of the 1(66*) excited state appears to be important in promoting multielectron reactivity but energy wasting quenching reactions, such as proton quenching, must be precluded. Additionally, attack by substrate in the equatorial plane of the metal-metal core appears to be a crucial feature for concerted multielectron reactivity. If these conditions are met, photochemical studies of M-n-M dimers should continue to open new avenues of excited state chemistry. 10. 11. 12. REFERENCES Adamson, A. W.; Sporer, A. H. J. Am. Chem. Soc. 1958,80, 3865. (a) Sutin, N.; Creutz, C. Adv. Chem. Ser. 1978.No. I68, 1.(b) Sutin, N.; Creutz, C. Pure Appl. Chem. 1980,52, 2717. Balzani,V.; Scandola, F. In. Photochemital Conversion and Storage of Solar Energy, Connolly, J. S. Ed.; Academic: New York, 1981, Chapter 4. Julliard, M.; Chanon, M. Chem. Rev. 1983, 83, 425. Kavarnos, G. J.; Turro, N. J. Chem. Rev. 1986, 86, 401. (a) Marcus, R. A. J. Chem. Phys. 1956, 24, 966. (b) Marcus, R. A. Ann. Rev. Phys. Chem. 1964, 15, 155. (a) Meyer, T. J. Acc. Chem. Res. 1978, II, 94. (b) Meyer, T. J. Prog. Inorg. Chem. 1983,30, 389. Sutin, N. Prog. Inorg. Chem. 1983, 30, 441. Sutin, N.; Creutz, C. J. Chem. Educ. 1983, 60, 809. Marcus, R. A.; Sutin, N. Biochim. Biophys.Acta, 1985,81], 265. Brunschwig, B. S.; Sutin, N. Comments Inorg. Chem. 1987, 6, 209. Isied, S. S. Prog. Inorg. Chem. 1984,32, 443. 213 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 214 (a) Mayo, S. L.; Eills, W. R.; Coutchley, R. J.; Gray, B. H. Science. 1986, 233, 948. (b) Axup, A. W.; Albin, M.; Mayo, S. L.; Coutchley, R. J.; Gray, B. H. J. Am. Chem. Soc. 1988, 110, 435. Elias, H.; Chou, M. H.; Winkler, J. R. J. Am. Chem. Soc. 1988, 110, 429. (a) McLendon, G.; Guarr, T.; McGuire, M.; Mimolo, K.; Strauch, S.; Tayler, K. Coord. Chem. Rev. 1985, 64, 113. (b) McLendon, G.; Guarr, T. Coord. Chem. Rev. 1985, 68, l. (c) Mclendon, G. Acc. Chem. Res. 1988,21, 160. Stein, G. Israel J. Chem. 1975, 14, 213. Balzani, V.; Bolletta, F.; Gandolfi, M. T .; Maestri, M. Top. Curr. Chem. 1978, 75, 1. (a) Maverick, A. W.; Gray, H. B. Pure Appl. Chem. 1980,52, 2339. (b) Maverick, A. W.; Gray, H. B. Science, 1981,214, 1201. Creutz, C.; Sutin, N. Natl. Acad. Sci. U.S.A. 1975, 72, 2858. Sutin, N. J. Photochem. 1979, 10, 19. Whitten, D. G. Acc. Chem. Res. 1980, 13, 83. (a) Bock, C. R.; Meyer, T. J.; Whitten, D. G. J. Am. Chem. Soc. 1974, 96, 4710. (b) Young, R. C.; Meyer, T. J.; Whitten, D. G. J. Am. Chem. Soc. 1975,97, 4781. Lin, C. T.; Bottcher, W.; Chou, M.; Creutz, C.; Sutin, N. J. Am. Chem. Soc. 1976, 98, 6536. (a) Kiwi, J.; Gratzel, M. Angew. Chem. Int. Ed. Engl. 1978, 17, 860. (b) Kiwi, J.; Gratzel, M. Angew. Chem. Int. Ed. Engl. 1979, 18, 624. (c) Gratzel, M. Acc. Chem. Res. 1981, 14, 376. Kiwi, J. J. Chem. Soc., Faraday Trans. 11, 1982, 78, 339. Sprintschnik, G.; Sprintschnik, H. W.; Kirsch, P. P.; Whitten, D. G. J. Am. Chem. Soc. 1976, 98, 2337. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 215 (a) Lehn, J. M.; Sauvage, J. P. Nouv. J. Chim. 1977, 1, 449. (b) Lehn, J. M. In Photochemical Conversion Storage of Solar Energy, Connolly, J. S. Ed.; Academic Press: New York, 1981, Chapter 6. Zameraev, K. I.; Parmon, V. N. In Energy Resources Through Photochemistry and Catalysis, Gratzel, M. Ed.; Academic Press: New York, 1983, Chapter 5. (a) Duonghong, D.; Borgarello, E.; Gratzel, M. J. Am. Chem. Soc. 1981, 103, 4685. (b) Borgarello, E.; Kiwi, J.; Pelizzetti, E.; Visca, M.; Gratzel, M. J. Am. Chem. Soc. 1981, 103, 6324. Gratzel, M.;'Kalyanasundaram, K.; Kiwi, J. Struct Bonding, 1982, 49, 37. Mclendon, G. In Energy Resources Through Photochemistry and Catalysis, Gratzel, M. Ed.; Academic Press: New York, 1983, Chapter 4. Kalyanasundaram, K. In Energy Resources Through Potochemistry and Catalysis, Gratzel, M. Ed.; Academic Press: New York, 1983, Chapter 7. Kalyanasundaram, K.; Gratzel, M.; Pelizzetti, E. Coord. Chem. Rev. 1986,69, 57. (a) Brunschwig, B. S.; Chou, M. H.; Creutz, C.; Ghosh, P. K. Sutin, N. J. Am. Chem. Soc. 1983,105, 4832. (b) Ghosh, P. K. Brunschwig, B. S.; Chou, M. H.; Creutz, C.; Sutin, N. J. Am. Chem. S0c.1984, 106, 4772. (c) Liu, D. K. ;Brunschwig, B. S.; Creutz, C.; Sutin, N. J. Am. Chem. Soc. 1986, 108, 1749. Shafirovich, V. Ya.; Khannanov, N. K.; Strelets, V. V. Nouv. J. Chim. 1980, 4 , 81. (a) Wrighton, M. S.; Ginley, D. S. J. Am. Chem. Soc. 1975, 97, 2065. (b) Wrighton, M. S. Top. Curr. Chem. 1976, 65, 37. Fox, A. J.; Poe, A.; Ruminski, R. J. Am. Chem. Soc. 1982, 104, 7327. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 216 Kobayashi, T.; Yasufuku, K.; Iwai, J.; Yesaka, H.; Noda, M.; 0htani, H. Coord. Chem. Rev. 1985, 64 ,1. Meyer, T. J.; Caspar, J. V. Chem. Rev. 1985, 85, 187. Geoffroy, G. L.; Wrighton, M. S. Oragnometallic Chemistry; Academic Press: New York, 1979. Wrighton, M. S.; Graff, J. L.; Luong, J. C.; Reichel, C. L.; Robbins, J. L. ACS, Symposium Ser. 1980, No. 155, 85. (a) Sperline, R.; Dickson, M. K.; Roundhill, D. M. J. Chem. Soc., Chem. Commun.1977, 62. (b) Filomena dos Remediopinto, M.A.; Sadler, P. J.; Neidle, S.; Sanderson, M. R.; Subbiah, A.; Kuroda, R. J. J. Chem. Soc., Chem. Commun. 1980, 13. (c) Alexander, K. A.; Bryan, S.A.; Dickson, M. K.; Hedden, D.; Roundhill, D. M. Inorg. Synth. 1986,24, 211. (a) Coleman, A. W.; Eadie, D. T.; Stobart, S. R.; Zaworotko, M. J.; Atwood, J. L. J. Am. Chem. Soc. 1982, 104, 922. (b) Beveridge, K. A.; Bushnell, G. W.; Stobart, S. R.; Atwood, J. L.; Zaworotko, M. J. Organometallics 1983, 2, 1447. (a) Mann, K. R.; Gordon, J. G., 11; Gray, H. B. J. Am. Chem. Soc. 1975, 97, 3553. (b) Lewis, N. S.; Mann, K. R.; Gordon, J. G., 11; Gray, H. B. J. Am. Chem. Soc.1976, 98, 7461. Che, C-M.; Butler, L. G.; Gray, H. B. J. Am. Chem. Soc. 1981, 103, 7796. (a) Rice, S. F.; Gray, H. B. J. Am. Chem. Soc. 1983, 105, 4571. (b) Rice, S. F.; Gray, H. B. J. Am. Chem. Soc. 1981, 103, 1593. (c) Rice, S. F.; Milder, S. J.; Gray, H. B.; Goldbeck, R. A.; Kliger, D. A. Coord. Chem. Rev. 1982,43, 349. Marshall, J. L.; Stobart, S. R.; Gray, H. B. J. Am. Chem. Soc. 1984, 106, 3027. Stein, P.; Dickson, M. K.; Roundhill, D. M. J. Am. Chem. Soc. 1983, 105, 3489. Che, C-M.; Butler, L. G.; Gray, H. B.; Crooks, R. M.; Woodruff, W. H. J. Am. Chem. Soc. 1983, 105, 5492. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 217 Dallinger, R. P.; Miskowski, V. M.; Gray, H. B.; Woodruff, W. H. J. Am. Chem. Soc. 1981, 103, 1595. (a) Roundhill, D. M. J. Am. Chem. Soc. 1985, 107, 4354. (b) Roundhill, D. M.; Atherton, S. J.; Shen, Z-P. J. Am. Chem. Soc. 1987, 109, 6076. Caspar, J. V.; Gray, H. B. J. Am. Chem. Soc. 1984, 106, 3029. (a) Mann, K. R.; Lewis, N. S.; Miskowski, V. M.; Erwin, D. K.; Hammond, G. S.; Gray, H. B. J. Am. Chem. Soc. 1977, 99, 5525. (b) Miskowski, V. M.; Sigal, I. S.; Mann, K. R.; Gray, H. B.; Milder, S. J.; Hammond, G. S.; Ryason, P. R. J. Am. Chem. Soc. 1979, 101, 4383. (c) Sigal, I. S.; Gray, H. B. J. Am. Chem. Soc. 1981, 103, 2220. (a) Gray, H. B.; Mann, K. R.; Lewis, N. S.; Thich, J. A.; Richman, R. M. Adv. Chem. Ser. 1978,No. 168, 44. (b) Mann, K. R.; Gray, H. B. Adv. Chem. Ser. 1979,No. 173, 225. (c) Gray, H. B.; Miskowski, V. M.; Milder, S. J.; Smith, T. P.; Maverick, A. W.; Buhr, J. D.; Gladfelter, W. L.; Sigal, I. 8.; Mann, K. R. Fundamental Research in Homogeneous Catalysis; Tsutsui, M., Ed.; Plenum: New York, 1979; Vol. 3. (a) Cotton, F. A.; Walton, R. A. Multiple Bonds Between Metal Atoms; Wiley-Interscience: New York, 1982 and reference therein. (b) Cotton, F. A.; Walton, R. A. Structure and Bonding, 1985, 62, 1 and references therein. Mortola, A. P.; Moskowitz, J. W.; Rosch, N.; Cowman, C. D.; Gray, H. B. Chem. Phys. Lett. 1975, 32, 283. Norman, J. G.; Kolari, H. J. J. Am. Chem. Soc. 1975, 97, 33. Hay, P. J. J. Am. Chem. Soc. 1982, 104, 7007. Mathisen, K. B.; Wahlgren, U.; Pettersson, L. G. M. Chem. Phys. Lett. 1984, 104, 336. Stromberg, A.; Pettersson, L. G. M.; Wahlgren, U. Chem. Phys. Lett. 1985, 118, 389. Hall, M. B. Polyhedron, 1987, 6, 679. 62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 21 8 Ziegler, T.; Tschinke, V.; Becke, A. Polyhedron, 1987, 6, 685. Bursten, B. B.; Clark, D. L. Polyhedron, 1987, 6, 695. (a) Cowman, C. D.; Gray, H. B.; J. Am. Chem. Soc. 1973, 95, 8177. (b) Cowman, C. D.; Trogler, W. C.; Gray, H. B. Isr. J. Chem. 1977, 15 , 308. (d) Miskowski, V. M.; Goldbeck, R. A.; Kliger, D. S.; Gray, H. B. Inorg. Chem. 1979, 18, 86. Trogler, W. C.; Gray, H. B. Acc. Chem. Res. 1978, II, 232. (3) Cotton. F. A.; Martin, D. S.; Fanwick, P. B.; Peters, T. J .; Webb, T. R. J. Am. Chem. Soc. 1976, 98, 4681. (b) Fanwick, P. B.; Martin, D. S.; Cotton. F. A.; Webb, T. R. Inorg. Chem. 1977, 16, 2103. (c) Trogler, W. C.; Cowman, C. D.; Gray, H. B.; Cotton. F. A. J. Am. Chem. Soc. 1977, 99, 2993. Rice, S. R; Wilson, R. B.; Solomon, E. I. Inorg. Chem. 1980, 19, 3425. (a) Manning, M. C.; Trogler, W. C. Inorg. Chem. 1982,21, 2797-2800. (b) Manning, M. C.; Trogler, W. C. J. Am. Chem. Soc. 1983, 105, 5311. Martin, D. S.; Newman, R.A.; Fanwick, P. E. Inorg. Chem. 1982, 21, 3400. Fraser, 1. F.; Peacock, R. D. Chem. Phys. Lett. 1983, 98, 620. Clark, R. J. H.; Stead, M. J. Inorg. Chem. 1983, 22, 1214. Lichtenberger, D. L.; Blevins, C. H. J. Am. Chem. Soc. 1984, 106. 1636. Robbins, G. A.; Martin, D. S. Inorg. Chem. 1984, 23, 2086. (a) Martin, D. S.; Huang, H.-W.; Newman, R. A. Inorg. Chem. 1984,23, 699-701. (b) Huang, H.-W.; Martin, D. S. Inorg. Chem. 1985,24, 96. (a) Fanwick, P. E. Inorg. Chem. 1985,24,258-263. (b) Fanwick, P. B.; Bursten, B. B.; Kaufmann, G. B. Inorg. Chem. 1985, 24,1165. 76. 77. 78. 79. 80. 81. 82. 83. 84. 85. 86. 87. 219 Dallinger, R. F. J. Am. Chem. Soc. 1985, 107, 7202. (a) Morris, D. E.; Sattelberger, A. P.; Woodruff, W. H. J. Am. Chem. Soc. 1986, 108. 8270. (a) Zietlow, T. C.; Hopkins, M. D.; Gray, H. B. J. Solid State Chem. 1985,57, 112. (b) Hopkins, M. D.; Gray, H. B.; Miskowski, V. M. Polyhedron, 1987, 6,705. (c) Hopkins, M. D.; Schaefer, W. P.; Bronikowski, M. J.; Woodruff, W. H.; Miskowski, V. M.; Dallinger, R. F.; Gray, H. B. J. Am. Chem. Soc. 1987, 109, 408. ((1) Hopkins, M. D.; Miskowski, V. M.; Gray, H. B. J. Am. Chem. Soc. 1988,110, 1787. Hopkins, M. D.; Zietlow, T. C.; Miskowski, V. M.; Gray, H. B. J. Am. Chem. Soc. 1985, 107, 510. Miskowsi, V. M.; Goldbeck, R. A.; Kliger, D. S.; Gray, H. B. Inorg. Chem. 1979, 18, 86. Hopkins, M. D.; Gray, H. B. J. Am. Chem. Soc. 1984, 106, 2468. Winkler, J. R.; Nocera, D. G.; Netzel, T. L. J. Am. Chem. Soc. 1986, 108, 4451. Nocera, D. G.; Gray, H. B. J. Am. Chem. Soc. 1981, 103, 7349. (a) Zietlow, T. C.; Klendworth, D. D.; Nimry, T.; Salmon, D. J.; Walton, R. A. Inorg. Chem. 1981, 20, 947. (b) Tetrick, S. M.; Coombe, V. T.; Heath, G. A.; Stephenson, T. A.; Walton, R. A. Inorg. Chem. 1984,23, 4567. (c) Dunbar, K. R.; Walton, R. A. Inorg. Chem. 1985, 24, 5. Walton, R. A. Isr. J. Chem. 1985, 25, 196. (a) Conner, K. A.; Gennett, T.; Weaver, M. J.; Walton, R. A. J. Electroanal. Chem. 1985, 196, 69. (b) Root, D. R.; Blevins, C. H.; Lichtenberger, D. L.; Sattelberger, A. P.; Walton, R. A. J. Am. Chem. Soc. 1986, 108,953. (c) Anderson, L. B.; Bakir, M.; Walton, R. A. Polyhedron, 1987, 6, 1483. Cotton, F. A.; Dunbar, K. R.; Falvello, L. R.; Tomas, M.; Walton, R. A. J. Am. Chem. Soc. 1983, 105, 4950. 88. 89. 90. 91. 92. 93. 94. 95. 96. 97. 98. 99. 100. 101. 220 Cotton, F. A.; Powell, G. L. J. Am. Chem. Soc. 1984, 106. 3371. (a) Cotton, F. A.; Daniels, L. M.; Dunbar, K. R.; Falvello, L. R.; Tctrick, S. M.; Walton, R. A. J. Am. Chem. Soc. 1985, 107. 3524. (b) Cotton, F. A.; Dunbar, K. R.; Price, A. C.; Schwolzer, W.; Walton, R. A. J. Am. Chem. Soc. 1986, 108. 4843. (a) Bucknor, S.; Cotton, F. A.; Falvello, L. R.; Reid, A. H., Jr.; Schmnlbach, C. D. Inorg. Chem. 1987,26, 2954. (b) Canich, J. M.; Cotton, F. A.; Dunbar, K. R.; Falvello, L. R. Inorg. Chem. 1988, 27, 804. Geoffroy, G. L; Gray, H. B.; Hammond, G. S. J. Am. Chem. Soc. 1974, 96, 5565. Trogler, W. C.; Gray, H. B. Nouv. J. Chim. 1977, I, 475. Erwin, D. K.; Geoffroy, G. L; Gray, H. B.; Hammond, G. S.; Solomon, E. 1.; Trogler, W. L.; Zagars, A. A. J. Am. Chem. Soc. 1977, 99 , 3620. Trogler, W. L.; Erwin, D. K.; Geoffroy, G. L; Gray, H. B.; J. Am. Chem. Soc. 1978, 100 , 1160. (a) Cotton, F. A.; Kalbacher, B. J. Inorg. Chem. 1976, 15, 522. (b) Bino, A.; Cotton, F. A. J. Am. Chem. Soc. 1979, 101, 4150. Nocera, D. G.; Gray, H. B. Inorg. Chem. 1984, 23, 3686. (a) Stephenson, T. A.; Bannister, E.; Wilkinson, G. J. Chem. Soc. 1964, 4150. (b) Lawton, D.; Mason, R. J. Am. Chem. Soc. 1965, 87, 921 . (a) Brencic, J. V.; Cotton, F. A. Inorg. Chem. 1969, 8, 7. (b) Brencic, J. V.; Cotton, F. A. Inorg. Chem. 1970, 9, 351. Brencic, J. V.; Cotton, F. A. Inorg. Chem. 1970, 9, 346. San Filippo, J. Jr.; Sniadoch, H. J.; Grayson, R. L. Inorg. Chem. 1974, 13, 2121. Cotton, F. A.; Extine, M. W.; Felthouse, T. R.; Kolthammer, W. S.; Lay, D. G. J. Am. Chem. Soc. 1981, 103, 4040. 102. 103. 104. 105. 106. 107. 108. 109. 110. 111. 112. 113. 114. 115. 116. 221 San Filippo, J., Jr. Inorg. Chem. 1972, 11, 3140. Best, S. A.; Smith, T. J.; Walton, R. A. Inorg. Chem. 1978, 17, 99. Cotton, F. A.; Falvello, L. R.; Harwood, W. S.; Powell, G. L.; Walton, R. A. Inorg. Chem. 1986,25, 3949. Agaskar, P. A.; Cotton, F. A Inorg. Chem. 1984, 23, 3383. Cotton, F. A.; Powell, G. L. Inorg. Chem. 1983, 22, 1507. Taube, H.; Bowen, A. R. Inorg. Chem. 1974, I3, 2245. Bowen, A. R.; Taube, H. J. Am. Chem. Soc. 1971, 193, 3287. (a) Bino, A.; Cotton, F. A. Inorg. Chem. 1979, 18, 3562. (b) Bino, A.; Cotton, F. A. Angew. Chem. Int. Ed. Engl. 1979, 18, 462. (c) Bino, A.; Cotton, F. A. Inorg. Chem. 1981, 20, 623. Demas, N. J.; Crosby, G. A. J. Phys. Chem. 1971, 75 , 991. Winkler, J. R.; Netzel, T. L.; Creutz, C.; Sutin, N, J. Am. Chem. Soc. 1987, 109, 2381. Bard, A. J.; Faulkner, L. R. Electrochemical Methods. Fundamental and Applicationsl; Wiley-Interscience: New York, 1980. Gordon, A. J.; Ford, R. A. The Chemist's Companions A Handbook of Practical Data, Techniques, and References; Wiley-Interscience: New York, 1972. Calvert, J. G.; Pitts, J. N. Photochemistry; Wiley-Interscience: New York, 1966. Frenz, B. A. "The Enraf-Nonius CAD 4 SDP - A Real - time System for Concurent X—Ray Data Collection an Crystal Structure Determination," in Computing in Crystallography, Schenk, H.; Olthof-Hazelkamp, R.; canKonigsveld, H.; Bassi, G. C. Eds; Delft University Press: Delft, Holland, 1978. Zachariasen, W. H.; Acta Crystallogr., 1963, 16, 1139. __ if ‘ 117. 118. 119. 120. 121. 122. 123. 124. 125. 126. 127. 128. 129. 130. 222 Cromer, D. T.; Waber, J. T. "International Tables for X—Ray Crystallography", Vol. IV, The Kynoch Press, Birmingham, England, 1974, Table 2.28. Ibers, J. A.; Hamilton, W. C. Acta Crystallogr., 1964, I7, 781. Cromer, D. T. "International Tables for X—Ray Crystallography," Vol. IV, The Kynoch Press, Birmingham, England, 1974, Table 2.3.1. Cruickshank, D. W. J. Acta Crystallogr., 1949, 2, 154. Duschinky, F. Acta Phys. Chim. URSS, 1937, 7, 55. Abbott, E. H.; Bose, K. S.; Cotton, F. A.; Hall, W. T.; Sekutowski, J. C. Inorg. Chem. 1978, 17, 3240. Cotton, F. A.; Fanwick, P. E.; Fitch, J. W.; Glicksman, H. D.; Walton, R. A. J. Am. Chem. Soc. 1979, 101, 1752. (a) Campbell, F. L., 111; Cotton, F. A.; Powell, G. L. Inorg. Chem. 1985, 24, 177. (b) Campbell, F. L., 111; Cotton, F. A.; Powell, G. L. Inorg. Chem. 1985,24, 4384. Mulliken, R. S. J. Chem. Phys. 1939, 7, 20. Stiegman, A. E.; Miskowski, V. M. J. Am. Chem. Soc. 1988, 110, 4053. Tolman, C. A. Chem. Rev. 1977, 77, 313. Tolman, C. A.; Seidel, W.C.; Gosser, L. W. J. Am. Chem. Soc. 197 4, 96, 53. Hopkins, M. D.; Miskowski, V. M.; Gray, H. B. J. Am. Chem. Soc. 1986, 108. 959. Cotton, F. A.; Frenz, B. A.; Pedersen, B.; Webb, T. R. Inorg. Chem. 1975,14, 391. 131. 132. 133. 134. 135. 136. 137. 138. 139. 140. 141. 142. 143. 144. 223 (a) Cotton, F. A.; Pedersen, B.; Webb, T. R. Inorg. Chem. 1975, 14, 388. (b) Cotton, F. A.; Pedersen, E. J. Am. Chem. Soc. 1975, 97. 303. Fanwick, P. E.; Martin, D. S.; Webb, T. R.; Robbins, G. A.; Newman, R. A. Inorg. Chem. 1978, 17, 2723. (a) Clark, R. J. H.; Franks, M. L. J. Am. Chem. Soc. 1975, 97. 2691. (b) Clark, R. J. H.; Stead, M. J. Inorg. Chem. 1983, 22, 1214. (a) Angell, C. L.; Cotton, F. A.; Frenz, B. A.; Webb, T. R. J. Am. Chem. Soc., Chem. Commun. 1973, 399. (b) Reference 13, Chapter 8. Loewenschuss, A.; Shamir, J.; Ardon, M. Inorg. Chem. 1976, 15. 238. Baxendale, J. H. Rad. Res. Suppl. 1964, 4, 139. Bowen, A. R.; Taube, H. J. Am. Chem. Soc. 1971, 93, 3287. Sasaki, Y.; Skyes, A. G. Am. Chem. Soc., Chem. Commun. 1973, 767. Ardon, M.; Pemick, A. Inorg. Chem. 1974, 13, 2275. Richens, D. T.; Skyes, A. g. Inorg. Chem. 1982, 21, 418. Bino, A.; Ardon, M.; Maro, I.; Maftory, M.; Dori, Z. J. Am. Chem. Soc. 1976,98, 7093. Cotton, F. A.; Dori, Z.; Marler, D. 0.; Schwotzer, W. Inorg. Chem. 1984, 23, 4033. Bino, A.; Cotton, F. A.; Dori, Z. J. Am. Chem. Soc. 1978, 100, 5252. Murmann, R. K.; Shelton, M. E. J. Am. Chem. Soc. 1980, 102, 3984. 145. 146. 147. 148. 149. 150. 151. 152. 153. 224 (a) Ardon, M.; Pemick, A. Inorg. Chem. 1973, 12, 1484. (b) Sasaki, Y.; Skyes, A. G. Am. Chem. Soc., Dalton Trans. 1974, 1968 Richens, D. T.; Skyes, A. G. Comments Inorg. Chem. 1981, 1, 141. Anbar, M. Adv. Chem. Ser. 1965,No. 50, 55. (a) Norman, J. G., Jr.; Kolari, H. J. J. Am. Chem. Soc., Chem. Commun. 1974, 303. (b) Norman, J. G., Jr.; Kolari, H. J. J. Am. Chem. Soc. 1975, 97. 33. (c) Mathisen, K. B.; Wahlgren, U.; Pettersson, L. G. M.; Wahlgren, U. Chem. Phys. Lett. 1984, 104, 336. (d) Stromberg, A.; Pettersson, L. G. M.; Wahlgren, U. Chem. Phys. Lett. 1985, 118, 389. Mortola, A. P.; Moskowitz, J. W.; Rosch, N.; Cowman, C. D.; Gray, H. B. Chem. Phys. Lett. 1975,32, 283. Schmidt, W.; Taube, H. Inorg. Chem. 1963,2, 698. Ilan, Y. Inorg. Chem. 1985,24, 4223. (a) Bennett, M. J.; Brencic, J. V.; Cotton, F. A. Inorg. Chem. 1969, 8, 1060. (b) Cotton, F. A.; Kalbacher, B. J. Inorg. Chem. 1976, 15 , 522. (c) Bino, A.; Cotton, F. A. Angew. Chem. Int. Ed. Engl. 1979, 18, 332. Mocella, M. T.; Okamoto, M. S.; Barefield, E. K. Synth. React. Inorg. Met.-0rg. Chem, 1974,4, 69.