PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 5/08 K:IPrqIAoc&PresICIRC/Date0m.indd DIPOLAR ENERGY TRANSFER IN RHENIUM(I) POLYPYRIDYL- BASED DONOR-ACCEPTOR ASSEMBLIES By Troy Elvin Knight A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PI-HLOSOPHY Chemistry 2009 ABSTRACT DIPOLAR ENERGY TRANSFER IN RHENIUM(I) POLYPYRIDYL- BASED DONOR-ACCEPTOR ASSEMBLIES By Troy Elvin Knight Dipolar energy transfer is a through-space mechanism that occurs when a donor emission dipole non-radiatively couples to an absorptive dipole in the acceptor. In 1946, Theodor Forster published the theoretical framework for dipolar (Forster) energy transfer in solution. The classic l/R6 distance dependence of the dipolar energy transfer rate has been widely used as a “spectroscopic ruler” to determine the distance between fluorescently labeled donor and acceptor complexes. In addition to the 1/R6 distance dependence, the through-space orientation (K2) between the donor and acceptor transition moment dipoles also affects the observed rate of energy transfer, and plays a particular role in the energy transfer efficiency in covalently attached donor-acceptor systems that are unable to sample all possible orientations in solution. The majority of the research presented in this dissertation involves understanding and quantifying the dependence of 2 K on F firster energy transfer in covalently attached transition metal-based donor-acceptor assemblies. The donor-acceptor systems discussed throughout this work are comprised of ReI polypyridyl donor molecules covalently attached to various first-row transition metal complexes through an organic-based linker molecule, and are referred to as MRex (where M represents a first-row acceptor metal center and X = 2 or 3). The lowest energy excited-state of the ReI donor moieties involves a metal-to-ligand charge transfer transition (3MLCT). This 3MLCT state is located on an isolated portion of the ReI- based donor moiety, which simplified the location of the donor emission dipole and in turn aided in the determination of the through-space orientation between the donor emission and acceptor absorption transition moment dipoles in the MRex complexes. The acceptor moieties are all M(acac)3 complexes that possess charge-transfer (M = Fem) or ligand-field (M = Cu", Crm, Col", and Ni") acceptor states. Varying the nature of the acceptor transitions permitted fundamental investigations of the dependence of Forster energy transfer on through-space dipolar orientation and on the size of the transition moment dipoles involved in the energy transfer event. The synthesis and characterization, photophysical properties, and dipolar energy transfer behavior in the MRex complexes will be presented and discussed. Copyright by Troy Elvin Knight 2009 Dedicated to my two grannies: Uva Nell Handley and Lillie Belle Knight. I hope I live up to the potential that both of you always saw in me. ACKNOWLEDGEMENTS To begin the acknowledgements I must first give a short back story on my road to the chemistry department at Michigan State. I began my graduate work at The University of Alabama at Birmingham where my research focused on the photophysical properties of transition metal compounds, which also happens to be the focus of the McCusker group. During my third year at UAB my research advisor was unfortunately denied tenure, and I was presented with some unexpected choices, either move to another group at UAB (which I did not want to do) or transfer to another university. My former advisor always spoke very highly of Jim and his science and suggested I look into joining the McCusker group. After contacting Jim, I decided to transfer to MSU and continue my focus on inorganic spectroscopy. 1 have to say it was one of the best choices I have ever made and I would do it all over again if given the chance. I have learned so much fiom Jim and I want to sincerely thank him for teaching me how to think and act as a scientist. Achieving the high scientific aptitude that Jim possesses will be daily goals of mine as I continue my career. Thanks Jim. “SCIENCE!” The second round of people I would like to thank is my family - Danny and Jeanie Knight and Treva Hadley. Dad, you are the hardest vi working person I have ever known and I want to thank you for being an excellent example for me through out these years. The constant drive you have to better yourself educationally and professionally has had a profound influence on me. I hope I can one day know as much about chemistry as you know about electricity. Mom, you are the best mother a person could ever ask for and I truly thank you for being there for me throughout these years. You always gave me your best everyday while growing up and I am truly gratefiJl for that. I know I also inherited my aptitude for math and science from you, which I am very thankful for as well. Treva, thanks for being my sister and putting up with me all of these years! I have watched you grow into a great person, a caring social worker, and an excellent mother to Drew. Next, I want to give a huge thanks to my fiancé Dr. Nicole Torres. Nicki, we have survived graduate school and writing our PhD theses together, which if we can do those things means we are really meant for each other! Over the past 3+ years I have watched you grow into a very skilled organic chemist and an excellent scientist, and I am so very proud of you and your accomplishments. I also love the fact we can bounce scientific ideas off each other at night after a long trying day in the lab. I know our postdoctoral training together at UNC will be filled with many discoveries and great times. Lastly, you are the most caring partner a person could hope vii for and I very thankful to have found you. I love you more than you will ever know. Lastly, I would like to thank the McCusker group for all of the great times we have spent together in the synthetic lab, laser lab, group meetings, and hanging out at the bar. All of you are doing great science and I know you will all have very productive scientific careers. I look forward to seeing everyone at various meetings and conferences. A few of you have contributed positively to my graduate career in various capacities and deserve particular mention: Dr. D. G. (Dong Guo), Dr. Monica Soler, Uncle Rick Fehir, U. J. (Joel Schrauben), Johnny 0. C. Kouzelos, and Kate McCusker. Graduate school would have been much different without all of you. Thanks! viii TABLE OF CONTENTS LIST OF TABLES ........................................................................................ xii LIST OF FIGURES ....................................................................................... xv Chapter 1. Introduction to Forster Energy Transfer Theory and Transition Metal-Based Donor-Acceptor Systems ........................................ 1 1.1 Introduction to Forster Energy Transfer Theory ............................... 1 1.2 Transition Metal-Based Donor/Acceptor Complexes ....................... 9 1.3 Contents of Dissertation .................................................................. 14 1.4 References ....................................................................................... 21 Chapter 2. Quantifying Forster Energy Transfer in Charge Transfer Based Donor-Acceptor Complexes: FeRe3 Multinuclear Assemblies .. 26 2.1 Introduction ..................................................................................... 26 2.2 Experimental Section ...................................................................... 32 2.2.1 Synthesis and Characterization ............................................ 32 2.2.2 Physical Measurements ....................................................... 38 2.3 Results and Discussion .................................................................... 46 2.3.1 Synthesis and Characterization ............................................ 46 2.3.2 Single-Crystal X-ray Structures .......................................... 48 2.3.3 Electrochemistry .................................................................. 53 2.3.4 Electronic Absorption Spectroscopy ................................... 55 2.3.5 Steady-State and Time-Resolved Emission ........................ 59 2.4 Mechanistic Considerations ............................................................ 65 2.4.1 Electron versus Energy Transfer Quenching ....................... 65 2.4.2 Dexter versus Forster Energy Transfer ............................... 68 2.4.3 Quantifying Forster Energy Transfer .................................. 71 2.4.4 Rate Constant Calculations: Modeling Solution Phase Energy Transfer Dynamics .................................................. 82 2.5 Geometry Optimization Calculation ............................................... 89 2.6 Conclusions ..................................................................................... 99 2.7 References and Notes .................................................................... 101 Chapter 3. Orbital-Specific Energy Transfer: CuRez Assemblies .............. 109 3.1 Introduction ................................................................................... 109 3.2 Experimental Section .................................................................... 115 3.2.1 Synthesis and Characterization ......................................... 115 3.2.2 Physical Measurements ..................................................... 122 ix 3.3 Results and Discussion .................................................................. 127 3.3.1 Synthesis and Characterization ......................................... 127 3.3.2 Single-Crystal X-ray Structures ........................................ 131 3.3.3 Electronic Absorption Spectroscopy ................................. 137 3.3.4 Steady-State and Time-Resolved Emission ...................... 141 3 .4 Mechanistic Considerations .......................................................... 151 3.4.1 Electron versus Energy Transfer Quenching .................... 151 3.4.2 Spectral Overlap Analysis: State-Selective Energy Transfer ............................................................................. 1 56 3.5 Conclusions ................................................................................... 171 3.6 References and Notes .................................................................... 174 Chapter 4. Confirming F orster Energy Transfer Involving Ligand-Field Acceptor States: CrRe3 Complexes ........................................ 184 4.1 Introduction ................................................................................... 184 4.2 Experimental Section .................................................................... 188 4.2.1 Synthesis and Characterization .......................................... 188 4.2.2 Physical Measurements ..................................................... 193 4.3 Results and Discussion .................................................................. 196 4.3.1 Synthesis and Characterization .......................................... 196 4.3.2 Electronic Absorption Spectroscopy ................................. 198 4.3.3 Steady-State and Time-Resolved Emission ...................... 202 4.4 Mechanistic Considerations .......................................................... 208 4.4.1 Electron versus Energy Transfer Quenching ..................... 208 4.4.2 Dexter versus F brster Energy Transfer ............................. 214 4.4.3 Comparison with the CuRez Assemblies ........................... 216 4.5 Conclusions ................................................................................... 218 4.6 References and Notes .................................................................... 220 Chapter 5. Spin Dependent Energy Transfer: CoRe3 and NiRe2 Complexes ................................................................................. 223 5.1 Introduction ................................................................................... 223 5.2 Experimental Section .................................................................... 229 5.2.1 Synthesis and Characterization .......................................... 229 5.2.2 Physical Measurements ..................................................... 231 5.3 Results and Discussion for [Co(pyacac)3(Re(bpy)(CO)3)3]- (OTf)3 ............................................................................................ 234 5.3.1 Synthesis and Characterization .......................................... 234 5.3.2 Electronic Absorption Spectroscopy ................................. 235 5.3.3 Time-Resolved Emission ................................................... 238 5.3.4 Spin-Dependent Dipolar Energy Transfer ......................... 239 5.4 Results and Discussion for [Ni(pyacac)2(THF)2(Re(bpy)(CO)3)2]- (OTf); ............................................................................................ 244 5.4.1 Synthesis and Characterization .......................................... 244 5.4.2 Single-Crystal X-ray Structures ........................................ 245 5.4.3 Electronic Absorption Spectroscopy ................................. 249 5.4.4 Steady-State and Time-Resolved Emission ...................... 251 5.4.5 Mechanistic Considerations: Dexter versus F 6rster .......... 256 5.4.6 Spin Dependent Dipolar Energy Transfer ......................... 259 5.5 Conclusions ................................................................................... 265 5.5 References ..................................................................................... 267 Chapter 6.‘Future Work ............................................................................... 270 6.1 Overall Goals ................................................................................. 270 6.2 Goal 1: Monitoring 3MLCT Relaxation in Re(bpy,)(CO)3 ........... 270 6.2.1 Background ........................................................................ 270 6.2.2 Dicyano Derivatives of CuRez and BeRez ..................... 272 6.2.3 Synthesis and Characterization .......................................... 274 6.2.4 Electronic Absorption Spectroscopy ................................. 274 6.2.5 Time-Resolved Emission ................................................... 276 6.2.6 Proposed TRIR Experiments ............................................. 277 6.3 Goal 2: Additional Derivatives for the CoRe3 Series ................... 279 6.3.1 Background ........................................................................ 279 6.3.2 Proposed CoRe3 Derivatives ............................................. 281 6.4 References ..................................................................................... 284 xi Table 2-1. Table 2-2. Table 2-3. Table 2-4. Table 2-5a. Table 2-5b. Table 2-5c. Table 2-6. Table 2-7a. Table 2-7b. LIST OF TABLES Crystallographic Data for [F e(pyacac)3(Re(bpy)(CO)3)3]- (OH), (2) and [A1(Pyaca0)3(Re(bPY)(C0)3)3](OTfls (5)50 Selected Bond Distances (A) and Angles (deg) for [Fe- (pyacac)3(Re(bpy)(CO)3)3](OTD3 (2) and [A1(pyacac)3(Re- (bpy)(CO)3)3](OTf)3 (5) ...................................................... 51 Electrochemical and Infrared Data for Complexes 1-6 ...... 55 Photophysical Data of Complexes 1-6 ................................ 65 RDA (A), e)T (°), (9D (°), oA (°), and K2 values determined from the internal dimensions of complex 2 using donor/acceptor interactions A:D, B:D, and C:D ................. 78 RDA (A), eT (°), @130), (9A (°), and K2 values determined from the internal dimensions of complex 2 using donor/acceptor interactions AzE, BE, and C:E .................. 79 RDA (A), OT (°), OD (°), 6)», (°), and K2 values determined from the internal dimensions of complex 2 using donor/acceptor interactions A:F, B:F, and C:F ................... 80 Calculated Forster rate constants for [Fe(pyacac)3(Re(tmb)- (CO)3)3](OTf)3 (1): [Fe(pyacac)3(Re(bpy)(CO)3)3](OTf)3 (2), and [Fe(pyacac)3(Re(deeb)(CO)3)3](OTf)3 (3) .................... 84 Calculated Fbrster rate constants (s") for complexes 1-3 at all donor-acceptor interactions occurring at A:D, B:D, and C:D ...................................................................................... 85 Calculated Forster rate constants (3") for complexes 1-3 at all donor-acceptor interactions occurring at AzE, BE, and C:E ....................................................................................... 86 xii Table 2-7c. Table 2-8. Table 2-9a. Table 2-9b. Table 2-9c. Table 2-10. Table 2-1 1. Table 3-1. Table 3-2. Table 3—3. Table 3-4. Calculated Forster rate constants (3") for complexes 1-3 at all donor-acceptor interactions occurring at A:F, B:F, and C:F ....................................................................................... 87 Fen-Re distances and FeOz/acac dihedral angles for the optimized and single-crystal X-ray structure data for complex 2 ............................................................................ 92 RDA (A), OT (°), OD (°), (9A (°), and K2 values determined from the optimized geometry of complex 2 using donor/acceptor interactions A:D, B:D, and C:D ................. 93 RDA (A), aT (°), aD (°), (aA (°), and K2 values determined from the optimized geometry of complex 2 using donor/acceptor interactions AzE, B:E, and C:E .................. 94 RDA (A), OT (°), OD (°), @A (°), and K2 values determined from the optimized geometry of complex 2 using donor/acceptor interactions A:F, B:F, and C:F ................... 95 Calculated Fbrster rate constants for [Fe(pyacac)3(Re(bpy)- (CO)3)3](OTf)3 (2) from geometry optimization calculations .......................................................................... 97 Calculated Forster rate constants (3") for the optimized structure of complex 2 at all donor-acceptor interactions... 98 Crystallographic Data for [Cu(pyacac)2(Re(bpy)(CO)3)2]- (OTflz (3) and [BeCPyacaC)2(Re(tmb)(C0)3)2l(0T02 (6) ...................................................................................... 134 Selected Bond Distances (A) and Angles (deg) for [Cu- (pyacao)2(Re(bpy)(C0)3)2](0Tf)2 (3) and [Be(pyacaC)2(Re- (tmb)(CO)3)2](OTf)2 (6) .................................................... 135 Photophysical Data for [M(pyacac)2(Re(bpy’)(CO)3)2](OTf)2 (complexes 1-10) ............................................................... 150 Electrochemical Data and Calculated Electron Transfer Driving Forces (AGF‘T) for Complexes 1-10 ..................... 154 xiii Table 3-5. Table 3-6. Table 4-1. Table 4-2. Table 5-1. Table 5-2. Calculated Spectral Overlap Integrals and Energy Transfer Rate Constants for Complexes 1-5 .................................... 161 Calculated Orbital Composition of the Lowest-Energy Spin Allowed Absorptions of Cu(phacac)2 in CHzClz solution .............................................................................. 165 Photophysical Data of Complexes 1-6 .............................. 208 Calculated Overlap Integral Values (J) and Energy Transfer Rate Constants for the CrRe3 and CuRez Complexes ....... 218 Crystallographic Data for [Ni(pyacac)2(THF)2(Re(bpy)— (CO)3)2](OTf)2 ................................................................... 247 Selected Bond Distances (A) and Angles (deg) for [Ni(pyacac)2(THF)2(Re(bpy)(CO)3)2](OTf); .................... 248 xiv Figure 1-1. Figure 1-2. Figure 1-3. Figure 1-4. Figure 1-5. Figure 1-6. Figure 1-7. Figure 2-1. Figure 2-2. LIST OF FIGURES Simplified energy level diagram illustrating resonance between the donor (D) emission and the acceptor (A) absorption transition dipoles ................................................. 3 Spectral overlap plot showing energy conservation between an arbitrary donor emission and acceptor absorption spectra .................................................................................... 5 Definitions of the angles used for calculating the orientation factor (K2) between the donor emission (m) and acceptor absorption (rA) transition moment dipoles ............................ 7 Range of values possible for the orientation factor (K2) between the donor emission (blue vectors) and the acceptor absorption (red vectors) transition moment dipoles .............. 8 Illustration of spherically-shaped donor (D) and acceptor (A) complexes that possess an isotropic through-space interaction between their respective transition moment dipoles ................................................................................... 9 Simplified energy level diagram for 2nd and 31rd row (16 polypyridyl complexes ........................................................ 11 General structure of the FeRe3 assemblies discussed in Chapter 2. The FeRe3 donor-acceptor complex shows the thermalized 3MLCT excited state of the ReI donor moiety utilized in all of the donor-acceptor complexes discussed in this dissertation .................................................................... 13 Polypyridyl derivatives coordinated to the ReI metal centers .................................................................................. 32 Drawing of the cations of [F e(pyacac)3(Re(bpy)(CO)3)3]- (0T03 (2, mp) and [A1(PyacaC)3(Re(bPY)(C0)3)3](0T03 (5, bottom) obtained from single-crystal X-ray structure determinations ..................................................................... 52 XV Figure 2-3. Figure 2-4. Figure 2-5. Figure 2-6. Figure 2-7. Figure 2-8. Figure 2-9. Electronic absorption spectra of [M(pyacac)3(Re(bpy’)- (CO)3)3](OTf)3 assemblies, where M = FeIII (black traces) or A1111 (blue traces) acquired in CH2C12 solution .................... 58 Electronic absorption spectrum of Fe(phacac)3 acquired in - CH2C12 solution at 298 K .................................................... 59 Corrected steady-state emission spectra for [M(pyacac)3(Re- (bpy’)(CO)3)3](OTf)3 assemblies in deoxygenated room- temperature CHzClz solution, where M = A1111 (blue traces) and FeIII (red traces) ............................................................ 61 Left. Nanosecond time-resolved emission data for the AlRe3 model complexes: 4 (tobs= 2260 i 100 ns), 5 (tobs= 560 i 30 ns), and 6 (robs = 235 i 20 ns). Right. TCSPC data for the corresponding FeRe3 complexes: 1 (robs = 450 i 30 ps), 2 (1:01,S = 755 i 40 ps), and 3 (rob, = 2.5 i 0.1 ns). Both spectra were measured in deoxygenated room-temperature CH2C12 solution ................................................................................ 64 T 0p. Nanosecond time-resolved differential absorption spectrum of [Al(pyacac)3(Re(tmb)(CO)3)3](OTf)3 (4) in room temperature CH2C12 solution (Apump = 355 nm). Bottom. Time-resolved absorption data for [Fe(pyacac)3(Re(tmb)- (CO)3)3](OTf)3 (1) in room-temperature CHzClz solution at Apmbe = 700 nm following ~100 fs excitation at Apump = 400 nm ........................................................................................ 68 Overlay of the emission spectra of [Al(pyacac)3(Re(tmb)- (C0)3)3](0Tf)3 (4, blue), [Al(pyacac)3(Re(bpy)(CO)3)3]- (0T03 (5, green), and [Al-(PyacaC)3(Re(deeb)(C0)3)3](OTfls (6, red) with the electronic absorption spectrum of F e(phacac)3 .......................................................................... 70 Through-space interactions of the Rel-based 3MLCT emission dipoles (blue) with the Fem-based 6LMCT absorption dipoles (red) ....................................................... 75 xvi Figure 2-10. Figure 2—1 1. Figure 2-12. Figure 2-13. Figure 3-1. Figure 3-2. Figure 3-3. Figure 3-4. Figure 3-5. Figure 3-6. Single Rel-based emission dipole moment spatially interacting with the three 6LMCT absorption dipole moments .............................................................................. 76 Point-dipole approximation used for calculating the through- space distances (RDA) and orientation factors (K2) needed for quantifying F orster energy transfer theory in the FeRe3 assemblies ............................................................................ 77 Energy level diagram depicting the excited-state dynamics of the FeRe3 assemblies ....................................................... 89 Drawing of the geometry optimized structure of [Fe(pyacac)3(Re(bpy)(CO)3)3](OTf)3 (2) ............................ 91 Structures of the CuRez and BeRez multinuclear assemblies........- .................................................................. 114 Drawing of the cation [(Be3(OH)3(pyacac)3)((Re(bpy)- (CO)3)3)](OTf)3 obtained from single-crystal X-ray structure determinations ................................................................... 13 1 Drawings of [Cu(pyacac)2(Re(bpy)(CO)3)2](OTf)2 (3, Top) and [Be(pyacac)2(Re(tmb)(CO)3)2](OTf)2 (6, Bottom) obtained from single-crystal X-ray structure determinations ................................................................... 1 36 Electronic absorption spectra of [M(pyacac)2(Re(bpy’)- (CO)3)2](OTf)2 assemblies in CH2C12 solution, where M = CuII (red traces) and Be" (blue traces) .............................. 139 Left. Electronic absorption spectrum of Cu(phacac)2 in showing the higher energy charge transfer and organic-based transitions. Right. Electronic absorption spectrum of a concentrated solution of Cu(phacac)2 showing the two mid- visible ligand-field bands. Both spectra were acquired in CH2C12 solution ................................................................. 141 Systematic titration of a CHzClz solution of Cu(phacac)2 with (NEt4)(CF3SO3) (2 - 40 eq.) ...................................... 141 xvii Figure 3—7. Figure 3-8. Figure 3-9. Figure 3-10. Figure 3-11. Corrected steady-state emission spectra for [M(pyacac)2(Re- (bpy’)(CO)3)2](OTf)2 where, M = cu" (red traces) and Be" (blue traces). All spectra were acquired in room-temperature deoxygenated CH2C12 solution .......................................... 144 Left. Nanosecond time-resolved emission data for the BeRez model complexes in room-temperature deoxygenated CHzClz solutions: 6 (robS = 1990 :t 100 ns), 7 (tot,s = 645 i 30 ns), 8 (robs: 540 i 30 ns), 9 (robS = 110 2!: 10 ns), and 10 (rob, = 250 :i: 20 ns). The data were acquired by monitoring at the emission maximum of each compound. Right. Time correlated single-photon counting (TCSPC) emission data for the CuRez complexes in room-temperature deoxygenated CHzClz solutions: 1 (1'1: 14.9 :1: 0.7), 2 (121: 8.1 i 0.4), 3 (n = 8.2 i 0.4), 4 (1:1= 5.6 i 0.3), and 5 (111: 5.0 d: 0.3) ........ 148 Illustration of oxidative (eq. 2a) and reductive (eq. 2b) electron transfer (kET) quenching processes out of the Rel- based 3MLCT excited-state to produce Cu1 or Cum as photo- products, respectively ........................................................ 153 Overlay of the emission spectra of [Be(pyacac)2(Re(tmb)- (C0)3)2](0T02 (6, 131111316), [B6(PyacaC)2(Re(dmb)(C0)3)2]- (0T02 (7, blue), [Be(pyacaC)2(Re(bpy)(C0)3)2](0T02 (8, green), [Be(pyacac)2(Re(dc1b)(CO)3)2](OTf); (9, orange), and [Be(pyacac)2(Re(deeb)(CO)3)2](OTf); (10, red) with the electronic absorption spectrum of Cu(phacac)2 (black trace) .................................................................................. 157 A. The ground state absorption spectrum of Cu(phacac)2 in CHzClz (black) fit with a series of three Gaussians (G1 (solid trace), G2 (dotted trace), and G3 (dashed trace)). B and C. Overlay of the emission spectra of [Be(pyacac)2(Re(tmb)- (C0)3)2](0T02 (6, Purple), [Be(Py3030)2(Re(dmb)(CO)3)2]- (0T02 (7, blue), [Be(PyacaC)2(Re(bpy)(C0)3)2](0T02 (8, green), [Be(pyacac)2(Re(dclb)(C0)3)2](OTflz (9, orange), and [Be(pyacac)2(Re(deeb)(CO)3)2](OTi); (10, red) with G2 and G3, respectively .......................................................... 163 xviii Figure 3-12. Figure 3-13. Figure 4-1. Figure 4-2. Figure 4-3. Figure 4-4. Figure 4-5. Figure 4-6. The four lowest energy transitions of Cu(phacac)2 determined from a TD-DFT calculation plotted against the ground state absorption spectrum of Cu(phacac)2 in CHzClz (black) ................................................................................ 166 Drawings of the orbitals involved in the lowest-energy spin allowed transitions of Cu(phacac)2 based on a TD-DFT calculation ......................................................................... 1 68 One electron-orbital description of the 4A2 —> 4T2 electronic transition in Crm. This diagram depicts the electronic configurations from which the multielectronic wavefimction corresponding to the 4A2 and 4T2 are derived .................... 186 Structure of the [M(pyacacl)3(Re(bpy’ )(CO)3)3](OTi)3 assemblies (where M= Cr or A!III and bpy’-— — —tmb, bpy, or deeb) .................................................................................. 188 Electronic absorption spectra of [M(pyacac)3(Re(bpy’)- (CO)3)3](OTf)3 assemblies, where M = CrIII (red trace) and (blue trace). All spectra were acquired in room- temperature CH2C12 solution ............................................. 200 Electronic absorption spectrum of Cr(phacac)3 acquired in room-temperature CHzClz solution ................................... 201 Corrected steady-state emission spectra for [M(pyacac)3(Re- (bpy’)(CO)3)3](OTf)3 assemblies, where M = A1111 (blue traces) and CrIII (red traces). All six spectra were acquired in deoxygenated room-temperature CH2C12 solutions following excitation at 355 nm (complexes 1, 2, 4, and 5) and 400 nm (complexes 3 and 6) .......................................................... 204 Corrected steady-state emission spectra for [Cr(pyacac)3(Re- (bpy)(CO)3)3](OTf)3 (2) monitored over the course of 4 hrs in room-temperature CHzClz solution. The various emission traces are represented as follows: the initial spectrum (black trace), after 1 hr (red trace), after 2 hrs (blue trace), after 3 hrs (green trace), and after 4 hrs (orange trace) ................ 205 xix Figure 4-7. Figure 4-8. Figure 4-9. Figure 4-10. Figure 4-1 1. Figure 5—1. Figure 5-2. Figure 5-3. Time correlated single-photon counting (TCSPC) data for the CrRe3 complexes: 1 (tot,S = 6.8 i 0.1 ns), 2 (robS = 4.8 d: 0.2 ns), and 3 (t'= 7.0 a: 0.3 ns) and r2 = 190 a 15 ns). All data were collected in deoxygenated room-temperature CHzClz solution ................................................................. 207 Steady-state emission spectrum of Cr(phacac)3 in 4:1 EtOH/MeOH at 82 K following excitation at 375 nm ...... 211 The black trace is the corrected steady-state emission spectra of [Cr(PyacaC)3(Re(bPY)(CO)3)3](0Tf)3 (2) in 411 EtOH/MeOH at 82 K following excitation at 375 nm. The red trace corresponds to the intensity of the emission spectrum scaled by the relative absorbances of the Rel-bpy and Crm(pyacac)3 chromophores ....................................... 212 Ground-state absorption spectrum of [Cr(pyacac)3(Re(bpy)- (CO)3)3](OTt)3 (2) in 4:1 EtOH/MeOH solution at 298 K (black trace), and the excitation spectrum of 2 in 4:1 EtOH/MeOH at 82 K monitoring at item = 795 nm (red trace) .................................................................................. 213 Overlay of the emission spectra of [Al(pyacac)3(Re(tmb)- (C0)3)3](0Tf)3 (4, blue), [Al(pyaca0)3(Re(bpy)(C0)3)3l- (OTD3 (5, green), and [Al(pyacac)3(Re(deeb)(CO)3)3](OTi)3 (6, red) with the electronic absorption spectrum of Cr(phacac)3. Data were acquired in room-temperature CHzClz solution ................................................................. 216 Illustration of total spin angular momentum conservation in the energy transfer quenching of a 3MLCT excited-state by 02 ....................................................................................... 225 Structure of the CoRe3 assembly ....................................... 227 Structure of the NiRe2 assembly ........................................ 228 XX Figure 5-4. Figure 5-5. Figure 5-6. Figure 5-7. Figure 5-8. Figure 5-9. Figure 5-10. Electronic absorption spectra of [Co(pyacac)3(Re(bpy)- (CO)3)3](OTf)3 (red trace) and [Al(pyacac)3(Re(bpy)- (CO)3)3](OTf)3 (blue trace) acquired in room-temperature CHzClz solution ................................................................. 236 Top. Electronic absorption spectrum of Co(phacac)3 showing the higher energy charge transfer and organic-based transitions. Bottom. Electronic absorption spectrum of a concentrated solution of Co(phacac)3 showing the mid- visible 1A] —> 1T1 ligand-field transition. Both spectra were acquired in room-temperature CH2C12 solution ................ 237 Left. Nanosecond time-resolved emission data for [Al- (pyacac)3(Re(bpy)(CO)3)3](OTf)3 (robS = 530 i 20 ns). Right. Nanosecond time-resolved emission data for [Co(pyacac)3- (Re(bpy)(CO)3)3](OTf)3 (tot,S = 595 :i: 20 ns). Both traces (A1 and Co) were acquired in deoxygenated room-temperature CHzClz solution monitoring at the emission maximum (565 nm). The solid red lines correspond to fits to mono- exponential decay models for both complexes ................. 239 Overlay of the emission spectrum of [Al(pyacac)3(Re(bpy)- (CO)3)3](OTf)3 (blue trace) with the electronic absorption spectrum of Co(phacac)3 (red trace) ................................. 241 Top. Spin conservation diagram illustrating the spin-allowed energy transfer pathway between the 3MLCT excited-state of the ReI donor and the CrIII acceptor. Bottom. Spin conservation diagram illustrating the spin-forbidden pathway between the 3l\fl_.CT excited-state of the ReI donor and the CoIII acceptor ........................................................ 243 Drawing of the cation [Ni(pyacac)3(THF)2(Re(bpy)(CO)3)2]- '(OTt); obtained from single-crystal X-ray structure determination ..................................................................... 249 Electronic absorption spectra of [Ni(pyacac)2(THF)2(Re- (bpy)(CO)3)2](OTf)2 (red trace) and [Be(pyacac)2(Re(bpy)- (CO)3)2](OTf)2 (blue trace) acquired in room-temperature CH3CN solution ................................................................. 250 xxi Figure 5-11. Figure 5-12. Figure 5-13. Figure 5-14. Figure 5-15. Figure 5-16. Figure 6-1. Electronic absorption spectra of Ni(phacac)2 acquired in room-temperature CH2C12 solution ................................... 251 Corrected steady-state emission spectra for [Ni(pyacac)2- (THF)2(Re(bpy)(CO)3)2](OTf); (red trace) and [Be(pyacac)2- (Re(bpy)(CO)3)2](OTf)2 (blue trace). Acquired in deoxygenated room-temperature CHZCIZ solution following excitation at 355 nm .......................................................... 253 A. Corrected emission spectra of [Ni(pyacac)2(TI-IF)2(Re- (bpy)(CO)3)2](OTt)2 in CH2C12 solution over a period of 27 hours (Aer = 355 nm) exhibiting a steady increase in emission intensity. B. Corrected emission spectra of [Ni(pyacac)2(TI-IF)2(Re(bpy)(CO)3)2](OTf)2 in THF solution over a 5 hour time period (lax = 355 nm) showing a constant emission intensity over the course of data collection ....... 254 Left. Nanosecond time-resolved emission data for [Be- (pyacac)2(Re(bpy)(CO)3)2](OTt)2 (robs = 530 i 30 ns). Right. TCSPC data for [Ni(pyacac)2(TI-IF)2(Re(bpy)(CO)3)2](OTf)2 (121: 5.0 :l: 0.4 ns and 12 = 25 :i: 3 ns). Both traces (Be and Ni) were acquired in deoxygenated room-temperature CH2C12 solution monitoring at the emission maximum (565 nm) ......................................................... 256 Overlay of the emission spectrum of [Be(pyacac)2(Re(bpy)- (CO)3)2](OTf)2 (blue trace) with the electronic absorption spectrum of Ni(phacac)2 (red trace) .................................. 259 Top. Spin conservation diagram illustrating the spin-allowed energy transfer pathway between the 3MLCT excited-state of the ReI donor and the pseudo-octahedral NiO6 acceptor. Bottom. Spin conservation dia am illustrating the spin- forbidden pathway between the MLCT excited-state of the ReI donor and the pseudo-D2,l NiO4 acceptor .................... 262 Structures of [Cu(pyacac)2(Re(dcnb)(CO)3)2](OTf); and [Be(pyacac)2(Re(dcnb)(CO)3)2](OTf)2 .............................. 273 xxii Figure 6-2. Figure 6-3. Figure 6-4. Figure 6-5. Figure 6-6. Electronic absorption spectra of [Cu(pyacac)2(Re(dcnb)- (CO)3)2](OTf)2 (red trace) and [Be(pyacac)2(Re(dcnb)- (CO)3)2](OTf)2 (blue trace) acquired in room-temperature CH3CN solution ................................................................. 275 Left. Nanosecond time-resolved emission data for [Be- (pyacac)2(Re(dcnb)(CO)3)2](OTf)2 (robs = 110 d: 10 ns) acquired in room-temperature deoxygenated CH2C12 solution (Apump = 400 nm; )‘Iprobe = 625 nm). Right. Time correlated single-photon counting (TCSPC) emission data for [Cu(pyacac)2(Re(dcnb)(CO)3)2](OTi)2 acquired in room- temperature deoxygenated CH2C12 solution (11mm, = 400 nm; limb, = 625 nm) ................................................................. 277 Overlay of the emission spectra of [Al(pyacac)3(Re(tmb)- (C0)3)3I(OTf)3 (blue), [A1(pyacac)3(Re(bpy)(CO)3)3](OTf)3 (green), and [Al(pyacac)3(Re(deeb)(CO)3)3](OTf); (red) with the electronic absorption spectrum of Cr(phacac)3. Plot is reproduced from Figure 4-12 ............................................ 280 Left. Steady-state emission spectrum of [Al(pyacac)3(Re- (dclb)(CO)3)3](OTf)3 in room-temperature CHzClz solution (rpm, = 375 nm); Am = 595 nm for 3MLCT —> 1A1 emission. Right. Nanosecond time-resolved emission data for [Al(pyacac)3(Re(dclb)(CO)3)3](OTt)3 (tob,= 120 :t 10 ns) acquired in deoxygenated room-temperature CHzClz solution (Apump = 415 nm; Apmbe = 595 nm) ........................ 282 Overlay of the emission spectra of [Al(pyacac)3(Re(bpy)- (C0)3)3](0Tf)3 (blue), [A1(PyacaC)3(Re(dclb)(C0)3)3](0T03 (green), and [Al(pyacac)3(Re(deeb)(CO)3)3](OTf)3 (red) with the electronic absorption spectrum of Co(phacac)3 .......... 283 Images in this dissertation are presented in color. xxiii Chapter 1: Introduction of Fiirster Energy Transfer Theory and Transition Metal-Based Donor/Acceptor Systems. 1.1 Introduction to Fiirster Energy Transfer Theory The initial descriptions of through-space energy transfer phenomena were introduced in the early 1900’s by the Nobel prize winning physicist Jean Baptiste Perrin. He was the first to theorize that energy between two closely spaced molecules in solution can be transferred by the interaction of their oscillating electric dipoles, as long as a resonance condition was satisfied between the two dipoles.1 This seminal work of Perrin ultimately led to the development of the well known fluorescence resonance energy transfer theory, or more commonly referred to as Fbrster theory. Ever since Theodor Fdrster first introduced his seminal paper in 1948,2 F firster energy transfer theory has become one of the most ubiquitous photophysical processes and has been used in biology, chemistry, and physics. The UK6 distance dependence between two interacting dipoles has been widely utilized as a spectroscopic ruler3'7 to determine the distance between donor and acceptor molecules, and has been shown to correspond very closely with distances determined from structural data.8 Examples of various areas impacted by Forster theory include measurements of distances between fluorescently labeled donor and acceptor molecules attached to proteins,9 determination of the conformational dynamics of RNA,10 and improving the light-harvesting capabilities of solar energy conversion devices.”15 In general, excited-state energy transfer processes involve the transfer of excited-state energy from a donor molecule (D) to an acceptor molecule (A), with the process beginning with the donor molecule absorbing a photon, and can be represented by equation 1-1: D+A+hv——>D*+A "W >1D+A* (1-1) Forster energy transfer is a through-space mechanism that occurs when a donor transition moment dipole non-radiatively couples to an acceptor transition moment dipole.”l7 The dipolar nature of this interaction gives rise to a R'6 (R = distance) dependence, allowing this mechanism to be operative over very long donor-acceptor separations. More specifically, Fbrster energy transfer involves the non-radiative coupling of donor emission dipoles with energetically matched acceptor absorption dipoles and is represented in Figure 1-1. A Dipole-Dipole Energy Transfer D l A V M Figure 1-1. Simplified energy level diagram illustrating resonance between the donor (D) emission and the acceptor (A) absorption transition dipoles. It can be seen from Figure 1-1 that the initially produced electronic excited states of the donor complex (Di) non-radiatively relax to the lowest energy vibronic level of D*, which can radiatively couple to various ground state vibrational levels of the donor (D) (blue, green, and red transitions). These emissive dipoles can then interact with energetically matched absorptive transitions of the acceptor molecule to produce acceptor excited states (A‘), which may be thought of as a virtual photon exchange between D’ and A. The transitions initiated in the acceptor molecule are equivalent to the acceptor absorbing a photon directly to generate A*. The picture of the two-state coupling shown in Figure 1-1 is a minimal model and represents the interaction of molecules in the gas phase. Donor emission and acceptor absorption transitions in solution interact with the surrounding medium and are usually observed as broad featureless spectra. Forster’s ingenious contribution to dipole-dipole energy transfer theory was his realization that coupled transitions in solution can be quantified by the spectral overlap of the donor emission spectrum with the acceptor absorption spectrum, which are both easily accessible experimental observables.18 The area of spectral overlap, which represents resonance matching of donor emission and acceptor absorption transitions, is shown qualitatively in Figure 1-2 (black lines). Forster’s spectral overlap model depends on small electronic communication between the donor and acceptor molecules (the so-called “weak coupling” regime),19 which is usually experimentally observed by the linear combination of the electronic absorption spectra of separated donor and acceptor molecules corresponding to the absorption spectrum of the intact donor-acceptor assembly. Specifically, the definition of the “weak coupling” regime given by F firster relies on two characteristics: 1) equilibration of the solvent bath with D“ occurs on a much faster time-scale than the excited-state energy transfer dynamics, and 2) the interaction of the donor and acceptor molecules with the solvent is much greater than the electronic coupling between them. Donor Emrssron uondrosqv Jotdooov Figure 1-2. Spectral overlap plot showing energy conservation between an arbitrary donor emission and acceptor absorption spectra. As mentioned above, Forster theory has a direct connection to parameters that can be accessed experimentally. The energy transfer rate constant (km) is described by the Forster equation given in equation 1-2,20 _ 90001n(10)K2(I)DJ EnT 1287T5774NATDR6 (1'2) where K2 is the dipole orientation factor, (DD is the radiative quantum yield of the donor, 1] is the refractive index of the solvent, NA is Avogadro’s number, ID is the excited-state lifetime of the donor, R is the donor-acceptor separation, and J the spectral overlap integral. This latter term, which essentially quantifies the resonance condition necessary for dipole-dipole coupling, can be evaluated from the spectroscopic properties of the system according to equation 1-3, Ho] D(V)8A(V)— adv (13) C where Fe is the (normalized) emission spectrum of the donor and EA is the absorption profile of the acceptor in units of molar absorptivity. From equation 1-2, knowledge concerning the geometric positions of the donor emission and acceptor absorption transition dipoles, which is known as the orientation factor (K2), must also be elucidated to accurately model excited- state kinetics. The orientation factor relates the angle between the donor emission and the acceptor absorption transition dipoles relative to each other and the angle the donor and acceptor transition dipoles each create with the donor-acceptor connection line, and is given in equation 1-4: = (cosG')T — 3cos®Dcos® A )2 (1-4) where @1- is the angle between the donor transition dipole moment (rD) and the acceptor transition dipole moment (rA), and OD and @A are the angles between the donor-acceptor connection line (RDA) and the donor transition dipole moment and the acceptor transition dipole moment, respectively. The angles are defined in Figure 1-3. GD ,5 (,9 ®Tf (N DA” / RDA / YD YA Figure 1-3. Definitions of the angles used for calculating the orientation factor (K2) between the donor emission (rD) and acceptor absorption (rA) transition moment dipoles. The values for K2 can range from 0 to 4 depending on the relative angles between the two transition dipoles, with K2 = 0 being the orthogonal interaction between the two dipoles, K2 = 1 being the parallel interaction, and K2 = 4 being the highest possible interaction with both dipoles aligned along the donor-acceptor connection line (Figure 1-4).21 K2=0 hoe...- ‘3» K _ 1 000.00 K2=4 {3.1—coco.“ Figure 1-4. Range of values possible for the orientation factor (1(2) between the donor emission (blue vectors) and the acceptor absorption (red vectors) transition moment dipoles. An isotropic value has also been determined for the dipolar orientation factor (K2 = %), and represents the average of all possible dipole-dipole through-space interactions between donor and acceptor species in solution.22 The isotropic value of 2/3 can be thought of as the donor emission and the acceptor transition moment dipoles existing within two spherically shaped molecules in solution (Figure 1-5), with these two systems possessing unrestricted movement within all three degrees of freedom (x, y, and z). The isotropic value works well for bimolecular donor-acceptor systems in solution such as lanthanide ions,23 or systems that behave like spheres that rapidly diffuse through all possible orientations,24 but this simplification can introduce a significant amount of error between the theoretical and observed energy transfer rate constants with covalently attached donor-acceptor systems that are unable to sample all degrees of fieedom.”26 Covalently attached donor-acceptor complexes, such as the Rel-based compounds discussed throughout this dissertation, negates the possibility of freely rotating donor and acceptor molecules and thus rules out the possibility of using K2 values of 2/3. Theoretically modeling and quantifying the orientation factor in covalently attached donor-acceptor systems is one of the overall themes of the work presented in this dissertation. Figure 1-5. Illustration of spherically-shaped donor (D) and acceptor (A) complexes that possess an isotropic through-space interaction between their respective transition moment dipoles. 1.2 Transition Metal-Based Donor/Acceptor Complexes Developing artificial energy-conversion systems is arguably the most important goal in the area of donor-acceptor photochemistry. Mother Nature has incorporated transition metal ions into various areas of the energy conversion sequences of photosynthesis,”30 which has led many researchers to follow in her footsteps and develop transition metal-based artificial photosynthetic systems.”32 Two major design lessons that can be taken from natural systems are 1) the initial light absorption step must produce a charge separation in order to generate high energy reagents that perform useful chemistry, such as excited state energy transfer or an electron transfer cascade and 2) spatial control of the excited-state energy flow must be achieved in order to deliver the useful energy to certain areas of the supramolecular array.33 Metal-based donor-acceptor assemblies consisting of III - have received d6 polypyridyl complexes of Rel, Ru", Os", and Ir considerable attention in this area due to possessing long-lived (nanosecond) charge-separated excited states in the form of metal-to-ligand charge transfer transitions (MLCT), relative stability of those charge transfer excited states, and well documented ground and excited-state electronic properties.”42 The ground and excited state electronic structure of d6 polypyridyl complexes deserves some focus in the section, due to the charge transfer excited states of ReI polypyridyl complexes utilized as the donor electronic states throughout this entire dissertation. Figure 1-6 shows a simplified energy level diagram for 2nd and 3rd row (16 polypyridyl complexes.43 The 10 absorption of a photon initially produces a 1MLCT excited state (t2g —-> at (polypyridyl)), which then undergoes extremely fast relaxation (~ 100 fs) through intersystem crossing and Vibrational cooling dynamics to the lowest energy 3MLCT excited state.44’45 It is out of this thermalized 3MLCT state that all of the radiative and energy transfer dynamics occur.46'52 The lifetime of the 3MLCT state in these complexes exist in the nanosecond to microsecond time regime,53'58 which is a timescale very amenable to performing useful excited-state chemistry processes and also to probe with typical photophysical instrumentation. ll 1MLCT 3MLCT Nuclear coordinate (Q) Figure 1-6. Simplified energy level diagram for 2nd and 3rd row (16 polypyridyl complexes. 11 One example of the Rel-based supramolecular assemblies (FeRe3) discussed in this dissertation and possessing a charge transfer excited state in the donor moieties is shown in Figure 1-7. Forster-type energy transfer reactivity can be envisioned to occur out of the thermalized 3MLCT state of the Re1 donor complex to produce an absorptive transition in the acceptor, and in the FeRe3 example the acceptor states are LMCT (ligand-to-metal charge transfer) transitions associated with the F e(acac)3 core (acac = acetylacetonate) (Figure 1-7). The charge transfer nature of the 3MLCT donor emission and the LMCT acceptor absorptions in the FeRe3 complexes induce the movement of charge within the individual donor and acceptor moieties, and these electronic transitions have been modeled as the blue (donor) and red (acceptor) vectors in Figure 1-7. The assignment of a specific excited state quenching mechanism depends strongly on the redox properties of the metal and the ligands, the strength of electronic coupling mediated by the bridging ligand (ie. whether the excited electron resides on the bridging or peripheral ligands), and the energy of the lowest energy 3MLCT state of the donor.59 The structural and electronic characteristics of the FeRe3 assemblies (Figure 1-7), along with the other Rel-based polynuclear assemblies reported throughout this work, are very amenable to Féirster energy transfer quenching of the ReI donor state. 12 o \\ ///° c\ \\ R mn--R§ Rel(bpy); 6LMCT: Fem(acac) —-> F e"(acac+)]. Metric details obtained from single- crystal X-ray structural data allowed for a quantitative application of F firster 15 energy transfer theory by systematically calculating the donor-acceptor separation (RDA) and spatial orientation of the charge transfer-based donor and acceptor transition moment dipoles (K2). The results of this analysis and agreement with experimentally derived energy transfer rate constants will be discussed. The unambiguous assignment of Fb’rster energy transfer as the dominant quenching pathway in the F eRe3 assemblies, led to the design of structurally similar complexes that possess acceptor metal centers with low- intensity ligand-field absorptions that overlap with the Rel-based donor emission spectra. The goal of the research contained in Chapter 3 was to examine if dipole-dipole energy transfer could operate in systems that possess very small transition dipole moments. Chapter 3 focuses on the synthesis, characterization, and photophysical properties of a family of donor-acceptor complexes with the general formula [Cu(pyacac)2(Re(bpy’)- (CO)3)2](OTi)2 (where bpy’ = 4,4’-5,5’-tetramethyl-2,2’-bipyridine, 4,4’- dimethyl-2,2’-bipyridine, 2,2’-bipyridine, 4,4’-dichloro-2,2’-bipyridine, or 4,4’-diethylester-2,2’-bipyridine). The CuRez complexes contain a d9 CuII acceptor core coordinated through two 3-(4-pyridyl)-acety1acetone ligands to two Re(bpy’)(CO)3 donor moieties. Steady-state as well as time-resolved emission measurements acquired in room-temperature CH2C12 solutions 16 indicate quenching of the Rel-based 3MLCT excited state when compared to isostructural BeRez model compounds. Electron transfer was found to be significantly endothermic for all five CuRez complexes; this fact, coupled with the ca. 10 A donor-acceptor distance and favorable spectral overlap between the 3MLCT emission profile and ligand-field absorptions of the CuII center implicated dipolar energy transfer as the dominant quenching pathway in these compounds. Gaussian deconvolution of the ground-state absorption spectrum of Cu(phacac)2 (phacac = 3-phenyl-acetylacetonate) allowed for a differential analysis of the spectral overlap between the donor emission spectra with the two observed ligand-field absorption bands of the Cu" ion. The rate of energy transfer was found to increase with increasing overlap for coupling to the lower-energy ligand-field band, consistent with expectations fi'om Ft'irster theory. These results were supported by time- dependent DFT calculations on Cu(phacac)2, and indicated preferential coupling to a particular ligand-field transition of the Cu" center. These results will be discussed in terms of an orbitally-specific energy transfer process occurring in the CuRez series. Observation of emission fi'om an acceptor moiety provides unequivocal proof of an excited state energy transfer mechanism. Although, the structural and electrochemical properties of the FeRe3 (Chapter 2) and 17 CuRez (Chapter 3) complexes rule out electron transfer as a viable quenching mechanism, the goal of the work contained in Chapter 4 was to introduce an acceptor metal center that emits in response to quenching the Rel-based 3MLCT excited state. Chapter 4 focuses on the synthesis, characterization, and photophysical properties of a series of chromophore- quencher complexes with the general formula [Cr(pyacac)3(Re(bpy’)- (CO)3)3](OTf)3 (with bpy’ defined above). The CrRe3 complexes contain a d3 CrIII acceptor core coordinated through three 3-(4-pyridyl)-acetylacetone ligands to three Re(bpy’)(CO)3 donor moieties. Steady-state as well as time- resolved emission and absorption measurements acquired in room- temperature CH2C12 solutions indicate quenching of the Rel-based 3MLCT excited state when compared to isostructural AlRe3 model compounds. Acquisition of the excitation spectrum of [Cr(pyacac)3(Re(bpy)(CO)3)3]- (OTf)3 at km, of the Crm-based 2E —> 4A2 emission in fiozen EtOH/MeOH (4:1) solution reproduced it’s ground-state absorption spectrum, which proves that all of the photons absorbed by [Cr(pyacac)3(Re(bpy)- (CO)3)3](OTt)3 results in Crm-based emission and unequivocally proves an energy transfer mechanism. A comparison between the nature of the ligand- field acceptor states in the CrRe3 and the CuRez complexes, along with their effect on dipolar energy transfer reactivity will also discussed. 18 In Chapter 5, the Rel-based compounds - [Co(pyacac)3(Re(bpy)- (C0)3)3](0Tf)3 and [Ni(PyacaC)2(THF)2(Re(bPY)(CO)3)2](0T02 - were synthesized in order to investigate the spin-state dependence of dipolar energy transfer. The first section of Chapter 5 contains the synthesis, characterization, and photophysical properties of a CoRe3 assembly that possesses a low-spin (16 Col" metal center ('Al, S = 0). Steady-state and time-resolved emission measurements acquired in room-temperature CHzClz solution revealed no quenching of the Rel-based 3MLCT excited state, with the observed excited state relaxation behavior comparable to an isostructural AlRe3 compound. Favorable spectral overlap between the Rel-based 3MLCT emission spectrum and the 1A] —+ 'T, ligand-field absorption profile of the Co"I ion was observed, suggesting dipolar energy quenching should be active in the CoRe3 assembly. Determination of the total spin angular momentum (ST) values possible in the CoRe3 complex revealed that the 'A, —-> 1T1 acceptor transition represents a spin-forbidden relaxation pathway for the 3MLCT donor state. This suppressed energy transfer reactivity between the 3MLCT donor and the CoIII core is compared with the results received in Chapter 3 on a structurally similar CrRe3 complex the exhibits significant quenching of the Rel-based 3MLCT state. The CrRe3 system possesses identical overall charge and structural characteristics with the CoRe3 19 complex, but simply differs in the spin-state of the acceptor metal center. The results on the CoRe3 system highlight the ability of controlling excited state quenching reactivity by simply varying the spin-state of the reagents. The second section of Chapter 5 will focus on the synthesis, characterization, and photophysical characteristics of [Ni(pyacac)2(THF)2- (Re(bpy)(CO)3)2](OTf)2, referred to as the NiRez assembly, which contains a pseudo-octahedral NiII (d8) acceptor metal center. Determination of the total spin angular momentum (ST) in the NiRez assembly showed the ligand-field absorption of the Ni06 core (3A2 —> 3T2) to be a spin-allowed relaxation pathway for the Rel-based 3MLCT donor state. Comparison of the steady- state emission of [Ni(pyacac)2(THF)2(Re(bpy)(CO)3)2](OTf); in THF and in CHzClz solutions at room-temperature revealed an on/off 3MLCT quenching mechanism, with the dissociation of the axially coordinating THF molecules in CH2C12 solution generating a NiO6.x (x = 1 or 2) complex and subsequent spin-forbidden energy transfer pathway. Future directions of the research are included as a sixth chapter. Various MRen (n = 2 or 3) compounds that have been synthesized and characterized, proposed complexes and photophysical experiments, and the scientific questions the molecules will address comprise the content of Chapter 6. 20 1.4 References (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13) (14) Perrin, J. C. R. Acad. Sci 1927, 184, 1097. F firster, T. Annalen der Physik 1948, 2, 55. Bates, M.; Blosser, T. R.; Zhuang, X. W. Phys. Rev. Lett. 2005, 94, 108101. Khan, Y. R.; Dykstra, T. E.; Scholes, G. D. Chem. Phys. Lett. 2008, 461, 305. Lakowicz, J. R.; Johnson, M. L.; Wiczk, W.; Bhat, A.; Steiner, R. F. Chem. Phys. Lett. 1987, I38, 587. Munoz-Losa, A.; Curutchet, C.; Krueger, B. P.; Hartsell, L. R.; Mennucci, B. Biophys. J. 2009, 96, 4779. Sahoo, H.; Roccatano, D.; Hennig, A.; Nau, W. M. J. Am. Chem. Soc. 2007, 129, 9762. Lakowicz, J. R. Principles of Fluorescence Spectroscopy; 3rd ed.; Springer: New York, 2006. Stryer, L. Ann. Rev. Biochem. 1978, 47, 819. Millar, D. P.; Klostermeier, D. Methods 2001, 23, 240. Venturi, M.; Serroni, S.; Juris, A.; Campagna, S.; Balzani, V. Top. Curr. Chem. 1998, 197, 194. Wasielewski, M. R. Chem. Rev. 1992, 92, 435. Steed, J. W.; Atwood, J. L. Supramolecular Chemistry; John Wiley and Sons: Chichester, 2000. Echegoygan, L.; Kaifer, A. E., Eds. Physical Supramolecular Chemistry; Kluwer Academic Publishers: Dordrecht, 1996. 21 (15) (16) (17) (13) (19) (20) (21) (22) (23) (24) (25) (26) (27) (28) (29) Adronov, A.; Gilat, S. L.; Frechet, J. M. J .; Ohta, K.; Neuwahl, F. V. R.; Fleming, G. R. J. Am. Chem. Soc. 2000, 122, 1175 and references therein. Speiser, S. Chem. Rev. 1996, 96, 1953. Wong, K. F.; Bagchi, B.; Rossky, P. J. J. Phys. Chem. A. 2003, 108, 5752. Scholes, G. D. Ann. Rev. Phys. Chem. 2003, 54, 57. Kasha, M. Radiation Res. 1963, 20, 55. Van Der Meer, B. W.; Coker, G. 1.; Chen, S.-Y. Resonance Energy Transfer, Theory and Data; VCH Publishers: New York, 1994. DeMember, J. R.; Filipescu, N. J. Am. Chem. Soc. 1968, 90, 6425. Wang, X. F.; Herman, B., Eds. Fluorescence Imaging Spectroscopy and Microscopy; John Wiley: New York, 1996. Thomas, D. D.; Carlsen, W. F.; Stryer, L. Proc. Natl. Acad. Sci. U. S. A. 1978, 75, 5746. Dale, R. 13.; Eisinger, J.; 13., B. w. Biophys. J. 1979, 26, 161. Furue, M.; Ishibashi, M.; Satoh, A.; Oguni, T.; Maruyama, K.; Sumi, K.; Kamachi, M. Coord. Chem. Rev. 2000, 208, 103. Grosshenny, V.; Harriman, A.; Hissler, M.; Ziessel, R. J. Chem. Soc, Faraday Trans. 1996, 92, 2223. Mukhopadhyay, S.; Manda], S. K.; Bhaduri, S.; Armstrong, W. H. Chem. Rev. 2004, 104, 3981. Ferreira, K. N.; Iverson, T. M.; Maghlaoui, K.; Barber, J .; Iwata, S. Science 2004, 303, 1831. Stryer, L. Biochemistry; 4th ed.; W.H. Freeman and Co.: New York, 1995. 22 (30) (31) (32) (33) (34) (35) (36) (37) (38) (39) (40) (41) (42) (43) Hoganson, C. W.; Babcock, G. T. Science 1997, 277, 1953. Balzani, V.; Scandola, F. Supramolecular Photochemistry; Horwood: Chichester, U. K., 1991. Alstrum—Acevedo, J. H.; Brennaman, M. K.; Meyer, T. J. Inorg. Chem. 2005, 44, 6802. Fabbrizzi, L.; Poggi, A., Eds. Transition Metals in Supramolecular Chemistry; Klewar Academic Publishers: Dordrecht, 1994. Balzani, V.; Ceroni, P.; Juris, A.; Venturi, M.; Campagna, S.; Puntoriero, F.; Serroni, S. Coord. Chem. Rev. 2001, 219, 545. Baranoff, E.; Barigelletti, F.; Bonnet, S.; Collin, J. P.; Flamigni, L.; Mobian, P.; Sauvage, J. P. In Photofunctional Transition Metals Complexes; Springer-Verlag Berlin: Berlin, 2007; Vol. 123, p 41-78. Blan, Z. Y.; Sumi, K.; Furue, M.; Sato, S.; Koike, K.; Ishitani, O. Inorg. Chem. 2008, 4 7, 10801. Easun, T. L.'; Alsindi, W. Z.; Towrie, M.; Ronayne, K. L.; Sun, X. 2.; Ward, M. D.; George, M. W. Inorg. Chem. 2008, 4 7, 5071. Otsuki, J.; Imai, A.; Sato, K.; Li, D. M.; Hosoda, M.; Owa, M.; Akasaka, T.; Yoshikawa, I.; Araki, K.; Suenobu, T.; Fukuzumi, S. Chem-Eur. J. 2008, I4, 2709. Pelleteret, D.; Fletcher, N. C. Eur. J. Inorg. Chem. 2008, 3597. Walther, M. E.; Wenger, O. S. Dalton Trans. 2008, 6311. De Cola, L.; Belser, P. Coord. Chem. Rev. 1998, I 7 7, 301. Welter, S.; Lafolet, F.; Cecchetto, E.; Vergeer, F.; De Cola, L. ChemPhysChem 2005, 6, 2417. McCusker, J. K. Acc. Chem. Res. 2003, 36, 876. 23 (44) (45) (46) (47) (48) (49) (50) (51) (52) (53) (54) (55) (56) (57) (58) Liard, D. J .; Busby, M.; Matousek, P.; Towrie, M.; Vléek, A. J. Phys. Chem. A 2004, 108, 2363. Damrauer, N. H.; Cerullo, G.; Yeh, A.; Boussie, T. R.; Shank, C. V.; McCusker, J. K. Science 1997, 275, 54. Knight, T. E.; Guo, D.; Claude, J. P.; McCusker, J. K. Inorg. Chem. 2008, 4 7, 7249. Wrighton, M.; Morse, D. L. J. Am. Chem. Soc. 1974, 96, 998. MacQueen, D. B.; Eyler, J. R.; Schanze, K. S. J. Am. Chem. Soc. 1992, 114, 1897. Sun, S.-S.; Lees, A. J. J. Am. Chem. Soc. 2000, 122, 8956. Vléek, A.; Busby, M. Coord. Chem. Rev. 2006, 250, 1755. Kalyanasundaram, K. Photochemistry of Polypyridine and Porphyrin Complexes; Academic Press: London, 1992. Lees, A. J. Chem. Rev. 1987, 87, 711. Caspar, J. V.; Meyer, T. J. J. Phys. Chem. 1983, 87, 952. Claude, J. P.; Omberg, K. M.; Williams, D. S.; Meyer, T. J. J. Phys. Chem. A 2002, 106, 7795. Schoonover, J. R.; Strouse, G. F. Chem. Rev. 1998, 98, 1335. Dattelbaum, D. M.; Omberg, K. M.; Schoonover, J. R.; Meyer, T. J. Inorg. Chem. 2001, 41, 6071. Dattelbaum, D. M.; Martin, R. L.; Schoonover, J. R.; Meyer, T. J. J. Phys. Chem. A. 2004, 108, 3518. Worl, L. A.; Duesing, R.; Chen, P.; Della Ciana, L.; Meyer, T. J. J. Chem. Soc., Dalton Trans. 1991, 849. 24 (59) Brown, W. R.; O'Boyle, N. M.; McGarvey, J. J.; Vos, J. G. Chem. Soc. Rev. 2005, 34, 641. 25 Chapter 2. Quantifying Fiirster Energy Transfer in Charge Transfer Based Donor/Acceptor Complexes: FeRe3 Multinuclear Assemblies. 2.1 Introduction Elucidating the mechanism of excited-state reactivity is a necessary first step for understanding and ultimately manipulating complex photo- induced chemical processes."4 Accordingly, numerous fundamental studies of excited-state dynamics have been reported in the literature. Assemblies based on d6 polypyridyl complexes of Rel, Ru", and OsII have garnered particular attention due to relative stability of their excited states, well documented ground and excited-state electronic properties, and the ability one has to tune these properties through synthetic means. Both electron and energy transfer processes have been the subject of intense scrutiny. As a result, much has been learned about the factors that govern both of these types of excited-state reactions in transition metal-based systems.5'14 The two most widely occurring energy transfer mechanisms are electron superexchange (Dexter)15 and through-space dipole-dipole coupling (F brster).16 Dexter energy transfer is subject to a distance dependence that falls off as exp(-2R) due to its reliance on orbital overlap. As such, it is usually relegated to covalently linked systems in which the donor and 26 acceptor are in close proximity (e.g., _<_ 5 A)”22 Fbrster transfer is a through-space mechanism that occurs when the donor emission dipole non- radiatively couples to an absorptive dipole in the acceptor.”25 The dipolar nature of this interaction gives rise to a shallower R'6 dependence, allowing this mechanism to be operative over much longer distances. Féirster-type reactivity is therefore usually dominant in systems that place the lowest energy excited-state on an electronically isolated portion of the donor or 26-29 between metal centers that are separated over long distances. The F6rster energy transfer rate constant (km) is described by the equation given in equation 2-1,30 _ 90001n(10)rc2DJ EnT 12871'5774NA2'DR6 (2-1) where x2 is the dipole orientation factor, (1),, is the radiative quantum yield of the donor, 77 is the refractive index of the solvent, NA is Avogadro’s number, I'D is the excited-state lifetime of the donor, R is the donor-acceptor separation, and J is a spectral overlap integral that essentially quantifies the resonance condition necessary for dipole-dipole coupling. This latter term can be evaluated from the spectroscopic properties of the system according to equation 2-2, 27 Ho] D (V)8A (V)€1V - 0 (2-2) where Fo is the (normalized) emission spectrum of the donor and BA is the absorption profile of the acceptor in units of molar absorptivity. From equation 2-1, knowledge concerning the geometric positions of the donor emission and acceptor absorption transition dipoles, which is known as the orientation factor (K2), must be elucidated to accurately model excited-state kinetics. The orientation factor relates the angle between the donor emission and the acceptor absorption transition dipoles relative to each other and the angle the donor and acceptor transition dipoles each create with the donor- acceptor connection line, and is given in equation 2-3: 2 = (cosG')T — 3cos®Dcos® A) (2-3) where @T is the angle between the donor transition dipole moment and the acceptor transition dipole moment, and @D and @A are the angles between the donor-acceptor connection line and the donor transition dipole moment and the acceptor transition dipole moment, respectively. Although equations 2-1, 2-2, and 2-3 constitute a complete description of the rate of dipolar energy transfer, quantifying the donor-acceptor distance (R) and the orientation factor (K2) can be quite difficult in metal-based donor-acceptor 28 complexes, due to R typically being approximated as the metal-metal separation and K2 usually assigned as 273.3140 Taking R as the metal-metal distance is a reasonable assumption when structural (data are not available, but in the point-dipole approximation of Fbrster theory this may or may not accurately reflect the relevant distance in systems comprised of donor and acceptor states that are charge-transfer in nature. In addition, the orientation factor of K2 = 2/3 typically invoked represents an isotropic value for species sampling all possible angular distributions.“ While appropriate for bimolecular energy transfer processes, this approximation may not be reasonable given the rotational barriers that likely exist in covalently attached donor-acceptor complexes. The resulting ambiguities that can arise concerning these two variables often lead to a large variance between experiment and theory. Moore et a1. recently applied F firster theory in conjunction with molecular modeling calculations utilizing naphthalene and anthracene, which possess known directional emission and absorption transitions, covalently linked through a ZnII containing macrocycle that holds the separation distance constant.”43 The studies revealed that the minimized conformers, with quantitatively determined donor-acceptor distances and orientation factors, very accurately describe the observed Forster energy 29 transfer kinetics when all of the possible conformers were taken into account. The donor-acceptor systems to be discussed in this chapter possess known donor-acceptor separations and directional electronic transitions as well, which utilize a ReI polypyridyl based 3MLCT state as the donor and F em tris-acetylacetonate based 6LMCT (ligand-to-metal charge-transfer) states as the acceptor. The current systems also have moderate spectral overlap between the donor emission and acceptor absorption profiles, which was instrumental in systematically studying and solely modeling the donor emission quenching dynamics as Fdrster energy transfer. Specifically, the synthesis, structure, and photophysical properties of a series of isostructurally related molecules are reported, with the general formula [Fe(pyacaC)3(Re(tmb)(C0)3)3](0T03 (1), [Fe(pyacaC)3(Re(bPY)- (C0)3)3](0Tf)3 (2). and [F6(PyacaC)3(Re(deeb)(C0)3)3](0T03 (3) (Where pyacac = 3-(4-pyridyl)-acetylacetonate, tmb = 4,4’-5,5’-tetramethyl-2,2’- bipyridine, bpy = 2,2’-bipyridine, deeb = 4,4’-diethylester-2,2’-bipyridine, and OTf = CF 3SO3'), and with the entire series referred to as the FeRe3 assemblies. The Fe"I metal centers are covalently attached to three fac- Re(bpy’)(CO)3 (bpy’ = tmb, bpy, and deeb) (Figure 2-1) moieties through three pyridyl-acetylacetonate bridging ligands. It was observed that the Re]- based 3MLCT excited-states for complexes 1, 2, and 3 are significantly 30 CllleIICth in the presence of the Fe111 metal center relative to structurally analogous AlRe3 analogs. The AlRe3 model systems with the general formula [A1(PyacaC)3(Re(tmb)(C0)3)3](0T03 (4), [A1(pyacaC)3(Re(bPY)- (C0)3)3l(0T1)3 (5). and [A1(pyacaC)3(Re(deeb)(C0)3)3](0T03 (6) were synthesized in order to investigate the excited-state dynamics of the Re(bpy’)(CO)3 moieties in the absence of the emission quenching dynamics incurred by the F eIII metal center. Varying the bpy’ attached to the ReI metal center permitted a systematic investigation of the donor-acceptor spectral overlap by modifying the energy of the 3MLCT excited-state, with the emission quenching rate constant corresponding with the amount of observed spectral overlap. in addition, the single-crystal X-ray structure data for complex 2 enabled very accurate calculations of the RDA and K2 interactions within this system, with the same values applied to complexes 1 and 3 by assuming that variation in the polypyridyl substituents causes minor altering of the internal dimensions within this family of compounds. The rigid structure of these systems also enabled a very accurate modeling of the solution phase F orster energy transfer kinetics by assuming the RDA and K2 values remain unchanged from those calculated fiom the X—ray structural data, with the theoretical rate constants accurately reproducing the experimentally observed values. Lastly, one of the major hurdles in 31 quantifying Ftirster energy transfer is accurately determining the through- space separation and orientation factor relating donor and acceptor transition dipoles, and the work in this chapter will show that the unique structure of the FeRe3 analogs allow for a very accurate determination of these two variables. Figure 2-1. Polypyridyl derivatives coordinated to the ReI metal centers. 2.2 Experimental Section 2.2.1 Synthesis and Characterization General. All solvents used were purified and dried according to previously reported methods.44 Spectroscopic grade CH2C12 was used for all photophysical measurements; the solvent was dried under CaHz reflux until 32 no water was detected by 1H NMR and degassed using freeze-pump—thaw techniques. 3-(4--pyridyl)-2,4--pcntanedione,45 Al(pyacac)3,45 Re(tmb)(CO)3- (on),46 Re(bpy)(CO)3(OTt),46 Re(deeb)(CO)3(OTf),46 and fac- [Re(bpy)(CO)3(4-Etpy)](PF6)47 (4-Etpy = 4-ethylpyridine) were prepared following literature procedures. 3-phenyl-2,4-pentanedione was purchased from TCI America. Elemental analyses and F T-IR data were obtained through the analytical facilities at Michigan State University. Mass spectra were obtained through the analytical facilities at The University of South Carolina. Tris(3-(4-pyridyl)acetylacetonato)iron(III), Fe(pyacac)3. The synthesis of this compound has been reported previously by a different method.48 Amounts of 70.1 mg (0.432 mmol) of FeCl3 and 230 mg (1.30 mmol) of pyacac were dissolved in 30 mL of THF and stirred for 6 hrs. Sodium tert-butoxide (125 mg, 1.30 mmol) was then added to the reaction flask and the solution was stirred overnight. The reaction mixture was filtered over celite to remove excess salt and the solvent removed under vacuum. The product was recrystallized from CHzClz/hexanes (1:1 v/v). Yield: 131 mg (52%). Anal. Calcd for C30H30N306Fe: C, 61.65; H, 5.17; N, 7.19. Found: C, 61.49; H, 5.13; N, 7.07. 33 Tris(3-phenyl-acetylacetonato)iron(III), Fe(phacac)3. The synthesis of this compound has been previously reported by a different method.48 Amounts of 86 mg (0.53 mmol) of F eC13 and 277 mg (1.57 mmol) of phacac were dissolved in 30 mL of THF and stirred for 6 hrs. Sodium tert-butoxide (149 mg, 1.57 mmol) was then added to the reaction flask and the reaction mixture was stirred overnight. The solution was then filtered over celite to remove excess salt and the solvent removed under vacuum. The product was recrystallized from CH2C12/hexanes (1:1 v/v). Yield: 238 mg (78%). Anal. Calcd for C33H3306Fe: C, 68.16; H, 5.72. Found: C, 68.11; H, 5.79. [Fe(pyacac)3(Re(tmb)(CO)3)3](OTf)3 (1). An amount of 230 mg (0.360 mmol) of Re(tmb)(CO)3(OTf) was dissolved in 75 mL of hot THF, after which 70 mg (0.12 mmol) of F e(pyacac)3 was added and the solution purged with argon for 20 min. The reaction mixture was then fit with a condenser and stirred under argon for 3 days in hot THF in the dark, after which a dark red solution formed along with an orange precipitate. The precipitate was collected and the filtrate was concentrated under vacuum to yield additional orange solid. The combined precipitates were dissolved in CHzClz, filtered through celite, and the solvent removed under vacuum. The product was recrystallized several times from CHzClz/pentane (1:1 v/v). Yield: 155 mg (52%). Anal. Calcd for Cg4I-I73N9F9024S3FeRe3: C, 40.70; H, 34 3.17; N, 5.08. Found: C, 40.36; H, 3.23; N, 4.86. IR (KBr, cm'l): 2031 s, 1918 s, 1614 m, 1566 s, 1448 m, 1365 m, 1263 s, 1155 m, 1032 s, 638 m. MS: [ESL m/z (rel. int.)]: 677.3 (70) {[Fe(pyacac)3(Re(tmb)(CO)3)3]}3+, 1090.5 (23) {[Fe(pyacac)3(Re(tmb)(CO)3)3](OTf)}2+, 2330.1 (1) {[Fe(pyaeac)3(Re(tmb)(C0)3)31(OTf)2}”- [Fe(pyacac)3(Re(bpy)(CO)3)3](OTf)3 (2). Amounts of 70 mg (0.12 mmol) of Fe(pyacac)3 and 210 mg (0.365 mmol) of Re(bpy)(CO)3(OTf) were dissolved in 25 mL of THF and flushed with argon for 20 min. The reaction was stirred in the dark for 3 days at room temperature. An orange solid precipitated out of solution and was collected and washed with hexanes. The solid was then dissolved in CH3CN, filtered through celite, and the solvent removed under vacuum. The product was recrystallized several times from acetonitrile/ether (1:1 v/v). X-ray quality crystals were obtained by slow diffusion of ether into an acetonitrile solution of the compound. Yield: 166 mg (60%). Anal. Calcd for C72H54N9F9024S3FeRe3: C, 37.37; H, 2.35; N, 5.45. Found: C, 37.00; H, 2.23; N, 5.27. IR (KBr, cm" I): 2033 s, 1922 s, 1568 m, 1446 m, 1367 w, 1261 m, 1159 m, 1032 m, 771 m, 638 m. MS: [ESL m/z (rel. int.)]: 621 (100) 1[Fe(pyacac)a(Re(bpy)(CO)a)3113*. 1006 (18) {1Fe(pyacacb- (CO)3)3](OTD}2+9 2161-1 (1) {[F6(PyacaC)3(Re(bpy)(C0)3)3](0T02} ”- 35 [Fe(pyacac)3(Re(deeb)(CO)3)3](OTf); (3). Amounts of 29 mg (0.050 mmol) of Fe(pyacac)3 and 108 mg (0.150 mmol) of Re(deeb)(CO)3(OTf) were dissolved in 25 mL of THF and flushed with argon for 20 min. The ' reaction was stirred in the dark for 4 days at room temperature. The solution was then filtered over celite and the solvent removed under vacuum to give a red solid. The product was recrystallized several times using CH2C12/pentane (1:1 v/v). Yield: 65 mg (48%). Anal. Calcd for C90H73N9F9036S3FeRe3°3CH2C12: C, 38.00; H, 2.79; N, 4.15. Found: C, 38.00; H, 2.93; N, 4.32. IR (KBr, cm'l): 2036 s, 1923 s, 1732 s, 1566 s, 1462 m, 1263 s, 1153 m, 1032 s, 767 m, 638 m. MS: [ESL m/z (rel. int.)]: 765.4 (75) {[Fe(pyacac)3(Re(deeb)(CO)3)3]}3+, 1222.6 (19) {[Fe(pyacac)3- (Re(deeb)(CO)3)3](OTf)}2+, 2594.2 (1) {[Fe(pyacac)3(Re(deeb)(CO)3)3]- toms)”. [Al(pyacac)3(Re(tmb)(CO)3)3](0T1)3 (4). An amount of 224 mg (0.360 mmol) of Re(tmb)(CO)3(OTf) was dissolved in 75 mL of hot THF, after which 66 mg (0.120 mmol) of Al(pyacac)3 was added and the solution purged with argon for 20 min. The reaction mixture was then fit with a condenser and stirred under argon for 3 days in hot THF in the dark, after which a dark yellow solution formed along with a yellow precipitate. The precipitate was collected and the filtrate concentrated under vacuum to yield 36 additional yellow solid. The combined precipitates were dissolved in CHzClz, filtered through celite, and the solvent removed under vacuum. The product was recrystallized several times from CHzClz/pentane (1:1 v/v). Yield: 185 mg (64%). Anal. Calcd for C34H73N9F9024S3A1Re3°CH2C12: C, . 40.55; H, 3.17; N, 4.95. Found: C, 40.37; H, 3.12; N, 4.95. IR (KBr, cm'l): 2031 s, 1918 s, 1612 m, 1585 s, 1452 m, 1396 m, 1263 s, 1155 m, 1032 s, "638 m. MS: [ESI, m/z (rel. int.)]: 667.7 (65) {[AIQJyacac)3(Re(tmb)- (CO)3)3]}3+, 1076.1 (25) {[Al(pyacac)3(Re(tmb)(CO)3)3](OTf)}2+, 2301.3 (1) 1lAl 20(1)] u(Mo Koo/cml 4.342 4.278 Rim 0.0918 0.0899 R1a 0.0675 0.0666 wR2b 0.1373 0.1564 GOF 1.007 1.044 aRl = 2111:.)- chll/ZIFOL waz = [2w(F02 — F,2)2/2w(1=,2)2]“2, w )+ (aP)2 + bP], where P = [F02 + 2Fc2]/3, 1/[o 2(F, Table 2-2. Selected Bond Distances (A) and Angles (deg) for [Fe(pyacaC)3(Re(bpy)(C0)3)3](0T03 (2)21nd [A1(PyacaC)3(Re- (bPY)(CO)3)3l(OT1)3 (5). 2 5 Bond Distances (A) Fe(1)- 0(1) 1.992(6) Al(1)-O(1) 1.877(10) Fe(1) — 0(2) 1.970(7) Al(l) — 0(2) 1.861(10) Fe(l) — 0(6) 1.997(6) Al(1) — 0(6) 1.871(11) Fe(1) — 0(7) 1.963(6) Al(1)— 0(7) 1.894(11) Fe(l)—O(11) 1.981(7) Al(1)—O(11) 1.883(12) Fe(1) — 0(12) 1.986(6) A1(1)— 0(12) 1.852(10) Re(1)—N(1) 2.232(7) Re(1)—N(1) 2.215(12) Re(2) —N(4) 2.218(7) Re(2) -N(4) 2.225(12) Re(3) —N(7) 2.206(9) Re(3) —N(7) 2.191(14) Fe(l)mRe(1) 9.88 Al(1)mRe(1) 9.76 Fe(1)---Rc(2) 9.88 Al(1)---Re(2) 9.74 Fe(1)«-Re(3) 9.78 Al(1)---Re(3) 9.66 Bond Angles (deg) 0(1) — Fe(1) — 0(2) 85.4(3) 0(1) — Al(1) — 0(2) 90.0(4) C(23) - Re(1)— N(1) 92.1(4) C(23) — Re(1)— N(1) 93.0(6) C(21) — Re(1)— N(1) 178.7(4) C(21) — Re(1)— N(1) 177.9(7) aplane lu-plane 2 280.1 aplane lee-plane 2 79.0 bplane 100-plane 2 82.5 bplane 1°»plane 2 81.0 cplane 1°»plane 2 68.8 _ cplane louplane 2 69.2 aPlane 1 is defined by atoms 0(1), 0(2), C(1), C(2), C(3), C(4), C(5); plane 2 is defined by atoms N(1), C(6), C(7), C(8), C(9), C(10). bPlane 1 is defined by atoms 0(6), 0(7), C(24), C(25), C(26), C(27), C(28); plane 2 is defined by atoms N(4), C(29), C(30), C(31), C(32), C(33). cPlane 1 is defined by atoms O(11), 0(12), C(7), C(48), C(49), C(50), C(51); plane 2 is defined by atoms N(7), C(52), C(53), C(54), C(55), C(56). 51 t ‘ cs4 1 .'3 Rel ‘ C9 C10 C5] C5 7‘ '- ‘ * p . cr ' csz ‘ N7 S, ‘ N1 , ‘ C2 012 . 50 "9, C55 .1 O u. S 9.3 a . I a D O NO‘S ‘UI r— O N A .01 Figure 2-2. Drawing of the cations of [Fe(pyacac)3(Re(bpy)(CO)3)3]- (010. (2. top) and [A1(pyacaC)3(Re(bPY)(C0)3)3](0T03 (5, bottom) obtained from single-crystal X-ray structure determinations. Atoms are represented as 50% probability thermal ellipsoids. 52 Figure 2—2 (cont’d) '.“ ‘T 5‘ ‘3. _ Re3 — \ . . C5 7 ‘ 'e 4 C5 “C55 . 3" C9 C56 . 1 '4 .. a? . 10 C5 C7 052 s \ C6 4 C 2‘ a 49 5‘ l o . C3 '4 .1 2 ‘ 51 é.‘ C7 ‘ 50 3 C 5- 012 Cl‘ V‘ l] 020 07 C2 C27 C2 " '- C28 2.3.3 Electrochemistry. The electrochemical properties of complexes 1-6 were examined using cyclic voltammetry; the data are given in Table 2-3. The availability of the Al”1 model complexes greatly simplifies assigning the features observed for all six complexes due to the redox-inert nature of this ion. Accordingly, the single reduction waves seen for complexes 4-6 can be immediately ascribed to the bipyridyl ligand of the Re moiety in each case. 53 The positive shift in potential across the series is consistent with the more electron withdrawing nature of the substituents as one progresses from the methyls of tmb (4) to the diethylester groups in complex 6. Similarly, the oxidation waves seen for all three complexes are easily assigned to the Rel/Rel1 couple. The influence of the bipyridyl substitutents are apparent in these data as well, with the more electron deficient ligand giving rise to the most positive oxidation potential for the Re center. The results are all consistent with what has been observed for complexes of the general form fac-[Re(4,4'-X2-bpy)(CO)3(4-Etpy)](PF6) previously reported in the literature.46 Given these assignments, the reductions at ca. -0.9 V observed for "1 center.70 It can be seen complexes 1-3 are clearly associated with the Fe that modification of the bipyridyl ligand has no discernible influence on the redox properties of the central Fe ion. In a similar vein, the ligand reduction and ReI oxidation potentials of all three Fe-containing compounds are essentially identical to what was observed for the AlIII analogs. These data are indicative of (relatively) weak electronic coupling between the central metal ion and the peripheral chromophores. 54 Table 2-3. Electrochemical and Infrared Data for Complexes 1-6. electrochemical potential (V) v(CO) (cm-l)a complexes 13,x (Rel/u) E,,d(1=e“"“) 1:.red (bpy'o’) A'(1)A",A'(2) FeRe3 tmb(1) +1.47"~c .0.91‘Le -1.84b 1918 2031 bpy(2) +1 .41b -0904 -1594 1922 2033 deeb (3) +1 .60b -0.93d - 1 .20d 1923 2036 AlRe3 tmb (4) +1.37b -1.82" 1918 2031 bpy (5) +1.42b 4.57"!f 1920 2033 deeb (6) +1.54b 119" 1930 2036 aMeasured as pressed KBr pellets. 1:,Measured in CH3CN solution. °Potential in CHzClz solution is +1.41 V. dMeasured in CHzClz solution. ePotential in CH3CN solution is -0.86 V. fPotential in CH2C12 solution is -1.57 V. 2.3.4 Electronic Absorption Spectroscopy. The electronic absorption spectra of the FeRe3 (1-3) (black traces) and AlRe3 (4-6) (blue traces) analogs have been acquired in room-temperature CHzClz solutions and are shown in Figure 2-3. Transitions associated with these complexes were assigned based on previously reported analyses of ReI polypyridyl and F e(phacac)3 absorption spectra. ReI polypyridyl complexes exhibit a well known lMLCT (t2g ——> 71* (bpy’)) transition occurring from approximately 330 to 430 nm depending on the substituents of the bpy’ ligand,“72 with the 55 absorption profiles of the AlRe3 compounds solely exhibiting these lMLCT based characteristics. Amax for the 1MLCT excited-state reflects the electron donating/withdrawing ability of the bpy’ substituents, with Amax for complexes 4, 5, and 6 occurring at 344 (8 = 18,500), 364 (8 = 11,500), and 394 (a = 14,300) nm, respectively. As the substituents become progressively more electron donating (e.g., H for bpy and CH3 for tmb) this feature systematically shifts to the blue and begins to overlap with the ligand-based absorptions in the ultraviolet (Figure 2-3). The AlRe3 analogs allow for analysis of the ground-state absorption behavior of the Re1 donor moieties without overlapping with the Fem-based charge-transfer transitions contained in the FeRe3 analogs. The presence of Fe"I in complexes 1-3 gives rise to a new, broad absorption feature on the low-energy side of the Rel-based charge transfer band. Figure 2-4 shows the absorption spectrum of Fe(phacac)3 which exhibits two strong transitions centered at 370 and 460 nm assigned as 6A. —+ 6MLCT (t2g —-> 71*(acac)) and 6A1 —+ 6LMCT (7t (acac) —> tzg) transitions, respectively.73’74 In complexes 1-3 the higher energy MLCT absorption is obscured by the more intense 1A1 —> 1MLCT band of the ReI chromophore, but the low energy tail of the 6A, ——> 6LMCT transition can be seen extending out to approximately 600 nm. The linear combination of the corresponding 56 AlRe3 analog with Fe(phacac)3 is also shown in Figure 2-3 (red trace), which sufficiently models the absorptive characteristics of the FeRe3 analogs. The additive property of these constituent species reiterates the minimal ground- "I state electronic interaction between the ReI and Fe metal centers observed from the infrared data. 57 A W03 lvi Molar Absorpt Molar Absorptivity (8) Molar Absorptivity (8) 30000‘ 25000 20000- 15000- 10000- 5000- 0.. 25000; 20000 \ 15000r \ 10000- 5000~ o . 40000‘ 35000 30000- 25000r 20000- 15000- ‘ 10000- 5000- O '- L l 1 l 350 400 450 500 550 ‘ 600 Wavelength (nm) Figure 2-3. Electronic absorption spectra of [M(pyacac)3(Re(bpy’)- (CO)3)3](OT1)3 assemblies, where M = FeIII (black traces) or AllII (blue traces). All spectra were acquired in room-temperature CHzClz solution. A. [F e(pyacac)3(Re(tmb)(CO)3)3](OTf)3 (1) and [Al(pyacac)3- (Re(tmb)(C0)3)3l(0Tf)3 (4). 13- [Fe(pyacaC)3(Re(bPY)(C0)3)3](0T03 (2) and [A1(Pyacae)3(Re(bp>')(C0)3)3l(OTf)3 (5)- C- [Fe(pyacae)3(Re- (deeb)(C0)3)3l(0Tf)3 (3) and [A1(PyacaC)3(Re(deeb)(C0)3)3](0T1)3 (6)- The dashed red lines in all three graphs correspond to linear combinations of the molar absorptivity (8, (M'l, cm'1)) profiles of the AlRe3 complex with that of Fe(phacac)3. 58 4" 6000- 5000 4000- 3000- 2000- 1000 Molar Absorptivity (M 1cm 0 l 350 400 450 500 550 600 650 Wavelength (nm) Figure 2-4. Electronic absorption spectrum of Fe(phacac)3 acquired in room-temperature CHzClz solution. 2.3.5 Steady-State and Time-Resolved Emission. Emission spectra for the FeRe3 and AlRe3 complexes were obtained in room—temperature deoxygenated CH2C12. The AlRe3 complexes represent an ideal structural model for the dynamics associated with the Re1 based 3MLCT emission, due to the inability of engaging in quenching dynamics incurred by the F eIII core in the FeRe3 systems. The emission profiles are given in Figure 2-5, and show the emission intensity for the F eRe3 complexes are significantly attenuated compared to the AlRe3 model systems. The spectral profiles correspond to previously reported photophysical studies of ReI polypyridyl systems, with the emission originating from the 3MLCT —-> 1A1 phosphorescence.75 The emission maximum for the tmb (526 nm), bpy (566 nm), and deeb (624 nm) derivatives reflect the electron donating and 59 withdrawing behavior of the polypyridyl ligands. The radiative quantum yields ((1),) of 4 ((1),: 0.51), 5 ((1).: 0.16), and 6 (CI),= 0.07) were determined relative to [(bpy)Re(CO)3(4-Etpy)](PF6) ((1)r = 0.18 in CHzClz) (Table 2-4), and are comparable to the reported values for the corresponding mononuclear ReI polypyridyl derivatives.46 The (1), values for complexes 1-3 are analytically unreliable due to a small emissive impurity generated by dissociation of the complexes. The source of the dissociation was discovered to be residual amounts of water contained in the CHzClz, which generates the highly emissive fac- [Re(bpy’)(CO)3(pyacac)] complex. The drying and distilling procedures that were employed were exhausted until no water was detectable by ]H NMR, but despite our best efforts small traces remained in the solvent. It should also be noted that the extent of dissociation was small over the course of the radiative quantum yield determinations, but a steady increase in the emission signal for the three systems was observed when monitored over several hours. 60 1 A ‘3 Q) .5 78‘ O 5 3‘ E a E A 'l B "U Q) .5 i O 5 r? (I) C: a E. l C 6 O.) .5 "‘8 O 5 .2» 'a C 8 .E .500 600 700 800 Wavelength (nm) Figure 2-5. Corrected steady-state emission spectra for [M(pyacac)3- (Re(bpy’)(CO)3)3](OTt)3 assemblies, where M = AlIII (blue traces) and Fe‘" (rcd traces). A. [Al(pyacac)3Re(tmb)(CO)3])3](OT1)3 (4) and [Fe- (PyacaC)3(Re(tmb)(C0)3)3](0Tf)3 (1) 13- [A1(pyacaC)3Re(bP)/)- (C0)3])3l(0Tf)3 (5) and [Fe(pyacaC)3(Re(bPY)(C0)3)3](0T03 (2) C- [Al(pyacac)3(Re(deeb)(CO)3)3](OTf), (6) and [Fe(pyacac)3(Re(deeb)- (CO)3)3](OTf)3 (3). The spectra were acquired in CHZCIZ solutions at room temperature following excitation at 355 nm (complexes 1, 2, 4, and 5) and 400 nm (complexes 3 and 6). 61 Additional details concerning the excited states of the A1Re3 and FeRe3 complexes were obtained through nanosecond emission lifetime and time-correlated single photon counting (TCSPC) measurements. The nanosecond emission decay lifetime (robs) in room-temperature CHzClz solution for complexes 4, 5, and 6 are shown on the left side of Figure 2-6, and could be fit to single-exponential models with rob, = 2260 :l: 100, 560 :l: 30, and 235 :l: 20 ns (Table 2-4), respectively. The corresponding radiative and non-radiative decay rate constants are, k, = 2.3 i 0.1 x 105 s'1 and knr = 2.2 d: 0.1x105 s'1 for 4, k, = 2.9 i 0.13 x105 s'1 and k... = 1.5 a 0.07 x106 s'1 for 5, and k. = 3.0 :t 0.13 x105 s'1 and 1r... = 4.0 a 0.17 x106 s" for 6. As with the quantum yields, the observed excited-state lifetimes and rate constants are all consistent with the assignment of 3MLCT —> 1A1 emission . reported previously for the mononuclear ReI polypyridyl derivatives.46’76 The kinetics reveal that the reduction in quantum yield is due primarily to an increase in the non-radiative decay rate for 3MLCT relaxation (km) as opposed to significant variations in radiative coupling to the ground state (Table 2-4). The lack of emission signal observed for the FeRe3 systems during nanosecond lifetime experiments prompted the employment of time- correlated single photon counting (TCSPC) methods. Plots of the TCSPC 62 data obtained in room-temperature CH2C12 solution for the FeRe3 complexes are shown on the right side of Figure 2-6. All three decay traces could be fit with single-exponential models with robs = 450 i 20 ps (kobS = 2.3 :l: 0.13 x 109 s"), 755 r 40 pS (km. = 1.3 a: 0.06 x 109 s"), and 2.50 a 0.1 ns (km. = 4.0 :t 0.09 x 108 S”) for complexes 1, 2, and 3, respectively (Table 2-4). The signal-to-noise ratio is relatively poor owing to a combination of virtually complete quenching of the Rel-based 3MLCT states coupled with radiative rate constants that are on the order of 105 s']. The time constants of the A1m model complexes are several orders of magnitude shorter than what was observed for the Fe"I complexes, indicating that excited-state relaxation in [II complexes 1-3 is dominated by reaction with the Fe core. The rate constant for the reaction can be given by equation 2-5, k q (FeRe3) = k (FeRes) " kobs(A1Re3) (2'5) obs where the values of kr and km for the 3MLCT excited state of a given FeRe3 assembly are taken to be equivalent to the corresponding AlRe3 model complex. Given the extensive quenching of the 3MLCT state as indicated by the time constant for decay in complexes 1-3, the observed lifetimes effectively correspond to the quenching time constants in all three cases. 63 4 100» 1 5 ,. E g. 80~ 1‘ b v 7,; .E‘ 60- :: g . 1 8 8 ‘ ll 1 1‘ l1 5 £40 i. l l 0 l 1 I 1 20 1 1 I I I 0 3000 6000 9000 1200015000 0 1 2 3 4 Time (ns) Time (ps) 50— 5 2 A 40— 50? [ > ”a? ‘ ‘ 40 E 30- 3 i . r? a 30- . r: 20 a 3 a) _ E E 20 I‘ll 10— "‘ l 10 0 l I l 1 1 I 0 1 r I r 1 0 300 600 900 120015001800 0 1 2 3 4 Time (ns) Time (ps) 50" 6 30% A 40r; 25% > 1 7,; E 30 2 gzoor E‘ i‘ 9150- 8 20:1 a E 1 a 100 ._ O , c: l " b—t l 50; 0'1 1 1 1 1 0 1 1 1 l 1 I I 1 I O 300 600 .900 1200 1500 0 2 4 6 8 10 12 14 16 Time (ns) Time (ns) Figure 2-6. Left. Nanosecond time-resolved emission data for the AlRe3 model complexes: 4 (robs: 2260 i 100 ns), 5 (robs: 560 d: 30 ns), and 6 (rob, = 235 d: 20 ns). Right. TCSPC data for the corresponding FeRe3 complexes: 1 (robS = 450 i 30 ps), 2 (1:01,s = 755 :t 40 ps), and 3 (rob, = 2.5 d: 0.1 ns). All data were collected in deoxygenated CH2C12 at room temperature. The red solid lines correspond to fits to single- exponential decay models. 64 Table 2-4. Photophysical Data of Complexes 1-6. E00 k k complexes AW (cm")a (I), k0,,(s‘1) (><105s")d (x106s")e FeRe3 tmb(1) b b c 2.3i0.1x109 f f bpy(2) b b c 1.3:l:0.1x109 f f deeb(3) b b c 4.0i0.1x108 f f AIR83 tmb(4) 526 19,900 0.51 4.4i0.2x105 2.3101 0.21001 bpy(5) 566 18,700 0.16 1.8i0.1x106 2,910.1 1,510.1 deeb (6) 624 16,800 0.07 4.3 :l: 0.2 x 106 3.0 :1: 0.1 4.0 a 0.2 aZero-point energy difference between 3MLCT excited state and ground state based on spectral fitting analysis. bThis value is expected to be identical to the corresponding AlIII complex. °Values are not quoted due to the presence of an emissive impurity. See text for further details. dk, = kob,*(1),. ekm = kob, - k,. fExcited-state decay kinetics dominated by kobs- 2.4 Mechanistic Considerations. 2.4.1 Electron versus Energy Transfer Quenching. Both electron and energy transfer processes can be envisioned to occur out of the Rel-based 3MLCT excited state. Based on the presence of Fem in the ground states of these complexes, electron transfer would proceed as an oxidative quenching . . 11 reaction to produce a Ren/FeII charge separated spe01es. ReI —> Re 65 ”I —-+ FeII reduction potentials for complexes 1-3 (Table 2- oxidation and Fe 3), along with the zero-point energy gaps of the 3MLCT states (E00) determined from fits of the emission spectra of the corresponding A1Re3 analogs (Table 2-4), were used to determine the thermodynamic driving force for photoinduced electron transfer.”78 These calculations revealed that electron transfer is thermodynamically unfavorable for complexes 2 and 3 (~0 and +0.45 eV, respectively), and only slightly exothermic in the case of complex 1 (-0.1 eV). The magnitude of AGET for complex 3 is prohibitively large, particularly given the observed rate constant of nearly 109 s]. In the case of complex 2, electron transfer at room temperature is thermodynamically feasible; however, the fact that the quenching rate is only a factor of ~3 faster than what is observed for complex 3 suggests that both of these complexes are reacting via similar mechanisms, i.e., energy transfer. The fact that electron transfer is predicted to be exothermic in the case of complex 1 prompted further study. We therefore carried out time-resolved absorption measurements on [F e(pyacac)3(Re(tmb)(CO)3)3](OTf)3 (1) in an effort to identify whether a charge-separated species was being formed upon 1A1 —+ lMLCT excitation. The transient absorption spectroscopy of ReI pOlypyridyl complexes has been described by a number of workers.”86 66 Their excited-state spectra typically consist of a moderately intense feature in the ultraviolet corresponding to absorptions of the polypyridyl radical anion, as well as a transient bleach due to loss of the ground-state 'AI —-> lMLCT absorption. Additional absorptions can also be observed toward the red edge of the visible spectrum that are usually ascribed to bpy' transitions of the excited state species. Consistent with these expectations, the differential absorption spectrum of [Al(pyacac)3(Re(tmb)(CO)3)3](OTf)3 (4) (Figure 2-7, top) exhibits transient absorptions at 370 nm and 760 nm that we assign to tmb-based transitions. The hallmark for an excited-state electron transfer process in complex 1 would be a wavelength-dependence in the observed kinetics, namely the loss of the tmb' features coupled with a persistence of the ground-state 1A] —> 1MLCT bleach. Instead, what we observe is complete ground-state recovery at all probe wavelengths with a time constant that is within experimental error of what was measured via time-resolved emission spectroscopy (Figure 2-7, bottom). We have recently discussed the notion that this observation does not necessarily rule out a sequential electron transfer process.87 In the present case, however, the driving force for charge-recombination would place the back-reaction deep enough into the inverted region such that the relative rates necessary to satisfy this condition are not likely to be realized. 67 0.12 - 0.10 ~ 0.08 r 0.06 r 0.04 - 13,} 0.02 - 5194831113 5 0.00 — ................................................. Change in Absorbance 400 500 600 700 800 0.0014 0.0012 0.0010 - 0.0008 - 0.0006 — 0.0004 - 0.0002 - 0.0000 3 -------- . -------- . ------- ------:- 0 200 400 600 800 1000 1200 Time (ps) Change in Absorbance Figure 2-7. Top. Nanosecond time-resolved differential absorption spectrum of [Al(pyacac)3(Re(tmb)(CO)3)3](OTf)3 (4) in room- temperature CH2C12 solution. The spectrum was constructed from the amplitudes of fits to single-exponential decay kinetics at each probe wavelength following excitation at 355 nm. Bottom. Time-resolved absorption data for [Fe(pyacac)3(Re(tmb)(CO)3)3](OTt)3 (1) in room- temperature CHzClz solution at Aprobe = 700 nm following ~100 fs excitation at Awmp = 400 nm. The red solid line corresponds to a fit of the data to a single-exponential decay model with rob, = 400 :t 30 ps. 2.4.2 Dexter vs Ftirster Energy Transfer. In light of these observations, the most likely explanation for 3MLCT quenching in all three of the F eRe3 assemblies is excited-state energy transfer. As discussed in the Introduction, 68 the Dexter mechanism requires orbital overlap between the donor and acceptor involved in the energy transfer. The X-ray structure data for 111 complex 2 shows a Rel-«F e separation of nearly 10 A, a value that lies at the limit of what is typically considered for an exchange-based process.”90 "1 subunits The lack of significant electronic coupling between the ReI and Fe is also supported by the similarities in v(CO) frequencies of the FeRe3 and AlRe3 analogs, as well as the fact that the absorption spectra of the FeRe3 assemblies can be represented in terms of a simple linear combination of its constituents (Figure 2-3). The lack of electronic communication denoted by the structural and spectroscopic data greatly attenuates the plausibility of a Dexter energy transfer mechanism for the excited-state quenching reactivity. The applicability of the Forster mechanism is supported by the moderate degree of spectral overlap that exists between the Rel-based 3MLCT emission and the Fem-based 6A1 —) 6LMCT absorption for the FeRe3 assemblies:91 this is depicted graphically in Figure 2-8. The overlap between the 3MLCT —> 'A1 emission of the Alm-containing tmb (4) (blue trace), bpy (5) (green trace), and deeb (6) (red trace) analogs with the 6A1 —> 6LMCT absorption of Fe(phacac)3 is shown. In particular, the plot shows the amount of spectral overlap is greatest for the highest energy 3MLCT excited-state corresponding to the most electron donating group (tmb) and smallest for the 69 most electron withdrawing group (deeb). The area ascertained from the product of the normalized emission profile of an AlRe3 analog with the extinction coefficient spectrum of the corresponding FeRe3 complex was used to calculate the reported spectral overlap values. The magnitude of spectral overlap (J) also corresponds with the observed rate constants for energy transfer with J values of 1.0 x 10'14 (kENT = 2.3 x 109 s'l), 7.6 x 10']5 (kENT = 1.3 x 109 s"), and 1.2 x 10'15 (1n;NT = 4.0 x 108 s") M'cm3 for the tmb (1), bpy (2), and deeb (3) analogs, respectively. It is quite clear that the overlap requirement in these systems is satisfied, which strongly implicates F0rster transfer as the dominant quenching mechanism in the FeRe3 complexes. 1.! l .—- Absorbance (normalized) <2 0 (pozgeunou) [(1181191111 uolssrulg 400 500 600 700 Wavelength (nm) Figure 2-8. Overlay of the emission spectra of [Al(pyacac)3(Re(tmb)- (C0)3)3](OT1)3 (4. blue), [A1(PyacaC)3(Re(bPY)(CO)3)3](0Tf)3 (5, green), and [A1(pyacac)3(Re(deeb)(CO)3)3](OTt)3 (6, red) with the electronic absorption spectrum of Fe(phacac)3. Data were acquired in room-temperature CH2C12 solution. 70 2.4.3 Quantifying Fbrster Energy Transfer. The spectral properties of the FeRe3 family of complexes combined with the structural rigidity of the system provides a rare opportunity to quantitatively apply F 0rster theory and compare calculated rates with those obtained experimentally. In addition to the overlap factor alluded to above, the rate of energy transfer is also sensitive to the relative orientation of the donor and acceptor transition dipoles (eq 2-3). An accurate determination of this quantity can be quite challenging. In one noteworthy example, Fleming and coworkers utilized TD-DFT (time-dependent density functional theory) to enable them to visualize the transition dipoles of peridinin, which in turn provided them with tremendous insights into the role geometry plays in facilitating energy transfer from the singlet excited state(s) of that system.92 Unfortunately, the complicated electronic structures of transition metal containing systems do not easily lend themselves to a similarly detailed analysis, so more approximate methods must usually be employed. An excellent example of this is that of Harriman and coworkers, in which F0rster energy transfer dynamics in Ru11 and Osll polypyridine donor- acceptor complexes bridged by a rigid Spiro-based spacer moiety were investigated.93 In their approach, energy-minimized structures were calculated for each molecule with the donor and acceptor transition dipole 71 moments modeled along the six respective Ru—N and Os—N bond vectors. Using this geometric picture, calculations of donor-acceptor separations (RDA) and orientation factors (K2) afforded a theoretical F0rster rate constant that agreed very closely with the experimentally observed values. We have taken a similar approach for analyzing the FeRe3 family of complexes, albeit with slight differences in the physical description of the system. For example, we have chosen to approximate the donor and acceptor transition dipoles as bisecting the local C2 axes of the bpy’ and acac ligands as opposed to them lying coincident with the metal-ligand bond vectors. In addition, the point of origin for the transition moment dipoles has been modeled to originate solely from the ligands involved in the donor and acceptor transitions. These choices are based largely on Density Functional Theory (DFT) calculations on [Ru(bpy)3]2+ by Daul94 and Gorelsky‘)5 that suggest the lowest energy excited-state (3MLCT) is localized on the bpy ligands. In addition, Meyer and coworkers have shown that the majority of the amplitude of the thermalized 3MLCT wavefunction for complexes of the form fac-[Re(4,4’-X2bpy)(CO)3(4-Etpy)](PF6) (X = CH3, H, and COzEt) is concentrated within the 11* levels of the bpy’ ligand and less so along the Re—N bond vector.65 72 In light of these reported computational results, Figure 2-9 is drawn to depict the Rel-based 3MLCT emission dipoles as blue arrows and the Fem- based 6LMCT absorption dipoles as red arrows. The emission dipoles of the three Rel donors (Rel, Re2, and R83) relative to the Fe111 center are presented simultaneously for clarity, but only a single ReI moiety is considered involved during an excited-state quenching event due to the low excitation photon flux used for the steady-state and time-resolved emission measurements (Figure 2-10). Figure 2-10 shows a single Rel-based emission dipole moment spatially interacting with the three 6LMCT absorption dipole moments. The donor-acceptor distances (R1, R2, and R3) and the three angles used in the K2 equation (OT, OD, and CA) are also shown, and were evaluated for each donor-acceptor dipolar interaction using the single-crystal X-ray structure data of complex 2. The Rel, Re2, and Re3 donor dipoles were examined individually relative to the three possible 6LMCT absorption dipoles, and an RDA and K2 value was calculated for each case. The charge- transfer nature of the donor and acceptor transitions makes it difficult to place an exact point of origin for each, so distance and orientation factors were calculated for a range of possible loci for both the donor and acceptor (Figure 2-11). Figure 2-11 specifically illustrates the point-dipole model utilized for calculating the donor-acceptor distances and angles, with the 73 through-space interaction of three different donor origins at a single Rel emission dipole with the three 6LMCT absorption dipole origins located on each FeOz triangle of the acceptor. The points A, B, and C, were chosen to systematically vary the location of the 3MLCT wavefunction in order to investigate the origin of the donor dipole. The MLCT nature of the donor emission makes it difficult to place an exact point of origin on the bpy ligand, so donor-acceptor distances and orientation factors for the three points (A, B, and C) along each of the donor moieties (Rel, Re2, and Re3) interacting with the 6LMCT transition dipoles were investigated. In addition, Figure 2-11 also shows the acceptor origin was systematically varied by placing it between the oxygen atoms of each acac ligand (points D1, D2, and D3), at the center of the three F e02 triangles (points E1, E2, and E3), and at the Fem metal center (point F), again due to the charge-transfer nature associated in this case with the acceptor absorption dipoles. The 6LMCT state can be envisioned as occurring fiom either the non-bonding electrons of the oxygen atoms or from the delocalized 7: electrons of the acac ligands, and we feel these three points of origin adequately represents the charge-transfer character of the acceptor with geometric considerations of both the FeIII metal center and acac ligand. This approach yielded a total of 81 donor-acceptor interactions (27 for each ReI fluorophore), each being 74 defined by specific RDA and K2 values that were evaluated based on the single-crystal X-ray structure data of complex 2. A complete list of the values of RDA and K2 used in our analysis can be found in Tables 2—5a, 2-5b, and 2-5c. R R Figure 2-9. Through-space interactions of the Rel-based 3MLCT emission dipoles (blue arrows) with the Fem-based 6LMCT absorption dipoles (red arrows). 75 --..-_..--.., I I I / I Figure 2-10. Single Rel-based emission dipole moment spatially interacting with the three 6LMCT absorption dipole moments. Also shown are the donor-acceptor distances (R1, R2, and R3) and the three angles used in the K2 equation (OT, OD, and 6),.) evaluated from the single-crystal X—ray structure data of complex 2. 76 Figure 2-11. Point-dipole approximation used for calculating the through-space distances (RDA) and orientation factors (K2) needed for quantifying Fbrster energy transfer theory in the FeRe3 assemblies. 77 Table 2-5a. RDA (A), OT (°), OD (°), (9A (°), and K2 values determined from the internal dimensions of complex 2 using donor/acceptor interactions A:D, B:D, and C:D. R(ll) R(12) R(13) R(22) R(21) R(23) R(33) R(31) R(32y 8.46 77.81 84.68 154.1 0.213 10.73 35.87 91.22 46.84 0.729 10.41 56.34 86.53 61.42 0.219 8.40 80.28 86.30 158.7 0.122 10.65 35.52 92.54 49.62 0.810 10.41 54.45 87.17 59.73 0.257 8.38 60.72 87.26 147.2 0.372 10.89 63.14 86.28 28.49 0.079 9.70 39.82 95.31 82.65 0.646 8.64 77.81 77.07 147.5 0.604 10.75 35.87 85.24 51.00 0.427 10.53 56.34 80.30 60.52 0.093 8.55 80.28 78.49 150.9 0.479 10.63 35.52 86.36 53.34 0.490 10.51 54.45 81.05 58.47 0.114 8.3 1 60.72 82.96 142 0.606 10.82 63.14 83.97 23.37 0.026 9.44 39.82 92.33 87.5 0.598 8.99 77.81 69.69 141.4 1.050 10.93 35.87 78.93 54.9 0.230 10.81 56.34 74.03 60.25 0.021 8.82 80.28 71.49 144.1 0.884 10.74 35.52 80.59 57.16 0.300 10.73 54.45 75.28 57.58 0.03 8.48 60.72 75.78 136.4 1.050 10.95 63.14 78.56 19.35 0.012 9.42 39.82 85.98 92.27 0.603 78 Table 2-5b. RDA (A), OT (°), (99 (°), OA (°), and K2 values determined from the internal dimensions of complex 2 using donor/acceptor interactions AzE, B:E, and C:E. R(ll) R(12) R(13) RQZ) R(21) RQBL R(33) R(31) R(32) A RDA 9r @1) ®A K2 9.13 77.81 86.25 156.1 0.153 10.25 35.87 89.49 49.78 0.629 10.08 56.34 87.05 65.02 0.239 9.08 80.28 87.74 160.3 0.079 10.19 35.52 90.84 52.76 0.707 10.07 54.45 88.04 63.29 0.287 9.00 60.72 88.53 149.7 0.309 10.25 63.14 87.85 30.44 0.126 9.64 39.82 92.49 86.91 0.601 B RDA 9.26 77.81 79.17 149.9 0.488 10.31 35.87 83.21 54.12 0.363 10.20 56.34 80.63 64.06 0.116 9.18 80.28 80.52 153.1 0.371 10.21 35.52 84.4 56.63 0.426 10.15 54.45 81.63 61.94 0.141 8.89 60.72 84.84 144.9 0.503 10.15 63.14 85.15 25.01 0.049 9.44 39.82 89.41 91.87 0.591 9.57 77.81 72.16 144.1 0.913 10.53 35.87 76.71 58.12 0.199 10.47 56.34 74.16 63.69 0.037 9.42 80.28 73.95 146.7 0.743 10.37 35.52 78.42 60.56 0.268 10.36 54.45 75.65 60.95 0.049 9.02 60.72 78.13 139.6 0.92 10.26 63.14 79.33 20.71 0.005 9.47 39.82 83.06 96.63 0.656 79 Table 2-5c. RDA (A), OT (°), OD (°), GA (°), and K2 values determined from the internal dimensions of complex 2 using donor/acceptor interactions A:F, B:F, and C:F. R(l 1) R(12) R(13) R(22) R(21) R(23) R(33) R(31) R(32) 9.80 77.81 87.61 22.13 0.009 9.80 35.87 87.61 127 0.784 9.80 56.34 87.61 1 11.2 0.359 9.76 80.28 88.98 18.25 0.014 9.76 35.52 88.98 123.8 0.712 9.76 54.45 88.98 1 12.9 0.363 9.63 60.72 89.63 28.16 0.223 9.63 63.14 89.63 147.4 0.219 9.63 39.82 89.63 88.8 0.589 9.90 77.81 81.01 28.00 0.041 9.90 35.87 81.01 122.5 1.130 9.90 56.34 81 .01 112.2 0.535 9.83 80.28 82.30 24.98 0.038 9.83 35.52 82.30 119.8 1.030 9.83 54.45 82.30 114.3 0.558 9.49 60.72 86.49 32.62 0.112 9.49 63.14 86.49 153.1 0.379 9.49 39.82 86.49 83.77 0.560 10.17 77.81 74.35 33.5 0.215 10.17 35.87 74.35 118.4 1.430 10.17 56.34 74.35 112.7 0.750 10.03 80.28 76.11 31.09 0.200 10.03 35.52 76.11 115.8 1.270 10.03 54.45 76.11 1 15.4 0.793 9.58 60.72 80.21 37.6 0.007 9.58 63.14 80.21 157.8 0.853 9.58 39.82 80.21 79.08 0.451 The analysis we have carried out is predicated on two critical assumptions: (1) that the variations in substitutents on the peripheral bipyridine group do not significantly alter the metrics relevant for dipolar energy transfer (thereby allowing us to use the X-ray structure of complex 2 as the basis for analyzing all three FeRe3 assemblies), and (2) that the geometry of each compound in solution is essentially unchanged from that 80 determined by solid-state X-ray crystallography. The major influence of the bipyridyl substituents will be to shift the electron density associated with the excited state according to the electron donating/withdrawing ability of the group. This assertion is supported by structure minimizations which revealed virtually identical bond distances and angles for all three complexes.96 Given the distribution of anchoring points on the bipyridyl ring that we are evaluating, we believe this first issue is being adequately addressed. In terms of solution versus solid-state geometry, the only significant degree of freedom in these systems is rotation along the Re—N (pyridine) bond. We expect there will be some barrier to this motion, but it is unlikely to afford the same average angle in solution as found in the solid state. Even though this represents a possible difference in the structure of the compound between what we measure in the solid-state versus what exists in solution, an analysis of this motion revealed that the relative distances and orientations of the donor and acceptor transition dipoles (and therefore RDA and K2) do not change over the entire 360° that the system can sample. Therefore, while in principal using a solid-state structure to model geometric properties in solution can be problematic, the particular structural aspects of the F eRe3 family of complexes makes such a comparison very straightforward. 81 2.4.4 Rate Constant Calculations: Modeling Solution Phase Energy Transfer Dynamics. A given ReI donor can couple to any of the three Fe- pyacac acceptor dipoles: the energy transfer process in these compounds can thus be described in terms of three parallel reactions. The overall rate constant for such a kinetic model is given by a linear combination of the rate constants for each reaction pathway as shown in equation 2-6, kRe(ll) + kRe(lz) + kRe(13) = kRel (2‘63) kRe(zl) + kRe(22) + ch(23) = kReZ (2'6b) kRe(31) + kRe(32) + kRe(33) = ch3 (2'69) where kRe(ij) corresponds to the rate calculated for the ith Re donor coupling to the jth Fe—pyacac acceptor; each of the km“) values derives from the average of the nine possible donor-acceptor vectors defined by the point- dipole origins depicted in Figure 2-11. For example, kg,” represents the RelzLMCTl dipole-dipole interaction that is the average of the nine rate constants calculated for each of the point dipole origins along the Rel donor and the covalently attached FeOz triangle (A:Dl, AzEl, A:F, B:Dl, B:El, B:F, C:Dl, C:El, and C:F) (Figure 2-11). Analogous procedures were applied to the Rel :LMCT2 and Rel :LMCT3 interactions, with all three total interactions summed to yield kRel. ‘ This same analysis was applied to the Re2 and Re3 donor moieties to yield km and kRe3, which were then 82 averaged to give an overall theoretical rate constant () from our model. Table 2-6 lists all nine kRer) (Re#:LMCT#) values for complexes 1-3 along with their values. It can be seen that there are variations in the calculated rates of energy transfer within each group of interactions owing to slight geometric differences at each Re-bpy site. Nevertheless, despite the lack of a quantitative picture of wavefunctions for the donor and acceptor charge-transfer states, the level of agreement obtained between experiment and theory — less than a factor of three across the entire series — is quite good and further supports our assignment of F brster transfer. The complete list of calculated rate constants for complexes 1-3 are given in Tables 2-7a, 2-7b, and 2-7c. 83 Table 2-6. Calculated F0rster rate constants for [Fe(pyacac)3(Re(tmb)- (C0)3)3](0Tf)3 (1), 1FC(PyacaC)3(Re(bPY)(CO)3)3](0T1); (2). and [F e(pyacac)3(Re(deeb)(CO)3)3](OTf)3 (3). Interaction"l complex 1b complex 2 complex 3b kr (8") kr (s') kr (s‘) Re1:LMCT1 3.6 x108 3.3 x108 5.5 x107 Rel :LMCT2 2.9 x 108 2.7 x 108 4.4 x 107 RelzLMCT3 1.3x 108 1.2x 108 1.9x 107 1rRel (s")° 7.8 x 108 7.2 x 108 1.2 x 108 Re2zLMCT1 3.1 x 108 2.8 x 108 4.7 x 107 Re2zLMCT2 3.1 x 108 2.8 x 108 4.6 x 107 Re2:LMCT3 1.4 x 108 1.3 x 108 2.2 x 107 er,2 (s‘)6 7.6 x 108 6.9 x 108 1.2 x 108 Re3:LMCT1 1.2x108 1.3 x 108 1.8x 107 Re3:LMCT2 3.9 x 108 3.6 x 108 5.9 x 107 Re3zLMCT3 5.3 x 108 4.9 x 108 8.0 x 107 km (s")c 1.0 x 109 9.8 x 108 1.6 x 108 (s")d 8.5 x 108 8.0 x 10" 1.3 x 108 km, (s‘) 2.3 x 109 1.3 x 10" 4.0 x :108 aDonor-acceptor through-space interaction as defined in the text and in Figure 2-11. bRDA and K2 values derived from the single-crystal X-ray data for complex 2 using the geometries outlined in Figures 2-9 and 2- 10. A complete list of calculated rate constants for complexes 1-3 can be found in Tables 2-7a, 2-7b, and 2-7c. cRate of energy transfer calculated according to eqs 2-6. dOverall rate of energy transfer given by (kRel '1' kReZ + ch3)/3- 84 Table 2-7a. Calculated Fbrster rate constants (8") for complexes 1-3 at all donor-acceptor interactions occurring at A:D, B:D, and C:D. Donor complex R(ll) R(12) R(13) kp (Rel) tmb (1) 2.9 x 108 2.4 x 108 8.6 x 107 6.2 x 108 A bpy(2) 2.7x 108 2.2x 108 8.0x 107 5.7x 108 deeb (3) 4.4 x107 3.6 x 107 1.3 x 107 9.3 x 107 tmb(1) 7.3 x 108 1.4x108 3.4x 107 9.0x 108 B bpy(2) 6.7x 108 1.3x 108 3.2x 107 8.3x 108 deeb(3) 1.1x108 2.1x107 5.2x 106 1.4x 108 tmb(1) 1.0x 109 6.8x 107 6.6x 106 1.1x109 C bpy (2) 9.2 x 108 6.3 x 107 6.1x 106 9.9 x 108 deeb(3) 1.5x 108 1.0x107 1.0x106 1.6x 108 R(22) R(21) R(23) 1rF (Re2) tmb(1) 1.7x 108 2.8x 108 , 1.0x108 5.5 x 108 A bpy(2) 1.6x108 2.6x 108 9.4x107 5.1x108 deeb (3) 2.6 x 107 4.2 x 107 1.5 x 107 8.3 x 107 tmb (1) 6.2 x 108 1.7 x 108 4.2 x 107 8.3 x 108 13 bpy(2) 5.7x108 1.6x108 3.9x 107 7.7x 108 deeb (3) 9.3 x 107 2.6 x 107 6.4 x 106 1.3 x 108 tmb(1) 9.4x 108 9.8x 107 9.9x 106 1.0x109 C bpy(2) 8.7x 108 9.1x 107 9.1x106 9.7x 108 deeb(3) 1.4x108 1.5x107 1.5x106 1.6x108 R(33) R(31) R(32) k..- (Re3) tmb(1) 5.4x 108 2.4x 107 3.9x108 9.5 x 108 A bpy(2) 5.0x 108 2.2x107 3.6x 108 8.8x 108 deeb(3) 8.1x107 3.6x106 5.9x 107 1.4x108 tmb(1) 9.2x 108 8.1x106 4.2x 108 1.3 x 109 B bpy(2) 8.5x 108 7.5x 106 3.9x108 1.2x109 deeb(3) 1.4x108 1.2x106 6.4x 107 2.1x108 tmb(1) 1.4x109 3.5x 106 4.3x 108 1.8x 109 C bpy(2) 1.3x109 3.2x106 4.0x 108 1.7x109 deeb (3) 2.1x108 5.3 x 105 6.5 x 107 2.8 x 108 85 Table 2-7 b. Calculated F0rster rate constants (3") for complexes 1-3 at all donor-acceptor interactions occurring at AzE, B:E, and C:E. Donor complex R(ll) R(12) R(13) kp (Re1) A tmb(1) bpy(2) deeb (3) 1.3x108 1.2x108 2.0x 107 2.7x 108 2.5 x 108 4.1x107 1.1x108 1.1x108 1.7x107 5.1x108 4.8x 108 7.8x 107 tmb (1) bpy(2) deeb (3) 3.9x108 3.6x108 5.9x107 1.5 x 108 1.4x108 2.3 x 107 5.2 x 107 4.8 x107 7.8 x 106 5.9 x 108 5.5 x 108 9.0x 107 tmb(1) bpy(2) deeb (3) 6.0x 108 5.5 x 108 9.0 x 107 7.3 x 107 6.8x107 1.1x107 1.4x107 1.3x107 2.1x106 6.9x 108 6.3 x 108 1.0x108 tmb(1) bpy(2) deeb (3) R(22) R(21) R(23) 1rF (Re2) 7.1x 107 6.5x 107 1.1x107 3.2 x 108 2.9x 108 4.8 x 107 1.4x108 1.3x108 2.1x107 5.3 x 108 4.9x 108 8.0x107 tmb(1) bpy(2) deeb (3) 3.1x 108 2.9x108 4.7x 107 1.9x108 1.7x108 2.9x107 6.5 x 107 6.0 x 107 9.8 x 106 5.7 x 108 5.2 x 108 8.6 x 107 tmb(1) bpy(2) deeb (3) 5.3 x 108 4.9x108 8.1x 107 1.1x108 1.0x108 1.6x107 2.0 x107 1.8 x 107 3.0x 106 6.6x108 6.1x108 1.0x108 tmb(1) bpy(2) deeb (3) R(33) R(31) R(32) kp (Re3) 2.9 x 108 2.7 x 108 4.4 x 107 5.5 x 107 5.0 x 107 8.2 x 106 3.8x 108 3.5x108 5.7x107 7.3x 108 6.7x108 1.1x108 tmb(1) bpy(2) deeb (3) 5.1x108 4.7x108 7.7x107 2.3 x 107 2.1x 107 3.4x106 4.2 x 108 3.9x108 6.3 x 107 9.5 x 108 8.8x108 1.4x108 tmb(1) bpy(2) deeb (3) 8.6x 108 7.9x 108 1.3 x 108 2.2 x 106 2.0x106 3.3 x105 4.6 x 108 4.2 x 108 6.9 x 107 1.3 x 109 1.2x109 2.0x108 86 Table 2-7c. Calculated Fbrster rate constants (8‘) for complexes 1-3 at all donor-acceptor interactions occurring at A:F, B:F, and C:F. Donor compound R(l 1) R(12) R(13) kp (Re1) tmb(1) 5.1x106 4.4x 108 2.0x108 6.5 x 108 A bpy(2) 4.7 x 106 4.1 x 108 1.9 x 108 6.0 x 108 deeb(3) 7.7 x 105 6.7 x 107 3.1 x 107 9.9 x 107 tmb(1) 2.2 x 107 6.0 x 108 2.9 x 108 9.1 x 108 B bpy(2) 2.0 x 107 5.6 x 108 2.6 x 108 8.4 x 108 deeb(3) 3.3 x 106 9.1 x 107 4.3 x 107 1.4 x 108 tmb(1) 9.7x107 6.5x108 3.4x108 1.1x109 C bpy(2) 9.0x107 6.0x108 3.1x108 1.0x109 deeb(3) 1.5 x 107 9.8 x 107 5.1x107 1.6 x 108 R(22) R(21) R(23) kF (Re2) tmb(1) 8.1x 106 4.1x108 2.1x108 6.3x 108 A bpy(2) 7.5 x 106 3.8x108 1.9x108 5.8x 108 deeb(3) 1.2 x 106 6.2 x 107 3.2 x 107 9.5 x 107 tmb(1) 2.1x107 5.7x108 3.1x108 9.0x108 B bpy(2) 2.0 x 107 5.3 x 108 2.9 x 108 8.4 x 108 deeb(3) 3.2 x 106 8.6 x 107 4.7 x 107 1.4 x 108 tmb(1) 9.9x107 6.3 x 108 3.9x108 1.1x109 C bpy(2) 9.1x107 5.8 x 108 3.6x108 1.0x109 deeb(3) 1.5 x 107 9.5 x 107 5.9 x 107 1.7 x 108 R(33) R(31) R(32) 1rF (Re3) tmb(1) 1.4x108 1.4x108 3.7x 108 6.5 x 108 A bpy(2) 1.3 x 108 1.3 x108 3.4x108 6.0x108 deeb(3) 2.1 x 107 2.1 x 107 5.6 x 107 9.8 x 107 tmb(1) 7.7 x 107 2.6 x 108 3.8 x 108 7.2 x 108 B bpy(2) 7.1 x 107 2.4 x 108 3.6 x 108 6.7 x 108 deeb(3) 1.2x107 3.9x107 5.8x107 1.1x108 tmb(1) 4.5 x 106 5.5 x 108 2.9 x 108 8.4 x 108 C bpy(2) 4.2 x 106 5.1 x 108 2.7 x 108 7.8 x 108 deeb(3) 6.9 x 105 8.4 x 107 4.4 x 107 1.3 x 108 87 The results of this study allow us to construct a comprehensive picture of the excited-state energies and dynamics for these FeRe3 systems (Figure 2—12). The lefi side of Figure 2-12 is an energy level diagram for the relevant electronic states of the Rel-bpy’ chromophore, along with kinetic pathways associated with the various excited-states. Initial population of the IMLCT excited-state is followed by rapid intersystem crossing (kisc) to the 3MLCT excited-state.”98 The thermalized triplet state can then undergo radiative (k,) and non-radiative (km) transitions to the 'Al ground-state, or can be quenched by the Fem core via Fdrster energy transfer (kENT)- The right side of Figure 2-12 shows the electronic structure of the F e(pyacac)3 core, which contains charge-transfer (6LMCT) and ligand-field electronic excited-states that are thermodynamically accessible from the 3MLCT manifold of the Re chromophore. Dipolar energy transfer results in the formation of a 6LMCT excited state within the Fe(pyacac)3 core, followed by 111 non-radiative relaxation to the 6A1 ground-state of the Fe moiety. 88 E: IMLCT::~~-l(.i§9~10_l3 S'_1 ‘Ei:kvib ~1012 S.1 :E kEnT ~108-109 s'1_ ,,.r""3MLCT hv 1r,~105 s'l [Xx ~ 6 -1 —I6LMCT . km 10 s . . . . (N on-radlatlve relaxatlon 5 via ligand-field manifold to 6A1 ground-state) Rel-bpy' F eIH(pyacac)3 chromophore core Figure 2-12. Energy level diagram depicting the excited-state dynamics of the FeRe3 assemblies. The rate constants for lMLCT —> 3MLCT intersystem crossing and vibrational cooling within the 3MLCT state are based on the work of Vléek and co-workers (cf. 85), whereas the other time constants represent approximate values for complexes 1- 6. 2.5 Geometry Optimization Calculation As mentioned in the previous section, one of the more notable aspects of the FeRe3 series are the rigid building blocks (i.e., stiff pyacac bridging ligands and coordination environments) that constitute these supramolecular complexes. These structural features allowed the crystal structure of Complex 2 to be utilized in calculating the donor-acceptor distances (RDA) and orientation factors (1(2) for all three analogs based on the geometries defined in Figures 2—9, 2-10, and 2-11. The ability to directly use the single- 89 crystal X-ray data to model solution phase structures and by extension the solution phase energy transfer dynamics is a unique property of the FeRe3 series (Table 2-6). A potential drawback of this approach is the effect crystal packing forces may have on the internal structure of complex 2. From the X-ray structural data of complexes 2 and 5 (Figure 2-2), it can be seen that the pyacac bridging ligands are slightly bent away from the expected linear geometry. This structural variation represents a possible error in the donor- acceptor distances (RDA) and orientation factors (1(2) calculated from the model given in Figure 2-10. In order to investigate if this possible source of error has affected the theoretical energy transfer rate constants, a gas phase geometry optimization calculation of complex 2 was performed using the - 99,1 Gauss1an 03 00 software package to determine if any differences exist between the solution phase and confined crystal structure geometries. Figure 2-13 shows the optimized geometry of complex 2, along with the ill . . 1 . observed llnear connectlon between the Fe and Re metal centers Vla the relaxed geometry of the pyacac bridge. 90 Figure 2-13. Drawing of the geometry optimized structure of [Fe(pyacaC)3(Re(bPY)(C0)3)3l(0T03 (2)- 111"-Re1 distances and the dihedral angles Table 2—8 compares the Fe between the F e02 and acac planes for the three F em/ReI donor-acceptor arms for the optimized and solid-state structures of complex 2. From Table 2—8, the dihedral angles between the FeOz triangles and the corresponding acac planes are close to zero in the optimized geometry and are bent away from planarity in the solid-state, which indicates crystal packing forces are affecting the internal geometry of the F eRe3 complexes. The bent geometry also decreases the Fen-Re distances by 0.2 — 0.3 A compared to the optimized geometry. The elongation of the metal—metal distances do 91 represent a source of error in the F6rster calculations that were based on the X-ray structural data of complex 2, so an identical analysis of the donor- acceptor distances (RDA) and orientation factors (1(2) for the optimized geometry of complex 2 was performed. The analyses and labeling system for the various donor and acceptor point dipoles and through-space interactions are identical to the system used for the single-crystal X-ray data outlined in section 2.4.3. The complete list of RDA and K2 values calculated using the optimized geometry of complex 2 can be found in Tables 2-9a, 2- 9b, and 2-9c. Table 2-8. Fen-Re distances and F eOz/acac dihedral angles for the optimized and single-crystal X-ray structure data for complex 2. Optimized single-cgstal FeOz/acac (1) (°)a 2.70 15.60 FeOz/acac (2) (°)21 2.80 10.75 F eOz/acac (3) (°)a 3.39 23.04 Fen-Rel (A) 10.080 9.881 Fe-"Re2 (A) 10.081 9.879 F e-"Re3 (A) 10.092 9.775 aDihedral angle between the FeOz triangle and corresponding acac planes. 92 Table 2-9a. RDA (A), OT (°), OD (°), (9A (°), and K2 values determined from the optimized geometry of complex 2 using donor/acceptor interactions A:D, B:D, and C:D. R(l 1) nag RQ3L R(ZZJ RQI) R(23) R(33LR(31) R(32) A RDA 9T (")0 8.89 81.41 73.84 174.5 0.961 11.26 33.77 68.82 52.56 0.030 11.19 52.50 80.69 55.21 0.110 8.88 87.15 74.40 172.0 0.720 11.16 35.29 76.14 55.78 0.170 1 1.27 50.65 76.00 51.18 0.032 8.89 81.50 74.05 174.4 0.938 1 1.28 35.11 69.07 51.20 0.022 11.20 52.55 80.97 54.54 0.112 9.28 81.41 67.12 167.7 1.661 11.73 33.77 63.81 47.77 0.003 1 1.42 52.50 75.23 60.84 0.056 9.26 87.15 67.59 164.6 1.328 1 1.49 35.29 70.74 56.59 0.074 1 1.60 50.65 70.66 51.92 0.001 9.27 81.50 67.33 167.5 1.630 11.75 35.1 1 64.08 46.33 0.008 1 1.43 52.55 75.52 60.23 0.056 9.82 81.41 60.42 161.6 2.417 12.32 33.77 58.52 43.87 0.089 11.81 52.50 69.30 65.87 0.031 9.796 87.15 60.84 158.4 1.984 1 1.96 35.29 65.02 57.65 0.019 12.07 50.65 64.99 52.98 0.017 9.81 81 .50 60.62 161.4 2.379 12.33 35.11 58.79 42.35 0.109 11.81 52.55 69.59 65.33 0.029 93 Table 2-9b. RDA (A), OT (°), OD (°), OA (°), and K2 values determined from the optimized geometry of complex 2 using donor/acceptor interactions AzE, B:E, and C :E. R(ll) R(12) R(13) 9.65 81 .41 74.19 174.9 0.928 10.81 33.77 71.53 55.75 0.088 11.19 52.50 77.70 58.52 0.076 R(22) R(21) 9.63 87.15 75.01 172.7 0.671 10.75 35.29 75.83 59.13 0.193 R(23) R(33) 10.81 50.65 75.75 54.31 0.041 9.64 81 .50 74.41 174.8 0.904 RQI) 10.82 35.11 71.78 54.33 0.073 R(32) 10.78 52.55 77.96 57.82 0.076 10.02 81.41 68.01 168.6 1.564 1 1.23 33.77 66.18 50.64 0.004 11.07 52.50 72.10 64.26 0.043 9.99 87.15 68.74 165.8 1.220 1 1.09 35.29 70.22 59.87 0.094 1 1.15 50.65 70.18 54.99 0.003 10.01 81.50 68.24 168.4 1.531 11.24 35.1 1 66.43 49.13 0.001 1 1.07 52.55 72.36 63.64 0.042 10.55 81.41 61.70 162.9 2.276 11.78 33.77 60.56 46.42 0.034 11.52 52.50 66.13 69.32 0.032 10.51 87.15 62.36 159.9 1.840 1 1.57 35.29 64.33 60.83 0.034 1 1.63 50.65 64.32 55.93 0.009 10.53 81.50 61.92 162.7 2.238 11.78 35.11 60.81 44.84 0.048 11.51 52.55 66.39 68.77 0.030 94 Table 2-9c. RDA (A), 9r (°), (99 (°), @A (°), and K2 values determined from the optimized geometry of complex 2 using donor/acceptor interactions A:F, B:F, and C:F. R(ll) R112) R(13) R(22) R(21) R(23) 10.40 81 .41 74.49 4.70 0.423 10.40 33.77 74.49 120.8 1.543 10.40 52.50 74.49 117.9 0.969 10.38 87.15 75.54 6.81 0.482 10.38 35.29 75.54 117.3 1.345 10.38 50.65 75.54 122.3 1.070 R931 M31) R(32) 10.40 81.50 74.73 4.82 0.409 10.40 35.11 74.73 122.3 1.538 10.40 52.55 74.73 118.6 0.974 10.77 81.41 68.78 10.58 0.843 10.77 33.77 68.78 126.2 2.170 10.77 52.50 68.78 112.1 1.035 10.73 87.15 69.73 13.22 0.926 10.73 35.29 69.73 116.6 1.644 10.73 50.65 69.73 121.7 1.393 10.75 81 .50 69.02 10.78 0.823 10.75 35.11 69.02 127.8 2.180 10.75 52.55 69.02 112.7 1.047 11.27 81.41 62.82 15.96 1.365 11.27 33.77 62.82 130.8 2.981 11.27 52.50 62.82 107.1 1.023 11.22 87.15 63.69 18.77 1.462 11.22 35.29 63.69 115.8 1.945 1 1.22 50.65 63.69 120.8 1.730 11.26 81.50 63.05 16.18 1.341 11.26 35.11 63.05 132.4 3.012 11.26 52.55 63.05 107.6 1.040 Table 2-10 lists all nine kRe(ij) (Re#:LMCT#) values and the value calculated for the optimized geometry of complex 2 based on the analysis outlined in section 2.4.5. The value calculated from the single-crystal X—ray structure of 2 is also listed in Table 2-10 for comparison, with the optimized and solid-state geometries yielding 95 values of 9.6 x 108 s'l and 8.0 x 108 s", respectively. The optimized geometry does give a closer energy transfer rate constant to the observed value of 1.3 x 109 s", but only by a small factor considering the point-dipole approximations assumed in the rate calculations (RDA and K2). These results confirm the applicability of single-crystal X-ray data to structurally well- defined donor-acceptor complexes and particularly for the donor-acceptor separations and orientation factors used in the energy transfer rate calculations in the FeRe; family. Due to the structural homology of the FeRe3 complexes, geometry optimization calculations were not performed on complexes 1 (tmb) and 3 (deeb). The complete list of calculated rate constants for complex 2 are given in Table 2-11. 96 Table 2-10. Calculated Forster rate constants for [Fe(pyacac)3(Re- ' (bpy)(CO)3)3](OTf)3 (2) from geometry optimization calculations. Interactiona complex 2b kT (3-1) Rel :LMCTI 6.8 x 108 RelzLMCTZ 2.1 x 108 RelzLMCT3 1.1 x 108 kRel (S-l)C 1.0 x 109 Re2:LMCT1 1.8 x 108 Re2:LMCT2 5.6 x 108 Re2:LMCT3 1.4 x 108 km,2 (s")° 8.8 x 108 Re32LMCT1 2.2 x 108 Re3zLMCT2 1.1x 108 Re3zLMCT3 6.6 x 108 kRe3 (54).: 9.9 x 108 kT (s")d 9.6 x 108 kT (s'l) (X-ray) 8.0 x 108 k,.,,(s") 1.3 x 10" aDonor-acceptor through-space interaction as defined in the text and in Figure 2-11. bRDA and K2 values derived from the minimized gas phase structure using the geometries outlined in Figures 2-9 and 2-10. A complete list of the calculated rate constants for complex 2 can be found in Table 2-10. °Rate of energy transfer calculated according to eqs 2-6a, 2-6b, and 2-6c. dOverall rate of energy transfer given by (kRel + km + kRe3)/3. 97 Table 2-11. Calculated Forster rate constants (8") for the optimized structure of complex 2 at all donor- acceptor interactions. Interaction R(l 1) R(12) R(13) kp (Re1) A-D 9.0 x 108 6.8 x 106 2.6 x 107 9.3 x 108 B-D 1.2x109 6.1x105 1.2x107 1.2x109 C-D 1.3x109 1.2x107 5.3x 106 1.3x109 A-E 5.3 x108 2.6x107 2.2x107 5.8x108 B-E 7.2x108 9.1x 105 1.1x107 7.3 x 108 CE 7.7 x 108 5.9 x 106 6.4 x 106 7.8 x108 A-F 1.6x108 5.6x 108 3.6x108 1.1x 1o9 B-F 2.5x 108 6.5x108 3.1x108 1.2x109 C-F 3.1x108 6.7x108 2.3x 108 1.2x109 R(22) R(21) R(23) kF (Re2) A-D 6.8 x 108 4.1x107 7.3 x 106 7.3 x 108 B-D 9.8x108 1.5x107 8.7x104 1.0x108 C-D 1.0x109 3.0x106 2.5 x 106 1.0x109 A-E 3.9x108 5.8x107 1.2x107 4.6x 108 B-E 5.7 x 108 2.3 x 107 6.2 x 105 5.9 x 108 GE 6.4 x 108 6.5 x 106 1.7 x 106 6.5 x 108 A-F 1.8x108 5.0x108 4.0x 108 1.1x109 B-F 2.8x108 5.0x108 4.2x108 1.2x109 C-F 3.4x108 4.5x 108 4.0x108 1.2x109 R(33) R(31) R(32) kF (Re3) A-D 8.8 x 108 4.8 x 106 2.6x107 9.1x 1o8 B-D 1.2 x 109 1.4x106 1.2 x 107 1.2 x 109 C-D 1.2 x 109 1.4x107 5.0x106 1.2 x 109 A-E 5.2x108 2.1x 107 2.2x107 5.6x 108 B-E 7.1x108 2.5 x 105 1.1x107 7.2x 108 GB 7.6 x 108 8.3 x 106 6.0 x 106 7.7 x 108 A-F 1.5 x 108 5.7 x 108 3.6x108 1.1x109 B-F 2.5x 108 6.5x108 3.1x108 1.2x109 C-F 3.1x108 6.9x108 2.4x108 1.2 x 109 98 2.6 Conclusions The synthesis, structures, and photophysical properties of a series of donor-acceptor complexes based on Rel-bipyridine donors and Fem-acac acceptors have been described. Steady-state and time-resolved emission spectroscopies indicated that the strongly emissive Rel-based 3MLCT excited-state was significantly quenched when compared to model 1" center had been replaced by All". The complexes in which the Fe favorable overlap between the donor emission and acceptor absorption profiles coupled with a ca. 10 A donor—acceptor separation, unfavorable driving forces for electron transfer, and the absence of features characteristic of charge separation in the transient absorption spectra allowed for an assignment of F firster (dipolar) energy transfer as the dominant excited-state reaction mechanism. The well-defined structural aspects of this system permitted a quantitative geometric analysis of the dipole-dipole coupling giving rise to the observed dynamics. The calculated energy transfer rate constants differed from the experimental values by less than a factor of three, a level of agreement that is significantly better than what is typically encountered. In addition to providing quantitative support for F 6rster transfer in this system, this study also demonstrates the degree of accuracy 99 that can be achieved if the metric details concerning dipole-dipole interactions can be explicitly described. 100 2.7 References and Notes (1) Alstrum-Acevedo, J. H.; Brennaman, M. K.; Meyer, T. J. Inorg. Chem. 2005, 44, 6802. (2) Balzani, V.; Scandola, F ., Supramolecular Photochemistry. Horwood: Chichester, U. K.; 1991. (3) Ziessel, R.; Hissler, M.; El-ghayoury, A.; Harriman, A. Coord. Chem. Rev. 1998, 178, 1251. (4) Keefe, M. H.; Benkstein, K. D.; Hupp, J. T. Coord. Chem. Rev. 2000, 205, 201. (5) Balzani, V.; Ceroni, P.; Juris, A.; Venturi, M.; Campagna, S.; Puntoriero, F.; Serroni, S. Coord. Chem. Rev. 2001, 219, 545. (6) Balzani, V.; Juris, A.; Venturi, M.; Campagna, S.; Serroni, S. Chem. Rev. 1996, 96, 759. (7) Argazzi, R.; Bertolasi, E.; Chiorboli, C.; Bignozzi, C. A.; Itokazu, M. K.; Murakami Iha, N. Y. Inorg. Chem. 2001, 40, 6885. (8) Meyer, T. J. Acc. Chem. Res. 1989, 22, 163. (9) De Silva, A. P.; Gunaratne, H. Q. N.; Gunnlaugsson, T.; Huxley, A. J. M.; McCoy, C. P.; Rademacher, J. T.; Rice, T. E. Chem. Rev. 1997, 97, 1515. (10) De Cola, L.; Belser, P. Coord. Chem. Rev. 1998, 177, 301. (11) Barigelletti, F.; Flamigni, L. Chem. Soc. Rev. 2000, 29, 1. (12) Hammarstrom, L. Curr. Opin. Chem. Biol. 2003, 7, 666. (13) Reece, S. Y.; Nocera, D. G. J. Am. Chem. Soc. 2005, 12 7, 9448. 101 (14) (15) (16) (17) (18) (19) (20) (21) (22) (23) (24) (25) (26) (27) Xu, Y. H.; Eilers, G.; Borgstrom, M.; Pan, J. X.; Abrahamsson, M.; Magnuson, A.; Lomoth, R.; Bergquist, J.; Polivka, T.; Sun, L. C.; Sundstrom, V.; Styring, S.; Hammarstrom, L.; Akermark, B. Chem. Eur. J. 2005, 11, 7305. Dexter, D. L. J. Chem. Phys. 1953, 21, 836. F firster, T. Discuss. Faraday Soc. 1959, 27, 7. Brown, W. R.; O'Boyle, N. M.; McGarvey, J. J.; Vos, J. G. Chem. Soc. Rev. 2005, 34, 641. Barigelletti, F.; F lamigni, L.; Balzani, V.; Collin, J .-P.; Sauvage, J .-P.; Sour, A.; Constable, E. C.; Cargill Thompson, A. M. W. J. Am. Chem. Soc. 1994, 116, 7692. Goeb, S.; De Nicola, A.; Ziessel, R.; Sabatini, C.; Barbieri, A.; Barigelletti, F. Inorg. Chem. 2006, 45, 1173. Chiorboli, C.; Bignozzi, C. A.; Scandola, F.; Ishow, E.; Gourdon, A.; I Launay, J .-P. Inorg. Chem. 1999, 38, 2402. Indelli, M. T.; Scandola, F.; Collin, J.-P.; Sauvage, J.-P.; Sour, A. Inorg. Chem. 1996, 35, 303. Bilakhiya, A. K.; Tyagi, B.; Paul, P.; Natarajan, P. Inorg. Chem. 2002, 41, 3830. Scholes, G. D. Ann. Rev. Phys. Chem. 2003, 54, 57. Wong, K. F.; Bagchi, B.; Rossky, P. J. J. Phys. Chem. A. 2003, 108, 5752. Speiser, S. Chem. Rev. 1996, 96, 1953. Constable, E. C.; Handel, R. W.; Housecroft, C. E.; Morales, A. F .; Flamigni, L.; Barigelletti, F. Dalton Trans. 2003, 1220. Ward, M. D.; Barigelletti, F. Coord. Chem. Rev. 2001, 216, 127. l 02 (28) (29) (30) (31) (32) (33) (34) (35) (36) (37) (3 8) (39) (40) Benniston, A. C.; Harriman, A.; Li, R; Sams, C. A. J. Am. Chem. Soc. 2005, 127, 2553. Fleming, C. N.; Dupray, L. M.; Papanikolas, J. M.; Meyer, T. J. J. Phys. Chem. A. 2002, 106, 2328. Van Der Meer, B. W.; Coker, G. 1.; Chen, S.-Y., Resonance Energy Transfer, Theory and Data. VCH Publishers: New York, 1994. Weldon, F .; Hammarstrom, L.; Mukhtar, E.; Hage, R.; Gunneweg, E.; Hassnoot, J. G.; Reedijk, J .; Browne, W. R.; Guckian, A. L.; Vos, J. G. Inorg. Chem. 2004, 43, 4471. Hurley, D. J.; Tor, Y. J. Am. Chem. Soc. 2002, 124, 13231. Schmehl, R. H.; Auerbach, R. A.; Wacholtz, W. F. J. Phys. Chem. 1988, 92, 6202. Gholamkhass, B.; Nozaki, K.; Ohno, T. J. Phys. Chem. B 1997, 101, 9010. Venturi, M.; Marchioni, F .; Ferrer Ribera, B.; Balzani, V.; Opris, D. M.; Schluter, A. D. ChemPhysChem 2006, 7, 229. Furue, M.; Ishibashi, M.; Satoh, A.; Oguni, T.; Maruyama, K.; Sumi, K.; Kamachi, M. Coord. Chem. Rev. 2000, 208, 103. Furue, M.; Maruyama, K.; Kanematsu, Y.; Kushida, T.; Kamachi, M. Coord. Chem. Rev. 1994, 132, 201. Furue, M.; Naiki, M.; Kanematsu, Y.; Kushida, T.; Kamachi, M. Coord. Chem. Rev. 1991, 111, 221. Furue, M.; Yoshidzumi, T.; Kinoshita, S.; Kushida, T.; Nozakura, S.; Kamachi, M. Bull. Chem. Soc. Jpn. 1991, 64, 1632. Grosshenny, V.; Harriman, A.; Hissler, M.; Ziessel, R. J. Chem. Soc., Faraday Trans. 1996, 92, 2223. 103 (41) (42) (43) (44) (45) (46) (47) (48) (49) (50) (51) (52) (53) (54) Wang, X. F .; Herman, B., Eds., Fluorescence Imaging Spectroscopy and Microscopy. John Wiley: New York, 1996. Moore, E. G.; Bernhardt, P. V.; Riley, M. J.; Smith, T. A. Inorg. Chem. 2006, 45, 51. Moore, E. G.; Bernhardt, P. V.; Pigliucci, A.; Riley, M. J.; Vauthey, E. J. Phys. Chem. A. 2003, 107, 8396. Perrin, D. D.; Armarego, W. L. F ., Purification of Laboratory Chemicals. 3rd ed.; Pergamon Press: Oxford, U. K., 1988. Mackay, L. G.; Anderson, H. L.; Sanders, J. K. M. J. Chem. Soc, Perkin Trans. 1. 1995, 18, 2269. Hino, J. K.; Della Ciana, L.; Dressick, W. J.; Sullivan, B. P. Inorg. Chem. 1992, 31, 1072. Caspar, J. V.; Meyer, T. J. J. Phys. Chem. 1983, 87, 952. Vreshch, V. D.; Lysenko, A. B.; Chemega, A. N.; Howard, J. A. K.; Krautscheid, H.; Sieler, J .; Domasevitch, K. V. Dalton Trans. 2004, 2899. SAINT, ver 6. 02a, Bruker AXS, Inc.: Madison, WI, 2000. Sheldrick, G. M. SADABS, ver 2.03, Bruker AXS, Inc.: Madison, WI, 2000. Sheldrick, G. M. SHELXTL, ver 6.12, Bruker AXS, Inc.: Madison, WI, 2001. a) Damrauer, N. H.; Boussie, T. R.; Devenney, M.; McCusker, J. K. J. Am. Chem. Soc. 1997, 119, 8253. b) Juban, E.A.; McCusker, J.K. J. Am. Chem. Soc. 2005, 127, 6857. Claude, J. P.; Meyer, T. J. J. Phys. Chem. 1995, 99, 51. Parker, C. A.; Rees, W. T. Analyst (London) 1960, 85, 587. 104 (55) (56) (57) (58) (59) (60) (61) (62) (63) (64) (65) (66) (67) (68) (69) DeWitt, L.; Blanchard, G. J.; LeGoff, E.; Benz, M. E.; Liao, J. H.; Kanatzidis, M. G. J. Am. Chem. Soc. 1993, 115, 12158. Origin, 7.5 ed.; OriginLab Corp.: Northhampton, MA, 1991 - 2004. Brandenburg, K.; Bemdt, M. DIAMOND, ver 3.1d, Crystal Impact: Bonn, Germany, 2006. Lin, C. T.; Sutin, N. J. Am. Chem. Soc. 1975, 97, 3543. Lin, C. T.; Bottcher, W.; Creutz, C.; Sutin, N. J. Am. Chem. Soc. 1976, 98, 6536. Lin, C. T.; Sutin, N. J. Phys. Chem. 1976, 80, 97. Katz, N. E.; Creutz, C.; Sutin, N. Inorg. Chem. 1988, 2 7, 1687. Wilkinson, F.; Farmilo, A. J. Chem. Soc., Faraday Trans. 1976, 72, 604. Pettersson, K.; Kilsé, K.; Martensson, J .; Albinsson, B. J. Am. Chem. Soc. 2004, 126, 6710. Dattelbaum, D. M.; Omberg, K. M.; Schoonover, J. R.; Martin, R. L.; Meyer, T. J. Inorg. Chem. 2002, 41, 6071. Dattelbaum, D. M.; Martin, R. L.; Schoonover, J. R.; Meyer, T. J. J. Phys. Chem. A. 2004, 108, 3518. Lucia, L. A.; Abboud, K.; Schanze, K. S. Inorg. Chem. 1997, 36, 6224. Chen, P.; Curry, M.; Meyer, T. J. Inorg. Chem. 1989, 28, 2271. Wenger, O. S.; Henling, L. M.; Day, M. W.; Winkler, J. R.; Gray, H. B. Inorg. Chem. 2004, 43, 2043. Busby, M.; Liard, D. J.; Motevalli, M.; Toms, H.; Vléek, A. Inorg. Chim. Acta. 2004, 357, 167. 105 (70) (71) (72) (73) (74) (75) (76) (77) (78) (79) (80) (81) (82) (83) (84) (85) McCarthy, H. J .; Tocher, D. A. Polyhedron 1987, 6, 1421. Sacksteder, L.; Zipp, A. P.; Brown, E. A.; Streich, J .; Demas, J. N.; DeGraff, B. A. Inorg. Chem. 1990, 29, 4335. Worl, L. A.; Duesing, R.; Chen, P.; Della Ciana, L.; Meyer, T. J. J. Chem. Soc., Dalton Trans. 1991, 849. Lintvedt, R. L.; Kernitsky, L. K. Inorg. Chem. 1970, 9, 491. Barnum, D. W. J. Inorg. Nuclear Chem. 1961, 21, 221. Striplin, D. R.; Crosby, G. A. Chem. Phys. Lett. 1994, 221, 426. Kestell, J. D.; Williams, Z. L.; Stultz, L. K.; Claude, J. P. J. Phys. Chem. A. 2002, 106, 5768. Rehm, D.; Weller, A. Isr. J. Chem. 1970, 8, 259. The Rehm-Weller equation for this system is given as AGET = ReI (ox) - FeIII (red) - E00, where E00 is the energy gap between the ground and lowest energy 3MLCT excited-states. Chen, P.; Westmoreland, T. D.; Danielson, E.; Schanze, K. S.; Anthon, D.; Neveux, P. E.; Meyer, T. J. Inorg. Chem. 1987, 26, 1116. Lin, R.; Guarr, T. F.; Duesing, R. Inorg. Chem. 1990, 29, 4169. Tapolsky, G.; Duesing, R.; Meyer, T. J. Inorg. Chem. 1990, 29, 2285. Tapolsky, G.; Duesing, R.; Meyer, T. J. J. Phys. Chem. 1991, 95, 1105. Abbott, L. C.; Arnold, C. J.; Ye, T.; Gordon, K. C.; Perutz, R. N.; Hester, R. E.; Moore, J. N. J. Phys. Chem. A 1998, 102, 1252. Liard, D. J .; Vlcek, A. Inorg. Chem. 2000, 39, 485. Liard, D. J .; Busby, M.; Matousek, P.; Towrie, M.; Vlcek, A. J. Phys. Chem. A 2004, 108, 2363. 106 (86) (87) (88) (89) (90) (91) (92) (93) (94) (95) (96) (97) (98) Busby, M.; Gabrielsson, A.; Matousek, P.; Towrie, M.; Di Bilio, A. J .; Gray, H. B.; Vléek, A. Inorg. Chem. 2004, 43, 4994. Soler, M.; McCusker, J. K. J. Am. Chem. Soc. 2008, 130, 4708. Wang, Y.; Schanze, K. S. Inorg. Chem. 1994, 33, 1354. Bignozzi, C. A.; Bortolini, 0.; Chiorboli, C.; Indelli, M. T.; Rampi, M. A.; Scandola, F. Inorg. Chem. 1992, 31, 172. Berg, K. E.; Tran, A.; Raymond, M. K.; Abrahamsson, M.; Wolny, J .; Redon, S.; Andersson, M.; Sun, L.; Styring, S.; Hammarstrom, L.; Toftlund, H.; Akermark, B. Eur. J. Inorg. Chem. 2001, 1019. Yardley, J. T., Introduction to Molecular Energy Transfer. Academic Press: New York, 1980. Vaswani, H. M.; Hsu, C.-P.; Head-Gordon, M.; Fleming, G. R. J. Phys. Chem. B. 2003, 107, 7940. Juris, A.; Prodi, L.; Harriman, A.; Ziessel, R.; Hissler, M.; El- ghayoury, A.; Wu, F.; Riesgo, E. C.; Thummel, R. P. Inorg. Chem. 2000, 39, 3590. Daul, C.; Baerends, E. J .; Vemooijs, P. Inorg. Chem. 1994, 33, 3538. Gorelsky, S. 1.; Lever, A. B. P. J. Organomet. Chem. 2001, 635, 187. Spartan '02, Wavefunction, Inc.: Irvine, CA, 1991 - 2001. Busby, M.; Matousek, P.; Towrie, M.; Vlcek, A. J. Phys. Chem. A. 2005, 109, 3000. Damrauer, N. H.; Cerulla, G.; Yeh, A.; Boussie, T. R.; Shank, C. V.; McCusker, J. K. Science 1997, 275, 54. 107 (99) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, J ., J. A.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J .; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J .; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, 0.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J .; Gomperts, R.; Stratmann, R. E.; Yazyev, 0.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, 0.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, 1.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez, C.; Pople, J. A. Gaussian 03, Gaussian, Inc.: Wallingford, CT, 2004. (100) Geometry optimization was performed with the UB3LYP functional and LANLZDZ basis set. 108 Chapter 3. Orbital-Specific Energy Transfer: CuRez Complexes 3.1 Introduction The development of molecular assemblies that can serve as photo- active components of optical systemsl'9 requires exquisite control of their photophysical properties. This level of control, in turn, can only be achieved if the nature of the ground and excited states of the constituents can be elucidated.”12 Polypyridyl complexes of Re], Ru", Os", and IrIII have received considerable attention in this regard due to their synthetic accessibility as well as the structure-function correlations that have emerged from several decades of research on their photo-induced properties.”"24 As a result, compounds within this class have been utilized in a variety of settings ranging from fundamental studies of electron donor/acceptor interactions”32 33-36 to artificial photosynthetic systems and, most recently, as catalysts for water splitting reactions.37'40 Recent work from our own group focused on energy transfer dynamics in a tetranuclear assembly consisting of three ReI polypyridyl donors surrounding an Fem(acac)3 moiety.“ Emission from the Rel-based Ill 3MLCT excited state was efficiently quenched by the Fe core; dipolar coupling between the ReI fluorophore and absorptions associated with the 109 F eIII center was quantitatively established through a detailed metrical analysis of the relative orientations of the donor/acceptor transition dipoles. Energy transfer rate constants on the order of 109 s'] were observed in this system, driven in part by the charge-transfer nature of the optical transitions involved. It occurred to us that this basic compositional motif could provide a convenient platform for examining additional fundamental aspects of dipolar energy transfer. The notion of coupling to weakly absorbing acceptors was of particular interest in light of other work we have carried out on energy transfer involving spin-coupled di-iron(III) complexes.”43 Ligand-field states are intriguing candidates in this regard, due to their low oscillator strengths (8 ~ 10 — 100 M'lcm") while at the same time representing nearly ideal manifestations of the point-dipole approximation central to Forster theory. Low energy ligand-field states of first-row transition metal complexes have been implicated as energy acceptors in a number of studies involving multinuclear assemblies.“58 Recently, Ford and coworkers have shown that ligand-field states of Cr111 in trans-Cr(cyclam)X2+ (X = C1" or ONO') complexes quench the photoluminescence of CdSe/ZnS core/shell quantum dots (QDs) via a F 6rster energy transfer mechanism.”60 The investigation showed that highly absorbing QDs act as antennae which in turn 110 photosensitize the weakly absorbing Crm-based 4A -—> 4T ligand-field states of an attached Cr(cyclam)(ONO)2+ complex to trigger the release of NO; potential applications of this water soluble system include the delivery of bioactive agents to specific physiological targets. Other recent examples of metal-centered acceptor states come from the work of Ward and coworkers, who have demonstrated significant quenching of MLCT type emission by f-f acceptor states in Ln(III)-containing donor/acceptor systems.“65 The shielding of the 4f electron shell of the Ln(III) ions by the outer core 5s and 5d electrons allow lanthanide coordination complexes to retain atomic-like absorption and emission profiles,66 resulting in line-like spectra that allow for straightforward assignments of acceptor states that energetically overlap with the 3MLCT emission profiles. These studies constitute excellent examples of metal-centered transitions involved in excited-state energy transfer, albeit with some ambiguity as to the mechanism of energy transfer operative in these systems. The clear-cut assignment of a dipolar mechanism in the FeRe3 series mentioned above led us to design structurally analogous complexes that contain ligand-field states possessing favorable spectral overlap with the Rel-based 3MLCT emission spectra. The current study thus replaces the 6LMCT acceptor states of the FeRe3 series“ with ligand-field (d-d) states in an effort to quantitatively explore energy transfer 111 using an acceptor that represents a nearly ideal manifestation of the point- dipole approximation central to Féirster theory. The synthesis, structure, and photophysical properties of a new family of trinuclear CuRez chromophore-quencher complexes having the general form [Cu(pyacac)2(Re(bpy’)(CO)3)2](OT1)2 (where pyacac = 3-(4-pyridyl)- acetylacetonate and bpy’ = 4,4’-5,5’-tetramethyl-2,2’-bipyridine (tmb, 1), 4,4’-dimethyl-2,2’-bipyridine (dmb, 2), 2,2’-bipyridine (bpy, 3), 4,4’- dichloro-2,2’-bipyridine (dclb, 4), and 4,4’-diethylester—2,2’-bipyridine (deeb, 5), and OTf = CF3SO3') are reported. The CuII metal centers are covalently attached to two fac-Re(bpy’)(CO)3 (bpy’ = tmb, dmb, bpy, dclb, and deeb) moieties through two pyridyl-acetylacetonate bridging ligands (Figure 3-1). Variations in the substituents on the bipyridyl groups allowed for systematic energetic tuning of the emission profile of the Rel-based 3MLCT state relative to the absorption profile of the CuII center. Emission from the ReI fluorophores in complexes 1, 2, 3, 4, and 5 was found to be significantly quenched in the presence of the CuII metal center relative to structurally analogous BeRez analogs. The BeRez model systems (Figure 3- 1) with the general formula [Be(pyacac)2(Re(tmb)(CO)3)2](OTf)2 (6), [Be(PyacaC)2(Re(dmb)(C0)3)2](OTfiz (7), [136(PyacaC)2(Re(bPY)(C0)3)2]- (0102 (8), [136(Pyaca0)2(Re(d01b)(CO)3)2](0T02 (9), and [36(PyacaC)2(Re- 112 (deeb)(CO)3)2](OTf)2 (10) were synthesized in order to investigate the excited-state relaxation kinetics of the Re(bpy’)(CO)3 moieties in the absence of the emission quenching dynamics incurred by the CuII metal center. The confluence of data acquired on the CuRez systems (donor- acceptor separation, redox properties, and spectral overlap analyses) unequivocally establishes dipolar energy transfer as the dominant quenching mechanism. Time-dependent density functional theory (TD-DFT) calculations on the Cu11 core - Cu(phacac)2 (phacac = 3-pheny1- acetylacetonate) - in a dichloromethane solvent bath revealed the orbital composition and transition energy of the Cu11 ligand-field acceptor states. The unique electronic properties of the Cu11 acceptor coupled with the computational work have allowed identification of an orbitally-specific pathway for energy transfer, a result that illustrates the role of relative transition dipole orientation operating at the atomic-orbital level. 113 " ‘0 [1,, m“ 0' ‘ — ’4’" / y; ‘ H-co or? co oc co c‘ .2 -Re— cl 8 —N N R2 \/ R1 R, R1 R1721" oc go a; N— oc— —12e— —N \ <"'O\Be-““‘°"‘> _N—1Re/—co Ia, — “'O/ ‘0" \ ’ s'\ —N /N_ 00 co R1 R1 1 (Cu) and 6 (Be): R, = R2 = CH3 2 (Cu) and 7 (Be): R, = CH3, R2 = H 3 (Cu) and 8 (Be): R, = R2 = H 4 (Cu) and 9 (Be): R, = Cl, R2 = H 5 (Cu) and 10 (Be): R, = COzEt, R2 = H Figure 3-1. Structures of the CuRez and BeRez multinuclear assemblies. 114 3.2 Experimental Section 3.2.1 Synthesis and Characterization General. All solvents used were purified and dried according to previously reported methods.67 Spectroscopic grade CHzClz was used for all photophysical measurements and was dried under CaHz reflux until no water was detected by 1H NMR. Solvents for both steady-state and time-resolved emission measurements were thoroughly degassed using freeze-pump-thaw techniques. 3-(4-pyridyl)-2,4-pentanedione,68 Cu(pyacac)2,69 Cu(phacac)2,7O Be(pyacac)2,7' Re(tmb)(CO)3(OTf),72 Re(dmb)(CO)3(OTf),72 Re(bpy)(CO)3- (orr),72 Re(dclb)(CO)3(OTf),72 Re(deeb)(CO)3(OTf),72 and fac- [Re(bpy)(CO)3(4-Etpy)](PF6)73 (4-Etpy = 4-ethylpyridine) were prepared according to literature procedures. 3-phenyl-2,4-pentanedione was purchased fiom TCI America. Elemental analyses and F T-IR data were obtained through the analytical facilities at Michigan State University; mass spectra were obtained through the analytical facilities at The University of South Carolina. [Cu(pyacac)2(Re(tmb)(CO)3)2](0T1); (1). An amount of 99 mg (0.24 mmol) of Cu(pyacac)2 was dissolved in 40 mL of hot THF, after which 300 mg (0.475 mmol) of Re(tmb)(CO)3(OTf) was added and the solution was purged with argon for 20 min. The reaction mixture was then fit with a 115 condenser and stirred under argon for 24 hours in hot THF in the dark, after which time a blue-green solid formed. The precipitate was collected, washed with dry THF, and recrystallized from acetonitrile/ether. Yield: 129 mg (32%). Anal Calcd for C56H52N6F6OI6SZCuRe2: C, 40.06; H, 3.12; N, 5.00. Found: C, 40.19; H, 3.30; N, 4.79. IR (KBr, cm'l): 2030 s, 1925 s, 1902 3,1614 m, 1575 s, 1417 m, 1261m, 1157 m, 1032 s, 639 m. MS: [ESI+, m/z (rel. int.)]: 690.5 (62) {[Cu(pyacac)2(Re(tmb)(CO)3)2]}2+, 1530.1 (2) {[Cu21<0To}“1 [Be(pyacac)2(Re(tmb)(CO)3)2](OTf); (6). An amount of 67 mg (0.19 mmol) of Be(pyacac)2 was dissolved in 40 mL of hot THF, after which 233 mg (0.369 mmol) of Re(tmb)(CO)3(OTf) was added and the solution was purged with argon for 20 min. The reaction mixture was then fit with a condenser and stirred under argon in hot THF for 24 hours in the dark. The volume of the mixture was reduced to ~10 mL, and hexanes were slowly added to precipitate a yellow solid. The yellow solid was collected, washed with dry THF, and recrystallized from acetonitrile/ether. X-ray quality crystals were obtained from slow diffusion of ether into an acetonitrile solution of the compound. Yield: 135 mg (45%). Anal Calcd for C56HszN6F6OI6S2BeRe2: C, 41.40; H, 3.23; N, 5.17. Found: C, 41.53; H, 3.44; N, 5.00. IR (KBr, cm“): 2031 s, 1916 3,1613 m, 1575 m, 1417 m, 1261 m, 1155 m, 1031 m, 847 m, 638 m. MS: [ESI+, m/z (rel. int.)]: 663.2 (86) {[Be(pyacac)2(Re(tmb)(CO)3)2]}2+, 1475.5 (2) {[Be(pyacac)2(Re(tmb)- (C0)3)2](OTf)} ”- [Be(pyacac)2(Re(dmb)(CO)3)2](OTf); (7). An amount of 69 mg (0.19 mmol) of Be(pyacac)2 was dissolved in 40 mL of hot THF, after which 227 mg (0.3 82 mmol) of Re(dmb)(CO)3(OTt) was added and the solution 119 was purged with argon for 20 min. The reaction mixture was then fit with a condenser and stirred under argon in hot THF for 24 hours in the dark. The volume of the mixture was reduced to ~10 mL, and hexanes were slowly added to precipitate a yellow solid. The yellow solid was collected, washed with dry THF, and recrystallized from acetonitrile/ether. Yield: 123 mg (41%). Anal Calcd for C52H44N6F6016S2BeRe2: C, 39.82; H, 2.83; N, 5.36. Found: C, 40.00; H, 2.94; N, 5.47. IR (KBr, cm'l): 2031 s, 1915 s, 1620 m, 1576 s, 1418 m, 1260 m, 1159 m, 1031 s, 847 m, 638 m. MS: [ESI+, m/z (rel. int.)]: 635.2 (100) {[Be(pyacac)2(Re(dmb)(CO)3)2]}2+, 1419.4 (2) {[Be(pyacac)2(Re(dmb)(CO)3)2](0T1)1“i [Be(pyacac)2(Re(bpy)(CO)3)2](OTf); (8). An amount of 72 mg (0.20 mmol) of Be(pyacac)2 was dissolved in 40 mL of hot THF, after which 228 mg (0.396 mmol) of Re(bpy)(CO)3(OTf) was added and the solution was purged with argon for 20 min. The reaction mixture was then fit with a condenser and stirred under argon in hot THF for 24 hours in the dark, after which time a crystalline yellow solid formed in the yellow reaction solution. The mixture was allowed to cool to room temperature and the yellow precipitate and yellow solution were separated by filtration. The precipitate was collected, washed with dry THF, and recrystallized from acetonitrile/ether. ESI-MS and single crystal X-ray analysis of the yellow 120 crystalline material revealed the formation of a Be3Re3 hexamer that contains a previously reported 36303 core”76 that will be discussed in section 3.3.1. The volume of the yellow reaction solution was reduced to ~10 mL, and hexanes were added to produce additional yellow solid. The solid was collected, washed with dry THF, and recrystallized from acetonitrile/ether. Yield: 31 mg (10%). Anal Calcd for C43H36N6F6016S2- BeRez: C, 38.12; H, 2.40; N, 5.56. Found: C, 38.25; H, 2.61; N, 5.39. IR (KBr, cm'l): 2033 s,1919 s, 1604 m, 1576 s, 1419m, 1261 m, 1161 m, 1031 m, 849 m, 771 m, 636 m. MS: [ESI+, m/z (rel. int.)]: 607.1 (83) {1Be(pyacac)2(Re(bpy)(C0)3>2112+ [Be(pyacac)2(Re(dclb)(CO)3)2](OTf); (9). An amount of 66 mg (0.18 mmol) of Be(pyacac)2 was dissolved in 40 mL of hot THF, after which 234 mg (0.363 mmol) of Re(dclb)(CO)3(OTf) was added and the solution was purged with argon for 20 min. The reaction mixture was then fit with a condenser and stirred under argon in hot THF for 24 hours in the dark, after which time a dark yellow solid formed. The mixture was allowed to cool to room temperature and hexanes were slowly added to precipitate additional solid. The dark yellow solid was collected, washed with dry THF, and recrystallized from acetonitrile/ ether. Yield: 199 mg (66%). Anal Calcd for C43H32N6C14F6OI6S2BeRe2: C, 34.94; H, 1.95; N, 5.09. Found: C, 35.06; H, 121 2.22; N, 4.81. IR (KBr, cm'l): 2035 s, 1919 s, 1575 s, 1467 m, 1258 s, 1158 m, 1030 m, 846 m, 755 w, 638 111. MS: [ESI+, m/z (rel. int.)]: 676.0 (100) {[Be(pyacac)2(Re(dc1b)(CO)3)2]}2+, 1501.1 (1) {[Be(pyacac)2(Re(dclb)- (C0)3)2](OTf)} ”- [Be(pyacac)2(Re(deeb)(CO)3)2](OTf); (10). An amount of 60 mg (0.17 mmol) of Be(pyacac)2 was dissolved in 40 mL of hot THF, after which 240 mg (0.334 mmol) of Re(deeb)(CO)3(OTi) was added and the solution was purged with argon for 20 min. The reaction mixture was then fit with a condenser and heated under argon for 24 hours in darkness. The volume of the mixture was reduced to ~10 mL, and hexanes were slowly added to precipitate an orange solid. The orange solid was collected, washed with dry THF, and recrystallized using acetonitrile/ether. Yield: 136 mg (45%). Anal Calcd for C60H52N6F6OZ4S2BeRe2: C, 40.02; H, 2.91; N, 4.67. Found: C, 39.63; H, 2.92; N, 4.45. IR (KBr, cm-l): 2035 s, 1925 s, 1732 m, 1576 m, 1464 m, 1323 m, 1263 s, 1155 m, 1031 m, 849 m, 767 m, 638 m. MS: [ESI+, m/z (rel. int.)]: 751.3 (79) {[Be(pyacac)2(Re(deeb)(CO)3)2]}2+, 1651.6 (2) {[BetpyacaoxRemeebxc013>21(0T01‘fi 3.2.2 Physical Measurements X-ray Structure Determinations. Single-crystal X-ray diffraction data for complexes 3 and 6 were acquired at the X-ray facility of Michigan 122 State University. Diffraction data were collected on a Siemens SMART diffractometer with graphite-monochromatic Mo K01 radiation (A. = 0.71073A). Data were collected at -100 °C by using an Oxford Cryosystems low temperature device. Crystallographic data are summarized in Table 1; selected bond distances and angles are listed in Table 2. Lattice parameters were obtained fi'om least-squares analyses and data were integrated with the program SAINT.77 The integration method employed a three dimensional profiling algorithm and all data were corrected for Lorentz and polarization factors, as well as for crystal decay effects. The absorption correction program SADABS78 was employed to correct the data for absorption effects. The structures were solved by direct methods and expanded using Fourier techniques. All structure calculations were performed with the SHELXTL 6.12 software package.79 Anisotropic thermal parameters were refined for all non-hydrogen atoms. Hydrogen atoms were localized in their calculation positions and refined by using the riding model. Further details concerning the structure determinations may be found in Supporting Information. Cyclic Voltammetry. Electrochemical measurements were carried out in a Nz-filled drybox (Vacuum Atmospheres) using a BAS CV-SOW electrochemical analyzer. A standard three-electrode arrangement was utilized consisting of a Pt working electrode, graphite counter electrode, and 123 a Ag/AgNO3 reference electrode. Measurements were carried out in CH3CN solution that was 0.1 M in NBu4PF6. Potentials are reported versus the ferrocene/ferrocenium couple, which was used as an internal standard. Electronic Absorption and Steady-State Emission Spectroscopies. Extinction coefficients for all compounds were acquired in room- temperature CHzClz solution using a Varian Cary 50 UV-Visible spectrophotometer. Steady-state emission spectra were acquired using a Spex Fluoromax fluorimeter and corrected for instrumental response using a NIST standard of spectral irradiance (Optronic Laboratories, Inc., OL220M tungsten quartz lamp).80 Spectra were acquired on samples dissolved in thoroughly degassed CHzClz under optically dilute conditions (o.d. ~ 0.1) and sealed under an argon atmosphere in 1 cm path length quartz cuvettes. Radiative quantum yields ((1),) were determined relative to fac- [Re(bpy)(CO)3(4-Etpy)](PF6) ((1)r = 0.18 in CH2C12).73 Quantum yields were calculated according to equation 5-1, _ (I) (lunk /Aunk) (flunk )2 unk — s d t (Istd /Astd) ”std (5-1) where (Dunk and (1),“, are the radiative quantum yields of the sample and the standard, respectively, lunk and 1,,d represent the areas of the corrected emission profiles for the sample and the standard, Am and A5“, are the 124 absorbance values of the sample and the standard at the excitation wavelength, and nlmk and nsrd correspond to the indices of refi'action of the sample and standard solutions (taken to be equal to the neat solvents). Excitation wavelengths were 355 nm for the tmb, dmb, and bpy analogues, 375 nm for the dclb analogues, and 400 nm for the deeb analogues. The corrected excitation spectrum of fac-[Re(bpy)(CO)3(4-Etpy)](PF6) in CH2C12 compared favorably with the compound’s absorption spectrum over the range of wavelengths examined (355 - 400 nm), implying that the radiative quantum yield for fac-[Re(bpy)(CO)3(4-Etpy)](PF6) does not vary significantly over this spectral window; the reported value of (I), = 0.18 was therefore used for determining the radiative quantum yields at 71,, = 355 nm, 375 nm, and 400 nm. Values for the zero-point energy gap (E00) of the Re]- based 3MLCT excited states were determined by fitting emission profiles based on the approach described by Claude and Meyer.81 Wavelength data were converted to energy units employing the correction of Parker and Rees;82 the best fit was determined by visual inspection of the results of a least-squares minimization routine. Time-Resolved Emission Spectroscopy. Nanosecond time-resolved emission data for the BeRez model complexes 6-10 were collected using a NszAG-based laser spectrometer that has been described previously.80 125 Data were acquired at room temperature in thoroughly degassed CHzClz solutions having absorbances of ~0.1 at the excitation wavelengths. Samples were sealed under an argon atmosphere in 1 cm path length quartz cuvettes. The decay traces correspond to an average of 250 shots of the signal probed at the emission maximum of each compound. Picosecond time-resolved emission data for the CuRez complexes 1-5 were collected using a time-correlated single photon counting (TCSPC) apparatus that has been described previously.83 Data were acquired in thoroughly degassed CH2C12 solutions having absorbances of ~0.1 at the excitation wavelength. Samples were sealed under an argon atmosphere in 1 cm path length quartz cuvettes. Each reported decay trace corresponds to a signal average of six data sets, with each data set resulting from ca. 30 minutes of data acquisition time. The decay traces for all five complexes manifest a small baseline offset within the ca. nanosecond data acquisition window due to the presence of a small amount of an emissive impurity identified as fac-[Re(bpy’)(CO)3(pyacac)] based on their similarities to the long-lived kinetics observed for the corresponding BeRe2 model complexes. Decay traces for all five Cu-containing complexes were therefore fit with bi- exponential kinetic models by fixing the long-lived component to the 126 lifetime of the corresponding BeRez complex. Data were fit using the OriginPro 7.5 software package.84 DFT Calculations. The Gaussian 0385 software package was used for all calculations. The nuclear coordinates used for Cu(phacac)2 in the calculations were based on the single-crystal X-ray structure reported by Carmichael et al.70 The geometry optimization was performed with the UB3LYP functional and LANLZDZ basis set. The effect of CH2C12 solvent on the optimization was modeled using the polarizable continuum model (PCM);86'89 no symmetry restrictions were imposed. Time-dependent density functional theory (TD-DF T) calculations were performed with the UB3LYP functional and the 6-311G** basis set, also incorporating CHzClz using the PCM model. All calculations used tight convergence criteria,90 and assumed a molecular charge of 0 and a ground-state spin of S = 1/2. The molecular orbitals were generated using GaussView.” 3.3 Results and Discussion 3.3.1 Synthesis and Structural Characterization. Our interest in developing these systems was to investigate dipole-dipole energy transfer processes in structurally well-defined assemblies that possess weakly absorbing acceptors. The choice of using ReI and CuII was based on the well 127 known MLCT-based reactivity of ReI and the low oscillator strength absorptions associated with CuII ligand-field states. The utilization of the M(pyacac)2 core (M = CuII and Be") as a ligand for Re(bpy’)(CO)3(OTf) allowed the pyridyl group to displace the weakly coordinating triflate anion and generate the di-cationic CuRez and BeRez complexes. The formation of the binuclear assembly was facilitated by the low steric crowding afforded by the 180° separation of the two pyacac ligands, along with the propensity of both metal ions to form 4-coordinate complexes. The ESI-MS data for complexes 1-10 in acetonitrile solution are consistent with the formation of the CuRez and BeRez assemblies. For example, complex 3 shows peaks corresponding to [Cu(pyacac)2(Re(bpy)- (CO)3)2]2+ and [Cu(pyacac)2(Re(bpy)(CO)3)2](OTf)1+. X-ray quality crystals for complexes 3 and 6 were generated by diffusion of ether into an acetonitrile solution of the complexes over the course of approximately one week; this in turn provides additional evidence for the general robustness of complexes 1-10 in solution. As mentioned in the synthetic procedure of [Be(pyacac)2(Re(bpy)- (CO)3)2](OTf)2 (8) (section 3.2.1), a crystalline byproduct consisting of a Be3Re3 hexamer was formed during the course of the reaction. The Be3Re3 complex was confirmed by elemental analysis, ESI-MS, and single-crystal 128 X-ray structural analysis. Yield: 39 mg (15%). Anal Calcd for C72H57N9F9- OZ7S3Be3Re3: C, 37.06; H, 2.46; N, 5.40. Found: C, 36.67; H, 2.79; N, 5.21. MS: [ESI+, m/z (rel. int.)]: 628.6 (100) [(Be3(OH)3(pyacac)3)((Re(bpy)- (C0)3)3)l3+, 1017-5 (20) [(Bea(0H)3(PyacaC)3)((Re(bpy)(C0)3)3)](0102+. The structure of the complex is given in Figure 3-2, and shows the core of the complex consisting of three Bell ions bridged by three hydroxo groups that form the Be3O3 core of the molecule, along with the Be3O3 core connected through three pyacac ligands to three Re(bpy)(CO)3 moieties that contain pseudo-C3,, symmetry commonly observed in ReN3C3 coordination environments. Synthesis procedures and single-crystal X-ray structural data for derivatives of the Be3O3 core - Be3(OH)3L3 (L = bis-chelate complex) - complexes have been reported in the literature.74 The geometry of the Be3O3 core and the Be—O bond distances of [(Be3(OH)3(pyacac)3)((Re(bpy)- (CO)3)3)](OTf)3 are very similar to these reported systems. The first steps in the synthesis of the Be(pyacac)2 starting material needed to synthesize complex 8 is to make an aqueous solution of pyacac and pyridine and then add Be(SO4)2 to the alkaline mixture, which could possibly generate both Be(pyacac)2 and the Be3(OH)3(pyacac)3 complex needed to form the [Be3Re3](OTf)3 system. ' Direct probe MS measurements on solid Be(pyacac)2 and in THF solution exhibited only m/z = 361.4 expected for 129 Be(pyacac)2, and not m/z = 606.6 expected for the Be3(OH)3(pyacac)3 complex, which rules out the possibility of Be3(OH)3(pyacac)3 being formed during the synthesis of Be(pyacac)2. In addition, the formation of the crystalline Be3Re3 hexamer was only observed during the synthesis of complex 8 and not for the four other BeRe2 compounds (6, 7, 9, and 10). This behavior suggests the most likely reason for the formation of the [(Be3(OH)3(pyacac)3)((Re(bpy)(CO)3)3)](OTf)3 cystalline material is due to the THF solubility differences between this complex and the four other possible Be3Re3 derivative comprised of tmb, dmb, dclb, or deeb. The four other Be3Re3 complexes were not observed in the ESI-MS analysis, which suggests that either they were never formed in the reaction solutions of complexes 6, 7, 9, and 10 or were removed during the purification (recrystallization) processes. 130 Figure 3-2. Drawing of the cation [(Be3(OH)3(pyacac)3)((Re(bpy)- (CO)3)3)](OT1)3 obtained from single-crystal X-ray structure determinations. 3.3.2 Single-Crystal X-ray Structures. X-ray quality crystals of [CU(Pya080)2(Re(bPY)(C0)s)2](0T02 (3) and [Be(pyacaC)2(Re(tmb)(C0)3)2]- (OTDZ (6) were obtained following diffusion of ether into acetonitrile solutions of the complexes. The complexes both crystallize in the monoclinic space group P2(1)/c. Crystallographic details are given in Table 3—1 with selected bond distances and angles for the two complexes listed in Table 3-2. The coordination environment about the central metal ion in complex 3 (Figure 3-3, Top) is formed from the four oxygen atoms of the 131 acac ligands. The Cu—O bond distances of ca. 1.917 :t 0.002 A and a 0.0 A deviation of the Cu" ion from the O4 mean-plane are both consistent with pseudo-D2,, square-planar CuII and compare favorably with other structurally characterized examples of Cull—acac systems.69’70’92’93 The structure of complex 3 also shows Cu-—-O distances of 2.702 and 2.545 A corresponding to axial interactions of the CF3SO3' counter ions with the Cu11 core. The axial Cu---O distances are notably longer than typically observed for tetratragonally distorted CuO6 complexes such as [Cu(HzO)6]2+,94'96 suggesting that the two Cu—--O associations are at best weakly covalent metal-ligand interactions; spectroscopic evidence indicates that these interactions are not retained in solution (Vida infra). As expected, significantly shorter metal-oxygen bonds (ca. 0.3 A) are observed for the tetrahedral coordination environment of complex 6 (Figure 3-3, Bottom) due to the smaller ionic radius of Be".71 The internal geometries of the Re(bpy)(CO)3 and Re(tmb)(CO)3 moieties in complexes 3 and 6 exhibit the pseudo-C3,, coordination environment common to Re1 complexes in this Classm"00 and are also insensitive to changes in the substituents of the p01ypyridyl ligand. Differences are noted in the Rel-«MII distances, with cOrnplex 6 being uniformly shorter by ~0.3 A, which is again due to the SII‘laller BeII ion and the concomitant decrease in metal-oxygen bond lengths 132 relative to complex 3. Despite these minor differences there is considerable structural homology between the two complexes, underscoring the appropriateness of using BeII as a structurally and electronically benign replacement for Cu11 in these systems. 133 Table 3-1. Crystallographic Data for [Cu(pyacac)2(Re(bpy)(CO)3)2](OTf)2 (3) and [Bemyacac)2(Re(tmb)(C0)3)2l(0T1)2 (6)- 3 6 formula C43H36N6F6016SZCUR62 C 56H52N6F 601682BCR€2 MW 1566.9 1624.6 cryst syst Monoclinic Monoclinic space group P2(1)/c P2(1)/c T/K 173 (2) 173 (2) a/A 10.3213(12) l8.7639(3) b/A 24.532(3) 16.9382(3) c/A 12.1608(14) 21.0542(4) a/° 9O 90 B/° 104.027(2) 106.241(1) 'y/° 90 90 V/A3 2987.3(6) 6424.5(2) Z 2 4 Dc/g cm" 1.833 1.696 20max 50 136.4 reflns measured 291 10 72385 independent reflns 5253 1 1601 observed reflns 4827 10856 M(Mo Koo/cm" 4.556 8.63 Rim 0.026 0.041 R1 3 0.019 0.048 wR2b 0.0443 0.132 GDP 1 .005 1 .050 aR1 = ZIIFOI- |F,||/2|F,|. wa2 = [Zw(F02 — F.2)2/2w(F,2)2]“2, w = 14620202) + (21>)2 + bP], where P = [F02 + 2F,2]/3. 134 Table 3-2. Selected Bond Distances (A) and Angles (deg) for [Cu(pyacac)2- (Re(bpy)(C0)3)2l(OTf)2 (3) and [B6(pyacaC)2(Re(tmb)(C0)3)2l- (OTflz (6)- 3 6 Bond Distances (A) Cu(1)—0(1) 1.920(2) Be(1)—O(1) 1.618(9) Cu(1)—0(2) 1.914(2) Be(1)—O(2) 1.590(9) Cu(1)—O(1A) 1.920(2) Be(1)—O(6) 1.601(9) Cu(1)—O(2A) 1.914(2) Be(1)—O(7) 1.610(10) Re(1)-—N(1) 2.168(2) Re(1)—N(1) 2.220(5) Re(1)—N(2) 2.170(2) Re(1)—N(2) 2.156(6) Re(1)—N(3) 2.214(2) Re(1)—N(3) 2.161(5) Re(1)—C(22) 1.918(4) Re(1)—C(25) 1.921(7) Re(1)—C(23) 1.927(4) Re(1)—C(26) 1.922(7) Re(1)—C(24) 1.915(4) Re(1)—C(27) 1.913(7) Cu(1) 0(8) 2.702 Cu(1) O(8A) 2.545 Cu(l)°"Re(1) 9.707 Be(1)°“Re(1) 9.466 Cu(1)-“Re(1A) 9.707 Be(1)°-'Re(2) 9.416 Bond Angles (deg) O(1)—Cu(1)—O(2) 92.0(8) O(1)—Be(1)—O(2) 105.6(5) O(1)—Cu(1)—O( 1 A) 180.0(13) O(1)—Be(1)—O(6) 108.8(6) N(1)—Re(1)—N(2) 7502(9) N(2)—Re(1)—N(3) 75.3(2) C(22)—Re(1)—N(3) 95.22(11) C(26)—Re(l)—N(l) 92.0(2) C(23)—Re(1)—N(3) 175.8(5) C(25)—Re( 1 )—N( 1) 176.9(3) aplane 1'°°plane 2 87.414 cplane 1°"plane 2 60.448 bplane 1mp1ane 2 87.414 dplane 1-«p1ane 2 75.75 aPlane 1 is defined by atoms 0(1), 0(2), C(1), C(2), C(3), C(4), C(5); plane 2 is defined by atoms N(3), C(6), C(7), C(8), C(9), C(10). bGiven by the atoms that define the two planes in the adjacent pyacac ligand. °Plane 1 is defined by atoms 0(1), 0(2), C(1), C(2), C(3), C(4), C(5); plane 2 is defined by atoms N(1), C(6), C(7), C(8), C(9), C(10). dGiven by the atoms that define the two planes in the adjacent pyacac ligand. 135 . ' . 02A 1 O1 . 024 . u \ R 1A - ‘- ' .9 fi N3A ’ fl ‘ N3 '. v Re1 023 01A 02 , . :. Figure 3-3. Drawings of [Cu(pyacac)2(Re(bpy)(CO)3)2](OTf)2 (3, Top) and [Be(pyacac)2(Re(tmb)(CO)3)2](OTDZ (6, Bottom) obtained from single-crystal X-ray structure determinations. Atoms are represented as 50% probability thermal ellipsoids. 136 3.3.3 Electronic Absorption Spectroscopy. The electronic absorption spectra of complexes 1-10 were acquired in room-temperature CH2C12 solution and are given in Figure 3-4. Rel polypyridyl complexes exhibit a 1A, —> lMLCT (t2g —-> 7:1 (bpy’)) transition in the range of 330 to 430 nm depending on the substituents of the bpy’ ligand,“102 with electron donating and withdrawing groups shifting Xmax for this feature to the blue and red, respectively. In the present series, the absorption maximum systematically shifts from ca. 350 nm for the tmb-containing complexes (1 and 6) to ca. 400 nm in the Rel-deeb analogues (5 and 10). The presence of charge transfer transitions associated with the CulI core in the ultraviolet (Figure 3-5, Left) gives rise to differences in this region between these adducts and their corresponding Be'l analogues. Nevertheless, the general similarities among these spectra, particularly with regard to the 1A, —-> 1MLCT absorption of the Rel-bpy’ chromophores, is suggestive of minimal electronic coupling between these groups and the CuII center in complexes 1-5. In addition to the Rel-based charge-transfer transitions, the CuRez complexes also possess ligand-field transitions associated with the square- planar CuO4 core. The low-energy tails of the charge-transfer transitions associated with the Re1 chromophores largely obscures these absorptions in compounds 1-10, however, the electronic absorption spectrum of 137 Cu(phacac); allows for an examination of these mid-visible d-d bands (Figure 35, Right). The two peaks at 530 nm (e = 47 M'lcm'l) and 652 nm (8 = 50 M'lcm'l) arise fiom four spin-allowed ligand-field transitions corresponding to dxz—> dxy, dyz—> dxy, dzz—> dxy and dx2 _y2—> dxy based on the polarized crystal spectrum of Cu(phacac)2103 and the coordinate system shown in Figure 3-5 (Right). This absorption spectrum, which accurately reflects the optical properties of the CuHOa core of complexes 1-5, will therefore provide the basis for the spectral overlap analysis to be described later. It should be noted that the structure of complex 3 shown in Figure 3-3 reveals two CF3SO3' counter ions interacting with the Cu11 core that could in principle yield 5- or 6-coordinate CuII metal centers for the five CuRez complexes (1-5) in solution. These interactions would be expected to alter the absorption profile of the CuII core,104 thereby negating the utility of the Cu(phacac); spectrum as a surrogate for the acceptor in the CuRez series. Systematic titration of a CHzClz solution of Cu(phacac); with (NEt4)(CF3SO3) (2 - 40 eq.) did not cause any discemable change in the absorption spectrum of Cu(phacac); (Figure 3-6). This result indicates that the triflate group is not associating with the CuII center in solution, thereby 138 validating the use of the Cu(phacac)2 ground-state absorption spectrum in spectral overlap analyses of the CuRez assemblies. 50000 40000 300004 20000- 10000 60000- 50000- 4000O~ 30000~ 20000- 10000- Molar Absorptivity (M 10m!) Molar Absorptivity (M'lcm'l) 300 400 500 600 Wavelength (nm) Figure 3-4. Electronic absorption spectra of [M(pyacac)2(Re(bpy’)- (CO)3)2](OT1)2 where, M = Cu" (red traces) and Be'1 (blue traces). All spectra were acquired in CHzClz solution at room temperature. A. [Cu(pyacac)2(Re(tmb)(CO)3)2](OTf)2 (l) and [BeQayacac)2(Re(tmb)- (CO)3)2](OT1)2 (6). B. [Cu(pyacac)2(Re(dmb)(CO)3)2](OTf)2 (2) and [BC(PyacaC)2(Re(dmb)(C0)3)2](OTflz (7)- C- [CUQDyacaC)2(R6(bPY)- (C013)2](OTf)2 (3) and [BC(pyaca0)2(Re(bp>')(C0)3)2](0102 (3) D- [Cu(pyacac)2(Re(dclb)(CO)3)2](OTf)2 (4) and [Be(pyacac)2(Re(dclb)- (C0)3)2](OT1)2 (9). E. [Cu(pyacac)2(Re(deeb)(CO)3)2](OTf)2 (5) and [Re(py acac)2(Re(deeb)(CO)3)2](OTDZ (10). 139 Molar Absorptivity (M'lcm'l) Molar Absorptivity (M"cm") Molar Absorptivity (M'cm'l) Figure 3-4 (Cont’d). 50000 40000 30000 20000 10000 60000 50000 40000 30000 20000 10000 60000 50000 40000 30000 20000 10000 300 400 500 600 Wavelength (nm) 140 “‘5 r" 60- v 0 E “'3 20000 _0 50 ‘ 3‘ E 40) °\c-/ 1‘ E 15000 g. o/ \ ‘é- ~55 30 8 10000- g: “g g 20~ 5 5000 B 10_ O .—r 2 0" 1 1 1 1 1 x g 0L 1 1 1 J_ 1 ‘ 300 350 400 450 500 550 500 600 700 800 900 Wavelength (nm) Wavelength (nm) Figure 3-5. Left. Electronic absorption spectrum of Cu(phacac)2 showing the higher energy charge transfer and organic-based transitions. Right. Electronic absorption spectrum of a concentrated solution of Cu(phacac)2 showing the two mid-visible ligand-field bands. Both spectra were acquired in room-temperature CHzClz solution. 0.07 ~ 0.06 0.05 0.04 ~ 0.03 0.02 - 0.0] ~ 0.00 - Absorbance 500 600 700 800 900 Wavelength (nm) Figure 3-6. Systematic titration of a CHzClz solution of Cu(phacac); with (NEt4)(CF3SO3) (2 - 40 eq.) 3.3.4 Steady-State and Time-Resolved Emission. Emission spectra for the CuRez and BeRez complexes were obtained in room-temperature deoxygenated CH2C12 solutions and are shown in Figure 3-7. The spectral 141 profiles for all of the compounds correspond well to previously reported photophysical studies of ReI polypyridyl systems, with the emission originating fi'om the 3MLCT —> 1A, transition.'05 Emission maxima for the tmb (524 nm), dmb (550 nm), bpy (568 nm), dclb (600 nm), and deeb (620 nm) derivatives reflect the expected trend in zero-point energy of the 3MLCT state based on the electron donating and withdrawing behavior of the substituents on the polypyridyl ligands. Likewise, the radiative quantum yields ((1),) for complexes 6-10 (Table 3-3) are all comparable to the reported values for the corresponding mononuclear ReI polypyridyl derivatives.72 The data on complexes 1-5 reveal that the emission intensities for the Cull-containing compounds are significantly attenuated compared to the BeRez model complexes. Unfortunately, radiative quantum yields for complexes 1-5 were found to be analytically unreliable due to a small amount of an emissive impurity present in solutions of these compounds. Cu(phacac); was determined to be non-emissive in room-temperature CHZCIZ solution: the impurity is therefore likely due to the presence of fac- [Re(bpy’)(CO)3(pyacac)] generated by displacement of the pyacac ligand by residual amounts of H20 contained in the CH2C12. The drying and distilling procedures that were employed were exhausted until no water was detectable by 1H NMR, but the low concentration of the CuRez compound 142 used for the emission measurements means that even trace amounts of H20 could be sufficient to generate a small amount of dissociated species. A similar problem was encountered in our previous study of FeRe3 assemblies;4 as in that system, the presence of this emissive component proves to be inconsequential for the ensuing analysis.106 143 1 A :2? VJ 8 a H S: .2 (I) .E :2 LL] n V A I 1 B r? m 3': 8 :2 1—4 x: .2 m .59. E LL] 0 500 600 700 800 Wavelength (nm) Figure 3-7. Corrected steady-state emission spectra for [M(pyacac)2- (Re(bpy’)(CO)3)2](OTf)2 where, M = CuIl (red traces) and Be11 (blue traces). All spectra were acquired in room-temperature deoxygenated CHzClz solution. A. [Cu(pyacac)2(Re(tmb)(CO)3)2](OTDZ (1) and [Be- (pyacac)2(Re(tmb)(CO)3)2](OTf); (6). B. [Cu(pyacac)2(Re(dmb)- (C0)3)2](0Tf)2 (2) and [BC(PyacaC)2(Re(dmb)(C0)3)2](OTflz (7)- C- [CU(PyacaC)2(Re(bPY)(C0)3)2](OTf)2 (3) and [Be(PyacaC)2(Re(bPY)- (CO)3)2](OT1)2 (8). D. [Cu(pyacac)2(Re(dclb)(CO)3)2](OTf)2 (4) and [Be(pyacac)2(Re(dclb)(CO)3)2](OTf)2 (9). E. [Cu(pyacac)2(Re(deeb)- (C0)3)2](0T1)2 (5) and [Be(PyacaC)2(Re(deeb)(C0)3)2](0T1)2 (10) The spectra were acquired following excitation at 355 nm (complexes 1, 2, 3, 6, 7, and 8), 375 nm (complexes 4 and 9), and 400 nm (complexes 5 and 10). For each plot, the emission profiles have been normalized with respect to the absorbance of each sample at its excitation wavelength. The relative magnitudes of the signals are therefore accurate representations of their relative intensities. The emission corresponding to the CuReg analogues are due largely to the presence of Re(bpy’)(CO)3(pyacac) in solution. See main text for further details. 144 Figure 3-7 (cont’d) Emission Intensity C ~ I Emission Intensity .2 {I} c: B E. c: .2 .2 E Lu 0 l _L 1 L 500 600 700 800 Wavelength (nm) Quantitative information concerning emission quenching by the Cu" center in complexes 1-5 was obtained via time-resolved emission spectroscopy. Data for the BeRez model complexes could be fit to single-exponential decay kinetic models; emission traces and values for the observed decay 145 rates for all five BeRez complexes are given on the left side of Figure 3-8 and Table 3-3, respectively. As with the quantum yields, the observed excited-state lifetimes are consistent with an assignment of 3MLCT —+ 1A, emission.”107 The kinetics reveal that the reduction in quantum yield across the series is due primarily to an increase in the non-radiative decay rate for 3MLCT relaxation (km) as opposed to significant variations in radiative coupling to the ground state. The only deviation from this trend is the smaller value of km for complex 10. This is most likely due to the additional conjugation present in the 71 system of the deeb ligand, which allows for increased delocalization of the excited electron relative to other members of the series. This increased delocalization is expected to give rise to a smaller net displacement of the ground- and excited-state potential energy surfaces, which in turn leads to a decrease in the rate of non-radiative decay.73 ’80’81’108 Analogous measurements on the nanosecond time-scale for the CuRez complexes failed to reveal any signal beyond that of the trace impurity mentioned above. Consequently, time-correlated single-photon counting (TCSPC) was employed to measure the excited-state lifetime of the Cu"- containing complexes (1-5). A plot of the TCSPC data obtained for complexes 1-5 in deoxygenated CHzClz solution is shown on the right side of Figure 3-8; data for all five CuRez complexes (1-5) are listed in Table 3-3. 146 The observed time constants for excited-state decay for complexes 1-5 are all significantly larger than the corresponding BeRez model complexes, ranging from a factor of ca. 20 in the case of complex 4 to more than two orders of magnitude for complex 1. The signal-to-noise ratio for the TCSPC data is relatively poor owing to a combination of significant quenching of the Rel-based 3MLCT states coupled with radiative rate constants for emission on the order of 105 8". Nevertheless, these observations clearly indicate the presence of a very efficient quenching process stemming from a reaction between the Rel-based 3MLCT excited state and the Cu11 core of the CuRez assemblies. 147 50- 6 .5 ‘ 1 {3,2000-1 A 40- 8, > a» l "—1 I500 ', E 30- E” l G) 'E E 1000—1 I: 20' g 1 (D ‘a ,2 l "" 10» 3, 500~ E ‘ 0 I I; I 1 I I I I m 0 L A; I I L I I l I 0 2 4 6 8 10 12 14 0 10 20 30 4O 50 60 70 80 Time (ns) x103 Time ("8) 40- 7 A1000- 2 U) > 30- V f E a? . v g 600-‘ .‘E‘ 20—; 3 { E E 400- 8 g 1 c: - _ H 10 § 200.4 E 0 1‘4 1 1 I I I l m 0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 0 5 10 15 20 25 30 35 40 Time (ns) x 103 Time (ns) Figure 3-8. Left. Nanosecond time-resolved emission data for the BeRez model complexes in room-temperature deoxygenated CH2C12 solutions: 6 (rob, = 1990 i 100 ns), 7 (robs = 645 :t 30 ns), 8 (rob, = 540 i 30 ns), 9 (robs: 110 i 10 ns), and 10 (robs: 250 :t 20 ns). The solid red lines correspond to fits to single-exponential decay models. The data were acquired by monitoring at the emission maximum of each compound. Right. Time correlated single-photon counting (TCSPC) emission data for the CuRez complexes in room-temperature deoxygenated CHzClz solutions: 1 (1'1: 14.9 i 0.7), 2 (131: 8.1 :t 0.4), 3 (1:1: 8.2 i 0.4), 4 (11: 5.6 i 0.3), and 5 (1,: 5.0 i 0.3). The data were acquired by monitoring at the emission maximum of each compound. The solid red lines correspond to fits to bi-exponential decay models; in each case the second component (12) corresponds to emission from trace amounts of Re(bpy’)(CO)3(pyacac) present in solution. I; was taken to be the emission lifetime of the corresponding BeRez model complex. 148 Intensity (mV) Intensity (mV) Intensity (mV) 40 30 20 4O 30 20 50 40 30 20 Figure 3-8 (cont’d) _ 8 A U) a. 3 .2 U) s: - B l .2 s: _‘ .2 .3. E I I I | I I L I m 0 I I I I I I l I J_I 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 0 5 10 15 20 25 30 35 40 Time (ns) x 103 Time (ns) _ 9 A (I) a. 1‘ 8 .2 (I) a: _ 53 I E. a - .2 .8 B I I I I I l 1 1 LL] 0 I 1 I I l 1 1 l 1 0 100 200 300 400 500 600 700 0 5 10 15 20 25 30 35 40 Time (ns) Time (ns) ' 10 A 100- 5 U) 8 ' v 80- .2 - 8 60- 8 .5 - 1:: 40- .2 - .8 20- E m 0 I I I I I 4| I I L 0 250 500 750 1000 1250 ‘0 5 IO 15 20 25 30 35 40 Time (ns) Time (ns) 149 Table 3-3. Photophysical Data for [M(pyacac)2(Re(bpy’)(CO)3)2](OTf)2 (complexes 1-10). E00 krc knrd,e complex Mum) (cm'l)a (I)r k(,1,s(s")b (X105 5") (><106 s'l) tmb (1) f f f 6.7 r 0.3 x 107 f 67 a 4 tmb (6) 524 19,900 0.63 5.0 :1: 0.3 x 105 3.2 a 0.2 0.19 i .01 dmb(2) f r f 1.21:0.1x108 f 120:1:6 dmb (7) 550 19,200 0.25 1.6 i 0.1 x 106 4.0 :1: 0.2 1.2 a 0.1 bpy(3) f f f 1.2ao.1><108 f 120i6 bpy(8) 568 18,750 0.18 1.9:r0.1><106 3.4102 1.6i0.1 dclb(4) f f f 1.8:t0.1><108 f 180110 dclb (9) 600 17,650 0.03 8.8 a 0.5 x 106 2.6 :1: 0.1 8.5 a 0.4 deeb (5) f f f 2.0 a: 0.1 x 108 f 200 i 10 deeb (10) 620 16,900 0.07 4.0 a 0.2 x 106 2.8 i 0.2 3.7 a 0.2 aZero-point energy difference between 3MLCT excited state and ground state derived from a spectral fitting analysis. bError bars on kob, represent the standard deviation (20) of five independent measurements. ckr = kob,*(I),. dk,,,(CuRe2) = kobs(CuRe2) — k,(BeRe2). ekm(BeRe2) = kobs(BeRe2) — k,(BeRe2). fThis value is not quoted due to the presence of an emissive impurity, but is expected to be identical to the corresponding BeII complex. See text for details. 150 3.4 Mechanistic Considerations 3.4.1 Electron versus Energy Transfer Quenching. Time-resolved emission data, while clearly revealing the presence of an excited-state reaction, provides no information as to the mechanistic origin of that reaction. Indeed, both electron and energy transfer processes can be envisioned to occur out of the Rel-based 3MLCT excited state. Based on the presence of Cu" in the ground states of these CuRe2 complexes, electron transfer (kET) could in principle proceed either by oxidative or reductive quenching to produce Cu1 or Cu!II as photo-products (Figure 3-9). In order to determine if electron transfer is thermodynamically viable in our system, the electrochemical properties for the CuRez as well as the BeRez analogues were measured. Cyclic and differential pulse voltammograms for complexes 1-10 were recorded in CH3CN solution; the data are listed in Table 34.109 The availability of the Bell model complexes considerably simplifies assigning many of the electrochemical properties of all ten complexes due to the redox-inert nature of this ion. Accordingly, the single reduction waves seen for complexes 6-10 can be immediately ascribed to the bipyridyl ligand of the Re moiety in each case. The positive shift in potential across the series is consistent with the more electron withdrawing nature of the substituents as one progresses from the methyls of tmb (6) to the diethylester 151 groups in complex 10. Similarly, the oxidation waves seen for complexes 6- 10 are easily assigned to the Rel/Re" couple. The influence of the bipyridyl substitutents are apparent in these data as well, with the more electron deficient ligand giving rise to the most positive oxidation potential for the Re center. The results are all consistent with what has been observed for complexes of the general form fac-[Re(4,4'-X2-bpy)(CO)3(4-Etpy)](PF6) previously reported in the literature.72 In addition to the Rel-based features, complexes 1-5 should also exhibit redox chemistry associated with the central CuIl ion, which prompted the investigation of the electrochemical properties of Cu(phacac)2. The electrochemical properties of Cu(acac)2 and related complexes in CH3CN solution have been previously reported,”0’m with the Cu11 metal center observed to undergo highly irreversible reductive (Cuu —-> CuI) and oxidative (CuII —> Cum) processes. Cu(phacac); exhibits a wave at -1.48 V in CH3CN solution which can be assigned to the CuII ——> CuI reduction; this potential shifts to and at -1.73 V in CHzClz. The CuII —> Cum oxidation occurs at +1.31 V in CH3CN and lies outside the solvent window of CHzClz (i.e., > 1.5 V). Although the Cu" potentials are masked in certain cases by the Rel- and polypyridyl-based processes, the assignments reported in Table 3-4 for 152 the CuRez series are validated by their close correspondence to the data acquired on the BeRez systems. The ReI -—> ReII oxidation potentials for complexes 1-5, Cu11 —+ CuI - 11 reductlon and Cu —+ Cu"I oxidation potentials of Cu(phacac)2, along with the zero-point energy gaps of the 3MLCT states (E00) determined from fits of the emission spectra of the corresponding BeRez analogs (Table 3-3) were used to determine the thermodynamic driving force for photoinduced electron transfer.”2’“3 The values of AGET listed in Table 3-4 reveal that electron transfer is significantly uphill for both oxidative and reductive quenching processes in all five CuRez complexes. Significantly, these reactions become more endothermic in CHzClz (the solvent used for the photophysical measurements) due to the shifts in the Cull-based redox processes indicated above. These electrochemical data therefore effectively rule out electron transfer as a viable quenching mechanism for the CuRe2 series. km (bpy')Re"-Cu' ox. (bPY')Re'NCu" —> (bpy")Re"-Cu“ —C red. .- l m kETZ (bpy )Re NCu Figure 3-9. Illustration of oxidative (kg—n) and reductive (kg-r2) electron transfer quenching processes out of the Rel-based 3MLCT excited-state to produce Cu1 or Cu"I as photo-products, respectively. 153 Table 3-4. Electrochemical Data and Calculated Electron Transfer Driving Forces (AGET) for Complexes 1-10. AC‘ETox ACIETred electrochemical potential (V) (eV)a (eV)b complex B... (Re"”)° E... (bpy'w'f tmb (1) +1.30 -1.83 +0.56 +0.67 tmb (6) +1.28 -1.81 -—- --- dmb (2) +1.31 -1.71 +0.66 +0.64 dmb (7) +1 .31 -1.68 --- --- bpy (3) +1.36 -1.61 +0.77 +0.60 bpy (8) +1.34 -1.57 --- --- dclb (4) +1.41 -1.38 +0.95 +0.50 dclb (9) +1.44 -1.34 m --- deeb (5) +1.51 -1.23 +1.14 +0.44 deeb (10) +1.50 -1.19 --- --- on (Cull/"If Ered (Cum) Cu(phacac); +1.31 -1.73d (-1.48)c aCalculated using the AGETox equation given in Ref. 113 and using the Cu" —+ CuI reduction in Cu(phacac)2 in CHzClz as the reduction potential of the acceptor. bCalculated with the AGETm, equation given in Ref. 113 and using the Cu" —> CulIl oxidation in Cu(phacac); in CH3CN as the oxidation potential of the acceptor. cMeasured in CH3CN solution. dMeasured in CH2C12 solution. Based on the analysis presented above, the most likely explanation for 3MLCT quenching in all five of the CuRez assemblies is excited-state energy transfer. The two most important mechanisms are electron superexchange 154 (Dexter)114 and dipole-dipole coupling (Féirster).115 Dexter energy transfer is subject to a distance dependence that falls off as e'2R due to its reliance on orbital overlap. As such, it is usually relegated to covalently linked systems in which the donor and acceptor lie in close proximity and are electronically coupled. F6rster transfer is a through-space mechanism that occurs when the donor emission non-radiatively couples to an absorptive feature of the acceptor. The dipolar nature of this interaction gives rise to a shallower R'6 distance dependence, allowing this mechanism to be operative over much . 1 longer dlstances.l 6 The X-ray structure data for complex 3 shows a Rel-"CuII separation of nearly 10 A, a value that lies at the limit of what is typically considered for an exchange-based process.46" 17" '9 The similarity in the charge-transfer region between the absorption spectra of complexes 1-5 and their respective BeRez analogues, as well as their electrochemical properties, suggests minimal electronic exists coupling between the Re1 and Cu'1 subunits. This point is amplified by the computational results of Meyer and coworkers which reveal that the majority of the amplitude associated with the thermalized 3MLCT wavefunction for complexes of the form fac- [Re(4,4’-X2bpy)(CO)3(4-Etpy)](PF6) (X = CH3, H, and COzEt) is concentrated within the 75 levels of the 4,4’-X2bpy derivatives and less so on the 4-Etpy ligand.120 This situation decreases the possibility of the pyacac 155 bridging ligands in the CuRez series imparting significant electronic coupling between the ReI and Cu" metal centers. This combination of the structural and electronic characteristics of the CuRez series make it unlikely that Dexter-type exchange is playing a dominant role in the energy transfer dynamics of these systems. 3.4.2 Spectral Overlap Analysis: State-Selective Energy Transfer. The classic equation describing the Ffirster energy transfer rate constant (kEnTr) is given in equation 3-3,m _ 90001n(10)x2k,J 5”" 1287:5774NAR6 (3‘3) where K2 is the dipole orientation factor, kr is the radiative rate constant of the donor, J is the spectral overlap integral, 1] is the refractive index of the solvent, NA is Avogadro’s number, and R is the donor-acceptor separation. The spectral overlap integral (J), which essentially quantifies the resonance condition for energy transfer, can be evaluated fiom the spectroscopic properties of the system and is given in equation 3-4, (3-4) where F0 is the normalized emission spectrum of the donor and 8A is the absorption profile of the acceptor in units of molar absorptivity. A plot of the 156 emission spectra of complexes 6 through 10 along with the visible absorption spectrum of Cu(phacac)2 is shown in Figure 3-10. It can be seen that the systematic shift in the emission maximum across the series results in a modulation of the overlap between the donor and the acceptor. As mentioned previously, the absorption spectrum of the CuII acceptor consists of a ligand-to-metal charge transfer feature in the blue/near—ultraviolet region, as well as ligand-field transitions in the mid-visible; the substituent changes on the bpy ligands serve to tune the resonance between the 3MLCT emission of the Rel-bpy’ fluorophore and these various acceptor state(s). _. 1 _‘ Absorbance (normalized) O O (pozneuuou) misuaiul uoissiwg #— N 212 210 1'8 16 14 Energy (x103 cm") Figure 3-10. Overlay of the emission spectra of [Be(pyacac)2(Re(tmb)- (C0)3)2](OTf)2 (6, purple), [136(pyacaC)2(Re(dmb)(C0)3)2](0T02 (7, blue), [BC(Pyaca0)2(Re(bpy)(C0)3)2](0102 (8, green), [Be(PyacaC)2- (Re(dclb)(CO)3)2](OTf)2 (9, orange), and [Be(pyacac)2(Re(deeb)- (CO)3)2](OTf)2 (10, red) with the electronic absorption spectrum of Cu(phacac)2 (black trace). 157 The Forster model defines a proportionality between the spectral overlap integral and the rate of energy transfer subject to variations in the donor-acceptor distance, the intrinsic radiative lifetime of the donor, as well as the relative orientation of the coupled transition dipoles (1(2). In viewing the series of complexes 1 through 5 as a whole, several of these variables can be accounted for explicitly as a means of examining trends in the rate of energy transfer. The availability of the BeRe2 model complexes affords values of kr for each compound in the series. We can therefore write an expression for the rate of energy transfer normalized for the rate of radiative decay as equation 5a: %£=C'K2'J (3-5a) r where C represents the collection of constants from the Féirster equation (including the R'6 donor-acceptor distance term). Our previous study on Fem-containing assemblies“ demonstrated that one could use metrical data afforded fiom single-crystal X-ray structures in conjunction with reasonable estimates for the positions of the relevant transition dipole vectors to calculate K2 explicitly. In the present case, the metal-localized nature of the d-d States on the Cun center makes this approach more difficult (vide infra), . . 2 however, we can 1nvertth1s process and calculate values for K based on our 158 experimental data. Values for kEnTr/kr, the overlap integral (Jmml), as well as K2 for compounds 1-5 are listed in Table 3-5. These numbers reveal the surprising result that the value of K2 is not constant across the series. Using a donor-acceptor distance of 10.37 A (a value corresponding to the distance between the Cu11 center and the midpoint of a vector that bisects the N-Re-N bond angle“), K2 for compound 1 is found to be 2.8 whereas for compound 5 the corresponding value is 5.7; this latter result exceeds the maximum value of K2 = 4 calculated for head-to-tail parallel transition dipoles.l '5 The underlying reason for the apparent disconnect between the predictions of eq 3-5a and the results obtained for compounds 1 through 5 can be understood by examining the normalized rates of energy transfer across the series in a slightly different manner. For any two members of the series i and j, we can write equation 3-5b for the relative change in the rate of energy transfer as a function of spectral overlap: 1w 2 kr 1' _Cj"‘j'Jj kw, _ Cr K124: (3-5b) The isostructural nature of compounds 1 through 5 implies that, to a reasonable degree of precision, the donor-acceptor distance is essentially 159 constant across the series.‘22 Cancellation of these terms affords equation 3- [kEnVJ k, . K? J _ _.l =__ _1' [kE"%J Kf J1. (3-50) This expression indicates that the relative rates of (kr-normalized) energy 5c: transfer should correlate with the ratio of their spectral overlap integrals weighted by a proportionality constant equal to the relative magnitudes of their orientation factors: dipolar coupling involving the same two states (i.e., the 3MLCT state of the Re-bpy’ moieties and the Cu11 acceptor) should therefore yield a slope of 1 since their orientation factors will cancel. Using the data for compound 1 as our im reference state we find that sz/Kzl increases from 1.18 for compound 2 to 2.03 for compound 5. These data confirm the trend suggested by the individual K2 values listed in Table 3-5, namely that the orientation factor associated with energy transfer in this system is different in one region of the spectrum versus another. Specifically, our analysis reveals a net increase in the rate of energy transfer as the donor emission comes into resonance with the low-energy portion of the CulI absorption profile. 160 Table 3-5. Calculated Spectral Overlap Integrals and Energy Transfer Rate Constants for Complexes 1-5. complex Jtotala JGla Joza 1038l kEnT (x 108 S'l)b kEnT/krc K2d tmb (I) 4.29 0.127 2.26 1.90 0.67 208 2.8 dmb (2) 5.14 0.046 2.07 3.03 1.18 295 3.3 bpy (3) 5.52 0.027 1.92 3.57 1.18 347 3.7 dclb (4) 6.68 0.0045 1.29 5.39 1.71 658 5.7 deeb (5) 7.09 0.0014 0.97 6.12 1.96 700 5.7 aReported in units of 10'16 M'Icm3 . bkEnT = kob,(CuRe2) — kob,(BeRe2) (Table 3-3). 0Values for k, were taken from the corresponding BeRez complex (Table 3-3). dCalculated based on eq. 3-5 for R = 10.37 A. In order to determine whether there are, in fact, differential contributions fiom the excited-state manifold of Cu" to the energy transfer dynamics of this system, we deconvolved the absorption profile of Cu(phacac)2 into a sum of Gaussians as a means of examining each portion of the compound’s excited-state structure individually (Figure 3-11A). As expected, the spectrum is well-described by three bands corresponding to the high-energy charge—transfer feature (G1) and two, lower-intensity bands assigned to ligand-field transitions (G2 and G3).'03 Using this fitted spectrum, the total spectral overlap integral (how) can be apportioned according to its various components (where 1m, = J G] + J02 + J03) (Figure 3- 113 and Figure 3-11C). The data in Table 3-5 reveals that the charge- transfer state contributes minimally to the overall spectral overlap, with 161 contributions ranging from a high of ca. 3% in the case of compound 1 to less than 0.02% in compound 5. This is a clear, quantitative indication that the ligand-field terms are the dominant acceptor states in this system. As the emission spectrum is tuned toward the red the total spectral overlap increases, however, the contribution from the higher energy ligand-field transition(s) decreases fi'om 53% in compound 1 to 14% in compound 5. At the same time, fractional overlap with the G3 component increases from 44% to 86% of the total. Since the value of sz/Kzl from eq 3-50 is observed to increase across the series, we can infer fiom this analysis that the rate of energy transfer from the 3MLCT state of the Re-bpy’ moiety is intrinsically faster when coupling to the G3-component of the Cu11 absorption profile as compared to G2. This conclusion is further supported by noting that the ratio of orientation factors as defined by eq 3-5c yields the same value of 2.03 for compounds 4 and 5, implying that 135/184 = 1. Inspection of the data in Table 5 reveals an increase in Jm, for compound 5 versus compound 4 but comparatively little change in the relative contributions of G2 and G3. In this context, the variation in K2 across the entire series can be interpreted in terms of differences in the orientation of the transition dipoles associated with G2 and G3: once changes in spectral overlap are isolated to G3 (i.e., compounds 4 and 5), the proportionality factor in eq 3-5c collapses to unity. 162 Molar Absorptivity (M'lcm'l) OJ C --;_ ___ ........... 22 20 l8 l6 14 12 Energy (x103 cm'l) O (pazrleuuou) Atsruoiul uorssttug Molar Absorptivity (normalized) 22 20 118 1'6 14 12 Energy (x103 cm'l) Figure 3-11. A. The ground state absorption spectrum of Cu(phacac)2 in CHzClz (black) fit with a series of three Gaussians (G1 (solid trace), G2 (dotted trace), and G3 (dashed trace)). B and C. Overlay of the emission spectra of [Be(pyacac)2(Re(tmb)(CO)3)2](OTf)2 (6, purple), [Be(pyacac)2(Re(dmb)(CO)3)2](OTDZ (7, blue), [Be(pyacac)2(Re(bpy)- (C0)3)2l(0Tf)2 (8, green), [B6(PyacaC)2(Re(dC1b)(C0)3)2](0T02 (9, orange), and [Be(pyacac)2(Re(deeb)(CO)3)2](OTDZ (10, red) with G2 and G3, respectively. 163 Figure 3-1 1 (cont’d) Molar Absorptivity (normalized) (pozneuuou) misuatul uorsstuig 22 20 1 8 16 14 12 Energy (x103 cm'l) The analysis presented above clearly points to a substantive difference in the nature of the excited states associated with the two observed absorption features of the Cull core in terms of their ability to engage in energy transfer from the Re-bpy’ fluorophore. We therefore carried out a time-dependent DFT calculation on Cu(phacac)2 in CH2C12 solution in order to obtain a more detailed theoretical description of these states. The transition energies, oscillator strengths, and orbital compositions of the four lowest energy excited states that are predicted from this calculation are listed in Table 3-6. The low oscillator strengths (fca,c = 0.0000) calculated for all four transitions are consistent with the Laporte-forbidden nature of d-d absorptions. The calculated transition energies along with the absorption spectrum of Cu(phacac)2 acquired in CH2C12 solution are plotted in Figure 3- 12. The results of the calculation are in reasonably good agreement with 164 experiment, which we take as an indication of the general validity of the wavefunctions comprising these four absorptive features. Table 3-6. Calculated Orbital Composition of the Lowest- Energy Spin Allowed Absorptions of Cu(phacac)2 in CH2C12 8011111011. E (cm'l) fcalca orbital transitions (Ci)b Excited State 1c 18,831 0.0000 Excited State 2C 18,703 0.0000 Excited State 3d 16,280 0.0000 Excited State 4d 14,778 0.0000 860 H 1080 (0.17) 870 H 10813 (0.12) 900 H 1088 (0.23) 9613 H 1080 (0.94) 1008 H 1080 (0.11) 8613 H 1088 (0.23) 870 H 1088 (0.31) 970 H 1080 (0.92) 860 H 1080 (0.42) 870 H 1080 (0.33) 900 H 1088 (0.23) 1000 H 1080 (0.85) 840 H 1080 (0.17) 950 H 1080 (0.62) 1070 H 1080 (0.75) aCalculated oscillator strength. bc, corresponds to the coefficient for the specific orbital transition in the calculated absorption feature. The contribution of each component to the total wavefunction is given by c,. cExcited states 1 and 2 comprise the G2 transition of the deconvolved spectrum in Figure 3-11A. dExcited states 3 and 4 comprise the G3 transition of the deconvolved spectrum in Figure 3-11A. 165 ()1 Ch 0 O O I I I ._r O O I I Molar Absorptivity (M'lcm‘l) N 15 A O 22 20 18 l6 14 12 Energy (x 103 cm'l) Figure 3-12. The four lowest energy transitions of Cu(phacac)2 determined from a TD-DF T calculation plotted against the ground state absorption spectrum of Cu(phacac)2 in CHzClz (black). The green and red dotted lines represent the G2- and G3-based transitions, respectively. The calculated oscillator strengths for all four transitions are zero (Table 3-6); the intensities of the calculated absorptions have therefore been arbitrarily scaled to facilitate comparison with the experimental spectrum. GaussView renderings of the orbitals involved in the four calculated excited states are shown in Figure 3-13. Excited States 1 (18,831 cm'l), 2 (18,703 cm'l), and 4 (14,778 cm'l) are dominated by dz2 —> dxy (960 —> 1080), dyz —> dxy (970 —> 10813), and dxz —> dxy (10713 —> 1086) ligand- field transitions, respectively. Excited State 3 (16,280 nm) has a small contribution from the 9013 —> 10813 (dx2_y2 —> dxy) transition, but is dominated by the 10013 —» 108B (73%) component that contains contributions from both charge-transfer and metal-centered excited states. A closer look at 1000 reveals a large metal-ligand bonding component to the wavefunction, along with density located solely on the Cu[1 metal center 166 buried within this bonding contribution. The metal-centered density suggests a significant fiaction of d-d character exists in the 10013 —+ 10813 transition, which would explain the lack of oscillator strength associated with Excited State 3. TD-DFT calculations were also performed on Cu(phacac)2 in frozen Dzh symmetry in a CHzClz environment in order to compare with the results from the minimized solution phase geometry. Results of the two TD-DFT analyses exhibited close correspondence in both energy ordering and orbital compositions for the four low-energy transitions, suggesting that the structure of Cu(phacac)2 doesn’t vary significantly between the solution phase and idealized geometries. 167 "'" i" " a? ;%.%w «nu-a baa-0a «a. .~ . ’1» 9613 ——> 10813 Exited State 1 A. a 4 ’ '3 u 14““ #031310 we» “as; Q 3133—”; a. 9. .10 9). J J 97B —+ 10813 Excited State 2 J ‘ ’ ‘\ ‘ +0} 04 +0} WW: I a J a J —_) 10813 10013 Excited State 3 JG; 40': T 10’ «.300: “at”! 0". I" . 10713 —» 10813 Excited State 4 Figure 3-13. Drawings of the orbitals involved in the lowest-energy spin allowed transitions of Cu(phacac)2 based on a TD-DFT calculation. The coordinate system used is that shown in Figure 3-5 with the CuO4 coordination environment drawn in the xy plane. See Table 3-6 for further details. 168 The analysis of the energy transfer dynamics of compounds 1 through 5 indicated a significant difference in the rate of energy transfer for coupling to the two absorption bands of the Cu11 absorption profile, a difference that appears to be linked to the dipole orientation factor K2. At first glance this would seem to be an unlikely scenario given the metal-centered nature of ligand-field absorptions; as mentioned in the Introduction, d-d acceptor states represent a nearly ideal manifestation of the point-dipole approximation inherent to Forster theory. Fleming and co-workers have demonstrated how one can use density matrix formulations to obtain detailed representations of transition dipoles for energy transfer processes involving carotenoids.123 The transition metal-based nature of the excited states involved in the present study makes an analogous approach difficult, however, visual inspection of the ground- and excited-state orbital configurations reveals two interesting features. A comparison of the calculated transition energies from TD-DFT with the Cu(phacac); ground state absorption spectrum shows the higher energy G2 band is dominated by the 9613 —+ 10813 (dz2 ——> dxy) and 9713 —+ 10813 (dyz -—> dxy) transitions, whereas the lower energy G3 band is comprised mainly of the 10013 —-* 10813 and the 10713 ——> 10813 (dxz —-> dxy) transitions. One striking difference between these transitions is the lack of an x-axis component in the 169 wavefunctions corresponding to the occupied orbitals of the two highest- energy excited states. Ligand-field transitions — at least those arising from either (i1 or d9 configurations where a one-electron orbital picture can be invoked — can be thought of in terms of a transfer of charge between the different planes defined by the orbital’s azithumal quantum number (m.). In this regard, the transitions contributing to the G2 feature (dz2 —> dxy and dyz —> dxy) will be distinct in a geometric sense from the dxz —> dxy transition that characterizes the lowest-energy feature in G3. Absent more detailed information it’s difficult to quantify whether this should give rise to an increase or decrease in dipolar coupling to the Rel-based excited state, but it seems reasonable to expect a difference. A second distinction between the G2 and G3 bands that could be playing a role in the observed dynamics is the nature of Excited State 3. Despite its low oscillator strength and all indications from experimental data of its ligand-field nature,‘03 Figure 3-13 illustrates that the 10013 —> 10813 transition is qualitatively different from the other three excited states comprising the visible absorption spectrum of Cu(phacac)2. Based on the spatial distribution of orbital 10013 there appears to be some charge-transfer character to this transition, although clearly not sufficient to impact the oscillator strength for the absorption. Nevertheless, the larger change in 170 charge distribution associated with Excited State 3 could be having a subtle influence on the magnitude of the transition dipole, which in turn would enhance the rate of energy transfer when coupling to this state. It is interesting to note the fact that the G3 feature is composed of distinct excited states explains at least qualitatively why the value of K2 calculated from eq 3-5a for compounds 4 and 5 was found to exceed the theoretical maximum of 4 (Table 3-5). Given the above analysis, each of these excited states should be viewed as an independent acceptor: coupling to both of these states as part of the overlap factor for G3 would therefore lead to an “effective” value of K2 (e.g., a sum of contributions) as opposed to the value one would calculate assuming a single donor-acceptor interaction. A more detailed theoretical treatment of this system would clearly be required in order to explore these ideas fiirther. 3.5 Conclusions The synthesis, structures, and photophysical properties of a series of covalently linked assemblies containing Rel-bipyridyl donors and a Cull—acac acceptor have been described. Steady-state and time-resolved emission spectroscopies indicated that the strongly emissive Rel-based 3MLCT excited-state was significantly quenched in the presence of the Cull center 171 relative to electronically benign BelI analogues. Favorable overlap between the donor emission and the visible absorption spectrum of the acceptor, coupled with a ca. 10 A donor-acceptor separation and unfavorable driving forces for electron transfer, allowed for an assignment of Forster energy transfer as the dominant mechanism for excited-state reactivity in this system. A detailed examination of the rates of energy transfer across the series revealed that the degree of dipolar coupling between the donor and acceptor was not constant across the absorption envelope of the Cu11 acceptor, but instead exhibited preferential coupling to the lower-energy portion of the ligand-field manifold. Fitting the absorption spectrum of the Cull chromophore with a series of Gaussians allowed for a differential analysis of the spectral overlap that quantified differences in the dipole orientation factor K2 as the emission profile was tuned across the absorption spectrum of the acceptor. A TD-DFT analysis of the central Cull species revealed the composition of the Cull-based ligand-field acceptor states and permitted the identification of the specific molecular orbitals responsible for the dipolar energy transfer quenching as well as the likely origins of their distinct roles in energy transfer. This study, in addition to providing a unique example of orbitally-specific energy transfer, also illustrates the importance 172 of considering the substructure of absorption bands in terms of its potential impact on dipolar energy transfer dynamics. 173 3.6 References and Notes (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13) (14) Balzani, V.; Credi, A.; Venturi, M. Chem. Soc. Rev. 2009, 38, 1542. Adronov, A.; Gilat, S. L.; Frechet, J. M. J .; Ohta, K.; Neuwahl, F. V. R.; Fleming, G. R. J. Am. Chem. Soc. 2000, 122, 1175 and references therein. Cline, E. D.; Adamson, S. E.; Bernhard, S. Inorg. Chem. 2008, 4 7, 10378. Cooke, M. W.; Hanan, G. S. Chem. Soc. Rev. 2007, 36, 1466. Elvington, M.; Brown, J.; Arachchige, S. M.; Brewer, K. J. J. Am. Chem. Soc. 2007, 129, 10644. Fihri, A.; Artero, V.; Pereira, A.; Fontecave, M. Dalton Trans. 2008, 5567. Gust, D.; Moore, T. A.; Moore, A. L. Acc. Chem. Res. 2001, 34, 40. Wasielewski, M. R. Chem. Rev. 1992, 92, 435. Aubert, V.; Guerchais, V.; Ishow, E.; Hoang-Thi, K.; Ledoux, 1.; Nakatani, K.; Le Bozec, H. Angew. Chem. Int. Ed. 2008, 4 7, 577. Balzani, V.; Juris, A.; Venturi, M.; Campagna, S.; Serroni, S. Chem. Rev. 1996, 96, 759. Balzani, V.; Scandola, F. Supramolecular Photochemistry; Horwood: Chichester, U. K., 1991. Brown, W. R.; O'Boyle, N. M.; McGarvey, J. J.; Vos, J. G. Chem. Soc. Rev. 2005, 34, 641. Balzani, V.; Ceroni, P.; Juris, A.; Venturi, M.; Campagna, S.; Puntoriero, F.; Serroni, S. Coord. Chem. Rev. 2001, 219, 545. Welter, S.; Lafolet, F.; Cecchetto, E.; Vergeer, F.; De Cola, L. ChemPhysChem 2005, 6, 2417. 174 (15) (16) (17) (18) (19) (20) (21) (22) (23) (24) (25) (26) Fleming, C. A.; Maxwell, K. A.; DeSimone, J. M.; Meyer, T. J.; Papanikolas, J. M. J. Am. Chem. Soc. 2001, 123, 10336. Baranoff, E.; Barigelletti, F.; Bonnet, S.; Collin, J. P.; Flamigni, L.; Mobian, P.; Sauvage, J. P. In Photofunctional Transition Metals Complexes; Springer-Verlag Berlin: Berlin, 2007; Vol. 123, p 41-78. Blan, Z. Y.; Sumi, K.; Furue, M.; Sato, S.; Koike, K.; Ishitani, O. Inorg. Chem. 2008, 4 7, 10801. Walther, M. E.; Wenger, O. S. Dalton Trans. 2008, 6311. Pelleteret, D.; Fletcher, N. C. Eur. J. Inorg. Chem. 2008, 3597. Ventura, B.; Barbieri, A.; Barigelletti, F.; Seneclauze, J. B.; Retailleau, P.; Ziessel, R. Inorg. Chem. 2008, 4 7, 7048. Easun, T. L.; Alsindi, W. Z.; Towrie, M.; Ronayne, K. L.; Sun, X. Z.; Ward, M. D.; George, M. W. Inorg. Chem. 2008, 4 7, 5071. Otsuki, J.; Imai, A.; Sato, K.; Li, D. M.; Hosoda, M.; Owa, M.; Akasaka, T.; Yoshikawa, I.; Araki, K.; Suenobu, T.; Fukuzumi, S. Chem-Eur. J. 2008, I4, 2709. Larsen, J .; Puntoriero, F.; Pascher, T.; McClenaghan, N.; Campagna, S.; Akesson, E.; Sundstrom, V. ChemPhysChem 2007, 8, 2643. Goeb, S.; De Nicola, A.; Ziessel, R.; Sabatini, C.; Barbieri, A.; Barigelletti, F. Inorg. Chem. 2006, 45, 1173. Shih, C.; Museth, A. K.; Abrahamsson, M.; Blanco-Rodriguez, A. M.; Di Bilio, A. J.; Sudhamsu, J.; Crane, B. R.; Ronayne, K. L.; Towrie, M.; Vléek, A.; Richards, J. H.; Winkler, J. R.; Gray, H. B. Science 2008, 320, 1760. Abrahamsson, M.; Jéiger, M.; Kumar, R. J .; Gsterman, T.; Persson, P.; Becker, H. C.; Johansson, 0.; Hammarstrom, L. J. Am. Chem. Soc. 2008, 130, 15533. 175 (27) (28) (29) (30) (31) (32) (33) (34) (35) (36) (37) (38) (39) Fleming, C. N.; Brennaman, M. K.; Papanikolas, J. M.; Meyer, T. J. Dalton Trans 2009, 3903. Meylemans, H. A.; Lei, C. F.; Damrauer, N. H. Inorg. Chem. 2008, 4 7, 4060. Fan, Y.; Zhang, L. Y.; Dai, F. R.; Shi, L. X.; Chen, Z. N. Inorg. Chem. 2008, 4 7, 2811. Wenger, O. S. Coord. Chem. Rev. 2009, 253, 1439. Benniston, A. C.; Harriman, A.; Li, P.; Patel, P. V.; Sams, C. A. Chem-Eur. J. 2008, 14, 1710. Yamamoto, Y.; Sawa, S.; Funada, Y.; Morimoto, T.; Falkenstrom, M.; Miyasaka, H.; Shishido, S.; Ozeki, T.; Koike, K.; Ishitani, O. J. Am. Chem. Soc. 2008, 130, 14659. Bian, Z. Y.; Sumi, K.; Furue, M.; Sato, S.; Koike, K.; Ishitani, O. Dalton Trans. 2009, 983. Alstrum-Acevedo, J. H.; Brennaman, M. K.; Meyer, T. J. Inorg. Chem. 2005, 44, 6802. Flamigni, L.; Collin, J. P.; Sauvage, J. P. Acc. Chem. Res. 2008, 41, 857. Balzani, V.; Credi, A.; Venturi, M. ChemSusChem 2008, I, 26. Concepcion, J. J .; Jurss, J. W.; Templeton, J. L.; 'Meyer, T. J. J. Am. Chem. Soc. 2008, 130, 16462. Liu, F.; Concepcion, J. J .; Jurss, J. W.; Cardolaccia, T.; Templeton, J. L.; Meyer, T. J. Inorg. Chem. 2008, 4 7, 1727. Xu, Y. H.; Akermark, T.; Gyollai, V.; Zou, D. P.; Eriksson, L.; Duan, L. L.; Zhang, R.; Akermark, B.; Sun, L. C. Inorg. Chem. 2009, 48, 2717. 176 (40) (41) (42) (43) (44) (45) (46) (47) (48) (49) (50) (51) (52) Youngblood, W. J .; Lee, S. H. A.; Kobayashi, Y.; Hernandez-Pagan, E. A.; Hoertz, P. G.; Moore, T. A.; Moore, A. L.; Gust, D.; Mallouk, T. E. J. Am. Chem. Soc. 2009, 131, 926. Knight, T. E.; Guo, D.; Claude, J. P.; McCusker, J. K. Inorg. Chem. 2008, 47, 7249. Picraux, L. B.; Smeigh, A. L.; Guo, D.; McCusker, J. K. Inorg. Chem. 2005, 44, 7846. Picraux, L. B.; Weldon, B. T.; McCusker, J. K. Inorg. Chem. 2003, 42, 273. Benniston, A. C.; Harriman, A.; Li, P.; Sams, C. A. J. Am. Chem. Soc. 2005, 127, 2553. Weldon, F.; Hammarstrom, L.; Mukhtar, E.; Hage, R.; Gunneweg, E.; Hassnoot, J. G.; Reedijk, J .; Browne, W. R.; Guckian, A. L.; Vos, J. G. Inorg. Chem. 2004, 43, 4471. Bignozzi, C. A.; Bortolini, 0.; Chiorboli, C.; Indelli, M. T.; Rampi, M. A.; Scandola, F. Inorg. Chem. 1992, 31, 172. Tamilarasan, R.; Endicott, J. F. J. Phys. Chem. 1986, 90, 1027. Kanemaguire, N. A. P.; Allen, M. M.; Vaught, J. M.;-Hallock, J. S.; Heatherington, A. L. Inorg. Chem. 1983, 22, 3851. Lei, Y. B.; Buranda, T.; Endicott, J. F. J. Am. Chem. Soc. 1990, 112, 8820. Langford, V. S.; von Arx, M. E.; Hauser, A. J. Phys. Chem. A 1999, 103, 7161. Wolcan, E.; Alessandrini, J. L.; F eliz, M. R. J. Phys. Chem. B 2005, 109, 22890. Neuman, D.; Ostrowski, A. D.; Mikhailovsky, A. A.; Absalonson, R. 0.; Strouse, G. F .; Ford, P. C. J. Am. Chem. Soc. 2008, I30, 168. 177 (53) (54) (55) (56) (57) (58) (59) (60) (61) (62) (63) (64) (65) (66) Otsuka, T.; Sekine, A.; Fujigasaki, N.; Ohashi, Y.; Kaizu, Y. Inorg. Chem. 2001, 40, 3406. Albinsson, B.; Martensson, J. J. Photochem. Photobiol. C-Photochem. Rev. 2008, 9, 138. Ward, M. D.; McCleverty, J. A. Dalton Trans. 2002, 275. Laye, R. H.; Couchman, S. M.; Ward, M. D. Inorg. Chem. 2001, 40, 4089. Scandola, F .; Indelli, M. T.; Chiorboli, C.; Bignozzi, C. A. Top. Curr. Chem. 1990, 158, 73. Scandola, F.; Bignozzi, C. A.; Chiorboli, C.; Indelli, M. T.; Rampi, M. A. Coord. Chem. Rev. 1990, 97, 299. Neuman, D.; Ostrowski, A. D.; Mikhailovsky, A. A.; Absalonson, R. 0.; Strouse, G. F.; Ford, P. C. J. Am. Chem. Soc. 2008, 130, 168. Ford, P. C. Acc. Chem. Res. 2008, 41, 190. Shavaleev, N. M.; Accorsi, G.; Virgili, D.; Bell, D. R.; Lazarides, T.; Calogero, G.; Armaroli, N.; Ward, M. D. Inorg. Chem. 2005, 44, 61. Davies, G. M.; Pope, S. J. A.; Adams, H.; Faulkner, S.; Ward, M. D. Inorg. Chem. 2005, 44, 4656. Kennedy, F.; Shavaleev, N. M.; Koullourou, T.; Bell, Z. R.; Jeffery, J. C.; Faulkner, S.; Ward, M. D. Dalton Trans. 2007, 1492. Baca, S. G.; Adams, H.; Sykes, D.; Faulkner, S.; Ward, M. D. Dalton Trans. 2007, 2419. Baca, S. G.; Pope, S. J. A.; Adams, H.; Ward, M. D. Inorg. Chem. 2008, 4 7, 3736. Buono-Core, G. E.; Li, H. Coord. Chem. Rev. 1990, 99, 55. 178 (67) (68) (69) (70) (71) (72) (73) (74) (75) (76) (77) (78) (79) (80) Perrin, D. D.; Armarego, W. L. F. Purification of Laboratory Chemicals; 3rd ed.; Pergamon Press: Oxford, U. K., 1988. Mackay, L. G.; Anderson, H. L.; Sanders, J. K. M. J. Chem. Soc., Perkin Trans. 1. 1995, 18, 2269. Chen, 3.; Fronczek, F. R.; Maverick, A. W. Chem. Commun. 2003, 2166. Carmichael, J. W., Jr.; Steinrauf, L. K.; Belford, R. L. J. Chem. Phys. 1965, 43, 3959. Vreshch, V. D.; Chemega, A. N.; Howard, J. A. K.; Sieler, J.; Domasevitch, K. V. Dalton Trans. 2003, 1707. Hino, J. K.; Della Ciana, L.; Dressick, W. J.; Sullivan, B. P. Inorg. Chem. 1992, 31, 1072. Caspar, J. V.; Meyer, T. J. J. Phys. Chem. 1983, 87, 952. Barbaro, P.; Cecconi, F.; Ghilardi, C. A.; Midollini, S.; Orlandini, A.; Alderighi, L.; Peters, D.; Vacca, A.; Chinea, E.; Mederos, A. Inorg. Chim. Acta. 1997, 262, 187. Klein, R. M.; Bailar, J. C. Inorg. Chem. 1963, 2, 1187. Sohrin, Y.; Matsui, M.; Hata, Y.; Hasegawa, H.; Kokusen, H. Inorg. Chem. 1994, 33, 4376. SAINT; ver 6.02a ; Bruker AXS, Inc.: Madison, WI, 2000. Sheldrick, G. M. SADABS ver 2.03; Bruker AXS, Inc.: Madison, WI, 2000. Sheldrick, G. M. SHELXT L ver 6.12; Bruker AXS, Inc.: Madison, WI, 2001. Damrauer, N. H.; Boussie, T. R.; Devenney, M.; McCusker, J. K. J. Am. Chem. Soc. 1997, 119, 8253. 179 (81) (82) (83) (84) (85) (86) (87) (88) (89) (90) Claude, J. P.; Meyer, T. J. J. Phys. Chem. 1995, 99, 51. Parker, C. A.; Rees, W. T. Analyst (London) 1960, 85, 587. DeWitt, L.; Blanchard, G. J.; LeGoff, E.; Benz, M. E.; Liao, J. H.; Kanatzidis, M. G. J. Am. Chem. Soc. 1993, 115, 12158. Origin; 7.5 ed.; OriginLab Corp.: Northhampton, MA, 1991 - 2004. Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, J ., J. A.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J .; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J .; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J .; Gomperts, R.; Stratmann, R. E.; Yazyev, 0.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. J .; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, 1.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, H; Chen, W.; Wong, M. W.; Gonzalez, C.; Pople, J. A.; Gaussian, Inc.: Wallingford, CT, 2004. Miertus, S.; Scrocco, S.; Tomasi, J. Chem. Phys. 1981, 55, 117. Tomasi, J.; Bonaccorsi, R.; Cammi, R.; Valle, F. J. J. Mol. Struc. (T heochem) 1991, 234, 401. Tomasi, J .; Bonaccorsi, R. Croat. Chem. Acta 1992, 65, 29. Tomasi, J .; Persico, M. Chem. Rev. 1994, 94, 2027. Frisch, M. J.; Frisch, A. Gaussian 98 User's Reference; Gaussian, Inc.: Pittsburgh, 1998. 180 (91) (92) (93) (94) (95) (96) (97) (98) (99) Dennington II, R.; Keith, T.; Millam, J. GaussView ver 4.1.2; Semichem, Inc.: Shawnee Mission, KS, 2003. Maverick, A. W.; Fronczek, F. R.; Maverick, E. F.; Billodeaux, D. R.; Cygan, Z. T.; Isovitsch, R. A. Inorg. Chem. 2002, 41 , 6488. Chen, B.; Fronczek, F. R.; Maverick, A. W. Inorg. Chem. 2004, 43, 8209. Gallucci, J. C.; Gerkin, R. E. Acta Crystallogr. 1989, C45, 1279. Blackburn, A. C.; Gallucci, J. C.; Gerkin, R. E. Acta Crystallogr. 1991, C47, 2019. Cotton, F. A.; Daniels, L. M.; Murillo, C. A.; Quesada, J. F. Inorg. Chem. 1993, 32, 4861. Lucia, L. A.; Abboud, K.; Schanze, K. S. Inorg. Chem. 1997, 36, 6224. Chen, P.; Curry, M.; Meyer, T. J. Inorg. Chem. 1989, 28, 2271. Wenger, O. S.; Henling, L. M.; Day, M. W.; Winkler, J. R.; Gray, H. B. Inorg. Chem. 2004, 43, 2043. (100) Busby, M.; Liard, D. J.; Motevalli, M.; Toms, H.; Vléek, A. Inorg. Chim. Acta. 2004, 35 7, 167. (101) Sacksteder, L.; Zipp, A. P.; Brown, E. A.; Streich, J .; Demas, J. N.; DeGraff, B. A. Inorg. Chem. 1990, 29, 4335. (102) Worl, L. A.; Duesing, R.; Chen, P.; Della Ciana, L.; Meyer, T. J. J. Chem. Soc., Dalton Trans. 1991, 849. (103) Belford, R. L.; Carmichael, J. W., Jr. J. Chem. Phys. 1967, 46, 4515. (104) Funck, L. L.; Ortolano, T. R. Inorg. Chem. 1968, 7, 567. (105) Striplin, D. R.; Crosby, G. A. Chem. Phys. Lett. 1994, 221, 426. 181 (106) It should be noted that the presence of this emissive impurity is explicitly incorporated in the kinetic models used to fit the emission data. (107) Kestell, J. D.; Williams, Z. L.; Stultz, L. K.; Claude, J. P. J. Phys. Chem. A. 2002, 106, 5768. (108) Barqawi, K. R.; Murtaza, Z.; Meyer, T. J. J. Phys. Chem. 1991, 95, 47. (109) Although the photophysical experiments for complexes 1-10 were measured in CHZCIZ solution, the cyclic voltammetry measurements were performed in CH3CN due to insufficient solubility of the CuRe2 complexes in CHzClz solution to carry out the electrochemical measurements. The use of redox potentials measured in CH3CN to evaluate driving forces for electron transfer in CH2C12 solutions is based on our previously reported data on structurally similar FeRe3/A1Re3 complexes which reveal that the redox properties of the Rel-bpy. fragment differ by less than 10 mV between the two solvents. The impact of the change in solvent on the redox properties of the Cull center is addressed in the text. (110) Gritzner, G.; Murauer, H.; Gutmann, V. J. Electroanal. Chem. Interfacial Electrochem. 1979, 101, 177. (111) Bradbury, J. R.; Hampton, J. L.; Martone, D. P.; Maverick, A. W. Inorg. Chem. 1989, 28, 2392. (112) Rehm, D.; Weller, A. Isr. J. Chem. 1970, 8, 259. (113) The form of the Rehm-Weller equation employed to characterize the oxidative and reductive quenching processes were AGET"x = on (Rel/H) ' Ered (cull/I) ' E00 and AGEde = on (Cull/Ill) ' Ered (bpyflfi) ' E009 respectively, where E00 is the energy gap between the ground and lowest energy 3MLCT excited state of the donor. In both cases, the small work term typically employed for bimolecular systems was not included in the analysis. (114) Dexter, D. L. J. Chem. Phys. 1953,21, 836. 182 (115) Forster, T. Discuss. Faraday Soc. 1959, 27, 7. (116) Yardley, J. T. Introduction to Molecular Energy Transfer; Academic Press: New York, 1980. (117) Wang, Y.; Schanze, K. S. Inorg. Chem. 1994, 33, 1354. (118) Berg, K. E.; Tran, A.; Raymond, M. K.; Abrahamsson, M.; Wolny, J .; Redon, S.; Andersson, M.; Sun, L.; Styring, S.; Hammarstrom, L.; Toftlund, H.; Akermark, B. Eur. J. Inorg. Chem. 2001, 1019. (119) Soler, M.; McCusker, J. K. J. Am. Chem. Soc. 2008, 130, 4708. (120) Dattelbaum, D. M.; Martin, R. L.; Schoonover, J. R.; Meyer, T. J. J. Phys. Chem. A. 2004, 108, 3518. (121) Van Der Meer, B. W.; Coker, G. 1.; Chen, S.-Y. Resonance Energy Transfer, Theory and Data; VCH Publishers: New York, 1994. (122) The degree to which a given substituent is electron donating or electron withdrawing is expected to subtly modulate the spatial characteristics of the 3MLCT excited state of the Re-bpy' group. However, our previous study of FeRe3 assemblies (cf. 41) demonstrated that these variations are minor given the scale of the distances involved in energy transfer in these types of systems. (123) Vaswani, H. M.; Hsu, C.-P.; Head-Gordon, M.; Fleming, G. R. J. Phys. Chem. B. 2003, 107, 7940. 183 Chapter 4. Confirming Forster Energy Transfer Involving Ligand—Field Acceptor States: CrRe3 Complexes. 4.1 Introduction The theoretical formalism of dipole-dipole energy transfer is described by the well-known Forster equation (equation 4-1), that elegantly relates experimental observables to the rate constant for energy transfer (kEnT): 1,2 szrJ kEnT = C R6 (4-1) where C is a collection of constants, and K2, k,, J, and R are as defined previously (eq. 1-1). The spectral overlap integral (J) describes the resonance condition for energy transfer and is by far the most essential variable in the Forster equation, and is given in equation 4-2. We}: 510’) (4—2) The spectral overlap integral (J) quantifies the energetic correspondence between the donor emission and the acceptor absorption transition moment dipoles, and is described in detail in Chapter 1. The degree of spectral overlap is directly proportional to the magnitude of the Forster rate constant 184 (kEnr), and must be non-zero in order for dipole-dipole energy transfer to be considered as a possible excited-state quenching pathway. Results from the CuRez assemblies presented in Chapter 3 showed that ligand-field acceptor states of Cu(pyacac)2 (pyacac = 3-(4-pyridy1)-2,4- pentanedione) can quench the 3MLCT (metal-to-ligand charge transfer) excited states of Rel-polypyridyl donor complexes. The spectral overlap ' values between the ReI donor emission spectra with the ligand-field absorption profile of Cu11 did not manifest a direct correlation with the magnitude of the energy transfer rate constants within the structurally homologous series. The ground state absorption spectrum of the d9 CuII core exhibits two broad absorption bands that are attributed to four spin-allowed ligand-field transitions,3 which did not allow for the spectral overlap integrals of each individual d-d transition with the ReI emission to be determined. The ground state absorption profile of Cu(phacac); (phacac = 3-phenyl-2,4-pentanedione) was fit with a series of Gaussians in order to evaluate the spectral overlap of the individual Cull—based ligand-field acceptor states with the emission spectra of the Rel-polypyridyl donor moieties. The trend in spectral overlap values with each of the Gaussians, along with computational efforts on Cu(phacac)2 revealed that the through- space orientation (1(2) of the 3d orbitals relative to the 3MLCT donor 185 emission dipoles also play a role in the Forster energy transfer dynamics. In contrast to the four distinct orbital based transitions corresponding to the ligand-field bands of Cu“, the 4A2 —» “T2 ligand-field absorption of Cr‘" is characterized by a multielectronic wavefunction deriving from a single configuration (Figure 4-1).4 .,,_1_l_1__"11 Figure 4-1. One electron-orbital description of the 4A2 —> 4T2 electronic transition in Crm. This diagram depicts the electronic configurations from which the multielectronic wavefunction corresponding to the 4A2 and 4T2 are derived. The focus of the work contained in this chapter is two fold, 1) utilize the nature of the Cr"l ligand-field transition to examine if the rate of 3MLCT emission quenching can directly correlate with the spectral overlap between ligand-field-type transitions and Rel-based emission spectra and 2) utilize the emissive properties of CrIII ligand-field excited states to provide unequivocal proof that the Rel-based donor complexes undergo energy transfer dynamics with ligand-field acceptor states. The synthesis, structure, and photophysical properties of a series of isostructurally related molecules, with the general formula 186 [Cr(Py3036)3(Re(tmb)(C0)3)3](0T03 (1), [Cr(PyacaC)3(Re(bPY)(C0)3)3]- (OTf)3 (2), and [Cr(pyacac)3(Re(deeb)(CO)3)3](OTf); (3) (where pyacac = 3- (4-pyridyl)-acetylacetonate, tmb = 4,4’-5,5’-tetramethyl-2,2’-bipyridine, bpy = 2,2’-bipyridine, deeb = 4,4’-diethylester-2,2’-bipyridine, and OTf = CF3SO3') are described. The Crm metal centers are covalently attached to three fac-Re(bpy’)(CO)3 (bpy’ = tmb, bpy, and deeb) moieties through three pyridyl-acetylacetonate bridging ligands (Figure 4-2). It was observed that the Rel-based 3MLCT excited-states for complexes 1, 2, and 3 are "1 metal center relative to significantly quenched in the presence of the Cr structurally analogous AlRe3 analogues (complexes 4, 5, and 6 from Chapter 2). Varying the bpy’ attached to the ReI metal center permitted a systematic spectral overlap analysis with the Crm ligand-field absorption profile. Emission and excitation spectra for complex 2 and Cr(phacac)3 were also acquired in frozen 4:1 EtOH/MeOH, the results of which serve to confirm quenching of the Rel-based 3MLCT excited state by an energy transfer mechanism. 187 73+ cl /// \\ O 12R1=R2=CH3 22R1=R2=H 3: R1 = COzEt, R2 = H Figure 4-2. Structure of the [M(pyacac)3(Re(bpy’)(CO)3)3](OTf)3 assemblies (where M = Cr'll or All" and bpy’ = tmb, bpy, or deeb). 4.2 Experimental Section 4.2.1 Synthesis and Characterization General. All solvents used were purified and dried according to previously reported methods.5 Spectroscopic grade CHzClz was used for all photophysical measurements; the solvent was dried under CaHz reflux until no water was detected by lH NMR and degassed using freeze-pump-thaw 188 techniques. 3-(4-pyridyl)-2,4-pentanedione,6 Cr[N(SiMe3)2]2(TI-IF)2,7 Al(pyacac)3,6 Re(tmb)(CO)3(OTf),8 Re(bpy)(CO)3(OTf),8 Re(deeb)(CO)3- (OTf),8 and fac-[Re(bpy)(CO)3(4-Etpy)](PF6) (4-Etpy = 4-ethylpyridine)9 were prepared following literature procedures. The synthetic procedures for [A1(PyacaC)3(Re(tmb)(C0)3)3](0T03 (4), [A1(PyacaC)3(Re(bPY)(C0)3)3l- (OTD3 (5), and [A1(pyacac)3(Re(deeb)(CO)3)3](OTO3 (6) are reported in Chapter 2. 3-phenyl-2,4-pentanedione was purchased from TCI America. Elemental analyses and F T-IR data were obtained through the analytical facilities at Michigan State University. Mass spectra were obtained through the analytical facilities at The University of South Carolina. [Cr(pyacac)3(Re(tmb)(CO)3)3](OTf)3 (1). Preparation of [Re(tmb)- (CO)3(pyacac)](OTf) precursor was prepared as follows: 193 mg (0.306 mmol) of Re(tmb)(CO)3(OTf) and 185 mg (1.04 mmol) of pyacac were dissolved in 40 mL of THF. The reaction solution was flushed with argon for 20 min and then refluxed for 6 hrs under argon. The volume of the reaction mixture was then reduced to 10 mL and hexanes were added to precipitate a yellow solid. The solid was then washed with ~ 100 mL of diethyl ether to remove excess pyacac, dried under vacuum overnight, and then used as is for the synthesis of compound 1. The synthesis of compound 1 is given as follows: 78 mg (0.151 mmol) of Cr[N(SiMe3)2]2(TI-IF)2 was 189 dissolved in 5 mL of THF and cooled to —78°C . in a Nz-filled drybox. 368 mg (0.455 mmol) of [Re(tmb)(CO)3(pyacac)](OTf) in 20 mL of THF was added dropwise to the cold solution over 5 min. A light green precipitate was observed immediately upon addition, and the mixture was allowed to stir for 20 min. The reaction mixture was removed from the drybox and additional green solid was precipitated by addition of hexanes. The green precipitate was dissolved in ACN, filtered through celite, and the solvent removed under vacuum. The product was recrystallized several times from acetonitrile/ether (1:1 v/v). Yield: 166 mg (44%). Anal. Calcd for C34H73N9F9024S3CrRe3'2HzO: C, 40.17; H, 3.29; N, 5.02. Found: C, 39.99; H, 3.56; N, 5.02. IR (KBr, v(CO) (cm")): 2030 s, 1912 s, 1565 s, 1031 s, 638 m. MS: [ESL m/z (rel. int.)]: 676 (100) {[Cr(pyacac)3(Re(tmb)- (C0)3)3]}3+, 1088-6 (33) {[Cr(py3030)3(Re(tmb)(CO)3)3](OTf)}2+, 2326-3 (1) 1[Cr(pyacac)3(Re(tmb)(c0m(our: and [Crtpyacac)3(Re(bpy)(CO)3)31- (OTf)2H. The general robustness of the CrRe3 complexes is also evidenced by the relative invariance of the steady-state emission intensity of complex 2 in CHzClz solution over the course of several hours (Vida infi'a), in stark contrast to what was observed for the FeRe3 complexes described in Chapter 2. The v(CO) stretching bands of the Re(bpy’)(CO)3 moieties reported in the Synthesis and Characterization section (4.2.1) were assigned based on previously reported data for fac-[Re(4,4’-X2-bpy)(CO)3(4-Etpy)](PF6) complexes.17 The characteristic spectral profile consists of two very intense 197 peaks. The broad band at lower energy corresponds to two overlapping transitions assigned to the A'(2) and A" modes (Cs symmetry), whereas the sharper, higher energy band is assigned as A'(l).18 The similarity of the carbonyl frequencies observed for each CrRe3/AlRe3 pair is indicative of "1 metal minimal direct electronic communication between the Re1 and Cr centers in the ground states of the CrRe3 assemblies.19 4.3.2 Electronic Absorption Spectroscopy. The electronic absorption spectra of complexes 1-6 were acquired in room-temperature CH2C12 solutions; spectra for all 6 complexes are shown in Figure 4-3. The observed features were assigned based on previously reported analyses of Re1 polypyridyl absorption spectra. ReI polypyridyl complexes exhibit a 1A. -—> lMLCT (t2g —> n1 (bpy’)) transition occurring from approximately 330 to 2021 d. 430 nm depending on the substituents of the bpy’ ligan The absorption profiles of the AlRe3 compounds permitted analysis of the donor-based IMLCT transitions without overlapping transitions associated with the Crm- containing systems (1-3). The absorption maximum for the lMLCT excited- state reflects the electron donating/withdrawing ability of the bpy’ substituents, with Max for complexes 4, 5, and 6 at occurring at 344 (a = 18,500 M’lcm'l), 364 (a = 11,500 M'lcm'l), and 394 (a = 14,300 M‘cm“) nm, respectively. As the substituents become progressively more electron 198 donating (e.g., H for bpy and (CH3)4 for tmb) this feature systematically shifts to the blue and begins to overlap with the ligand-based absorptions in the ultraviolet (Figure 4-3). 199 75000 W A 60000 4 45000 T 30000- 15000- 75000 62500 50000- 37500- 25000 12500b 75000- 62500~ 50000- 37500- 25000- 12500- Molar Absorptivity (M'lcm'l) Molar Absorptivity (M'lcm’l) Molar Absorptivity (M‘lcm') 300 350 400 450 500 550 600 Wavelength (nm) Figure 4-3. Electronic absorption spectra of [M(pyacac)3(Re(bpy’)- (CO)3)3](OTf)3 assemblies, where M = CrIII (red traces) and All" (blue traces). All spectra were acquired in room-temperature CHzClz solution. A. [Cr(pyacac)3(Re(tmb)(CO)3)3](OTf)3 (1) and [Al(pyacac)3- (Re(tmb)(CO)3)3](OTf)3 (4)- R [Cr(pyacac)3(Re(bpy)(CO)3)3](OTf)3 (2) and [A1(Pyaca0)3(Re(bpy)(C0)3)3](OTf): (5)- C- [Cr(pyaca0)3(Re- (deeb)(C0)3)3l(0Tf)3 (3) and [A108/3030):;(Re(deeb)(C0)3)3](OTfls (6)- 200 l I 7 v U" 51’ ‘ . The presence of Crlll in complexes 1-3 gives rise to a new, broad absorption feature on the high-energy side of the Rel-based charge-transfer band. Figure 4-4 shows the absorption spectrum of Cr(phacac)3 in room— temperature CHzClz which exhibits a strong transition centered at 390 nm assigned to a 4A2 ——> 4LMCT (71: (acac) -—> tzg) transition. The 4A2 —-> 4LMCT absorption in complexes 1-3 is somewhat obscured by the 1A1 -—> lMLCT band of the Re1 chromophore, but a small shoulder is apparent in the three CrRe3 absorption spectra (Figure 4-3). Figure 4-4 also shows a broad peak at 570 nm (e = 113 M'lcm'l) which is assigned to the 4A2 —-> 4T2 ligand-field III transition of the Cr metal center. These assignments are based on the well- known ground state absorption properties of Cr(acac)3 derivatives.”24 700 600 500_ 400» 300- 200~ 100~ Molar Absorptivity (a) 400 500 600 700 800 Wavelength (nm) Figure 4-4. Electronic absorption spectrum of Cr(phacac)3 acquired in room-temperature CH2C12 solution. 201 4.3.3 Steady-State and Time-Resolved Emission. Emission spectra for the CrRe3 and AlRe3 assemblies were obtained in room-temperature deoxygenated CH2C12 solution. As discussed in Chapter 2, the AlRe3 complexes represent an ideal structural model for the dynamics associated with the ReI based 3MLCT emission due to the inability of the AlIII ion to engage in either electron or energy transfer quenching. The emission profiles for the CrRe3 and AlRe3 complexes are given in Figure 4-5. The spectral profiles for both the Crm- and Alm-containing complexes correspond to previously reported photophysical studies of Re1 polypyridyl systems, with the emission originating from the 3MLCT —> 1A. phosphorescence.25 The emission maximum for the tmb (526 nm), bpy (566 nm), and deeb (624 nm) derivatives reflect the electron donating and withdrawing behavior of the polypyridyl ligands. The radiative quantum yields ((1),) for complexes 1-6 were determined relative to [(bpy)Re(CO)3(4- Etpy)](PF6) (<1)r = 0.18 in CHzClz) and are given in Table 4-1. The (I)r values for the AlRe3 model complexes, which were originally reported and discussed in Chapter 2, are comparable to the reported values for the corresponding mononuclear ReI polypyridyl derivatives.8 As can be seen in Figure 4-5, the emission intensity for all three CrRe3 complexes are significantly attenuated compared to the AlRe3 model systems indicating 202 extensive quenching of the Rel-based 3MLCT excited-state. In addition, smaller radiative quantum yields ((1),) calculated for the CrRe3 complexes (1- 3) compared to the All" model complexes (Table 4-1) indicated significant quenching 3MLCT excited-state as well. It should also be noted that (I), values for previously reported chromophore-quencher complexes of FeRe3 (Chapter 2) and CuRez (Chapter 3) were analytically unreliable due to small amounts of H20 contained in the CHzClz that caused slight dissociation of the complexes over time, which produced Re(bpy’)(CO)3(pyacac) species as emissive impurities. The emission intensity of complex 2 remained constant when monitored over the course of 4 hrs in CHzClz solution (Figure 4-6), which attests to the robustness of the CrRe3 complexes. Therefore, quantitative radiative quantum yield values for complexes 1-3 could be reported (Table 4-1). 203 Emission Intensity Emlss10n Intensny 500 600 700 800 500 600 700 800 Wavelength (nm) Wavelength (nm) 1F (: Emission Intensity 500 600 700 800 Wavelength (nm) Figure 4-5. Corrected steady-state emission spectra for [M(pyacac)3_ (Re(bpy’)(CO)3)3](OTf)3 assemblies, where M = All" (blue traces) and Cr‘” (red traces). A. [Al(pyacac)3Re(tmb)(CO)3])3](OTt)3 (4) and [Cr(pyacac)3(Re(tmb)(CO)3)3](OTD3 (1). B. [Al(pyacac)3Re(bpy)- (C0)3])3](0Tf)3 (5) and [Cr(pyacaC)3(Re(bPY)(C0)3)3](0T03 (2)- C- [Al(pyacac)3(Re(deeb)(CO)3)3](OTD3 (6) and [Cr(pyacac)3(Re(deeb)- (CO)3)3](OTf)3 (3). All six spectra were acquired in deoxygenated room-temperature CH2C12 solutions following excitation at 355 nm (complexes 1, 2, 4, and 5) and 400 nm (complexes 3 and 6). For each plot, the emission profiles have been normalized with respect to the absorbance of each sample at its excitation wavelength. The relative magnitudes of the signals are therefore accurate representations of their relative intensities. 204 300000 - 250000 - 200000 - 150000- 100000 - N 50000 - Whig» 0 Intensity (cps) <1, *4“ 2 500 550 600 650 700 Wavelength (nm) Figure 4-6. Corrected steady-state emission spectra for [Cr(pyacac)3- (Re(bpy)(CO)3)3](OTt)3 (2) monitored over the course of 4 hrs in room- temperature CHzClz solution. The various emission traces are represented as follows: the initial spectrum (black trace), after 1 hr (red trace), after 2 hrs (blue trace), after 3 hrs (green trace), and after 4 hrs (orange trace). Additional details concerning the excited states of the AlRe3 and CrRe3 complexes were obtained through nanosecond emission lifetime and time-correlated single photon counting (TCSPC) measurements. The nanosecond emission lifetime data in room-temperature CH2C12 solution for complexes 4, 5, and 6 were initially reported in Chapter 2, with rob, = 2260 i 100, 560 i 30, and 235 i 20 ns, respectively. The corresponding radiative and non-radiative decay rate constants for 4, 5, and 6 are given in Table 4-1. As with the quantum yields, the observed excited-state lifetimes and rate 8.26 constants are all consistent with Rel-based 3MLCT —+ 'Al emission, which were discussed in Chapter 2. 205 The lack of discemable signal observed for the CrRe3 systems during nanosecond lifetime experiments prompted the employment of time- correlated single photon counting (TCSPC) methods. Plots of the TCSPC data obtained in room-temperature CHzClz solution for the CrRe3 complexes are shown in Figure 4-7. Complexes 1 and 2 could be fit with single- exponential models with rob, = 6.8 :1: 0.1 ns and 4.8 i 0.1 ns, respectively (Table 4-1). Complex 3 exhibited bi-exponential kinetics with lifetimes of 1'1 = 7.0 i 0.3 ns and 12 = 190 :t 15 ns (Table 4-1). Due to the similar values of T2 and the AlRe3 model complex (deeb, 6), the slower component is assigned to residual amounts of [Re(deeb)(CO)3(pyacac)](OTf) starting material remaining after purification of complex 3. The signal-to-noise ratio for the three TCSPC data sets (1-3) are relatively poor owing to a combination of virtually complete quenching of the Rel-based 3MLCT states coupled with radiative rate constants that are on the order of 105 s". As can be seen from the 3MLCT lifetimes in the A1Re3 and CrRe3 compounds mentioned above, 1:01,, for the three Alm assemblies are significantly longer 111 than what was observed for the Cr complexes, indicating that 3MLCT relaxation in complexes 1-3 is dominated by reaction with the Cr"I core. 206 60 E 1,? 15~ e, 40 3 b as 10 i E 8 20 .3 5- E. E 0 5 10 15 20 25 30 Time (ns) 1400: 1200 1000 800 Intensity (cps) 600 400 0 215 510 715 100 125 150 115 Time (ns) Figure 4-7. Time correlated single-photon counting (TCSPC) data for the CrRe3 complexes: 1 (rob, = 6.8 i 0.1 ns), 2 (robS = 4.8 i 0.2 ns), and 3 (11: 7.0 i 0.3 ns) and 1:2 = 190 i 15 ns). All data were collected in deoxygenated room-temperature CH2C12 solution. The red solid lines correspond to fits to single-exponential decay models for 1 and 2, and a bi-exponential decay model for 3 due to the presence of [Re(deeb)- (CO)3(pyacac)](OTf) as an emissive impurity. See text for further details. 207 Table 4-1. Photophysical Data of Complexes 1-6. E00 kobs (SJ) kt knr complexes km, (cm'l)a (I)r (><108 s") (X 106 s")b (X 106 s'l)c CI'R63 tmb (1) 526 19,700 0.024 1.5 :t 0.1 3.6 i 0.2 150 i: 10 bpy(2) 566 18,600 0.009 2.1401 19401 210410 deeb(3) 624 17,000 0.003 1,440.1d 04840.02 140410 AlRe3 tmb(4)c 526 19,900 0.51 0.0044 0.234001 02240.01 bpy (5)6 566 18,700 0.16 0.018 0.29 3: 0.01 1.5 3: 0.1 deeb (6)6 624 16,800 0.07 0.043 0.30 i 0.01 4.0 d: 0.2 aZero-point energy difference between 3MLCT excited state and ground state based on spectral fitting analysis. bkr = kobs*(1),. ck“, = kob, - k,. dValue reported is the faster component (1'1) of the bi-exponential behavior observed in the TCSPC data for complex 3. See text for further details. eThese data were reproduced from Chapter 2 for comparison purposes. 4.4 Mechanistic Considerations 4.4.1 Electron versus Energy Transfer Quenching. Both electron and energy transfer processes can be envisioned to occur out of the Rel-based 3MLCT excited state. Based on the presence of Cr"1 in the ground states of the CrRe3 complexes, electron transfer would proceed as either an oxidative quenching reaction to produce (bpy’)Re"—CrII or a reductive quenching 208 reaction to produce (bpy")Re"—Crw. Based on previously reported electrochemical data by Murray and Hiller on Cr(acac)3,27 a CrIV redox- product that would be produced in the reductive quenching mechanism would be highly unlikely and will not be discussed further. In order to determine if the oxidative quenching process is a thermodynamically viable 1" ——> CrII reduction potential of mechanism in the CrRe3 systems, the Cr Cr(acac)3,27 the Re1 —> RelI oxidation potentials for the AlRe3 model complexes (4-6), and the zero-point energy gaps of the 3MLCT states (E00) determined from fits of the emission spectra of the corresponding CrRe3 analogues (Table 4-1), were used to calculate the thermodynamic driving force for photoinduced electron transfer (AGET) in complexes 16.2839 The calculated AGET values are all significantly uphill with values of +1.18 V (tmb, 1), +1.56 V (bpy, 2), and +1.87 V (deeb, 3). The large endothermic AGET values effectively rule out electron transfer as a viable 3MLCT quenching mechanism in the CrRe3 series. It should also be noted that cyclic voltammetry on Cr(phacac)3 in CH2C12 solution was performed with m . . —> CrII reductlon peaks observed 1n neither the Cr-III —» Crlv oxidation or Cr the solvent window, agreeing with the very large AGET values calculated from the Cr(acac)3 literature values. 209 Direct experimental evidence for energy transfer in the CrRe3 series was obtained from low-temperature emission spectroscopy. The lowest energy excited state of Cr(acac)3-type complexes (the 2E ligand-field state) does not emit at room-temperature but exhibits a sharp emission spectrum with km, from approximately 750 to 850 nm in frozen solution.22 The narrowness of the band, which is assigned as the 2E —> 4A2 emission, results from the minimal nuclear displacement associated with the intraconfigurational nature of the excited state.”33 Figure 4-8 shows the emission spectrum of Cr(phacac)3 in 4:1 EtOH/MeOH at 82 K that exhibits the sharp 2E —> 4A2 emission. It should also be noted that this low- temperature 2E —> 4A2 emission of Cr(phacac)3 represents the expected result of Forster energy transfer quenching of the 3MLCT donor state by the Crm- based 4A2 —-* 4T2 ligand-field transition in the CrRe3 series, and not by the 4A2 —-> 2E state. The extremely small spectral overlap calculation between _ Rel-based emission and the spin-forbidden 4A2 —> 2E transition negates the possibility of dipolar energy transfer coupling between the two states. 210 Emission Intensity (au) Wavelength (nm) Figure 4-8. Steady-state emission spectrum of Cr(phacac)3 in 4:1 EtOH/MeOH at 82 K following excitation at 375 nm. , . 4 . . 111 As dlscussed above, observatlon of 2E —> A2 em1ss10n fi'om the Cr core (Figure 4-8) in the CrRe3 assemblies would be direct experimental evidence for energy transfer quenching. Steady-state emission data for complex 2 in 4:1 EtOH/MeOH at 82 K is given in Figure 4-9. Based on the Cr(phacac)3 emission spectrum shown in Figure 4-8, the resulting emission characteristics for complex 2 have been assigned to the Crm-based 2E —> 4A2 transition. In order to distinguish between direct excitation of the Crm-based 4A2 —+ 4T2 absorption giving rise to the observed 2E —> 4A2 emission, and 2E- based emission signal resulting from energy transfer dynamics in complex 2, the extinction coefficient values at Mun“, = 375 nm in CH2C12 were compared between complex 2 and Cr(phacac)3. The comparison showed that ~ 9% of the 375 nm photons absorbed by complex 2 directly excite the Cr(acac)3 core, while the rest are absorbed by the IA] —> 1MLCT transition of the Rel 211 donor. The red trace in Figure 4-9 represents the emission spectrum of complex 2 if only absorption by the Cr"I acceptor (9% contribution) was responsible for producing the 2E ——> 4A2 emissive photons (ie. no energy transfer). The emission spectrum of complex 2 (black trace) clearly shows a much larger intensity than would be expected from simple Cr(acac)3 excitation, which implicates energy transfer fi'om the Rel-based 3MLCT lll excited state into the Cr core as being responsible for the increase in 2E —> 4A2 emission intensity. ~ 00 l 6 Emission Intensity (x 10 cps) Wavelength (nm) Figure 4-9. The black trace is the corrected steady-state emission spectra of [Cr(pyacac)3(Re(bpy)(CO)3)3](OTf)3 (2) in 4:1 EtOH/MeOH at 82 K following excitation at 375 nm. The red trace corresponds to the intensity of the emission spectrum scaled by the relative absorbances of the Rel-bpy and Cr'"(pyacac)3 chromophores. See text for further details. Additional evidence for excited state energy transfer was obtained by measuring the excitation spectrum of complex 2 at an energy corresponding 111 to emission solely from the Cr fluorophores. Figure 4-10 shows the 212 absorption spectrum of complex 2 at 298 K along with the excitation spectrum of complex 2 at 82 K resulting from monitoring the Crm-based 2E -—> 4A2 emission maximum (km, = 795 nm). As can be seen in Figure 4-10, the overall shape of the absorption and excitation spectra are conserved, which demonstrates that emission produced from the 2E —> 4A2 transition of the Cr"I center originates from both the Rel-based lMLCT and CrIH-based absorptions. These spectra provide additional evidence that the 2E —+ 4A; emission results from energy transfer quenching of the Rel-based 3MLCT excited state. A noticeable deviation in the two spectra are noted in the near UV region of Figure 4-10, and results from poor photon-to-current efficiency for the photo-diode employed to adjust for lamp fluctuations in excitation spectrum. p.- 1 p—l Absorbance (normalized) (pazrleuuou) Kirsuaml O 0 4L 350 400 450 500 Wavelength (nm) Figure 4-10. Ground-state absorption spectrum of [Cr(pyacac)3(Re- (bpy)(CO)3)3](OTf)3 (2) in 4:1 EtOH/MeOH solution at 298 K (black trace), and the excitation spectrum of 2 in 4:1 EtOH/MeOH at 82 K monitoring at item = 795 nm (red trace). 213 4.4.2 Dexter vs Fiirster Energy Transfer. In light of the prohibitively large electron transfer driving forces and the low-temperature emission and excitation spectra for complex 2, the 3MLCT quenching in all three of the CrRe3 assemblies has been assigned as excited-state energy transfer. The two most important mechanisms for energy transfer are electron superexchange (Dexter)34 and through-space dipole-dipole coupling (Fiirster).35 Dexter energy transfer is subject to a distance dependence that falls off as e'2R due to its reliance on orbital overlap. As such, it is usually relegated to covalently linked systems in which the donor and acceptor are in close proximity (S 5 A)”37 Due to the structural and electronic similarities between the CrRe3 series and the FeRe3 assemblies (Chapter 2), Dexter energy transfer has been ruled out as a possible quenching mechanism in the CrRe3 complexes. See the detailed comparison between the two energy transfer mechanisms given in Chapter 2 (2.4.2) for further details. In addition to the similarities between the CrRe3 and FeRe3 assemblies mentioned above, the applicability of the Ffirster mechanism in the CrRe3 series is supported by the moderate degree of spectral overlap that exists between the Rel-based 3MLCT emission and the Crm-based 4A2 —-) 4T2 absorption: this is depicted graphically in Figure 4-11. The overlap between 214 the 3MLCT —> [A] emission of the Alm-containing tmb (4) (blue trace), bpy (5) (green trace), and deeb (6) (red trace) analogues with the 4A2 —> 4T2 absorption of Cr(phacac)3 (black trace) is shown, with the bpy derivative possessing the greatest amount of spectral overlap. The area ascertained from the product of the normalized emission profile of an AlRe3 analogue with the extinction coefficient spectrum of Cr(phacac)3 was used to calculate the reported spectral overlap values. Forster theory predicts that as the spectral overlap (J) increases across a structurally homologous series, the observed rate constant for energy transfer should increase by that same factor. The magnitude of spectral overlap (J) shows excellent agreement with the trend in k,.,, values of the CrRe3, with J values of 6.92 x 10'16 (k,,, = 1.5 x 108 s"), 8.07 x 10'16 (kos. = 2.1 x 108 s"), and 6.28 x 10''6 (ks.s = 1.4 x 108 s") M"cm3 for the tmb (1), bpy (2), and deeb (3) analogs, respectively. It is quite clear that the overlap requirement in these systems is satisfied, which strongly implicates Férster transfer as the dominant quenching mechanism in the CrRe3 complexes. 215 Absorbance (normalized) (pazneuuou) Ansuarul uoyssgwg 22 20 l8 16 14 Energy (x 103 cm'l) Figure 4-11. Overlay of the emission spectra of [Al(pyacac)3(Re(tmb)- (C0)3)3](0Tf)3 (4. blue), [A1(Pyacac)3(Re(bPY)(C0)3)3](OTfls (5, green), and [Al(pyacac)3(Re(deeb)(CO)3)3](OTf)3 (6, red) with the electronic absorption spectrum of Cr(phacac)3. Data were acquired in room-temperature CH2C12 solution. 4.4.3 Comparison with the CuRez Assemblies. A plot of the spectral overlap between the Rel-based 3MLCT emission and the Cull-based ligand- field absorptions for the CuRez assemblies was reported and discussed in Chapter 3 (Figure 3-10). The ground state absorption spectrum of Cu(phacac)2 in CHzClz solution shows a two-band structure, with these bands representing four discrete d-d transitions. Table 4-2 gives the spectral overlap integrals, energy transfer rate constants (kEnT), and kEnT/k, values for the CuRez and the CrRe; families. As discussed in Chapter 3, kEnT/k, values are used as the energy transfer rate constants due to variations in kr for the Rel donor moieties. Dividing kEhr by kr essentially normalizes out the variation in the radiative component and allows for sole evaluation of the 216 energy transfer coupling. As can be seen from Table 4-2, the differences in spectral overlap (J) among the CrRe3 series directly corresponds to the observed trend in the kEnT/k, values, but an analogous correlation for the CuRez series is not observed. In addition to spectral overlap considerations in the CuRez series, variations in through-space coupling (1(2) of the 3MLCT transition moment dipoles with the four separate ligand-field absorption dipoles of the CulI acceptor had to be taken into account to interpret the dipolar energy transfer dynamics (Chapter 3). It was found that the two CuRez complexes possessing the lowest energy 3MLCT excited states (dclb and deeb) possess very favorable energetic (J) and through-space interactions (K2) with the lowest energy Cuu d-d transition, which in turn yielded the fastest observed energy transfer rate constants for the two derivatives (Table 4-2). The direct dependence of the trend in kEnT/k, with the amount of spectral overlap seen for the CrRe3 complexes essentially validates the analysis of the Cull-containing complexes presented in Chapter 3. This validation comes from the fact that the 4A2 —) 4T2 ligand-field acceptor transition derives from a single configuration of the Cr"I ion, which results in identical through-space interactions with the three 3MLCT emission dipoles within the series and produces a direct spectral overlap dependence of the energy transfer rate constants. 217 Table 4-2. Calculated Overlap Integral Values (J) and Energy Transfer Rate Constants for the CrRe3 and CuRez Complexes. kEnr compound J a (X 108 s' 1)b anT/kr [Cr(pyacac)3(Re(tmb)(CO)3)3](OT1)3 (1) 6.92 1.50 650 [Cr(pyacac)3(Re(bpy)(CO)3)3](OTf)3 (2) 8.07 1.70 718 [Cr(pyacac)3(Re(deeb)(CO)3)3](OTf)3 (3) 6.28 1.40 452 [Cu(pyacac)2(Re(tmb)(CO)3)2](OT1)2c 4.29 0.67 208 [Cu(pyacac)2(Re(dmb)(CO)3)2](OTf); 5.14 1.18 295 [Cu(pyacac)2(Re(bpy)(CO)_~,)2](OT1)2c 5.52 1 . 18 347 [Cu(pyacac)2(Re(dclb)(CO)3)2](OTf);c 6.68 1.71 658 [Cu(pyacac)2(Re(deeb)(CO)3)2](OTf);c 7.09 1 .96 700 alReported in units of 10'16 M'lcm3. bkEnT = kob,(CrRe3) — kob, (AlRe3). °These data were reproduced from Chapter 3 for comparison purposes. 4.5 Conclusions The synthesis and photophysical properties of a series of donor- acceptor complexes based on Rel-bpyridine donors and Crm-acac acceptors have been described. Steady-state and time-resolved emission spectroscopies indicated that the strongly emissive Rel-based 3MLCT excited state was significantly attenuated when compared to model Ill 1111 complexes in which the Cr center had been replaced by A . Observation 111 of enhanced 2E —) 4A2 emission of the Cr acceptor in complex 2 in frozen 218 4:1 EtOH/MeOH unequivocally identified energy transfer as the 3MLCT quenching mechanism. The favorable overlap between the 3MLCT donor emission and the 4A2 —> 4T2 acceptor absorption profile, coupled with a ca. 10 A donor-acceptor separation, allowed for an assignment of F6rster (dipolar) energy transfer as the dominant excited-state reaction mechanism. Lastly, enhancement of the Cr"I acceptor emission observed in complex 2 gives direct evidence for energy transfer as the 3MLCT excited state quenching mechanism assigned in all of the Rel-based polynuclear donor- acceptor assemblies discussed throughout this dissertation. 219 4.6 References and Notes (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13) (14) (15) (16) Ftirster, T. Discuss. Faraday Soc. 1959, 27, 7. Van Der Meer, B. W.; Coker, G. 1.; Chen, S.-Y. Resonance Energy Transfer, Theory and Data; VCH Publishers: New York, 1994. Belford, R. L.; Carmichael, J. W., Jr. J. Chem. Phys. 1967, 46, 4515. Juban, E. A.; McCusker, J. K. J. Am. Chem. Soc. 2005, 127, 6857. Perrin, D. D.; Armarego, W. L. F. Purification of Laboratory Chemicals; 3rd ed.; Pergamon Press: Oxford, U. K., 1988. Mackay, L. G.; Anderson, H. L.; Sanders, J. K. M. J. Chem. Soc., Perkin Trans. 1. 1995, 18, 2269. Harada, Y.; Girolami, G. Polyhedron 2007, 26, 1758. Hino, J. K.; Della Ciana, L.; Dressick, W. J .; Sullivan, B. P. Inorg. Chem. 1992, 31, 1072. Caspar, J. V.; Meyer, T. J. J. Phys. Chem. 1983, 87, 952. Murakami, Y.; Nakamura, K. Bull. Chem. Soc. Jpn. 1966, 39, 901. Damrauer, N. H.; Boussie, T. R.; Devenney, M.; McCusker, J. K. J. Am. Chem. Soc. 1997, 119, 8253. Claude, J. P.; Meyer, T. J. J. Phys. Chem. 1995, 99, 51. Parker, C. A.; Rees, W. T. Analyst (London) 1960, 85, 587. DeWitt, L.; Blanchard, G. J.; LeGoff, E.; Benz, M. E.; Liao, J. H.; Kanatzidis, M. G. J. Am. Chem. Soc. 1993, 115, 12158. Origin; 7.5 ed.; OriginLab Corp.: Northhampton, MA, 1991 - 2004. Damrauer, N. H.; McCusker, J. K. Inorg. Chem. 1999, 38, 4268. 220 (17) (18) (19) (20) (7-1) (22) (23) (24) (25) (26) (27) (28) (29) (30) Dattelbaum, D. M.; Omberg, K. M.; Schoonover, J. R.; Martin, R. L.; Meyer, T. J. Inorg. Chem. 2002, 41, 6071. Dattelbaum, D. M.; Martin, R. L.; Schoonover, J. R.; Meyer, T. J. J. Phys. Chem. A. 2004, 108, 3518. Argazzi, R.; Bertolasi, E.; Chiorboli, C.; Bignozzi, C. A.; Itokazu, M. K.; Murakami Iha, N. Y. Inorg. Chem. 2001, 40, 6885. Sacksteder, L.; Zipp, A. P.; Brown, E. A.; Streich, J .; Demas, J. N.; DeGraff, B. A. Inorg. Chem. 1990, 29, 4335. Worl, L. A.; Duesing, R.; Chen, P.; Della Ciana, L.; Meyer, T. J. J. Chem. Soc., Dalton Trans. 1991, 849. Forster, L. S. Chem. Rev. 1990, 90, 331. Kirk, A. D. Chem. Rev. 1999, 99, 1607. Endicott, J. F.; Ramasami, T.; Tamilarasan, R.; Lessard, R. B.; Ryu, C. K.; Brubaker, G. R. Coord. Chem. Rev. 1987, 77, 1. Striplin, D. R.; Crosby, G. A. Chem. Phys. Lett. 1994, 221, 426. Kestell, J. D.; Williams, Z. L.; Stultz, L. K.; Claude, J. P. J. Phys. Chem. A. 2002, 106, 5768. Murray, R. W.; Hiller, L. K. Anal. Chem. 1967, 39, 1221. Rehm, D.; Weller, A. Isr. J. Chem. 1970, 8, 259. The form of the Rehm-Weller equation employed to characterize the oxidative quenching process is AGETox = Eox (Rel/H) - Bred (Crmm) - E00, where E00 is the energy gap between the ground and lowest energy 3MLCT excited state of the donor. The small work term typically employed for bimolecular systems was not included in the analysis. Krause, R. A.; Trabjerg, I.; Ballhausen, C. A. Acta. Chem. Scan. 1970, 24, 593. 221 (31) (32) (33) (34) (35) (36) (37) Chen, S.; Porter, G. B. J. Am. Chem. Soc. 1970, 92, 2189. Atanasov, M. A.; Schonherr, T. Inorg. Chem. 1990, 29, 4545. Yardley, J. T.; Beattie, J. K. J. Am. Chem. Soc. 1972, 94, 8925. Dexter, D. L. J. Chem. Phys. 1953, 21 , 836. F ’o'rster, T. Annalen der Physik 1948, 2, 55. Brown, W. R.; O'Boyle, N. M.; McGarvey, J. J .; Vos, J. G. Chem. Soc. Rev. 2005, 34, 641. Goeb, S.; De Nicola, A.; Ziessel, R.; Sabatini, C.; Barbieri, A.; Barigelletti, F. Inorg. Chem. 2006, 45, 1173. 222 Chapter 5. Spin Dependent Energy Transfer: CoRe3 and NiRez Complexes. 5.1 Introduction The area of focus contained in this chapter is on understanding the role of electronic spin in determining the rate of excited-state energy transfer in polynuclear transition metal-based donor-acceptor complexes. In addition to possessing favorable thermodynamics, excited-state energy transfer reactions, just as in simple bimolecular reactions, must conserve the total spin angular momentum of the reactants and products throughout the course of the reaction. From vector physics, the possible total spin angular — momentum (lsrl) values for a combination of two complexes with spins fi fl 1811 and 1 21 are “fl S1 +S2 1345—1 ,..., 151 - Sz . For example, if 1 11 2 and 1821 = 1 then the ISTI values are 3, 2, and 1. A generic representation of an excited-state energy transfer reaction between an _ — * SD excited donor with and a ground state acceptor with 1SA| is given in equation 5-1: 223 T “—T kn _— T 1801+ s, ——ET—>1s,,1+ s, (5-1) In addition to favorable thermodynamics, the reaction defined by equation 5- fl 1 must possess a common IST1 between the reactants and products. A classic example of spin conservation involved in energy transfer reactions is the quenching of 3MLCT (metal-to-ligand charge transfer) excited-states of low-spin d6 polypyridyl complexes.l Figure 5-1 illustrates total spin angular momentum conservation in an excited-state energy transfer reaction between # :1: SD a d6 polypyridyl-based 3MLCT state ( = 1), which describes the excited state properties of the Re1 donors utilized throughout this dissertation, and an fl fl 0; molecule that possesses a triplet ground state (1 AI = l). A ISTI value of 1 is conserved between the reactant complexes and product complexes involved in the energy transfer process, indicated by the connecting arrow, and therefore the excited state reaction is spin-allowed. Investigating this fundamental property of excited-state reactivity utilizing transition metal- based donor-acceptor systems, will be the focus of this chapter. 224 * Energy Transfer * D ——-A > D A 80:1 SA=1 80:0 SA=1 ST=2,1,0 ST: Figure 5-1. Illustration of total spin angular momentum conservation in the energy transfer quenching of a 3MLCT excited-state by Oz. The first section of this chapter will focus on the synthesis and photophysical properties of a tetranuclear complex with the general formula [Co(pyacac)3(Re(bpy)(CO)3)3](OTf)3. The CoIII metal center is covalently attached to three fac-Re(bpy)(CO)3 moieties through three pyridyl- acetylacetonate bridging ligands (Figure 5-2). It was observed that the relaxation dynamics of the Rel-based 3MLCT excited-state in the CoRe3 assembly is similar to a structurally analogous AlRe3 model complex. Based on spectral overlap determinations, the 1A1 ——> 'T1 ligand-field state of the Co"1 acceptor shows very favorable energetic matching with the Rel-based 3MLCT donor emission, which indicates the Col" ligand-field state should serve as an efficient quenching pathway out of the ReI excited-state. These results contrast with the excited-state dynamics described in the previous chapter for the isostructural CrRe3 complex - [Cr(pyacac)3(Re(bpy)(CO)3)3]- 225 (OTt)3. The difference in reactivity can be explained by considering the role of spin conservation in the energy transfer reaction: whereas the 4A2 ——> 4T2 ligand-field transition of the Crlll center in the CrRe3 assembly yields a spin- allowed energy transfer pathway out of the Rel-based 3MLCT excited-state. fl Evaluation of IST1 for excited-state energy transfer in the CoRe3 system shows that coupling to the Com-based 'Al -—+ 1T1 ligand-field transition represents a spin-forbidden relaxation pathway. This represents the first example of an energy transfer reaction in which spin polarization is dictating reactivity. 226 Figure 5-2. Structure of the CoRe3 assembly. The second section of this chapter will focus on the structural and photophysical properties of a chromophore-quencher molecule with the general formula [Ni(pyacac)2(THF)2(Re(bpy)(CO)3)2](OTDZ (where pyacac = 3-(4-pyridyl)—acetylacetonate, THF = tetrahydrofuran, bpy = 2,2’- bipyridine, and OTf = CF3SO3'). The NiII metal center is covalently attached to two fac-Re(bpy)(CO)3 moieties through two pyridyl-acetylacetonate bridging ligands (Figure 5-3). The X-ray structural data for the NiRez system shows two THF solvent molecules axially coordinated to the NilI 227 metal center that produces a pseudo-octahedral NiO6 core (3A2). It was observed that the Rel-based 3MLCT excited-state for the NiRez assembly is significantly quenched when compared to an electronically benign BeRez model complex. Based on spectral overlap considerations, the 3A2 —> 3T2 ligand-field state of NiIl was assigned as the thermodynamically viable quenching pathway for the Rel-based 3MLCT excited state. Measuring the 3MLCT emission properties of the NiRez compound in CHzClz and in THF solutions revealed an axial ligand dissociation mechanism. The NiRez assembly, and particularly chemistry associated with the Ni" ion, represents an excellent example of the utility of basic coordination chemistry in controlling excited-state reactivity. Figure 5-3. Structure of the NiRez assembly. 228 5.2 Experimental Section 5.2.1 Synthesis and Characterization General. All solvents used were purified and dried according to previously reported methods.2 Spectroscopic grade CHzClz was used for all photophysical measurements; the solvent was dried under CaHz reflux until no water was detected by 1H NMR and the solvents used in the emission experiments were degassed using freeze-pump-thaw techniques. 3-(4- pyridyl)-2,4-pentanedione,3 Ni(phacac)2,4 and fac-[Re(bpy)(CO)3(4- Etpy)](PF(,)5 (4-Etpy = 4—ethylpyridine) were prepared following literature procedures. The synthetic procedure for [Re(bpy)(CO)3(pyacac)](OTf) is presented in Chapter 4. The synthetic procedure for [Al(pyacac)3(Re(bpy)- (CO)3)3](OT1)3 is presented in Chapter 2 and the synthetic procedure for [Be(pyacac)2(Re(bpy)(CO)3)2](OTf)2 is presented in Chapter 3. 3-phenyl— 2,4-pentanedione was purchased from TCI America and Ni(1,5- cyclooctadiene); was purchased from Sigma-Aldrich. Elemental analyses and FT-IR data were obtained through the analytical facilities at Michigan State University. Mass spectra were obtained through the analytical facilities at The University of South Carolina. [Co(pyacac)3(Re(bpy)(CO)3)3](OT03. Amounts of 11 mg (0.04 mmol) of Co(acetate)2-4HzO and 98 mg (0.13 mmol) of [Re(bpy)(CO)3- 229 '. -3 4‘? 1 (pyacac)](OTf) were dissolved in 10 mL of ethanol and heated at 100°C for 20 min in air. The reaction solution turned a dark green color and a yellowish precipitate formed in the solution. The green filtrate was collected, the solvent removed under vacuum, and the resulting green solid placed under vacuum overnight. The product was recrystallized several times from acetonitrile/ether (l :1 v/v). Yield: 25 mg (25%). Anal. Calcd for C72H54N9F9024S3C0Re3-2H20: C, 36.80; H, 2.49; N, 5.36. Found: C, 36.40; H, 2.47; N, 4.98. MS: [ESL m/z (rel. int.)]: 622.3 (5) 1[CotpyacacMRetbpyXCOh)3113*, 1007.9 (2) {1Co(pyacac)s(Re(bpy)- (CO>3)31(0T0}2*. [Ni(pyacac)2(Re(bpy)(CO)3)2](OTf)2. In 30 mL of THF were dissolved 285 mg (0.379 mmol) of [Re(bpy)(CO)3(pyacac)](OTf) in a N2- filled drybox. 52 mg (0.189 mmol) of Ni(1,5-cyclooctadiene)2 were then added and the solution turned a deep red color, along with the formation of a green precipitate. The reaction mixture was allowed to stir for 24 hrs in the drybox, after which time the reaction mixture was removed from the drybox and the green precipitate collected. The green solid was washed with THF, placed under vacuum overnight, and then recrystallized several times from acetontrile/ether (1:1 v/v) to yield the product. X-ray quality crystals were obtained by slow diffusion of THF into an acetonitrile solution of the 230 01:“ compound. Yield: 41 mg (14%). Anal. Calcd for C43H36N6F6016S2NiRe2: C, 36.91; H, 2.32; N, 5.38. Found: C, 35.86; H, 2.61; N, 4.99. IR (KBr, v(CO) cm“): 2033 s, 1922 s. MS: [ESL m/z (rel. int.)]: 632.0 (35) {[Ni(pyacac)2(Re(bpy)(CO)s)2]}2‘1 1413.0 (1) {[Ni(pyacac)2(Re(bpy)- (COthOTDVi Tris(3-phenyl-acetylacetonato)cobalt(III), Co(phacac)3. The synthesis of this compound has been previously reported by a different method.6 Amounts of 85 mg (0.534 mmol) of Co(acetate)2-4HZO and 205 mg (1.16 mmol) of phacac were dissolved in 10 mL of H20 and heated at 100°C for 20 min in air. A dark green precipitate that formed in the reaction solution was collected, washed with water, and recrystallized from hot ethanol. Yield: 100 mg (50%). Anal. Calcd for C33H33O6Co-0.3EtOH: C, 67.45; H, 5.86. Found: C, 67.05; H, 5.45. 5.2.2 Physical Measurements X-ray Structure Determinations. Single-crystal X-ray diffraction data for [Ni(pyacac)2(TI-IF)2(Re(bpy)(CO)3)2](OTf)2 were acquired at the X- ray facility of Michigan State University. Diffraction data were collected on a Siemens SMART diffractometer with graphite-monochromatic Mo K01 radiation (A. = 0.71073A). Data were collected at -100 °C by using an Oxford Cryosystems low temperature device. Crystallographic data are summarized 231 in Table 5-1; selected bond distances and angles are listed in Table 5-2. Lattice parameters were obtained fiom least-squares analyses and data were integrated with the program SAINT.7 The integration method employed a three dimensional profiling algorithm and all data were corrected for Lorentz and polarization factors, as well as for crystal decay effects. The absorption correction program SADABS8 was employed to correct the data for absorption effects. The structures were solved by direct methods and expanded using Fourier techniques. All structure calculations were performed with the SHELXTL 6.12 software package.9 Anisotropic thermal parameters were refined for all non-hydrogen atoms. Hydrogen atoms were localized in their calculation positions and refined by using the riding model. Further details concerning the structure determinations may be found in Supporting Information. Electronic Absorption and Steady-State Emission Spectroscopies. Extinction coefficients for all compounds were acquired in CHzClz solutions using a Varian Cary 50 UV-Visible spectrophotometer. Steady-state emission spectra were acquired using a Spex F luoromax fluorimeter and corrected for instrumental response using a NIST standard of spectral irradiance (Optronic Laboratories, Inc., OL220M tungsten quartz lamp).10 Spectra were acquired on samples dissolved in thoroughly degassed CHzClz 232 under optically dilute conditions (o.d. ~ 0.1) and sealed under an argon atmosphere in 1 cm path length quartz cuvettes. Radiative quantum yields ((1),) were determined relative to fac- [Re(bpy)(CO)3(4-Etpy)](PF6) (ch, = 0.18 in Cinch).5 Quantum yields were calculated according to equation 5-2, 2 (I) (I) (Iunk/Aunk){nunk] unk Std (Istd /Astd) ”std (SI-2) where (Dunk and (1),“, are the radiative quantum yields of the sample and the standard, respectively, Iunk and Istd represent the areas of the corrected emission profiles for the sample and the standard, Aunk and Astd are the absorbance values of the sample and the standard at the excitation wavelength, and nunk and 1],“! correspond to the indices of refraction of the sample and standard solutions (taken to be equal to the neat solvents). An excitation wavelength of 355 nm was used for all the compounds studied in this chapter. Time-Resolved Emission Spectroscopy. Nanosecond time-resolved emission data for [Co(pyacac)3(Re(bpy)(CO)3)3](OTf)3, [Al(pyacac)3(Re- (bPY)(C0)3)3](0Tf)3. and [Be(PyacaC)2(Re(bp)')(C0)3)2](OTf)2 were collected using a NszAG-based laser spectrometer that has been described previously.10 Data were acquired at room temperature in thoroughly 233 degassed CH2C12 solutions having absorbances of ~0.1 at A.,, = 355 nm. Samples were sealed under an argon atmosphere in 1 cm path length quartz cuvettes. The decay traces correspond to an average of 250 shots of the signal probed at the emission maximum of each compound. Picosecond time-resolved emission data for [Ni(pyacac)2(THF)2(Re- (bpy)(CO)3)2](OTf)2 were collected using a time-correlated single photon counting (TCSPC) apparatus that has been described previously.ll Data were acquired in thoroughly degassed CHzClz solutions having absorbances of ~0.1 at A1,, = 370 nm. The sample was sealed under an argon atmosphere in 1 cm path length quartz cuvette. The reported decay trace corresponds to a signal average of three data sets, with each data set resulting from ca. 30 minutes of data acquisition time. Data were fit using the OriginPro 7.5 software package. '2 5.3 Results and Discussion for [Co(pyacac)3(Re(bpy)(CO)3)3](OTf)3 5.3.1 Synthesis and Characterization. The interest in the CoRe3 system was to synthesize a donor-acceptor complex with similar overall charge and structural characteristics to the CrRe3 compounds presented in Chapter 4. The 4A2 —+ 4T2 ligand-field acceptor state of Crm(acac)3 is replaced by a 1A, -—9 'Tl transition in the case of the CoRe3 system, allowing the comparison 234 of energy transfer rates simply on the differences in spin-state associated with the acceptor. As with the F eRe3 (Chapter 2) and CrRe3 (Chapter 4) complexes, the formation of the tetranuclear CoRe3 assembly was facilitated by the low steric crowding afforded by the 1200 separation of the three pyacac ligands 111 attached to the Co center. ESI-MS data for [Co(pyacac)3(Re(bpy)- (CO)3)3](OTf)3 in acetonitrile solution are consistent with the formation of the desired complex, with m/z peaks observed for [Co(pyacac)3(Re(bpy)- (count and lCotpyacachtRetbpyXCOhhl2". 5.3.2 Electronic Absorption Spectroscopy. The electronic absorption spectra of [Co(pyacac)3(Re(bpy)(CO)3)3](OTf)3 and [Al(pyacac)3(Re(bpy)- (CO)3)3](OTf)3 were acquired in room-temperature CHzClz solutions and are shown in Figure 5-4. Transitions associated with the AlIII complex were assigned in Chapter 2, with the shoulder centered at 360 nm assigned to the Rel-based 1A. -—> 1MLCT (t2g —> 11:. (bpy’)) transition (Figure 5-4). The AlRe3 complex allows for analysis of the ground-state absorption behavior of the ReI donor moieties without overlapping with the Com-based charge- transfer transitions associated with the CoRe3 assembly. 235 Absorbance (au) 300 400 500 600 Wavelength (nm) Figure 5—4. Electronic absorption spectra of [Co(pyacac)3(Re(bpy)- (CO)3)31(OTD3 (red trace) and [Al(pyacac)3(Re(bpy)(CO)3)3](OTf)3 (blue trace) acquired in room-temperature CHzClz solution. The presence of Co"1 in the CoRe3 complex gives rise to a new, broad absorption feature occurring at approximately 340 nm and overlapping with the Rel-based charge transfer band (Figure 5-4). Figure 5-5 (Top) shows the absorption spectrum of Co(phacac)3 that exhibits the 340 nm transition which is assigned to a 1A] —+ lLMCT (7t (acac) —> eg*) transition. Co(phacac)3 also possesses a transition at km, = 605 nm (Figure 5-5, Bottom) with a measured extinction coefficient of 193 M'lcm'l. This band is easily assigned to the 1A, —> 1T] transition, the lowest-energy spin-allowed d-d band of a pseudo-octahedral CoIn species.“3 It should also be noted that the IA. —> IT] transition occurs at approximately the same energy as the 4A2 -—> 4T2 transition of Cr(phacac)3 reported in Chapter 4, and should have similar spectral overlap with 3MLCT emission of the Rel-bpy fluorophore. 236 Comparing the effect on the rate of 3MLCT relaxation incurred by the Co"l and Cr"! ligand-field acceptor states will be the focus of the discussion section. 35000- 30000- 25000- 20000- 15000. 10000- 5000- 0 Molar Absorptivity (M'lcm'l) 250 300 350 400 450 500 550 Wavelength (nm) Figure 5-5. Top. Electronic absorption spectrum of Co(phacac)3 showing the higher energy charge transfer and organic-based transitions. Bottom. Electronic absorption spectrum of a concentrated solution of Co(phacac)3 showing the mid-visible 1A1 —> 1T1 ligand-field transition. Both spectra were acquired in room-temperature CHzClz solution. 237 Figure 5-5 (cont’d) 200 l60~\ 120-\ 4:. O // Molar Absorptivity (M'lcm'l) 00 O C 500 600 700 800 900 1000 Wavelength (nm) O r 5.3.3 Time-Resolved Emission. Insight into the Rel-based 3MLCT excited state relaxation behavior of the CoRe3 and AlRe3 complexes were obtained through nanosecond emission lifetime measurements. Nanosecond time- resolved emission data in room-temperature CHzClz solution for [A1(pyacac)3(Re(bpy)(CO)3)3](0T03 and [C0(pyacac)s(Re(bpy)(CO)3)3l- (OTf)3 are given in Figure 5-6. As discussed in Chapter 2, emission from the AlRe3 complex could be fit to a single-exponential model with rob, = 530 i 20 ns, with corresponding radiative (k,) and non-radiative (km) decay rate constants of 3.0 i 0.1 x 105 s'1 and 1.6 d: 0.1 x 106 s", respectively, corresponding to ground-state recovery for the Rel-based 3MLCT excited state. The time-resolved emission data for [Co(pyacac)3(Re(bpy)(CO)3)3]- (OTf); could likewise be fit to a single-exponential model with rob, = 595 i 20 ns. The 3MLCT emission lifetime in the CoRe3 assembly is ~60 ms 238 longer than the electronically benign AlRe3 model compound. The slightly longer observed time constant is not completely understood, but the data do illustrate the absence of 3MLCT quenching in the CoRe3 system. Additional CoRe3 material needs to be synthesized and the time constants re-measured before the longer emission lifetime in the CoRe3 complex can be scrutinized. 50- Al 50- Co A 40» A 40- > > E 30— E 30- 32‘ .E‘ E 20 - a 20 - 8 2 t: t: "‘ 10 - "‘ 10 - 0 J l l I l 0 I l I I 0 1 2 3 4 0 l 2 3 4 Time (x103 ns) Time (x103 ns) Figure 5-6. Lefi‘. Nanosecond time-resolved emission data for [Al- (pyacac)3(Re(bpy)(CO)3)3](OTD3 (robS = 530 i 20 ns). The data were originally presented in Chapter 2 and are reproduced here for comparison purposes. Right. Nanosecond time-resolved emission data for [Co(pyacac)3(Re(bpy)(CO)3)3](OTD3 (robs = 595 i 20 ns). Both traces (Al and Co) were acquired in deoxygenated room-temperature CH2C12 solution monitoring at the emission maximum (565 nm). The solid red lines correspond to fits to mono-exponential decay models for both complexes. 5.3.4 Spin-Dependent Dipolar Energy Transfer. As mentioned in the Introduction, in addition to possessing a downhill thermodynamic driving force (i.e. spectral overlap), the total spin angular momentum must be conserved during an energy transfer event. In Chapter 4, significant 239 quenching of the Rel-based 3MLCT excited state was observed for a 3 MLCT structurally similar CrRe3 complex ( 1'0133 = 4.8 :t 0.1 ns) that possesses a similar ligand-field absorption profile and maximum as the CoRe3 assembly, and therefore should produce similar spectral overlap with the 3MLCT emission profile and manifest corresponding rates of 3MLCT relaxation. The similarity in emission lifetimes between the CoRe3 and AlRe3 complexes indicates that the CoIll center is not engaging in excited- state reactivity with the Rel-based 3MLCT state, even though F firster energy transfer quenching is thermodynamically favorable due to a significant amount of spectral overlap observed between the 1A1 —> IT, ligand-field transition of Co(phacac)3 and the Rel-based 3MLCT donor emission spectrum (Figure 5-7). The only differing characteristic that could be giving rise to the large variance observed for the 3MLCT relaxation kinetics between the CoRe3 and CrRe3 assemblies is the spin-state of the Com ('Al) and CrIII (4A2) metal centers. 240 .—- —n O I l I I I 0 500 600 700 800 900 1000 Wavelength (nm) Molar Absorptivity (normalized) (pozueuuou) Ansuarul uogssrurg Figure 5-7. Overlay of the emission spectrum of [Al(pyacac)3(Re(bpy)- (CO)3)3](OTf)3 (blue trace) with the electronic absorption spectrum of Co(phacac)3 (red trace). Figure 5-8 shows spin conservation diagrams between the Rel-based 3MLCT excited state with the Cr"I (Top) and Com (Bottom) acceptor cores in 111 the CrRe3 and CoRe3, respectively. The Cr ground state in the CrRe3 complex is 4A2 (18A1= 3/2), and when coupled with the 3MLCT excited _. SIS state of the Re1 donor ( = 1) generates total spin angular momentum (18T1) values of 5/2, 3/2, and 1/2. As an energy transfer (EnT) reaction occurs out of the] Rel-based 3MLCT excited state to produce the Crm-based 4A2 —> 4T21igand-field transition, an 1ST1 = 3/2 is generated on the products side of the diagram from the coupling of the resulting ReI ground state 241 # A = 3/2). As can been (lSDl = 0) with the Crm-based 4T2 excited state ( fl seen in Figure 5-8, an 1311 value of 3/2 between the reagents and the products is conserved, which provides the spin-allowed energy transfer pathway. Conversely, Figure 5-8 (Bottom) shows the corresponding process for the CoRe3 assembly. The C0111 ground state in the CoRe3 complex is IA] (lSAl = O), and when coupled with the 3MLCT excited state of the ReI —' an: SD # donor ( = 1) generates a total spin angular momentum (ISTj ) value of III 1 on the reagents side. The acceptor state of the Co ion is the 1A1 —> 1T1 fl 32. ligand-field state ( = 0), and when coupled with the ReI ground state — _——' (ISDI = 0) an 18T1 = 0 is produced on the products side. As can'be seen in — # Figure 5-8 (Bottom), a mutual IST1 value between the reagents (ISTI = 1) —’ and the products (IST1 = O) is not conserved, therefore the energy transfer reaction is spin-forbidden. The spin statistics outlined in Figure 5-8 illustrates the theoretical framework behind the shorter 3MLCT emission 242 lifetime observed in the CrRe3 complex compared to the CoRe3 assembly. The results discussed throughout this section highlight the control one can have on excited-state reactivity by simply changing the spin state of a single reagent. EnT - II III I III * (bpy )Re mCr spimauowed (bpy)Re~’W(Cr ) 80:1 SA=3/2 80:0 SA=3/2 S1- = 5I2, 3/2, 1I2 S1- = 3I2 EnT b - R "MC ||| .................. , b Re'wv Col" * ( py) e o spin-forbidden ( py) ( ) 80:1 SA=O 30:0 SA=O ST = 1 ST = 0 Figure 5-8. Top. Spin conservation diagram illustrating the spin- allowed energy transfer pathway between the 3MLCT excited-state of the ReI donor and the CrlII acceptor. Bottom. Spin conservation diagram illustrating the spin-forbidden pathway between the 3MLCT excited- state of the ReI donor and the CoIII acceptor. 243 5.4 Results and Discussion for [Ni(pyacac)2(THF)2(Re(bpy)(CO)3)2]- (OTfh 5.4.1 Synthesis and Characterization. The interest in the NiRe2 complex was to explore the effect solvent coordination can play on Forster energy transfer dynamics in covalently attached donor-acceptor complexes, and whether the rate of energy transfer can be controlled by simply varying the coordination strength of the solvent. Unlike the square-planar Cu"(acac)2 center utilized in the CuRez series (Chapter 3), square-planar Ni"(acac)2 possesses a large propensity to bind molecules in the axial positions.”'7 These solvent interactions change the ligand-field environment from 4- to 6- coordinate, and more importantly to the work in this chapter, the spin-state of the Ni" center.‘8 The formation of the trinuclear NiRez assembly was facilitated by the low steric crowding afforded by the 180° separation of the two pyacac ligands attached to the Ni11 center. ESI-MS data for [Ni(pyacac)2(THF)2(Re- (bpy)(CO)3)2](OTf)2 in acetonitrile solution are consistent with the formation of the desired complex, with m/z peaks observed for [Ni(pyacac)2(Re- (bpy)(c0)3)212* and [NitpyacachtRe 20(1)] 4075 u(Mo K01)/cm" 3.738 Rim 0.0775 R1a 0.0509 wR2b 0.1351 GOF 0.995 3R1 = sum. chll/ZIFOI. waz = [2w(F02 — Fc2)2/2w(F02)2]'/2, w = l/[O' 2(13,") + (aP)2 + bP], where P = [F.3- + 21:31/3. 247 Table 5-2. Selected Bond Distances (A) and Angles (deg) for [Ni(pyacac)2(THF)2(Re(bpy)(CO)3)2](OTDZ. Bond Distances (A) Ni( 1 )—O( 1) 1 .971(5) Ni(1)—O(2) 1.981(5) Ni(1)—O(1A) 1.971(5) Ni(1)—O(2A) 1.981(5) Re( 1 )—N( 1) 2.23 8(5) Re( 1 )—N(2) 2.152(6) Re(l)—N(3) 2.112(10) Re(1)—C(12) 1.903(12) Re(1)—C(13) 1.901(10) Re(1)—C(14) 1.912(9) Ni(1)--- O(6) 2.155(6) Ni(1)--- O(6A) 2.155(6) Ni(1)°"Re(1) 9.868 Ni(1)°°°Re(1A) 9.868 Bond Angles (deg) O(1)—Ni(1)—O(2) 89.6(2) O(1)—Ni(l)—O(1A) 179.999(2) N(1)—Re(l)—N(2) 88.0(2) C(12)—Re(l)—N(1) 92.1(4) C(14)—Re(l)—N(1) 178.2(4) aplane Implane 2 87.539 bplane 1mplane 2 87.539 aPlane 1 is defined by atoms 0(1), 0(2), C(7), C(8), C(8), C(10), C(11); plane 2 is defined by atoms N(1), C(2), C(3), C(4), C(5), C(6). bGiven by the atoms that define the two planes in the adjacent pyacac ligand. 248 Figure 5-9. Drawing of the cation [Ni(pyacac)2(THF)2(Re(bpy)— (CO)3)2](OT1)2 obtained from single-crystal X-ray structure determination. Atoms are represented as 50% probability thermal ellipsoids. 5.4.3 Electronic Absorption Spectroscopy. The electronic absorption spectra of [Ni(pyacac)2(THF)2(Re(bpy)(CO)3)2](OTDZ and [Be(pyacac)2(Re- (bpy)(CO)3)2](OTf)2 were acquired in room-temperature CH3CN solutions and are shown in Figure 5-10. Transitions associated with these complexes were assigned based on previously reported analyses of ReI polypyridyl absorption spectra and”28 the ground state absorption data for the structurally similar CuRez complexes previously discussed (Chapter 3). The lowest energy absorption feature for BeRez is centered at 360 nm (a = 6,800 M'lcm") and is assigned as the Re[(bpy)-based 1A. —> lMLCT (t2g —-> rt* (bpy’)) transition. The BeRez complex allows for analysis of the ground— state absorption behavior of the Re1 donor moieties without overlapping with 249 the Niu-based charge-transfer absorptions associated with the NiRez analogue. These charge transfer transitions are very apparent in the NiRez absorption spectrum due to the sizable increase in the molar absorptivity values observed in the ~ 340 to 400 nm range. 50000 40000 ' 30000 ~ 20000 - 10000 - Molar Absorptivity (M'lcm'l) 250 300 350 400 450 500 550 Wavelength (nm) Figure 5-10. Electronic absorption spectra of [Ni(pyacac)2(THF)2(Re- (bPY)(C0)3)2](0Tf)2 (red trace) and [36(pyacaC)2(Re(bPY)(C0)3)21- (OTf); (blue trace) acquired in room-temperature CH3CN solution. In addition to the Rel-based charge transfer transition, the NiRez complex also possesses ligand-field transitions associated with the Nill core. The electronic absorption spectrum of Ni(phacac)2 in CHzClz solution, which serves as a model for the ligand-field transitions associated with the Ni11 center of the NiRez assembly is shown in Figure 5-11. The shape of these low energy Nin-based transitions with Max = 640 nm, as well as the observed amax = 17 M'lcm'l has been observed previously by Fackler and Cotton.29 250 N N O LI. 1 r t—l LII T U! r Molar Absorptivity (M'lcm'l) 8 O 500 600 700 800 900 Wavelength (nm) Figure 5-11. Electronic absorption spectra of Ni(phacac)2 acquired in room-temperature CHzClz solution. 5.4.4 Steady-State and Time-Resolved Emission. Emission spectra for [Ni(pyacac)2(THF)2(Re(bpy)(CO);)2](OT02 and [Be(pyacac)2(Re(bpy)- (CO)3)2](OT1)2 were obtained in deoxygenated room-temperature CH2C12 solution. The BeRe2 analogue represents an ideal structural model for the dynamics associated with the Re1 based 3MLCT emission, due to the inability of engaging in quenching dynamics incurred by the NiII core in the NiRez system. The emission profiles are given in Figure 5-12, and show the emission intensity for the NiRez is significantly quenched relative to the Ben-containing system. The spectral profile for the BeRe2 compound was first observed in Chapter 3, with the emission originating from the 3MLCT —-> 1A] phosphorescence.30 The emission maximum for both the Ni"- and Ben-containing derivatives is 565 nm suggesting similar emission characteristics for the NiRez complex. A radiative quantum yield ((1),) of 251 0.18 was calculated for the BeRe2 derivative (Chapter 3) when determined relative to [(bpy)Re(CO)3(4-Etpy)](PF6) ((1)r = 0.18 in CHzClz). The (I)r value exactly corresponds with the reported value of the mononuclear Rel(bpy) standard indicating the 3MLCT relaxation in the BeRe2 model complex can be solely assigned to the ReI donor moiety. A (I), value for the NiRez complex is analytically unreliable due to dissociation of the pseudo- octahedral NiO6 coordination environment in CHzClz solution. Figure 5-13A shows the emission spectrum of NiRez in CHzClz solution over a period of 27 hours, which shows a steady increase in the 3MLCT -—> 1A1 emission intensity occurring as the NiRez complex remains in the CHzClz solution over the course of the experiment. The increase in intensity is attributed not to dissociation of the Re(bpy)(CO)3(pyacac) donor moiety from the NiII metal center, as was seen for the F eRe3 (Chapter 2) and CuRez complexes (Chapter 3) due to residual amounts of H20 contained in the CH2C12, but loss of the axial THF ligands attached to the NiII center to generate a N106-x (x = 1 or 2) environment. This is confirmed by Figure 5-13B, which shows the emission intensity of the NiRez complex in THF solution remains constant when monitored over the course of 5 hours. If the NiRez complex was dissociating to generate Re(bpy)(CO)3(pyacac) species in solution, the dissociation process would be occurring to varying degrees regardless of 252 solvent. The possible NiO6.x structures formed in solution are discussed in the next section. Emission Intensity (normalized) Wavelength (nm) Figure 5-12. Corrected steady-state emission spectra for [Ni(pyacac)2- (THF)2(Re(bPY)(C0)3)2](OTf)2 (red tr8109) and [Be(PyacaC)2(Re(bPY)- (CO)3)2](OT1)2 (blue trace). Acquired in deoxygenated room- temperature CH2C12 solution following excitation at 355 nm. The emission profiles have been normalized with respect to the absorbance at 355 nm, and therefore the relative magnitudes of the signals are accurate representations of their relative intensities. 253 4000000- 3500000- 3000000- 2500000- 2000000- 1500000— 1000000- 500000- 0 Intensity (cps) 450 500 550 600 650 700 750 800 Wavelength (nm) 100000- 80000r 60000» 40000- lntensity (cps) 20000— O 500 550 600 650 700 Wavelength (nm) Figure 5-13. A. Corrected emission spectra of [Ni(pyacac)2(THF)2(Re- (bpy)(CO)3)2](OTf)2 in CH2C12 solution over a period of 27 hours (hex = 355 nm) exhibiting a steady increase in emission intensity. See text for further details. B. Corrected emission spectra of [Ni(pyacac)2(THF)2- (Re(bpy)(CO)3)2](OTf)2 in THF solution over a 5 hour time period (3.“ = 355 nm) showing a constant emission intensity over the course of data collection. Additional details concerning the excited-state states of the BeRe2 and NiRez complexes were obtained through nanosecond emission lifetime and time-correlated single photon counting (TCSPC) measurements. The nanosecond emission decay lifetime (robs) in room-temperature CHzClz 254 solution for [Be(pyacac)2(Re(bpy)(CO)3)2](OTf)2, which was originally reported and analyzed in Chapter 3, is shown on the'lefi side of Figure 5-14 and could be fit to a single-exponential model with rob, = 530 :1: 30 ns. The corresponding radiative (k,) and non-radiative (km) decay rate constants are 3.4 i 0.2 x 105 s'1 and 1.6 i 0.1 x 106 s", respectively, which were all consistent with the assignment of 3MLCT —-> 'A1 emission given in Chapter 3. A plot of the TCSPC data obtained in room-temperature CH2C12 solution for [Ni(pyacac)2(THF)2(Re(bpy)(CO)3)2](OTf); is shown on the right side of Figure 5-14. The decay trace could be fit with a bi-exponential model with 1'] = 5.0 i 0.4 ns and ”£2 = 25 i 4 ns, with the source for the bi-exponential behavior observed in the NiRe2 system proposed in the next section. Regardless of the bi-exponential behavior, the observed time constants for excited-state decay in the Nin-containing system are significantly larger than the corresponding BeRe2 model complex, indicating the presence of a very efficient quenching process stemming from a reaction between the Rel-based 3MLCT excited state and the Ni" core. 255 40- Be 1400~\ Ni 12001 9 30', ”g 1000—1‘11 é ' 3’ 800- 5 20-1 3? 1 g 1 2 600—1 +2 10_l E 400-. T 1 200:1 0 1 0 . . . . . . . . . 0.0 05 1:0 1:5 2:0 2:5 3:0 3:5 0 10 20 3o 40 50 60 7o 80 Time (ns) x 103 Time (ns) Figure 5-14. Lefi. Nanosecond time-resolved emission data for [Be(pyacac)2(Re(bpy)(CO)3)2](OTDZ (“cobs = 530 i 30 ns). Right. TCSPC data for [Ni(pyacac)2(THF)2(Re(bpy)(CO)3)2](OTf)2 (1:1= 5.0 :t 0.4 ns and 12 = 25 i 3 ns). Both traces (Be and Ni) were acquired in deoxygenated room-temperature CHzClz solution monitoring at the emission maximum (565 nm). The solid red lines correspond to fitting to a mono-exponential decay model for BeRe2 and to a bi-exponential decay model for NiReg. See text for discussion of the bi-exponential kinetics for NiRez. 5.4.5 Mechanistic Considerations: Dexter vs Ftirster Energy Transfer. Dexter energy transfer,“ which is described as a simultaneous double- electron exchange mechanism, requires orbital overlap between the donor and acceptor involved in the energy transfer.32 The X-ray structure data for [Ni(pyacac)2(THF)2(Re(bpy)(CO)3)2](OTf)2 (Figure 5—9) shows a Rel-"Ni" separation of nearly 10 A, a value that lies at the maximum of what is typically considered for an exchange-based process due to the exponential dependence (e'ZR) (R = donor-acceptor separation) of orbital overlap.33 The 10 A metal-metal distance, coupled with observed similarities in the v(CO) 256 frequencies of the NiRez and BeRe2 complexes (Section 5.2.1), largely negate the possibility of significant electronic coupling between the Re1 and Ni" subunits. This point is amplified by the computational results of Meyer and coworkers on similar ReI mononuclear complexes that show the 3MLCT wavefilnction would not exist on the bridging pyacac ligand,20 which drastically negates the possibility of the pyacac bridging ligands imparting significant electronic coupling between the ReI and NiII metal centers. This combination of the structural and electronic characteristics of the NiRez complex makes it unlikely that Dexter-type exchange is playing a dominant role in the energy transfer dynamics of these systems. Work discussed in Chapter 3 on the structurally similar [Cu(pyacac)2- (Re(bpy)(CO)3)2](OTt)2 assembly found Férster energy transfer to be the dominant excited-state quenching mechanism. The CuRez compound contains the same ReI donor species (ie. equivalent redox potentials and emission characteristics), pyacac bridging ligand, and similar donor-acceptor separations as the NiRez system. Although the results of the CuRez derivative can’t unequivocally assign the quenching mechanism for the NiRe2 complex, the findings do point to analogous excited-state reactivity out of the 3MLCT state of the Re1 donor. In addition, the spectral overlap between the Rel-based 3MLCT emission with the Nin ligand-field states also 257 implicates Féirster-type reactivity as the dominant excited-state quenching mechanism in the NiRez system. A qualitative picture of the spectral overlap of the Re(bpy) donor moiety with the NiII ligand-field manifold of Ni(phacac)2 is shown in Figure 5-15, which corresponds to a calculated spectral overlap value of J = 2.14 x 10'16 M'lcm3. The spectral overlap value (J = 5.52 x 10'16 M 1cm3 ) between the Rel-based 3MLCT emission and the CuIl ligand-field states in the CuRez complex (Chapter 3) is similar to the [Ni(pyacac)2(TI-IF)2(Re(bpy)(CO)3)2](OTf)2, which manifests as similar kCURez observed kinetics between the two complexes ( obs = 1.2 d: 0.1 x 108 kNiRez and obs = 2.0 i 0.2 x 108). The similar spectral overlap and energy transfer quenching dynamics observed between the CuRez and NiRez complexes provides additional proof of Forster-type reactivity occurring in [Ni(PyacaC)2(THF)2(Re(bPY)(C0)3)2](OTf)2- 258 Molar Absorptivity (normalized) (pozgeuuou) Ansuaiul uogssrulg Energy (x 103 cm'l) Figure 5-15. Overlay of the emission spectrum of [Be(pyacac)2(Re- (bpy)(CO)3)2](OTf)2 (blue trace) with the electronic absorption spectrum of Ni(phacac)2 (red trace). 5.4.6 Spin Dependent Dipolar Energy Transfer. As mentioned in the Introduction, in addition to possessing a downhill thermodynamic driving force, the total spin angular momentum - an abiding property of reactants and products - must be conserved during the course of a chemical reaction. Conservation of total spin angular momentum during an excited state energy transfer event must also be retained. From the structure of the NiRez complex given in Figure 5-9, two THF solvent molecules are coordinated in the axial positions of the Ni" metal center to generate a Ni(acac)2(THF)2 complex (i.e. Ni06 coordination sphere) in the core of the donor-acceptor assembly. The single-crystal X-ray structure data of the NiRez assembly (Table 5-2) shows the Ni—O (THF) distances are ~ 0.2 A longer than the Ni—O (acac) distances, suggesting a weaker metal-ligand covalent 259 interaction between Ni" and the oxygens of the THF ligands. The dissociation of both THF ligands from NiRe2 in solution would yield a ratio of donor-acceptor complexes with a pseudo-octahedral NiO6 and pseudo-D2}, NiO4 cores in solution depending on the propensity for NiRez to lose THF ligands. Figure 5-16 shows spin conservation diagrams between the Rel- based 3MLCT excited state with a pseudo-octahedral NiO6 (Top) and pseudo-D21| NiO4 (Bottom) acceptor cores. The NiII ground state in the NiO6 fl core is 3A2 (ISAI = 1), and when coupled with the 3MLCT excited state of fl SB — the ReI donor ( = 1) generates total spin angular momentum (ISTI) values of 2, 1, and 0. Based on the overlap integral plot given in Figure 5- 15, the acceptor state of the NiO(, core is the 3A2 —> 3T2 ligand-field. transition. As energy transfer (EnT) quenching occurs out of the 3MLCT _ excited state to produce the NiH-based 3A2 —> 3T2 transition, an ISTI of 1 is generated on the products side of the diagram from the coupling of the _p resulting ReI ground state (1801 = 0) with the Nin-based 3T2 excited state fl # A = 1). As can been seen in Figure 5-16, an 1311 = 1 between the ( 260 reagents and the products is conserved, which provides the spin-allowed energy transfer pathway. Conversely, Figure 5-16 (Bottom) shows coupling between the ReI excited state and the ligand-field states of the MO, core. The Nill ground fl state in MO. is 'A. (1S A1 = 0), and when coupled with the 3MLCT excited —.§ Sh state of the Re1 donor ( = l) generates a total spin angular momentum (ISTI ) value of 1 on the reagents side. On the products side, the acceptor states of the MO, core are S = O —> S = 0 ligand-field transitions (ISAI = O) — and the ground state of the ReI produced after the energy event is ISDI = 0 fl # A mutual ST value between the reagents (ISTI = 1) and the products (IST1 = 0) is not conserved, therefore energy transfer is spin-forbidden. 261 EnT - II ' I ' * (bpy )Re m NIOo spimauowed (hm/)Re M(N'06) 80:1 SA=1 SD=O SA=1 ST = 2,1,0 S1- = 1 EnT b - R "mN'o ------------------ > b Re'JW NiO * ( py ) e I 4 spin-forbidden ( DY) ( 4) 80:1 SA=O 80:0 SA=0 s1- = 1 s1- = 0 Figure 5-16. Top. Spin conservation diagram illustrating the spin- allowed energy transfer pathway between the 3MLCT excited-state of the Re1 donor and the pseudo-octahedral NiO6 acceptor. Bottom. Spin conservation diagram illustrating the spin-forbidden pathway between the 3MLCT excited-state of the ReI donor and the pseudo-D2}, NiO4 acceptor. — It is proposed that the differences in total spin angular momentum (ISTI) conservation in the energy transfer reaction between the Rel-based 3MLCT state and the ligand-field states of NiO6 and MO, cores, illustrated in Figure 5-16 (Top), give rise to the different behavior observed in the steady-state emission spectra of Ni(pyacac)2(TI-IF)2(Re(bpy)(CO)3)2](OTf); in CHzClz and THF solutions (Figure 5-13). This proposed axial ligand dissociation mechanism is manifested in the steady-state emission spectrum of the NiRez assembly in CH2C12 solution (Figure 5-13A), in which the emission signal is seen to steadily increase over the course of many hours in solution indicating 262 lesser quenching of the 3MLCT excited states over time. The THF dissociation mechanism has been attributed to a simple ligand exchange process between the NiRez complex and the very large concentration of CHzClz solvent molecules. This projected route is confirmed by the steady- state emission data for the NiRez complex in THF solution (Figure 5-13B), in which no change in the emission intensity was observed over the course of 5 hrs. The unvarying emission intensity of the NiRez system in THF solution is assigned to suppression of the axial ligand dissociation mechanism occurring in the presence of a large excess of THF molecules. The results of the steady-state emission data for the NiRez assembly in the two solvent systems reveal that dipolar energy transfer dynamics can be manipulated by simply varying the spin of the acceptor metal center, and furthermore, can be controlled by simply varying the coordination environment around a single metal center. Due to the above discussion on the steady-state emission data for Ni(pyacac)2(THF)2(Re(bpy)(CO)3)2](OTf)2 in CHzClz, the time-correlated single photon counting (TCSPC) data for the NiRez assembly (Figure 5-14, Right) requires further discussion. The two relaxation processes (rl= 5.0 ns and 12 = 25 ns) observed are both significantly faster than the rob, = 530 ns measured for the BeRe2 model complex (Figure 5-14, Left), which suggests 263 there are two competing quenching processes out the Rel-based 3MLCT excited state. If the NiO6 and NiO4 acceptor cores (discussed above) were the only two NiII coordination environments produced in the CH2C12 solution, a time-constant resulting from the spin-allowed pathway (13] = 5.0 ns) and a baseline offset representing the spin-forbidden coupling Nio4 ~ BeRe2 . . . . obs ~ obs ) would be the only observed klnetrc processes wrthrn the ( 13 time frame shown in the TCSPC data (Figure 5-14, Left). Another possible coordination environment for the NiII core in the NiRez complex is NiOs, which would result from dissociation of a single THF solvent molecule to form Ni(acac)2(THF). Based on the work of Raymond and coworkers in the 1960’s, the NiO5 structure could exist either in square-pyramidal or trigonal bipyramidal geometries due to the soft potential that exists between the two geometries.34 Existence of either the square-pyramidal (pseudo-C4,) or trigonal bipyramidal (pseudo-D3),) NiOs structures in solution could be responsible for the r; = 25 ns observed in the TCSPC data for the NiRez assembly (Figure 5-14). Further work, including photophysical and solution phase magnetic susceptibility data on the NiRez and Ni(acac)2(TI-IF)x complexes, will need to be performed to unmask the 25 ns component. 264 5.5 Conclusions The synthesis, structures, and photophysical properties of a series of donor-acceptor complexes based on Rel-bipyridine donors with a Com-acac acceptor (CoRe3; Section 5.3) and a Nin-acac acceptor (NiRez; Section 5.4) have been described. Steady-state and time-resolved emission spectroscopies indicated that the Rel-based 3MLCT excited state relaxation kinetics is significantly affected by the spin-state of the acceptor metal center. In the CoRe3 complex, the Com-based ligand-field transition (1A1 —) 1T1) represents a thermodynamically viable relaxation pathway out the 3MLCT donor state, but the similarity in 3MLCT emission lifetimes between "1 ligand- the CoRe3 assembly and an AlRe3 model complex ruled out the Co field state as a possible reaction pathway. The reason for the similar kinetics between the complexes was shown to derive from lack of total spin angular momentum conservation in the excited-state energy transfer reaction of CoRe3. In the NiRez system, the Nin-based ligand-field transition (3A2 —> 3T2) provided a spin-allowed quenching pathway out of the Rel-based 3MLCT excited-state when compared to a Ben-containing model complex, and a much faster 3MLCT relaxation was observed in the NiRez assembly. Favorable spectral overlap between the 3MLCT donor emission and acceptor ligand-field states coupled with a ca. 10 A donor-acceptor distance, allowed 265 for an assignment of Ftirster (dipolar) energy transfer as the dominant excited-state quenching mechanism in the Nin-containing complexes. These complexes represent excellent examples of the ability to manipulate excited- state reactivity by simply varying the spin-state of a single reagent involved. 266 5.5 References (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13) (14) (15) (16) Mulazzani, Q. G.; Sun, H.; Hoffman, M. 2.; Ford, W. E.; Rodgers, M. A. J. J. Phys. Chem. 1994, 98, 1145. Perrin, D. D.; Armarego, W. L. F. Purification of Laboratory Chemicals; 3rd ed.; Pergamon Press: Oxford, U. K., 1988. Mackay, L. G.; Anderson, H. L.; Sanders, J. K. M. J. Chem. Soc, Perkin Trans. 1. 1995, I8, 2269. r Doehring, A.; Goddard, R.; Jolly, P. W.; Krueger, C.; Polyakov, V. R. Inorg. Chem. 1997, 36, 177. Caspar, J. V.; Meyer, T. J. J. Phys. Chem. 1983, 87, 952. ,1 i Murakami, Y.; Nakamura, K. Bull. Chem. Soc. Jpn. 1966, 39, 901. l9: SAINT; ver 6.02a ed.; Bruker AXS, Inc.: Madison, WI, 2000. Sheldrick, G. M.; ver 2.03 ed.; Bruker AXS, Inc.: Madison, WI, 2000. Sheldrick, G. M.; ver 6.12 ed.; Bruker AXS, Inc.: Madison, WI, 2001. Damrauer, N. H.; Boussie, T. R.; Devenney, M.; McCusker, J. K. J. Am. Chem. Soc. 1997, 119, 8253. DeWitt, L.; Blanchard, G. J.; LeGoff, E.; Benz, M. E.; Liao, J. H.; Kanatzidis, M. G. J. Am. Chem. Soc. 1993, 115, 12158. Origin; 7.5 ed.; OriginLab Corp.: Northhampton, MA, 1991 - 2004. Barnum, D. W. J. Inorg. Nucl. Chem. 1961, 21, 221. Graddon, D. P. Nature 1956, 183, 1610. Hashagen, J. T.; Fackler, J. P. J. Am. Chem. Soc. 1965, 87, 2821. Holm, R. H.; Cotton, F. A. J. Phys. Chem. 1961, 65, 321. 267 (17) (18) (19) (20) (21) (22) (23) (24) (25) (26) (27) (28) (29) (30) (31) (32) Tsukahara, Y.; Kamatani, T.; lino, A.; Suzuki, T.; Kaizaki, S. Inorg. Chem. 2002, 41 , 4363. Fackler, J. P.; Cotton, F. A. J. Am. Chem. Soc. 1960, 82, 5005. Dattelbaum, D. M.; Omberg, K. M.; Schoonover, J. R.; Martin, R. L.; Meyer, T. J. Inorg. Chem. 2002, 41 , 6071. Dattelbaum, D. M.; Martin, R. L.; Schoonover, J. R.; Meyer, T. J. J. Phys. Chem. A. 2004, 108, 3518. Argazzi, R.; Bertolasi, E.; Chiorboli, C.; Bignozzi, C. A.; Itokazu, M. K.; Murakami Iha, N. Y. Inorg. Chem. 2001, 40, 6885. Cotton, F. A.; Wise, J. J. Inorg. Chem. 1966, 5, 1200. Montgomery, H.; Lingefelter, E. C. Acta. Cryst. 1964, 17, 1481. Busby, M.; Liard, D. J.; Motevalli, M.; Toms, H.; Vlcek, A. Inorg. Chim. Acta. 2004, 35 7, 167. Chen, P.; Curry, M.; Meyer, T. J. Inorg. Chem. 1989, 28, 2271. Lucia, L. A.; Abboud, K.; Schanze, K. S. Inorg. Chem. 1997, 36, 6224. Sacksteder, L.; Zipp, A. P.; Brown, E. A.; Streich, J.; Demas, J. N.; DeGraff, B. A. Inorg. Chem. 1990, 29, 4335. Worl, L. A.; Duesing, R.; Chen, P.; Della Ciana, L.; Meyer, T. J. J. Chem. Soc., Dalton Trans. 1991, 849. Fackler, J. P.; Cotton, F. A. J. Am. Chem. Soc. 1961, 83, 3775. Striplin, D. R.; Crosby, G. A. Chem. Phys. Lett. 1994, 221, 426. Dexter, D. L. J. Chem. Phys. 1953, 21, 836. Brown, W. R.; O'Boyle, N. M.; McGarvey, J. J.; Vos, J. G. Chem. Soc. Rev. 2005, 34, 641. 268 (33) Soler, M.; McCusker, J. K. J. Am. Chem. Soc. 2008, 130, 4708. (34) Aristides, T.; Raymond, K. N.; Spiro, T. G. Inorg. Chem. 1970, 9 2415. 269 Chapter 6. Future Work 6.1 Overall Goals Chapter 6 focuses on continuing the fundamental investigation of Féirster energy transfer dynamics in Rel-based donor-acceptor assemblies discussed throughout this dissertation. The majority of the discussion in this chapter will center on proposing various spectroscopic tools and additional donor-acceptor complexes that will be needed to address two remaining goals of the research effort: 1) the complete characterization of the dynamical behavior (i.e. vibrational) of the Re(bpy’)(CO)3 donor moieties undergoing 3MLCT energy transfer quenching processes and 2) the synthesis of additional CoRe3 derivatives to further understand the spin-dependent energy transfer processes proposed for the CoRe3 and CrRe3 complexes outlined in Chapter 5. 6.2 Goal 1: Monitoring 3’MLCT Relaxation in the Re(bpy’)(CO)3 Donors 6.2.1 Background. The 3MLCT excited state of the ReI fluorophores can be thought of as Re"(bpy')(CO)3, in which the Re1 metal has been formally oxidized and the polypyridyl ligand formally reduced to generate Re(tzg5 )—bpy(1r*'). A brief description of the electronic structure and 270 photophysical properties of d6 polypyridyl complexes is given in the Introduction. As one can imagine, the ground state vibrational properties of the Re(bpy)(CO)3 moiety will be significantly perturbed in the charge transfer excited state. The ground-state vibrational signatures for the polypyridyl-based C-C and C-N stretching modes and the CO stretching frequencies are the two particular areas of the Re(bpy)(CO)3 donor moiety that will be significant altered in the excited-state. The bond order of the C- C and the C-N bonds of the bpy will be decreased due to population of a bpy-based 71:1 orbital, and the bond order of the CO ligands will be increased due to less backbonding contribution from the oxidized Re metal center.1 Time-resolved infrared spectroscopy (TRIR) has been shown to be an excellent technique for probing the relaxation dynamics associated with transition metal-based excited states.2’3 Due to the high oscillator strengths and backbonding interactions associated with the CO ligands, TRIR has been used to monitor the excited-state CO stretching frequencies to characterize the 3MLCT excited state relaxation in [Re(Xz-bpy)(CO)3(4- Etpy)](PF6) (where X = CH3, H, or COzEt and 4-Etpy = 4-ethylpyridine).4’5 The v(CO) bands associated with [Re(Xz-bpy)(CO)3(4-Etpy)](PF6) were observed to shift to higher energy in response to an increase in the triple- bond character of the CO ligands due to less Re-CO backbonding. In 271 addition, the technique has been shown to monitor and assign the mechanism of excited-state reactivity out of the 3MLCT excited state of Rel- based donor-acceptor assemblies by monitoring the extent to which the v(CO) bands shift in energy.6'8 Based on the TRIR data reported for the mono- and poly-nuclear Re]- based complexes discussed above, application of TRIR spectroscopy to the MRex (x = 2 or 3) assemblies presented throughout in dissertation would be a nice compliment to the steady-state and time-resolved emission data. All of the MRex complexes possess Re(bpy’)(CO)3 donor moieties that possess v(CO) bands that could serve as spectroscopic tags to monitor relaxation out of the 3MLCT excited state. Additionally, even though the donor-acceptor distances, electrochemical properties, and spectral overlap considerations in the MRex complexes prove dipolar energy transfer as the dominant 3MLCT quenching process, the energies for the v(CO) bands in the 3MLCT excited state derived fiom TRIR would be different depending on whether energy or electron transfer quenching was operative (Vida infra). 6.2.2 Dicyano Derivatives of CuRez and BeRe2. With an eye toward developing a system that could be used for assigning the excited-state quenching mechanisms in the MRex complexes Via TRIR, I have synthesized and characterized [Cu(pyacac)2(Re(dcnb)(CO)3)2](OTf)2 and [Be(pyacac)2- 272 (Re(dcnb)(CO)3)2](OTf)2 (where dcnb = 4,4’-dicyano-2.2’-bipyridine). Just as with the complexes reported in Chapter 3, the Cu11 and Bell metal centers are covalently attached to two fac-Re(bpy’)(CO)3 moieties through two pyridyl-acetylacetonate bridging ligands (Figure 6-1). The BeRe2 complex was synthesized to investigate the v(CO) bands associated with the 3MLCT excited state of the Re(dcnb)(CO)3 moiety in the absence of emission quenching reactivity in the CuRez assembly. CN‘I 2+ Figure 6-1. Structures of [Cu(pyacac)2(Re(dcnb)(CO)3)2](OTf); and [Be(pyacac)2(Re(dcnb)(CO)3)2](OTf)2 273 6.2.3 Synthesis and Characterization The synthesis procedure for [Cu(pyacac)2(Re(dcnb)(CO)3)2](OTDZ is same as reported for the CuRez series in Chapter 3. Amounts of 75 mg (0.18 mmol) of Cu(pyacac); and 225 mg (0.360 mmol) of Re(dcnb)(CO)3(OTf) were used in the preparation of the CuRez compound. Yield: 83 mg (28%). Anal Calcd for C52H32N10F6OI6S2CuRe2 °4HzOz C, 35.92; H, 2.32; N, 8.05. Found: C, 35.74; H, 2.10; N, 7.84. MS: [ESI+, m/z (rel. int.)]: 684.4 (35) {[Cu(pyacac)2(Re(dcnb)(CO)3)2]}2+, 1517.9 (1) {[Cu(pyacac)2(Re(dcnb)- (COthorori The synthesis procedure for [Be(pyacac)2(Re(dcnb)(CO)3)2](OTf)2 is same as reported for the BeRez series in Chapter 3. Amounts of 67 mg (0.19 mmol) of Be(pyacac)2 and 233 mg (0.372 mmol) of Re(dcnb)(CO)3(OTf) were used in the preparation of the BeRe2 compound. Yield: 141 mg (47%). Anal Calcd for C52H32N10F6016S2BeRez °4HzOz C, 37.08; H, 2.39; N, 8.31. Found: C, 37.22; H, 2.18; N, 8.04. MS: [ESI+, m/z (rel. int.)]: 657.1 (41) {[Be(pyacac)2(Re(dcnb)(CO)3)2]}2+, 1463.3 (1) {[Be(pyacac)2(Re(dcnb)- (COthOToti 6.2.4. Electronic Absorption Spectroscopy. The electronic absorption spectra of [Cu(pyacac)2(Re(dcnb)(CO);)2](OTf); and [Be(pyacac)2(Re- (dcnb)(CO)3)2](OTf)2 were acquired in room-temperature CH3CN solutions 274 and are shown in Figure 6-2. Transitions associated with these complexes were assigned based on previously reported analyses of CuRez/BeRez complexes reported in Chapter 3. The Rel-based 1A. —> lMLCT (t2g ——> 71* (bpy’)) transition occurs at Am, = 395 nm for both complexes, with the absorption profile of [Be(pyacac)2(Re(dcnb)(CO)3)2](OTf); solely exhibiting these lMLCT based characteristics. The BeRe2 complex allows for analysis of the ground-state absorption behavior of the Re1 donor moieties without overlapping with the Cull-based charge-transfer transitions associated with the CuRez assembly. In addition, the similarity in the shape and energy of the 1MLCT absorption bands in the two complexes indicates minimal electronic communication between the Re1 and Cu'I metal centers in the ground state. 60000 1- 50000 - 40000 30000 - 20000 - 10000 Molar Absorptivity (M'cm'l) 300 400 500 600 Wavelength (nm) Figure 6-2. Electronic absorption spectra of [Cu(pyacac)2(Re(dcnb)- (CO)3)2](OTf)2 (red trace) and [Be(pyacac)2(Re(dcnb)(CO);)2](OTf)2 (blue trace) acquired in room-temperature CH3CN solution. 275 6.2.5 Time-Resolved Emission. Quantitative information concerning emission quenching by the Cu" center was obtained via time-resolved emission spectroscopy. Data for the BeRe2 model complex could be fit to a single-exponential decay kinetic model (robS = 110 d: 10 ns); emission trace for the observed decay rate in the BeRe2 complex is given on the left side of Figure 6-3. Due to a lack of signal observed with the nanosecond emission apparatus for the CuRez complex, time-correlated single-photon counting (TCSPC) was employed to measure the excited-state lifetime of the Cu"- containing complexes. A plot of the TCSPC data obtained for the CuRez complex in deoxygenated CHzClz solution is shown on the right side of Figure 6-3; the observed time constant (4.1 i 0.4 ns) for excited-state decay in the CuRez complex is significantly shorter than the BeRe2 model complex, and shows the Rel-based 3MLCT excited state is significantly quenched in the presence of CuII relative to Be". The analysis of the time- resolved emission data for [Cu(pyacac)2(Re(dcnb)(CO)3)2](OTf)2 and [Be(pyacac)2(Re-(dcnb)(CO)3)2](OTDZ were performed analogous to the CuRez complexes reported in Chapter 3 and the reader is directed there for further discussion. 276 250 - ' Cu 200- ‘ I—i Ur O I O O Intensity (mV) Intensity (cps) soul 0 200 400 600 800 0 10 20 30 40 Time (ns) Time (ns) Figure 6-3. Left. Nanosecond time-resolved emission data for [Be(pyacac)2(Re(dcnb)(CO)3)2](OTf)2 (robS = 110 d: 10 ns) acquired in room-temperature deoxygenated CHzClz solution (kpump = 400 nm; 31mm, = 625 nm). Right. Time correlated single-photon counting (TCSPC) emission data for [Cu(pyacac)2(Re(dcnb)(CO)3)2](OTf)2 acquired in room-temperature deoxygenated CH2C12 solution ()tpump = 400 nm; Ambe = 625 nm). The solid red line corresponds to a fit to a bi- exponential decay model with t, =4] :1: 0.4 ns and 1:, = 110 ns (fixed); the latter corresponds to emission from Re(dcnb)(CO)3(pyacac) present in solution as a trace impurity. 6.2.6 Proposed TRIR Experiments. The Re(dcnb)(CO)3 donor moieties of [Cu(pyacac)2(Re(dcnb)(CO)3)2](OTf)2 and [Be(pyacac)2(Re(dcnb)(C0);)2]- (OTf); represent ideal candidates for probing quenching dynamics of charge transfer-based excited states using TRIR. In addition to monitoring the dynamical shift in the CO ligands that occur in response to the RelI (d5) metal center in the 3MLCT excited state, the high oscillator strength of the CN substituents on the bipyridyl ligand should serve as ideal spectroscopic tags for measuring the excited state relaxation associated with the reduced 277 polypyridyl ligand. Measuring the energies of the v(CO) and v(CN) bands associated with the 3MLCT excited state in the BeRe2 complexes will be critical to the mechanistic analysis in the CuRez analogue. If excited-state energy transfer is the dominant mechanism in the CuRez complex, then the shifts (v(GS) to v(ES)) observed for the v(CO) and v(CN) bands in the Be" and Cull-containing systems will correspond. Due to the shorter excited- state lifetime measured for the CuRez complex (Figure 6-3), the only difference between the two experiments will be the shorter time window that will be needed to collect the CuRez data. The energy transfer process in the CuRez system is in essence, a faster pathway for the 3MLCT state to relax and therefore will exhibit similar v(CO) and v(CN) bands as seen for the charge transfer excited state in the BeRe2 complex. If electron transfer is the dominant quenching process in the CuRez assembly, the shifts in the v(CO) bands of the Ben- and Cull-containing systems will not correlate due to the “complete” oxidation of the ReI metal center that would occur during an electron transfer event in CuRez, compared with partial oxidation (~70%) occurring during a charge transfer excitation. In addition, the v(CN) observed in the BeRe2 complex should disappear due to electron transfer quenching removing the excited electron from the all level of the dcnb ligand. 278 6.3 Goal 2: Additional Derivatives for the CoRe3 Series 6.3.1 Background. In Chapter 5, a CoRe3 donor-acceptor complex - [Co(pyacac)3(Re(bpy)(CO)3)3](OTf); - was synthesized in order to explore the spin-state dependence of excited state energy transfer reactions. Based on the observation of significant spectral overlap between the 3MLCT emission profile of the Re(bpy) donor moiety and the 1A1 —) lT1 ligand-field absorption band in the Co111 acceptor, dipolar energy transfer quenching of the 3MLCT state is thermodynamically favorable and predicted to occur. Nanosecond emission lifetime measurements on the CoRe; complex in CHzClz solution exhibited similar 3MLCT relaxation kinetics as an electronically benign AlRe; model system. Determination of the total spin orbital angular momentum (ST) for the energy transfer quenching reaction in CoRe; revealed a spin-forbidden coupling mechanism between the photo- induced reagents and the energy transfer products. Furthermore, comparison of the 3MLCT emission lifetimes of CoRe; and a structurally identical CrRe; complex that possesses similar spectral overlap characteristics and a spin- allowed coupling mechanism between the reagents and products showed a ca. 100 fold decrease in the lifetime of CrRe3, and confirmed that dipolar energy transfer can be turned on or off depending on the spin-state of a single ion. 279 The trend in observed energy transfer rate constants for the series of CrRe; complexes presented in Chapter 4 showed a direct dependence on the magnitude of spectral overlap between the Rel-based 3MLCT emission spectra and the Crm—based 4A2—> 4T2 ligand-field acceptor transition. Figure 6-4 reproduces the spectral overlap plot for the CrRe; series originally reported in Chapter 4, with the bpy derivative (green trace) exhibiting the highest spectral overlap value and the largest energy transfer rate constant. To adequately compare the CrRe; (spin-allowed) and the CoRe; (spin- forbidden) 3MLCT relaxation processes, additional Re(bpy’) derivatives of the CoRe; complex will need to be prepared to generate an analogous spectral overlap plot for the CoRe; series. Absorbance (normalized) (pazgeuuou) Knsuarul uogssrurg 22 20 18 l6 14 Energy (x 103 cm") Figure 6-4. Overlay of the emission spectra of [Al(pyacac)3(Re(tmb)- (C0)3)3](0Tf)3 (b11112), [A1(pyacaC)3(Re(bPY)(CO)3)3l(OTf)3 (green), and [Al(pyacac)3(Re(deeb)(CO)3)3](OTf)3 (red) with the electronic absorption spectrum of Cr(phacac)3. Plot is reproduced from Figure 4- 12 (Chapter 4). The largest calculated spectral overlap (bpy; green) resulted in the fastest energy transfer rate constant within the series. 280 6.3.2 Proposed CoRe3 Derivatives. Based on the spectral overlap plot between the [Al(pyacac)3(Re(bpy)(CO)3)3](OTf); emission profile (Xmax = 565 nm) and the Co(phacac)3 1A; —> 1T; ligand-field manifold (Amax = 605 nm) reported in Chapter 5, additional CoRe; derivatives possessing lower energy Re(bpy’) emitters need to be synthesized (preferably one with an emission maximum of ~ 600 nm) to generate an analogous spectral overlap plot created for the CrRe; series. One obvious choice is to synthesize the deeb derivative (deeb = 4,4’-diethylester-2,2’-bipyridine), due mainly to already possessing [Al(pyacac)3(Re(deeb)(CO)3)3](OTf); as a structural model and secondly that the Amax for emission is 625 nm which lies just on the low energy side of the Co(phacac)3 'A; -—> 1T1 ligand-field band but still retains descent spectral overlap. Based on literature reports that show an emission maximum of ~ 600 nm in CH2C12 solution, the next CoRe; analogue proposed is the dclb derivative (dclb = 4,4’-dichloro-2,2’-bipyridine). The synthesis and photophysical characterization of the [Al(pyacac)3(Re(dclb)(CO)3)3](OTf); structural model have already been performed based on experimental protocols used for similar AlRe3 complexes reported in Chapter 2. The steady-state and nanosecond time-resolved emission data for [Al(pyacac)3- (Re(dclb)(CO)3)3](OTf)3 is shown in Figure 6-5. The photophysical 281 properties of [Al(pyacac)3(Re(dclb)(CO)3)3](OTf)3 correspond with the other AlRe; derivatives reported in Chapter 2, and the observed emission properties are assigned to the Rel-based 3MLCT —r 1A] transition. 40 - A? A 1 E E . :: §‘ 20 - 1 O m '8 E "8 E 10 -‘ LL) I I I A O I l l 1 I 500 600 700 800 0 200 400 600 800 Wavelength (nm) Time (ns) Figure 6-5. Left. Steady-state emission spectrum of [Al(pyacac)3(Re(- dclb)(CO)3)3](OTf)3 in room-temperature CH2C12 solution O‘rump = 375 nm); Am“ = 595 nm for 3MLCT —> 1A1 emission. Right. Nanosecond time-resolved emission data for [Al(pyacac)3(Re(dclb)(CO)3)3](OTf)3 (rob, = 120 d: 10 ns) acquired in deoxygenated room-temperature CH2C12 solution (kpump = 415 nm; Apmbe = 595 nm). Figure 6-6 shows the spectral overlap plot between the Alm- containing bpy, dclb, and deeb 3MLCT emission profiles with the IA] —> 1T1 ligand-field absorption band of Co(phacac)3. As can be seen from Figure 6- 6, ideal spectral overlap between the Re(dclb) moiety (green trace) and the CoIII 1A] —> 1T1 absorption band (black trace) is observed, and most importantly a spectral overlap plot analogous to the one in the CrRe; series has been generated. In other words, a spin-allowed spectral overlap plot and 282 a spin-forbidden spectral overlap plot have now been generated to compare two structurally identical donor-acceptor complexes that differ simply by the acceptor spin-state. (pozneuuou) Ausualul uogssrula Molar Absorptivity (normalized) 500 600 700 800 Wavelength (nm) Figure 6-6. Overlay of the emission spectra of [Al(pyacac)3(Re(bpy)- (C0)3)31(0Tf)3 (blue), [A1(PyacaC)3(Re(dclb)(C0)3)3](OTfla (green), and [Al(pyacac)3(Re(deeb)(C0););](OTf); (red) with the electronic absorption spectrum of Co(phacac)3. In closing, the final two compounds required to complete the CoRe; series — [Co(pyacac)3(Re(dclb)(CO)3)3](OTf); and [Co(pyacac)3(Re(deeb)- (CO)3)3](OTf)3 - need to be synthesized and characterized. The synthesis procedure for these two remaining complexes should follow the same procedure used for [C0(pyacac)3(Re(bpy)(CO)3)3](OTf); reported in Chapter 5. 6.4 References (1) (2) (3) (4) (5) (6) (7) (8) Glyn, P.; George, M. W.; Hodges, P. M.; Turner, J. J. J. Chem. Soc., Chem. Commun. 1989, 1655. Vlcek, A.; Farrell, I. R.; Liard, D. J.; Matousek, P.; Towrie, M.; Parker, A. W.; Grills, D. C.; George, M. W. J. Chem. Soc., Dalton Trans. 2002, 701. Schoonover, J. R.; Strouse, G. F. Chem. Rev. 1998, 98, 1335. Dattelbaum, D. M.; Martin, R. L.; Schoonover, J. R.; Meyer, T. J. J. Phys. Chem. A. 2004, 108, 3518. Dattelbaum, D. M.; Omberg, K. M.; Schoonover, J. R.; Martin, R. L.; Meyer, T. J. Inorg. Chem. 2002, 41, 6071. Chen, P.; Palmer, R. A.; Meyer, T. J. J. Phys. Chem. A 1998, 102, 3042. Claude, J. P. ,Omberg, K. M.; Williams, D. S.; Meyer, T. J. J. Phys. Chem. A 2002,106,7795. Alsindi, W. Z.; Easun, T. L.; Sun, X.-Z.; Ronayne, K. L.; Towrie, M.; Herrera, J.; George, M. W.; Ward, M. D. Inorg. Chem. 2007, 46, 3696. 284 ATE 111111 111111111111111111113 03063 1349 1293 ”11111111