Transition Metal Manganese and Main Group Gallium Complexes as Catalysts for Alkene Epoxidation by Wenchan Jiang A dissertation submitted to the Graduate Faculty of Auburn University in partial fulfillment of the requirements for the Degree of Doctor of Philosophy Auburn, Alabama May 04, 2012 Approved by Christian R. Goldsmith, Assistant Professor David M. Stanbury, Professor Anne E. V. Gorden, Associate Professor Eduardus C. Duin, Associate Professor Department of Chemistry and Biochemistry ii Abstract Epoxides are important organic synthesis precursors and widely used in industry. Alkene epoxidation has been under extensive investigation over decades. Synthetic models have been established for catalytic epoxidations, employing transition and main group metals. Epoxidation proceeds either through high valent metal-oxo species and/or a Lewis acid mechanism for oxygen transfer to substrate. Specifically, by high valent metal-oxo species, Jacobsen?s catalyst (MnIII-salen) and non-porphyrin MnII complexes can catalyze enantioselective epoxidation and electron-rich terminal alkene epoxidation respectively. Sharpless catalyst (TiIV-tartrate) and Al3+ centered complexes catalyze epoxidation through Lewis acid mechanisms. Manganese complexes with phenanthroline derivatives ([Mn(R-phen)2Cl2]Cl, R = NH2, Me, H, Cl, NO2) as catalysts are reported. The MnII complexes catalyze epoxidation of a variety of substrates with electron-rich alkenes being favored. The installation of functional groups on the 5-position of phenanthroline ligand plays important roles in influencing catalyst reactivity. The more electron-donating the functional group is, the more reactive it is. The reactivities with different substituent groups on the ligand have been correlated with Hammett constant, which showed a good linear relationship. EPR and spectroscopic studies indicated possible involvement of high valent MnIV-oxo species which decays quickly under experimental conditions. Gallium complexes are effective catalysts as well. [Ga(phen)2Cl2]Cl is reported to be the first homogeneous gallium catalyst for alkene epoxidation with peracetic acid. Compared with other group 13 metals, the activity of [Ga(phen)2Cl2]Cl is high with low loading of 1%. The epoxide is the only product, indicating good selectivity. Furthermore, six gallium(III) complexes with N-donor ligands were synthesized to study the mechanism of Ga(III)-catalyzed olefin epoxidation by peracetic acid. Although the complexes with relatively electron-poor phenanthroline derivatives display faster initial reactivity, the Ga(III) complexes with the polydentate pyridylamine ligands appear to be more robust, with less noticeable decreases in their catalytic activity over time. The more highly chelating trispicen and tpen are associated with markedly decreased activity. iii Acknowledgement This dissertation is completed under the supervision of Prof. Christian Goldsmith. I am very grateful for his help all through these years. As my advisor, he has set up an example for me about how to pursue science achievement in a hard working and intellectual way. I leant a lot from his expertise in chemistry. I owe thanks to Prof. David Stanbury, who taught me in two advanced inorganic courses and shared his profound knowledge in helpful research discussion. I extend my thanks to Prof. Anne Gorden, who taught me organic chemistry in my first year and gave me suggestions in my oral defense. I also want to thank Dr. John Gorden who guided me through all of my crystal data collection and analysis. I am indebted to Prof. Evert Duin who taught me biochemistry and helped with EPR experiments and data analysis, and Dr. Yonnie Wu for his assistance with MS experiments and interpretation. I am thankful to Prof. William Robert Ashurst for spending his precious time serving on the committee and providing advice on this dissertation. I appreciate their help and care. I will always remember the friendship I got in my lab with all lab mates Roger (Yu) He, Qiao Zhang, and Meng Yu. I would like to thank my fianc?e Wenting (Lucia) Wang, who moved from Boston to be with me in Auburn and has been supportive all through even during my toughest time. Last but not least, I dedicate this dissertation to my parents, without their support and encouragement I wouldn?t even have arrived in the United States to pursue my Ph.D. degree. iv Table of Contents Abstract??????????????????????????????..?ii List of Tables ?????????????????????????????vi List of Figures???????????????????????????..vii List of Schemes????????????????????????????..x Chapter 1 Review of Alkene Epoxidation Catalyzed by Main Group and Transition Metal Complexes??????????????????????...1 1.1 Value and Industrial Production of Epoxides ???????.???2 1.2 Development of Alternative Catalysts for Alkene Epoxidation???..2 1.2.1 Jacobsen?s Catalyst??????????????..???.3 1.2.2 Manganese Complexes with Neutral, Polydentate N-Donors?..9 1.2.3 Non-Heme Iron Complexes????????????...?12 1.2.4 The Sharpless Catalyst???????????????...15 1.2.5 Homogenous Aluminum Catalysts???????...???..18 1.3 Summary ???????.????????????????.20 References.?????????????????????.???.21 Chapter 2 Alkene Epoxidation Catalyzed by Manganese Complexes with Electronically Modified 1,10-Phenanthroline Ligands ??.???.??25 2.1 Introduction ????????????????????.?.....26 2.2 Experimental ??.???????????????????...27 2.3 Results ??????.??????????????????.30 2.4 Discussion ????.??????????????????...32 iv 2.5 Conclusions ????.??????????????????.35 References ????????????????????????..37 Chapter 3 A Homogeneous Gallium(III) Compound Selectively Catalyzes the Epoxidation of Alkenes ???????????????????..39 3.1 Introduction ??????????.????????????.40 3.2 Experimental ?????????????????????....40 3.3 Results and Discussion ??????????????????.44 3.4 Conclusions ???????????????????.???.55 References ????????????????????????..56 Chapter 4 Selective Epoxidation of Electron-Rich Alkenes Catalyzed by a Series of Gallium(III) Compounds ???????????????????58 4.1 Introduction ??????????????.????????.59 4.2 Experimental ??????????.???????????...60 4.3 Results ??????.????????????????.??66 4.4 Discussion???????????????????????77 4.5 Conclusions ?????????.?????????????.82 Appendix?????????????????????????.83 References ????..??????????????????...?.99 vi List of Tables Table 1.1 Asymmetric epoxidation of representative olefins by Jacobsen?s catalyst.........4 Table 1.2 Epoxidations catalyzed by [MnII(R,R-mcp)(CF3SO3)2] with PAA????.......10 Table 1.3 Epoxidations catalyzed by [Fe(mep)(CF3SO3)2]????????????..13 Table 1.4 Summary of epoxidation scope catalyzed by Beller?s in situ FeIII complex ?...14 Table 1.5 Selected asymmetric epoxidation of allylic alcohols??????????..17 Table 1.6 Epoxidation of ?, ?-unsaturated ketones by [Al(H2O)6]3+/H2O2 ????...?..19 Table 2.1 Yields of epoxides for Mn(II)-catalyzed olefin oxidation by PAA ???...?. 29 Table 3.1 Selected crystallographic data for [Ga(phen)2Cl2]Cl ???...??..???.....43 Table 3.2 Bond lengths (?) and bond angles (?) for [Ga(phen)2Cl2]+?..??................?46 Table 3.3 Yields/turnover numbers of gallium(III) catalyzed alkene epoxidation reactions by various grades and amounts of PAA ???.????????????..48 Table 3.4 Yields of epoxide products (%) in control reactions ????.??????..49 Table 3.5 Summary of ICP-OES analysis ??????????????.???.?.51 Table 4.1 Selected crystallographic data for [Ga(bispicen)Cl2]Cl (1) and [Ga(trispicen)Cl](GaCl 4)(Cl) (7)??????????????????..71 Table 4.2 Alkene Conversions/Yields of Epoxides (%) of Reactions Catalyzed by 1-6?.73 Table A1 Selected crystallographic data for [Ga(trispicen)Cl]Cl2?????????..83 vii List of Figures Figure 1.1 Structure of the prototypical Jacobsen catalyst ?????...?.???..4 Figure 1.2 The molecular structure of the R,R-mcp ligand and the crystal structure of [MnII(R,R-mcp)(CF 3SO3)2]???????????..?????.?..8 Figure 1.3 Illustration of the mep ligand (left) and the crystal structure of dinuclear [Fe 2(?-O)(?-CH3CO2)(mep)2] cation???????...?????.....12 Figure 1.4 The dimeric structure of the Sharpless catalyst??????????.16 Figure 2.1 Epoxidation of cyclooctene by peracetic acid catalyzed by [Mn(NO 2-phen)2Cl2] in MeCN at 0 oC under N2.????.?????31 Figure 2.2 X-band EPR spectra of a solution of 1.1mM [Mn(NO2-phen)2Cl2] in 3:2 MeCN/CHCl 3 ??????????????????..????.32 Figure 2.3 Plot of the reaction yields (as assessed at 1 h) for the oxygenation of each alkene as a function of the ? P value of the 5-substituent on the catalyst?s phenanthroline ligands ???..??????????????.?..34 Figure 3.1 ORTEP representation of the cation [Ga(phen)2Cl2]+ ???????..45 Figure 3.2 Yield of cyclohexene oxidation by 2 equiv of PAAR as a function of time???????????????????????????52 Figure 3.3 Mass spectrometric analysis of the reaction between cyclooctene and two equiv. of PAA R at t = 5 s (top) and t = 60 min (bottom) ????.??.53 Figure 4.1 ORTEP representation of the cation [Ga(bispicen)Cl2]+?.?..................70 Figure 4.2 ORTEP representation of the dication [Ga(trispicen)Cl]2+????...?71 Figure 4.3 Yields of alkene epoxidations by PAAR catalyzed by the bispicen complex 1 as a function of time????????????.???????..75 Figure 4.4 Yields of alkene epoxidation by PAAR catalyzed by the 1-6 and [Ga(phen) 2Cl2]Cl as a function of time?????????????.76 Figure 4.2 Yield of epoxides by Ga(bispicen)Cl3 as a function of time ???...?.67 Figure A1 ORTEP representation of [Ga(trispicen)Cl]Cl2?MeCN?3H2O grown from a solution of [Ga(tpen)Cl 2]Cl in MeCN?????????????82 viii Figure A2 Mass spectrum (ESI) of [Ga(tpen)Cl2]Cl??????????..??83 Figure A3 1H NMR spectrum of 1 in CD3OD at 294 K???????????.84 Figure A4 13C NMR spectrum of 1 in CD3OD at 294 K???????????84 Figure A5 IR spectrum of 1??????????????????????85 Figure A6 1H NMR spectrum of 2 in CD3OD at 294 K??????????.....85 Figure A7 13C NMR spectrum of 2 in CD3OD at 294 K???????????86 Figure A8 IR spectrum of a powder sample of 2????????..?????.86 Figure A9 IR spectrum of a crystalline sample of 7???????????.?..87 Figure A10 1H NMR spectrum of 3 in CDCl3 at 294 K?????????...??87 Figure A11 1H NMR spectrum of 3 in CD3CN at 294 K???????????.88 Figure A12 13C NMR spectrum of 3 in CDCl3 at 294 K???????????..88 Figure A13 IR spectrum of a powder sample of 3??????.???????..89 Figure A14 1H NMR spectrum of 4 in CD3OD at 294 K???????????.89 Figure A15 13C NMR spectrum of 4 in CD3OD at 294 K?????...?????.90 Figure A16 IR spectrum of a powder sample of 4???...??????????90 Figure A17 1H NMR spectrum of 5 in DMSO-d6 at 294 K??????????.91 Figure A18 1H NMR spectrum of 5 in CD3OD at 294 K???????????.91 Figure A19 13C NMR spectrum of 5 in DMSO-d6 at 294 K??..???????.92 Figure A20 IR spectrum of a powder sample of 5???????.??????..92 Figure A21 1H NMR spectrum of 6 in DMSO-d6 at 294 K?????.?????93 Figure A22 13C NMR spectrum of 6 in DMSO-d6 at 294 K????????...?.93 Figure A23 IR spectrum of a powder sample of 6????????...?????94 ix Figure A24 The yields of the organic products of [Ga(bispicen)Cl2]Cl-catalyzed cyclohexene oxidation by PAA R over time???????.????..94 Figure A25 Mass spectrum (ESI) of the gallium species present in solution 3 h after the beginning of the cyclohexene epoxidation reaction catalyzed by 1?95 Figure A26 Mass spectrum (ESI) of the gallium species present in solution 3 h after the beginning of the cyclohexene epoxidation reaction catalyzed by 2??..?????????????????????????.96 Figure A27 Mass spectrum (ESI) of the gallium species present in solution 3 h after the beginning of the cyclohexene epoxidation reaction catalyzed by 3....97 x List of Schemes Scheme 1.1 Substituent effects on enantioselectivity ???????????...???.5 Scheme 1.2 Possible pathways for epoxidation catalyzed by Jacobsen?s compound ?..?.6 Scheme 1.3 Hammett plots depicting enantiomeric ratios of epoxides generated with catalysts with various electronically modified ligands ???...??.?.??.7 Scheme 1.4 Parallel Lewis acid and redox pathways for [Mn(salen)]+ catalyzed epoxidations?????????????????????????..9 Scheme 1.5 Possible mechanistic pathways for Mn(II)-catalyzed olefin epoxidation?.....12 Scheme 1.6 Proposed catalytic cycle for non-heme iron catalyzed epoxidation????..15 Scheme 1.7 Proposed mechanism for Sharpless epoxidation (E refers to ester group)?...18 Scheme 1.8 Proposed mechanism for epoxidation catalyzed by [Al(H2O)6]3+???..?..20 Scheme 2.1 Illustration of Mn(R-phen)2Cl2 complexes ?????????????..27 Scheme 2.2 Mn(R-phen)2Cl2 catalyzed epoxidation reaction ??????????.?.27 Scheme 2.3 Substrate used in Mn(R-phen)2Cl2 epoxidation ????????.??..?31 Scheme 3.1 Illustration of epoxidation catalyzed by [Ga(phen)2Cl2]Cl ????.??.....40 Scheme 3.2 Proposed Sharpless type mechanism ????.??????????..?54 Scheme 4.1 Neutral N-donor ligands ????????????????...................59 Scheme 4.2 Possible solution structures of 2?????????????????....67 Scheme 4.3 Proposed solution structure of 3??????????..???????.68 Scheme 4.4 Possible conformations of the phen ligands in complexes 4 and 5????...69 Scheme 4.5 Illustration of possible mechanism for the epoxidation reactions?????.79 1 Chapter 1 Review of Alkene Epoxidation Catalyzed by Main Group and Transition Metal Complexes 2 1.1 Value and Industrial Production of Epoxides Epoxides are valuable for their capacity to be converted into a wide range of useful functional groups. Ethylene oxide and propylene oxide are particularly vital to industry. Ethylene oxide is a key precursor for ethylene glycol and surface-active agents.1 Further polymerization of ethylene glycol leads to polyethylene glycol for detergent use.2 Propylene oxide is used in the preparation of polyurethane-based plastics. Certain medicinal compounds are either prepared via epoxide intermediates or are themselves epoxides. The HIV protease inhibitor indinavir is made from 3,4-epoxytetrahydrofuran.3 The antibiotic (1R,2S)-(-)-(1,2)-epoxypropyl phosphonic acid is prepared by the epoxidation of cis-1-propenylphosphonic acid.4-5 Industrial epoxidation relies heavily on heterogeneous catalysts to accelerate the oxidation of alkenes by dioxygen or other oxidants. In the industrial preparation of ethylene oxide, ethylene is directly oxidized with air or pure oxygen at 230-280 oC under 10-35 atm in a packed-bed, multitubular reactor.6-7 Generally, these systems use cesium as a promoter and silver supported on ?-Al2O3 as a catalyst.8-9 The activity relies on silver?s ability to rapidly adsorb and desorb oxygen.10-11 Gold has also been used as a catalyst, albeit on different solid supports such as titanium/silicon.12 Propylene oxide is often produced using a gold catalyst supported on a titanium-containing silicate mesoporous material, with barium serving as a promoter.13 These reactions have several disadvantages, including the cost of the catalysts, the necessity of high reaction temperatures and pressures, and substantial side-reactivity. The high reaction temperatures, in particular, preclude the use of these systems for the production of more thermally sensitive epoxides. 1.2 Development of Alternative Catalysts for Alkene Epoxidation Much recent research in the catalysis of olefin epoxidation reactions is directed at identifying and developing cheaper alternatives to the aforementioned silver and gold species that can speed the oxidation of alkenes under milder conditions. Recently investigated 3 catalysts are remarkably diverse and include both main group and transition metal complexes. Boron,14 arsenic,15-16 selenium,17-19 silicon,20 and aluminum21 compounds have been reported to catalyze the epoxidation of alkenes by H2O2. Transition metal complexes have been explored more thoroughly. The successful catalysts can be further classified into two categories, based on whether they contain early or late transition metal ions. In the catalysts that contain early transition metals, such as molybdenum,22-25 vanadium,26-29 zirconium,30 and titanium,31-34 the metal ions are believed to remain in their highest possible oxidation states during the alkene epoxidation. The metals are believed to accelerate oxygen atom transfer from a metal-oxidant adduct to the substrate by acting as Lewis acids. Late transition metal species used to promote the oxidation of alkenes to epoxides include manganese35-42 and iron complexes.43-45 The proposed mechanisms for alkene epoxidation involving these more redox-active metals generally posit higher valent intermediates, such as MnV=O species, as the active oxidant.46 Although O2 is the most desirable oxygen source due to its ready availability and low cost, it isn?t commonly used in specialty epoxidation reactions because of the high free energy barrier for its activation and its involvement in autoxidation-related radical side-reactions.47 Because of these drawbacks, two-electron terminal oxidants, such as hydrogen peroxide (H2O2), isodosylbenzene (PhIO), sodium hypochlorite (NaOCl), meta-chloroperoxybenzoic acid (m-CPBA), N-methyl-morpholine-N-oxide,38 dimethyldioxirane,48 or peracetic acid (PAA),42 are used instead. The remainder of this Chapter will focus on five specific classes of homogeneous catalysts for alkene epoxidation. These have been chosen due to their relevance to the material presented in Chapters 2-4. 1.2.1 Jacobsen?s Catalyst Jacobsen?s catalyst consists of a MnIII-salen compound (Figure 1.1) and promotes the enantioselective epoxidation of non-functionalized olefins by either PhIO or NaOCl.49-51 Sample illustrative reactions are summarized in Table 1.1. 4 N N O O MnIII + Figure 1.1 Structure of the prototypical Jacobsen catalyst37 Table 1.1 Asymmetric epoxidation of representative olefins by Jacobsen?s catalysta entry olefin substrate yield (%)b ee (%)c equiv of catalyst required for complete reaction 1 Ph Me 84 92 0.04 2 Mep-ClC6H4 67 92 0.04 3 O 72 98 0.02 4 O NC 96 97 0.03 5 O O 63 94 0.15 6d Ph CO2Me 65e 89 0.10 aReactions were run at 4 ?C in CH2Cl2 with 0.25 mmol catalyst and 12.5 mmol substrate. bIsolated yield based on olefin. cEnantiomeric excess. dReaction carried out in the presence of 0.4 equiv of 4-phenylpyridine N-oxide. eYield determined by GC. All data are from reference 35. 5 Several trends were noted in the studies of Jacobsen?s catalyst and its close derivatives. First, electron-donating and withdrawing substituents on substrates influence both the cis/trans selectivity and the rate of the reaction, as summarized in Scheme 1.1. With the more electron-rich alkenes, the cis-epoxide product forms preferentially. It is hypothesized that electron-withdrawing groups stabilize radical intermediates that can form when the substrate binds to the MnV-oxo center. This facilitates C-C bond rotation necessary for the formation of trans-epoxides (Scheme 1.2). CO2R 5% Jacobsen catalystCH2Cl2, 4?C O Ar CO2R O Ar CO2R + X Scheme 1.1 Substituent effects on stereoselectivity 6 Scheme 1.2 Possible pathways for epoxidation catalyzed by Jacobsen?s compound46 The cis/trans ratios of the stilbene oxides produced by the oxidation to stilbene by various iodosylbenzene derivatives were found to be dependent on the terminal oxidant.52 The results suggest at least partial coordination of an equiv of terminal oxidant in the active oxidant, as opposed to the sole agency of a MnV=O species. The counterion also modulates the cis/trans ratio of the epoxide products.53 In the epoxidation of cis-stilbene, [Mn(salen)]+ catalysts with strongly coordinating counteranions favor trans-stilbene oxide as the major product, whereas non-ligating couteranions are associated with the preferential formation of cis-stilbene oxide.54 Computational studies suggest the involvement of different MnV=O spin-states, which are stabilized by coordinating ligands to varying degrees.55-56 7 The salen ligand in Jacobsen?s first epoxidation catalyst is chiral, opening the possibility for chiral epoxide products from prochiral alkenes. The enantioselectivity of MnIII(salen)-catalyzed asymmetric epoxidation is directly correlated to the electronic character of the ligand, with more strongly electron-donating ligands resulting in higher enantiomeric excesses. The ratio of enantiomers for the epoxidation of 2,2-dimethylchromene and cis-?-methylstyrene can be fit to a Hammett plot (Scheme 1.3).36 Jacobsen and co-workers proposed that the ligand?s electron-withdrawing groups enable the Mn(V) oxo to react with the substrate in an earlier transition state, which reduces the selectivity. Electron-donating groups, conversely, stabilize the Mn(V) oxidant, allowing a later transition state and higher enantioselectivity.57 Much effort has been spent towards evaluating the mechanism of the Jacobsen epoxidation. Four mechanisms have been proposed (Scheme 1.2 and Scheme 1.4). In the first proposed mechanism, the olefin substrate reacts directly with the MnV-oxo oxidant, proceeding through a transition state resembling a manganese-bound epoxide. Due to the concerted nature of this pathway, the oxygen atom transfer results in exclusively cis-epoxides. In the second proposed mechanism, metallaoxetane species result from the reaction between the alkene and the Mn(V) species. The subsequent release of the product from the organometallic intermediate can either occur in a single concerted step, resulting in cis-epoxides, or in two steps, which can allow the formation of trans-epoxides. In the third pathway, the MnV-oxo oxidant adds to the alkene as a radical species. The alkyl radical can rotate freely around the former C=C bond, allowing both cis- and trans-epoxides. The fourth pathway (pathway 2 in Scheme 1.4) differs from the previous three in that it does not involve oxidation or reduction of the manganese center.54 Instead, the manganese serves as a Lewis acid, analogous to what has been proposed for early transition metal-containing catalysts for olefin epoxidation. The relevant oxidant in this pathway is a MnIII adduct with the terminal oxidant (OLG in Scheme 1.4). 8 Scheme 1.3 Hammett plots depicting enantiomeric ratios of epoxides generated with catalysts with various electronically modified ligands. Data from reference 34. 9 MnIII Scheme 1.4 Parallel Lewis acid and redox pathways for [Mn(salen)]+ catalyzed epoxidations. 1.2.2 Manganese Complexes with Neutral, Polydentate N-Donors Manganese can also catalyze alkene epoxidation when bound to neutral N-donor ligands. Unlike the aforementioned salen complexes, the metal starts in the +2, rather than +3, oxidation state. Another key difference from the aforementioned salen complexes is that the tetradentate ligand does not coordinate the metal in a square planar fashion. Stack et al evaluated the epoxidation performance of a variety of MnII complexes, finding [MnII(R,R-mcp)(CF3SO3)2] to be the most active catalyst (Figure 1.2).41 Within 5 min, [MnII(R,R-mcp)(CF3SO3)2] efficiently catalyzes the epoxidation of a wide range of terminal olefins by PAA (Table 1.2), with turnover frequencies as high as 250 min-1. The reactivity proceeds with a catalyst loading as low as 0.1 mol%.41 Although terminal alkenes can be oxidize, electron-rich olefins are preferentially oxidized. 10 Figure 1.2 The molecular structure of the R,R-mcp ligand and the crystal structure of [MnII(R,R-mcp)(CF3SO3)2].41 Table 1.2 Epoxidations catalyzed by [MnII(R,R-mcp)(CF3SO3)2] with PAAa aGeneral reaction conditions: [olefin]o = 0.50 M, MeCN, 25 oC. Yields measured at 5 min. bYields determined by GC. The numbers in parentheses represent a standard deviation of a minimum of three experiments. cIsolated yields, 1 g scale. d0.25-mol scale, 88% yield. e98% cis-epoxide, 2% trans-epoxide. f97% trans-epoxide, 3% cis-epoxide. gtrans,trans- CH3CH2=CH2CH2=CH2-CO2CH2CH3. h4:1 mixture of 4,5-monoepoxide and 2,3-monoepoxide. i0 oC, diepoxide product, 20% disastereoselectivity. j-20 ?C, R-carvone-8,9-monoepoxide, 15% disastereoselectivity. All data are from reference 41. 11 In subsequent research, the influence of the solvent?s acidity and the catalyst?s ligand structure were further explored.40, 42 The pH greatly impacts the catalyst reactivity, and the addition of HClO4 slows the initial reaction rate. Conversely, the use of a less acidic grade of PAA results in faster reactivity and higher turnover numbers. The inhibitory effect of acid is attributed to the protonation of the N-donor ligand, which promotes its dissociation from the metal center. Various MnII salts (e.g. MnCl2, Mn(CF3SO3)2) are not competent catalysts for olefin epoxidation in the absence of the N-donor ligand. Neutral, tetradentate N-donor ligands capable of binding Mn(II) in a cis-? conformation best support the reactivity. The availability of cis vacant coordination sites associated with this conformation is believed to facilitate activation of the PAA.40 Electron paramagnetic resonance (EPR) studies were unable to identify high valent manganese species during the catalysis; the EPR spectrum appears unchanged from the Mn(II) starting material after the addition of oxidant. However, this does not necessarily preclude the possible involvement of such high valent intermediates, which simply may not accumulate to detectable levels. The possible contribution of a Lewis acid mechanism was tested by analyzing the reactivity of the redox-inactive [ZnII(R,R-mcp)]2+. The Zn(II) complex could not catalyze the oxidation of alkenes, suggesting that the manganese is accelerating the reactivity through alternative pathways, at least for the R,R-mcp system.41 As with the manganese-salen complexes, multiple mechanistic pathways have been proposed for the Mn(II)-catalyzed olefin epoxidations (Scheme 1.5).58 The leftmost pathway has the manganese remaining in the +2 oxidation state, with a Mn(II)-oxidant adduct transferring an oxygen atom to the substrate. Here, the Mn(II) is simply acting as a Lewis acid. The middle pathway features the Mn(II) being oxidized to a MnIV=oxo intermediate that oxidizes the substrate. Upon the oxygen atom transfer, the manganese gets reduced back to the +2 oxidation state. The oxygen atom transfer can occur in either a single concerted step or in two steps, analogous to that depicted for the manganese salen complexes in Scheme 1.2. In the rightmost pathway, the MnIV=oxo intermediate acts as a Lewis acid activator, binding an additional equiv of oxidant, which provides the transferred oxygen atom. The high-valent manganese oxo species does not change its +4 oxidation state during the actual catalysis. 12 Scheme 1.5 Possible mechanistic pathways for Mn(II)-catalyzed olefin epoxidation.58 1.2.3 Non-Heme Iron Complexes Many non-heme iron complexes have been found to accelerate the epoxidation of alkenes by H2O2. White et al., for instance, found that the complex of Fe(II) with the tetradentate ligand N,N?-dimethyl-N,N?-bis(2-pyridylmethyl)-ethane-1,2-diamine (mep, Figure 1.3) could catalyze the epoxidation of a wide variety of substrates, including terminal olefins, in under 5 min with isolated yields ranging from 60-90% (Table 1.3).45 They crystallized a ?-oxo, carboxylate bridged diiron(III) complex (Figure 1.3) from the reaction between the FeII-mep precursor, H2O2, and acetic acid. Whether this species is catalytically relevant remains unresolved . N N N N Figure 1.3 Illustration of the mep ligand (left) and the crystal structure of dinuclear [Fe2(?-O)(?-CH3CO2)(mep)2]3+ cation.45 13 Table 1.3 Epoxidations catalyzed by [Fe(mep)(CF3SO3)2]a aReaction conditions: olefin (2.0 mmol, 0.16 M in MeCN), mononuclear precursor (3.0 mol %), CH3CO2H, H2O2 (aqueous 50 wt %, 3.0 mmol, 1.5 M in MeCN, added dropwise over 2 min), 4 oC, 5 min. bIsolated yields based on an average of 3 runs. cGC yields were determined using nitrobenzene as an internal standard for especially volatile substrates. dReaction carried out at [olefin]o = 0.13 M due to poor solubility in MeCN. e5 mol % mononuclear precursor. f1.5 mol % mononuclear precursor/1.5 mol % CH3CO2H. g6 mol % CH3CO2H. Data from Reference 45. Beller et al. prepared a catalyst in situ using FeCl3.6H2O as the iron source and a protonated pyridine-2,6-dicarboxylate (H2Pydic) ligand which was proposed to mimic the carboxylate and histidine groups in the coordination sphere of a non-heme dioxygenase.59 Their system showed excellent reactivity and selectivity towards terminal and 1,2-disubstituted aromatic olefins and moderate reactivity towards 1,3-dienes at room temperature (Table 1.4). 14 Table 1.4 Summary of epoxidation scope catalyzed by Beller?s in situ FeIII complexa aReaction procedure: in a 25 mL Schlenk tube, FeCl3.6H2O (0.025 mmol), H2pydic (0.025 mmol), tert-amyl alcohol (9 mL), pyrrolidine (0.05 mmol), olefin (0.5 mmol) and dodecane (GC internal standard, 100 mL) were added in sequence at RT in air. To this mixture, a solution of 30% H2O2 (114 mL, 1.0 mmol) in tert-amyl alcohol (886 mL) was added over a period of 1 h (or 5 min) at RT by a syringe pump. bConversion and yield were determined by GC analysis. cSelectivity refers to the ratio of yield to conversion as percentage. dThe oxidant was added over a period of 5 min. e19% trans-?-methylstyrene oxide was observed. Data from Reference 59. The coordination geometry of the ligand strongly impacts the activity of iron catalysts. Catalysts with pentadentate ligands are found to have inferior stereoselectivity and epoxidation efficiency,60 whereas systems containing tetradentate ligands with a proclivity for binding metal ions in trans conformations generally lead to better reactivity.61-62 The counteranions on the metal salt also influence the activity, similar to what was observed for the manganese-salen systems. In a study of iron compounds with tetradentate N-donor ligands, increasing the amount of chloride present in solution decreases the alkene 15 epoxidation.62 Weakly coordinating anions like triflate and hexafluoroantimonate, conversely, lead to enhanced activity. In the epoxidation catalyzed by non-heme Fe(II) catalysts, two high-valent iron-oxo oxidants have been proposed: FeV=O and FeIV=O. The agency of FeV=O species was suggested as a means to explain how 18O atoms from added H218O were incorporated into the epoxide products.62-63 Such species have not, however, been isolated and spectroscopically characterized. FeIV=O species, conversely, are stable enough to be structurally and spectroscopically studied.64-66 Consequently, Fe(IV) species are currently believed to be more plausible intermediates (Scheme 1.6). FeII N N X X N N X: weakly coordinating anion FeII N N X O N N O H H FeIV N N O O N N H H FeIV N N OH2 O N N H2O2 heterolytic cleavage alkene, X epoxide, H2O Scheme 1.6 Proposed catalytic cycle for non-heme iron catalyzed epoxidation. 1.2.4 The Sharpless Catalyst Sharpless used a mixture of titanium(IV) tetraisopropoxide (Ti(Oi-Pr)4), (+/-)diethyl tartrate, and t-BuOOH to catalyze the asymmetric epoxidation of olefins.31-32 The major species in the titanium-tartrate mixture is a dimeric species (Figure 1.4), as supported by mass spectrometry, NMR, and IR measurements.33 16 Figure 1.4 The dimeric structure of the Sharpless catalyst.33 The Ti(IV) complex catalyzes the oxidation of a variety of substrates by t-BuOOH. (Table 1.5).31 The enantioselectivity is determined by the given tartrate enantiomer and is not dependent on the substituents on the allylic alcohol skeleton. The proposed mechanism of the Sharpless epoxidation is shown on Scheme 1.7. There are several key features about this route: 1) both the substrate and terminal oxidant bind to the Ti(IV) as ligands; 2) the tartrates direct the bound oxidant towards the face of the C=C bond, accounting for the observed enantioselectivity; 3) the epoxide and t-BuO group dissociate from Ti(IV) center to complete the catalytic cycle; 4) the titanium remains in the +4 oxidation state, and the redox-activity is limited to the two transiently bound ligands. 17 Table 1.5 Selected asymmetric epoxidation of allylic alcoholsa Allylic Alcohol Epoxyalcohol % yieldb % e.e.c Configuration OH H OH O 2 3 77 95 2 (S), 3 (S) OH H OH O 2 3 79 94 2 (S), 3 (R) OAc OH H OH O 7 6 70 >95 6 (S), 7 (S) Ph Ph OH Ph H Ph OH O 87 >95 2 (S), 3 (S) n-C9H19 OH H H OH O 2 3 79 >95 2 (S), 3 (S) n-C9H19 OH H H OH O 2 3 82 90 2 (S), 3 (R) n-C9H19 OH O 2 H H OH 3 80 90 2 (R), 3 (S) OH H H OH O 2 81 >95 2 (S) a5%-10% catalyst loading, -20 oC. bIsolated yield. cEnantiomeric excess determined by 1H NMR on the corresponding epoxy acetates. Data from Reference 31. 18 Scheme 1.7 Proposed mechanism for Sharpless epoxidation (E refers to ester group).67 1.2.5 Homogenous Aluminum Catalysts Although aluminum is typically not redox-active, it can support modest alkene epoxidation catalysis. Valentine used [AlIII(porphyrin)]Cl to promote the epoxidation of stilbene and cyclohexene by PhIO, with up to 9.3 turnovers.61 More recently, mixtures of [Al(H2O)6]3+ and H2O2 in MeCN-H2O were reported to catalyze the epoxidation of ?, ?-unsaturated ketones (Table 1.6).68 Much like the aforementioned Sharpless epoxidation, Al(III) catalysis is proposed to proceed through pathways that do not involve metal-centered redox or high-valent metal oxo intermediates. 19 Table 1.6 Epoxidation of ?, ?-unsaturated ketones by [Al(H2O)6]3+/H2O2 C: conversion, S: selectivity. Reaction conditions: substrate (25 mmol), di-n-butylether (internal standard, 12.5 mmol), 70 wt.% H2O2 (56 mmol), Al(ClO4)3?9H2O (0.5 mmol), THF (solvent), at the temperature indicated in this table. Data from Reference 68. The reactivity of (R)-carvone (compound 1 in Table 1.6) is used to assess the electronic character of the active oxidant. (R)-Carvone has two C=C bond sites. The external C=C bond is most nucleophilic, whereas the ring olefin is more electrophilic, as supported by DFT calculations. Epoxidation exclusively occurs at the ring C=C bond site, suggesting a nucleophilic oxidant, e.g. HOO- derived from H2O2. This differs from most transition metal-based systems, which prefer to oxidize more electron-rich olefins. A Sharpless-type mechanism is proposed, based on DFT calculations (Scheme 1.8).21 Much like the aforementioned Ti(IV) catalysis, the oxidant binds to the Al(III) center as a ligand, and the Al(III) serves to activate the oxidant through its Lewis acidity. The substrate is subsequently oxidized by the bound H2O2, generating the epoxide and H2O as a by-product. The [Al(H2O)6]3+/H2O2/MeCN-H2O system requires elevated temperatures and long reaction times, particularly relative to the aforementioned transition metal catalysts. Additionally, the reactivity becomes less selective as the catalyst degrades, leading to increasingly large amounts of side-products such as diols. 20 Scheme 1.8 Proposed mechanism for epoxidation promoted by [Al(H2O)6]3+. 69 1.3 Summary In Chapter 2 of this thesis, we use Mn(II) complexes with phenanthroline derivatives to assess the influence of the ligand?s electronic character on Mn(II)-catalyzed alkene epoxidation. We also study the impact of acidity on the epoxidation reactivity. In Chapters 3 and 4, we study olefin epoxidation catalyzed by Ga(III) complexes with various N-donor ligands. We systematically explore how the ligand influences the reactivity in Chapter 4. 21 References (1) Ullmann's Encyclopedia of Industrial Chemistry, Sixth Edition, John Wiley & Sons, Inc., 2003, Vol 12, Ethylene Glycol, 597. (2) Ullmann's Encyclopedia of Industrial Chemistry, Sixth Edition, John Wiley & Sons, Inc., 2003, Vol 28, Polyoxalkylenes, 497. (3) Senanayake, C. H., Roberts, F. E., DiMichele, L. M., Ryan, K. M., Liu, J., Fredenburgh, L. E., Foster, B. S., Douglas, A. W., Larsen, R. D., Verhoeven, T. R. and Reider, P. J., Tetrahedron Lett. 1995, 36, 3993-3996. (4) Wang, X.-Y., Shi, H.-C., Sun, C. and Zhang, Z.-G., Tetrahedron 2004, 60, 10993-10998. (5) Hendlin, D., Stapley, E. O., Jackson, M., Wallick, H., Miller, A. K., Wolf, F. J., Miller, T. W., Chaiet, L., Kahan, F. M., Foltz, E. L., Woodruff, H. B., Mata, J. M., Hernandez, S. and Mochales, S., Science 1969, 166, 122-123. (6) Mao, C.-F. and Albert Vannice, M., Appl. Catal., A 1995, 122, 61-76. (7) Ayame, A., Uchida, Y., Ono, H., Miyamoto, M., Sato, T. and Hayasaka, H., Appl. Catal., A 2003, 244, 59-70. (8) Linic, S. and Barteau, M. A., J. Catal. 2003, 214, 200-212. (9) Atkins, M., Couves, J., Hague, M., Sakakini, B. H. and Waugh, K. C., J. Catal. 2005, 235, 103-113. (10) Linic, S. and Barteau, M. A., J. Am. Chem. Soc. 2001, 124, 310-317. (11) Linic, S. and Barteau, M. A., J. Am. Chem. Soc. 2003, 125, 4034-4035. (12) Taylor, B., Lauterbach, J. and Delgass, W. N., Appl. Catal., A 2005, 291, 188-198. (13) Lu, J., Zhang, X., Bravo-Su?rez, J. J., Bando, K. K., Fujitani, T. and Oyama, S. T., J. Catal. 2007, 250, 350-359. (14) Wolf, P. F. and Barnes, R. K., J. Org. Chem. 1969, 34, 3441-3445. (15) Tosaki, S.-y., Tsuji, R., Ohshima, T. and Shibasaki, M., J. Am. Chem. Soc. 2005, 127, 2147-2155. (16) van Vliet, M. C. A., Arends, I. W. C. E. and Sheldon, R. A., Tetrahedron Lett. 1999, 40, 5239-5242. 22 (17) Hori, T. and Sharpless, K. B., J. Org. Chem. 1978, 43, 1689-1697. (18) Jacobson, S. E., Mares, F. and Zambri, P. M., J. Am. Chem. Soc. 1979, 101, 6946-6950. (19) ten Brink, G.-J., Fernandes, B. C. M., van Vliet, M. C. A., Arends, I. W. C. E. and Sheldon, R. A., J. Chem. Soc., Perkin Trans. 1 2001, 224-228. (20) Rebek, J. and McCready, R., Tetrahedron Lett. 1979, 20, 4337-4338. (21) Kuznetsov, M. L., Kozlov, Y. N., Mandelli, D., Pombeiro, A. J. L. and Shul?pin, G. B., Inorg. Chem. 2011, 50, 3996-4005. (22) Sobczak, J. and Zi??kowski, J. J., J. Mol. Catal. 1977, 3, 165-172. (23) Sharpless, K. B., Townsend, J. M. and Williams, D. R., J. Am. Chem. Soc. 1972, 94, 295-296. (24) Ledon, H. J., Durbut, P. and Varescon, F., J. Am. Chem. Soc. 1981, 103, 3601-3603. (25) Mitchell, J. M. and Finney, N. S., J. Am. Chem. Soc. 2001, 123, 862-869. (26) Chong, A. O. and Sharpless, K. B., J. Org. Chem. 1977, 42, 1587-1590. (27) Itoh, T., Jitsukawa, K., Kaneda, K. and Teranishi, S., J. Am. Chem. Soc. 1979, 101, 159-169. (28) Mimoun, H., Mignard, M., Brechot, P. and Saussine, L., J. Am. Chem. Soc. 1986, 108, 3711-3718. (29) Murase, N., Hoshino, Y., Oishi, M. and Yamamoto, H., J. Org. Chem. 1999, 64, 338-339. (30) Li, Z. and Yamamoto, H., J. Am. Chem. Soc. 2010, 132, 7878-7880. (31) Katsuki, T. and Sharpless, K. B., J. Am. Chem. Soc. 1980, 102, 5974-5976. (32) Woodard, S. S., Finn, M. G. and Sharpless, K. B., J. Am. Chem. Soc. 1991, 113, 106-113. (33) Finn, M. G. and Sharpless, K. B., J. Am. Chem. Soc. 1991, 113, 113-126. (34) Corey, E. J., J. Org. Chem. 1990, 55, 1693-1694. (35) Jacobsen, E. N., Deng, L., Furukawa, Y. and Mart?nez, L. E., Tetrahedron 1994, 50, 4323-4334. (36) Jacobsen, E. N., Zhang, W. and Guler, M. L., J. Am. Chem. Soc. 1991, 113, 6703-6704. (37) Jacobsen, E. N., Zhang, W., Muci, A. R., Ecker, J. R. and Deng, L., J. Am. Chem. Soc. 1991, 113, 7063-7064. (38) Palucki, M., Pospisil, P. J., Zhang, W. and Jacobsen, E. N., J. Am. Chem. Soc. 1994, 116, 9333-9334. 23 (39) Terry, T. J. and Stack, T. D. P., J. Am. Chem. Soc. 2008, 130, 4945-4953. (40) Murphy, A. and Stack, T. D. P., J. Mol. Catal. A: Chem. 2006, 251, 78-88. (41) Murphy, A., Dubois, G. and Stack, T. D. P., J. Am. Chem. Soc. 2003, 125, 5250-5251. (42) Murphy, A., Pace, A. and Stack, T. D. P., Org. Lett. 2004, 6, 3119-3122. (43) Dubois, G., Murphy, A. and Stack, T. D. P., Org. Lett. 2003, 5, 2469-2472. (44) Chen, K., Costas, M., Kim, J., Tipton, A. K. and Que, L., J. Am. Chem. Soc. 2002, 124, 3026-3035. (45) White, M. C., Doyle, A. G. and Jacobsen, E. N., J. Am. Chem. Soc. 2001, 123, 7194-7195. (46) Flessner, T. and Doye, S., J. Prakt. Chem. 1999, 341, 436-444. (47) Pozzi, G. and Cinato, F., Chem. Commun. 1998, 877-878. (48) Kass, S. R. and DePuy, C. H., J. Org. Chem. 1985, 50, 2874-2877. (49) Linker, T., Angew. Chem. Int. Ed. 1997, 36, 2060-2062. (50) Feichtinger, D. and Plattner, D. A., Angew. Chem. Int. Ed. 1997, 36, 1718-1719. (51) Feichtinger, D. and Plattner, D. A., Chem. Eur. J. 2001, 7, 591-599. (52) Collman, J. P., Zeng, L. and Brauman, J. I., Inorg. Chem. 2004, 43, 2672-2679. (53) Adam, W., Roschmann, Konrad J. and Saha-M?ller, Chantu R., Eur. J. Org. Chem. 2000, 2000, 3519-3521. (54) Adam, W., Roschmann, K. J., Saha-M?ller, C. R. and Seebach, D., J. Am. Chem. Soc. 2002, 124, 5068-5073. (55) Schr?der, D., Shaik, S. and Schwarz, H., Acc. Chem. Res. 2000, 33, 139-145. (56) Abashkin, Y. G. and Burt, S. K., Org. Lett. 2003, 6, 59-62. (57) Palucki, M., Finney, N. S., Pospisil, P. J., G?ler, M. L., Ishida, T. and Jacobsen, E. N., J. Am. Chem. Soc. 1998, 120, 948-954. (58) Ottenbacher, R. V., Bryliakov, K. P. and Talsi, E. P., Inorg. Chem. 2010, 49, 8620-8628. (59) Anilkumar, G., Bitterlich, B., Gelalcha, F. G., Tse, M. K. and Beller, M., Chem. Commun. 2007, 0, 289-291. (60) Bukowski, M. R., Comba, P., Lienke, A., Limberg, C., Lopez de Laorden, C., Mas-Ballest?, R., Merz, M. and Que, L., Angew. Chem. Int. Ed. 2006, 45, 3446-3449. (61) Nam, W., Ho, R. and Valentine, J. S., J. Am. Chem. Soc. 1991, 113, 7052-7054. 24 (62) Mas-Ballest?, R., Costas, M., van den Berg, T. and Que, L., Chem. Eur. J. 2006, 12, 7489-7500. (63) Oldenburg, P. D., Shteinman, A. A. and Que, L., J. Am. Chem. Soc. 2005, 127, 15672-15673. (64) Rohde, J.-U., In, J.-H., Lim, M. H., Brennessel, W. W., Bukowski, M. R., Stubna, A., M?nck, E., Nam, W. and Que, L., Science 2003, 299, 1037-1039. (65) Klinker, E. J., Kaizer, J., Brennessel, W. W., Woodrum, N. L., Cramer, C. J. and Que, L., Angew. Chem. 2005, 117, 3756-3760. (66) Costas, M., Mehn, M. P., Jensen, M. P. and Que, L., Chem. Rev. 2004, 104, 939-986. (67) Katsuki, T. and Martin, V. Asymmetric Epoxidation of Allylic Alcohols: the Katsuki?Sharpless Epoxidation Reaction, Org. React., 1996, 48, 1-299. (68) Rinaldi, R., de Oliveira, H. F. N., Schumann, H. and Schuchardt, U., J. Mol. Catal. A: Chem. 2009, 307, 1-8. (69) Kuznetsov, M. L., Kozlov, Y. N., Mandelli, D., Pombeiro, A. J. L. and Shul?pin, G. B., Inorg. Chem. 2011, 50, 3996-4005. 25 Chapter 2 Alkene Epoxidation Catalyzed by Manganese Complexes with Electronically Modified 1,10-Phenanthroline Ligands* * This chapter is a revision of a published paper: Christian R. Goldsmith, Wenchan Jiang, Inorg. Chim. Acta, 2012, 384, 340-344. Reprinted with permission. Copyright? 2012 by Elsevier B.V. 26 2.1 Introduction The epoxidation of terminal and electron-de?cient ole?ns remains a challenging task in synthetic chemistry. Although many alkenes react directly with peracids, these uncatalyzed reactions often require extended reactions times and/or elevated temperatures, both of which can reduce selectivity. The use of transition metal catalysts can enhance the selectivity, reduce reaction times, and allow reactivity at ambient temperatures 1-21. Many homogeneous and heterogeneous catalysts for epoxidation reactions employ manganese as the active metal.1-13 Electron-rich ole?ns are generally more reactive, to the extent that the development of catalysts capable of activating more electron-de?cient ole?ns, such as terminal alkenes, remains an intense area of research.1-4, 21 The terminal oxidant and the acidity of the reaction also strongly in?uence the activity.1-2, 13, 22 Systematic studies on the in?uence of the electronic nature of the catalyst, as opposed to that of the substrate,23 on alkene epoxidation are comparatively rare.24-25 In studies involving manganese?salen catalysts, the addition of electron-donating substituents onto the catalyst?s ligand were found to slow the rate and improve the stereoselectivity of ole?n oxidation.10, 24 This has been attributed to the ability of the electron donors to stabilize the Mn=O bond in the Mn(V) oxo species that is believed to be the active oxidant.25 We report three novel monomeric manganese complexes with electronically modi?ed derivatives of 1,10-phenanthroline (phen) that are related to two highly active catalysts reported by Murphy and Stack.2 1,10-phenanthroline is chosen as the ligand backbone based on the facts that with versatile derivatives it has strong chelating ability and is relatively oxidative robust during catalysis. The electronic perturbations are installed on the 5-position of the phen ligands in order to minimize the in?uence of steric effects on the alkene epoxidation. Scheme 2.1 shows the molecular structure of the phen derivatives and a sketch of the anticipated coordination around the Mn(II) ion.26-27 Each of the [Mn(R-phen)2Cl2] compounds (R = NH2, CH3, H, NO2) can catalyze the oxygenation of alkenes by peracetic acid (Scheme 2.2). The yields of the reactions are correlated with the electron-donating capacity of the substituent on the 5-position, with electron-withdrawing groups impeding, not hastening, the alkene epoxidation. 27 Scheme 2.1 Illustration of Mn(R-phen)2Cl2 complexes Scheme 2.2 Mn(R-phen)2Cl2 catalyzed epoxidation reaction 2.2 Experimental Materials Acetonitrile (MeCN), manganese(II) chloride (MnCl2), 1,10-phenanthroline (phen), 5-nitro-1,10-phenanthroline (NO2-phen), 5-chloro-1,10-phenanthroline (Cl-phen), cyclohexene, cyclohexene oxide, 3-cyclohexenol, 3-cyclohexenone, cyclooctene, 1-octene, 2,4-hexadienoic acid (sorbic acid), peracetic acid (32% in acetic acid, PAA), and 1,2-dichlorobenzene were purchased from Sigma?Aldrich and used as received. 5-Amino-1,10-phenanthroline (NH2-phen) and 5-methyl-1,10-phenanthroline (Me-phen) were acquired from Acros. Anhydrous ethanol (EtOH) and anhydrous diethyl ether (ether) were bought from Pharmco-Aaper and Fisher Scienti?c, respectively. Deuterated acetonitrile (CD3CN) was purchased from Cambridge Isotopes. Instrumentation All 1H magnetic resonance (NMR) spectra were acquired on a 400 MHz AV Bruker NMR spectrometer at 294 K. All NMR resonances were referenced to internal standards. A Shimadzu IR Prestige-21 FT-IR spectrophotometer was used to record infrared spectra on samples that were prepared as KBr pellets. All samples that were submitted for elemental analysis were powdered and dried before shipment; Atlantic Microlabs (Norcross, GA) per- formed all elemental analysis. The magnetic susceptibilities of solid samples were measured 28 on a Johnson Matthey balance (model MK I#7967). The diamagnetic contribution to each magnetic susceptibility was estimated using Pascal?s constants.28 Electron paramagnetic resonance (EPR) spectra were collected on a Bruker EMX-6/1 X-band EPR spectrometer operated in the perpendicular mode. Each EPR sample was run as a frozen 3:2 MeCN/CHCl3 solution in a quartz tube, and the program EasySpin was used to analyze the acquired EPR data. Gas chromatography (GC) was obtained on a ThermoScienti?c Trace GC Ultra spectrometer with a ?ame ionization detector (FID). Syntheses Dichloro(bis(5-nitro-1,10-phenanthroline))manganese(II), [Mn(NO2-phen)2Cl2] A solution of NO2-phen (0.105 g, 0.47 mmol) in 10 mL of EtOH was added to 0.046 g of MnCl2 (0.23 mmol). The resultant slurry was heated until the residual manganese salt had dissolved. The solution was ?ltered and cooled to precipitate the product as a yellow powder (0.113 g, 85%). Solid-state magnetic susceptibility (294 K): ?eff = 5.7 ?B. Elemental Anal. Calc. for C24H14Cl2MnN6O4: C, 50.02; H, 2.45; N, 14.58. Found: C, 49.80; H, 2.51; N, 14.54%. FT-IR (KBr, cm-1): 3422 (m), 3325 (m), 3226 (m), 3051 (m), 2951 (m), 2922 (m), 2852 (m), 2729 (w), 2694 (w), 1634 (s), 1628 (s), 1501 (s), 1480 (s), 1378 (m). Dichloro(bis(1,10-phenanthroline))manganese(II), [Mn(phen)2Cl2] The synthesis of this compound has been reported previously 17. Solid-state magnetic susceptibility (294 K): ?eff = 5.8 ?B. Elemental Anal. Calc. for C24H16Cl2MnN4: C, 59.28; H, 3.32; N, 11.52. Found: C, 58.41; H, 3.25; N, 11.35.FT-IR% (KBr, cm-1): 2951 (w), 2922 (w), 2852 (w), 2727 (w), 2673 (w), 1635 (m), 1544 (m), 1494 (m), 1300 (m). Dichloro(bis(5-methyl-1,10-phenanthroline))manganese(II), [Mn(Me-phen)2Cl2] Me-phen (0.100 g, 0.52 mmol) and MnCl2 (0.051 g, 0.26 mmol) were added to 10 mL of EtOH. The resultant slurry was re?uxed until the solids had completely dissolved. The solution was ?ltered and cooled to precipitate the product as a yellow powder (0.106 g, 79%). Solid-state magnetic susceptibility (294 K): ?eff = 5.6 ?B. Elemental Anal. Calc. for C26H20Cl2MnN4.H2O: C, 58.66; H, 4.17; N, 10.53. Found: C, 58.83; H, 3.83; N, 10.52%. FT-IR (KBr, cm-1): 3043 (w), 2951 (s), 2922 (s), 2852 (s), 2725 (w), 2673 (w), 1462 (m), 1377 (m). Dichloro(bis(5-amino-1,10-phenanthroline))manganese(II), [Mn(NH2-phen)2Cl2] NH2-phen (0.103 g, 0.51 mmol) was dissolved in 10 mL of EtOH. The resultant solution was added to 0.051 g of MnCl2 (0.25 mmol) and heated to dissolve as much of the residual manganese salt as possible. The hot solution was ?ltered, then cooled to deposit the product 29 as a yellow powder (0.112 g, 87%). Solid-state magnetic susceptibility (294 K): ?eff = 5.7 ?B. Elemental Anal. Calc. for C24H18Cl2MnN6: C, 55.83; H, 3.51; N, 16.28. Found: C, 55.24; H, 3.60; N, 15.97%. FT-IR (KBr, cm-1): 3385 (w), 2922 (s), 2854 (s), 2725 (w), 2671 (w), 1529 (m), 1514 (m), 1462 (m), 1377 (m), 1350 (m). Reactivity studies All reactions were run in MeCN at 0 oC under N2. The initial concentrations of the catalyst, ole?n substrate, and terminal oxidant (PAA) were 0.010, 1.00, and 2.00 mM, respectively. The reaction mixtures were stirred for 1 h, at which point excess ether was added to precipitate the manganese compounds. With all substrates except sorbic acid, the quenched reaction mixtures were sequentially ?ltered through plugs of neutral alumina, basic alumina, and silica gel to remove the remaining PAA and metal salts. For these substrates, the preceding workup did not remove organic starting materials or products from the reaction mixture. For reactions using sorbic acid as the substrate, the reaction mixtures were not run through the basic alumina plug; the workup was otherwise identical to that previously described. All products were identi?ed by both GC and NMR. The yields of the products were quanti?ed relative to both an internal standard of 1,2-dichlorobenzene, which does not react under these conditions, and the remaining alkene starting material. All reactions were run at least six times to ensure reproducibility; the yields were averaged to provide the values listed on Table 2.1. Table 2.1 Yields of epoxides for Mn(II)-catalyzed olefin oxidation by PAA. Substrate R = NH2 R = Me R = H R= NO2 Controla Cyclohexene 41 (?7) 40 (?7) 36 (?7) 24 (?5) 5 (?3) 1-Octene 39 (?4) 36 (?6) 29 (?7) 23 (?7) 0b Cyclooctene 87 (?6) 78 (?7) 76 (?7) 69 (?9) 20 (?3) Sorbic acid 30 (?3) 25 (?3) 23 (?5) 16 (?5) 0 Initial concentrations in MeCN: [substrate] = 1.00 mM, [PAA] = 2.00 mM, [Mn(R-phen)2Cl2] = 0.010 mM. Reactions ran at 0 oC for 1 h under N2. a: Experiments were run in the absence of Mn(II) with identical concentrations of substrate and PAA. b: No epoxide products were detected above the limit of detection. 30 2.3 Results Syntheses and characterization of the manganese complexes All four organic ligands are commercially available (Scheme 2.1), and the syntheses of the associated Mn(II) complexes are straightforward. The complex [Mn(phen)2Cl2] had been reported previously;29 the other three Mn(II) compounds can prepared through the same fundamental procedure. The yields for the metal complex syntheses range from 70% to 83%, with no correlation between the yield and the identity of the substituent on the 5 -position of the phen. Spectroscopically, the four compounds are typical Mn(II) complexes. Each is pale yellow, with no ligand-to-metal charge transfer bands past 300 nm. The magnetic moments of the compounds range from 5.6 to 5.8 ?B and are most consistent with high-spin d5 metal centers. The electronic modi?cations to the ligand are not great enough to force a change in the spin-state. Reactivity studies with manganese compounds The abilities of the four manganese compounds to act as catalysts for ole?n epoxidation were tested with four ole?n substrates: cyclohexene, cyclooctene, 1-octene, and sorbic acid (Scheme 2.3, Table 2.1). The latter two are relatively electron-de?cient and resistant to epoxidation. In the absence of a metal catalyst, no epoxidation above the detection limit of the GC is observed over the course of 1 h for sorbic acid and 1-octene. The more electron-rich cyclohexene and cyclooctene, conversely, will react directly with PAA, with non-negligible amounts of products observable at 1 h. With all four ole?n substrates, the yield of the manganese-catalyzed epoxidation dwarfs that of the uncatalyzed reaction. As shown in Figure 2.1, the epoxidation reaction is mostly ?nished at 1 h. The yield continues to increase past this point, but this additional activity can be attributed to the uncatalyzed reaction between the residual ole?n and the peracid. Only epoxides and unreacted alkene starting materials are observed in the product mixtures. Cyclohexene is often oxidized at the allylic position to 3-cyclohexenol and 3-cyclohexenone,11-12, 30 but neither of these side-products is observed. Diols, which have been observed as products in some transition metal-catalyzed epoxidation reactions,19-20, 31-33 are likewise not seen above the limit of detection. The diene sorbic acid is oxygenated at the more electron-rich alkene farther from the carboxylic acid functional group, yielding 4,5-sorbic acid oxide as the sole organic product. 31 EPR spectroscopy of [Mn(NO2-phen)2Cl2] The reaction between [Mn(NO2-phen)2Cl2] and PAA was followed by EPR (Figure 2.2). The addition of PAA to a solution of the Mn(II) compound in 3:2 MeCN/CHCl3 does not generate Mn(IV) species above the limit of detection, as had been observed by Ottenbacher et al. in a recent study of manganese complexes with neutral N-donor ligands.13 The spectra of the solutions with 0, 2, 4, and 8 equiv of PAA are nearly identical, with integrated signal intensities that are within error of each other. Although the solvent and the concentrations of PAA in these samples are different than those used in the catalytic epoxidation reactions, they were suf?cient to generate Mn(IV) species in the aforementioned studies of Mn(II) complexes with aminopyridyl ligands.13 When 50 equiv of PAA are added, the integrated signal intensity decreases by 15%. This is accompanied by a noticeable browning of the solution, which is typical of oxidation to either Mn(III) or Mn(IV) 34. Monitoring the reaction by optical spectroscopy reveals that the brown color is due to an absorption band at 365 nm. Scheme 2.3 Substrates used in this study Figure 2.1 Epoxidation of cyclooctene by peracetic acid catalyzed by [Mn(NO2-phen)2Cl2] in MeCN at 0 oC under N2. The results from four independent kinetic runs are superimposed. 32 Figure 2.2 X-band EPR spectra of a solution of 1.1 mM [Mn(NO2-phen)2Cl2] in 3:2 MeCN/CHCl3 (solid). To this solution was added 2, 4, 8, and 50 equiv of peracetic acid. The following signal intensities were calculated through the double integration of the ?rst-derivative spectra with subsequent normalization to the intensity of the 0 Equiv spectrum: 1.00 (0 Equiv), 0.99 (2 Equiv), 0.97 (4 Equiv), 1.01 (8 Equiv), 0.85 (50 Equiv). The temperature of all samples was 77 K. 2.4 Discussion Installing electronic modi?cations on organic ligands has been shown to impact the reactivity of transition metal catalysts. In the current work, Mn(II) complexes with electronically modi?ed phenanthroline (phen) ligands are prepared and used to catalyze the epoxidation of alkenes by peracetic acid (PAA). The electronic perturbations are installed on the 5-position of the phen heterocycle (Scheme 1); this placement minimizes the ability of the added functional groups to modulate the chemistry through steric effects. Previously, it was found that Mn(II) complexes with 2-derivatized phen ligands displayed severely curtailed oxidative catalysis.2 The functional groups present on the 2-position either block the access of substrate and/or terminal oxidant to the metal center or weaken the Mn-N bonds enough to promote the dissociation of the ligands from the manganous ions.2 Either of these events would account for the diminished reactivity. The reactivity was tested using commercially available PAA as the terminal oxidant. Commercially available PAA is not as effective as a previously reported, less concentrated grade of PAA;35 consequently, the yields (Table 2.1) and reaction times (Figure 2.1) of the 33 reported ole?n epoxidations compare poorly to previously reported epoxidation catalyzed by Mn(II)-phen species.2 The greatest number of catalytic turnovers is 87, corresponding to the epoxidation of cyclooctene to cyclooctene oxide by [Mn(NH2-phen)2Cl2], and 20 of those turnovers can be accounted for by the background reactivity (Table 2.1). The oxidative ef?ciency, de?ned as the yield with respect to the terminal oxidant, is also inferior, reaching a maximum of 44% in the aforementioned cyclooctene reaction. The selectivities of the catalyzed reactions for the epoxide, however, are retained,2 and no products corresponding to allylic oxidation or dihydroxylation are observed. The less extensive reactivity offers an advantage in that it allows us to differentiate the oxygenating activities of the four manganese catalysts. Murphy and Stack had previously compared the reactivity of manganese complexes with 1,10-phenanthroline and 5-chloro- -1,10-phenanthroline. With a 5 min reaction time, they did not observe a meaningful difference in the epoxidation activity using either commercial or custom-made PAA.2 Using a less ef?cient terminal oxidant and extending the reaction times to 1 h reveal the impact of such electronic modi?cations. When an electron-withdrawing substituent (NO2) is used, the yields of all four alkene epoxidation reactions decrease (Figure 2.3). Conversely, electron-donating substituents (NH2, Me) increase the yields of the reactions measured at 1 h. The decrease in the yield going from R = NH2 to NO2 is approximately 15% for each substrate, with a strong linear correlation between the yield at 1 h and the ?P parameter for the substituent on the 5-position of the phen. With cyclohexene, 1-octene, and sorbic acid, the reactivity is approximately halved upon switching the ligand from NH2-phen to NO2-phen. The R values for the ?ts in Figure 2.3 range from 0.95 (cyclohexene) to 0.98 (cyclooctene). These results differ from the manganese?salen chemistry developed by Jacobsen, in which electron-withdrawing substituents were found to increase the rate of epoxidation.10, 24-25 As with other reported systems, more electron-rich ole?ns (cyclohexene, cyclooctene) are oxygenated to a greater extent than more electron-de?cient ones (sorbic acid, 1-octene),1, 13, 24, 30, 36 and the diene sorbic acid is oxygenated at the more electron-rich 4,5-alkene. 34 Figure 2.3 Plot of the reaction yields (as assessed at 1 h) for the oxygenation of each alkene as a function of the ?P value of the 5-substituent on the catalyst?s phenanthroline ligands. There is compelling evidence that Jacobsen?s system proceeds through a Mn(V) oxo species,10, 37 which installs the oxygen atom onto the alkene through either a single two-electron or two one-electron steps.23 A study of Mn(II) complexes with aminopyridyl ligands also implicates higher-valent oxidants in epoxidation, speci?cally Mn(IV) oxo adducts with the terminal oxidant.13 The presence of electron-withdrawing groups would further destabilize these already electron-de?cient metal centers and provide a greater thermodynamic impetus for the reduction of the manganese oxidant, with concomitant oxidation of the alkene.24-25 That we see the opposite effect may indicate either a different rate-determining step or a fundamentally different mechanism. The greater negative charge on the salens used in the aforementioned work would be anticipated to better stabilize more positively charged metal centers (e.g. Mn(V)) relative to neutral ligands. The EPR spectra of [Mn(NO2-phen)2Cl2] and the [Mn (NO2-phen)2Cl2]/PAA mixtures differ only slightly, and we observe no unambiguous evidence for Mn(IV) even with a 50-fold excess of terminal oxidant. Parallel experiments monitored by NMR did not reveal any paramagnetic resonances that would be indicative of Mn(III). The EPR signal intensity of the Mn(II), however, does decrease 15% when a 50-fold excess of PAA is added. This loss of signal intensity and the concurrent browning of the solution under these circumstances suggest that the manganese is being oxidized. The data suggest that [Mn(R-phen)2Cl2] 35 complexes may be more difficult to oxidize than related salen and aminopyridyl manganese complexes but certainly cannot preclude higher-valent oxidants in the catalytic cycle. One mechanistic possibility is that the reactivity proceeds through a Mn(IV) oxo species but with the rate-limiting step being the oxidation of the Mn(II) rather than the oxygen atom transfer from the oxidized manganese species to the alkene. This would prevent the accumulation of Mn(IV) above the detection threshold of EPR and would explain the decreased activity with more electron-withdrawing ligands. In their analysis of non-porphyrin manganese catalysts, Murphy and Stack proposed that the acid in many epoxidation reactions quenches the reactivity by protonating the phenanthroline, which facilitates their removal from the manganese ions.1 In their control studies, these unbound manganese ions were found to be unproductive as catalysts for epoxidation. The more electron-de?cient phen derivatives are anticipated to bind to the metal ions less avidly and [Mn(R-phen)2Cl2] complexes with these ligands may be more susceptible to acid-induced deactivation. Although we do not observe any of the 1H NMR features associated with dissociated phenanthroline ligands in mixtures of the [Mn(R-phen)2Cl2] species with either acetic acid or peracetic acid, this remains a plausible alternative explanation for the reduced activity of [Mn(NO2-phen)2Cl2]. Another possibility consistent with the EPR data is that the active oxidant is a Mn(II)?PAA adduct and that higher-valent manganese species are not relevant to the hydrocarbon oxidation. Similar low-valent metal species have been mentioned as potential oxidants in other transition metal-based epoxidations.13, 38 Furthermore, complexes with metals that generally do not exhibit much redox activity (e.g. Ti(IV), Al(III), Zn(II)) have been found to act as competent catalysts for alkene epoxidation.30, 39-41 If the metal ion were merely acting as a Lewis acid, however, more strongly electron-withdrawing ligands would be anticipated to amplify the reactivity, which is contrary to what we observe. 2.5 Conclusions A series of Mn(II) complexes with electronically modi?ed phenanthroline ligands has been prepared. The three novel compounds catalyze the epoxidation of alkenes by commercially available peracetic acid. Unexpectedly, the more electron-rich ligands lead to faster reactivity with all alkenes investigated, as assessed by the yields of the ole?n oxidations measured at 1 36 h. The results may suggest that [Mn(phen)2Cl2] and related complexes may oxidize alkenes to epoxides through a fundamentally different free energy pathway than Mn(III)?salen catalysts. Further experimental and theoretical work is needed to fully elucidate these differences. 37 References (1) Murphy, A., Pace, A. and Stack, T. D. P., Org. Lett. 2004, 6, 3119-3122. (2) Murphy, A. and Stack, T. D. P., J. Mol. Catal. A: Chem. 2006, 251, 78-88. (3) Murphy, A., Dubois, G. and Stack, T. D. P., J. Am. Chem. Soc. 2003, 125, 5250-5251. (4) Ho, K.-P., Wong, W.-L., Lam, K.-M., Lai, C.-P., Chan, T. H. and Wong, K.-Y., Chem. Eur. J. 2008, 14, 7988-7996. (5) Jacobsen, E. N., Zhang, W., Muci, A. R., Ecker, J. R. and Deng, L., J. Am. Chem. Soc. 1991, 113, 7063-7064. (6) Terry, T. J. and Stack, T. D. P., J. Am. Chem. Soc. 2008, 130, 4945-4953. (7) Wu, M., Wang, B., Wang, S., Xia, C. and Sun, W., Org. Lett. 2009, 11, 3622-3625. (8) Fujii, Y., Ebina, F., Yanagisawa, M., Matsuoka, H. and Kato, T., J. Inorg. Organomet. P. 1994, 4, 273-288. (9) Collman, J. P., Lee, V. J., Kellen-Yuen, C. J., Zhang, X., Ibers, J. A. and Brauman, J. I., J. Am. Chem. Soc. 1995, 117, 692-703. (10) McGarrigle, E. M. and Gilheany, D. G., Chem. Rev. 2005, 105, 1563-1602. (11) Nehru, K., Kim, S. J., Kim, I. Y., Seo, M. S., Kim, Y., Kim, S.-J., Kim, J. and Nam, W., Chem. Commun. 2007, 4623-4625. (12) Yin, G., Buchalova, M., Danby, A. M., Perkins, C. M., Kitko, D., Carter, J. D., Scheper, W. M. and Busch, D. H., J. Am. Chem. Soc. 2005, 127, 17170-17171. (13) Ottenbacher, R. V., Bryliakov, K. P. and Talsi, E. P., Inorg. Chem. 2010, 49, 8620-8628. (14) Sharpless, K. B. and Michaelson, R. C., J. Am. Chem. Soc. 1973, 95, 6136-6137. (15) White, M. C., Doyle, A. G. and Jacobsen, E. N., J. Am. Chem. Soc. 2001, 123, 7194-7195. (16) Kamata, K., Yonehara, K., Sumida, Y., Yamaguchi, K., Hikichi, S. and Mizuno, N., Science 2003, 300, 964-966. (17) Gelalcha, F. G., Anilkumar, G., Tse, M. K., Br?ckner, A. and Beller, M., Chem. Eur. J. 2008, 14, 7687-7698. (18) B?hl, M., Schurhammer, R. and Imhof, P., J. Am. Chem. Soc. 2004, 126, 3310-3320. (19) Chen, K., Costas, M., Kim, J., Tipton, A. K. and Que, L., J. Am. Chem. Soc. 2002, 124, 3026-3035. (20) Mas-Ballest?, R. and Que, L., J. Am. Chem. Soc. 2007, 129, 15964-15972. 38 (21) Dubois, G., Murphy, A. and Stack, T. D. P., Org. Lett. 2003, 5, 2469-2472. (22) Zhang, J.-L. and Che, C.-M., Chem. Eur. J. 2005, 11, 3899-3914. (23) Linde, C., Kolia?, N., Norrby, P.-O. and ?kermark, B., Chem. Eur. J. 2002, 8, 2568-2573. (24) Jacobsen, E. N., Zhang, W. and Guler, M. L., J. Am. Chem. Soc. 1991, 113, 6703-6704. (25) Cavallo, L. and Jacobsen, H., J. Org. Chem. 2003, 68, 6202-6207. (26) Malinowski, T. J., Kravtsov, V. C., Simonov, Y. A., Lipkowski, J. and Bologa, O. A., J. Coord. Chem. 1996, 37, 187-193. (27) McCann, M., Casey, M. T., Devereux, M., Curran, M. and McKee, V., Polyhedron 1997, 16, 2741-2748. (28) Bain, G. A. and Berry, J. F., J. Chem. Educ. 2008, 85, 532. (29) Morcom, R. E. and Bell, C. F., J. Inorg. Nucl. Chem. 1973, 35, 1865-1874. (30) Nam, W. and Valentine, J. S., J. Am. Chem. Soc. 1990, 112, 4977-4979. (31) Matsumoto, T., Furutachi, H., Kobino, M., Tomii, M., Nagatomo, S., Tosha, T., Osako, T., Fujinami, S., Itoh, S., Kitagawa, T. and Suzuki, M., J. Am. Chem. Soc. 2006, 128, 3874-3875. (32) Mas-Ballest?, R., Costas, M., van den Berg, T. and Que, L., Chem. Eur. J. 2006, 12, 7489-7500. (33) Qui?onero, D., Musaev, D. G. and Morokuma, K., Inorg. Chem. 2003, 42, 8449-8455. (34) Cotton, F. A., Wilkinson, G., Advanced Inorganic Chemistry, 5th Edition, John Wiley & Sons, New York 1988. (35) Hawkinson, A.T., Schmitz, W. R., E.I. du Pont de Nemours & Co., US 2910504, 1959. (36) Indictor, N., J. Org. Chem. 1965, 30, 2074-2075. (37) Hamada, T., Fukuda, T., Imanishi, H. and Katsuki, T., Tetrahedron 1996, 52, 515-530. (38) Collman, J. P., Chien, A. S., Eberspacher, T. A. and Brauman, J. I., J. Am. Chem. Soc. 2000, 122, 11098-11100. (39) Katsuki, T. and Sharpless, K. B., J. Am. Chem. Soc. 1980, 102, 5974-5976. (40) Kuznetsov, M. L., Kozlov, Y. N., Mandelli, D., Pombeiro, A. J. L. and Shul?pin, G. B., Inorg. Chem. 2011, 50, 3996-4005. (41) Pescarmona, P. P., Janssen, K. P. F. and Jacobs, P. A., Chem. Eur. J. 2007, 13, 6562-6572. 39 Chapter 3 A Homogeneous Gallium(III) Compound Selectively Catalyzes the Epoxidation of Alkenes* * This chapter is a revision of a published paper: Wenchan Jiang, John D. Gorden, and Christian R. Goldsmith, Inorg. Chem., 2012, 51, 2725-2727. Reprinted with permission. Copyright ? 2012 by the American Chemical Society. 40 3.1 Introduction Group 13 elements have been explored intensively as catalysts for many reactions because of their low cost and ability to be readily extracted and recovered from reaction mixtures.1 The overwhelming majority of their catalytic capability derives from their strong Lewis acidity.2-6 Alkyl and hydrido complexes with these elements are also well-known for their ability to stoichiometrically reduce various functional groups, with the metal complex donating a carbanion or hydride, respectively.2-3, 7-9 Because of the limited redox activity exhibited by these elements, there are comparatively few reports of group 13 catalyzed oxidation?reduction reactions; notable examples include Meerwein?Verley?Ponndorf carbonyl reduction and Oppenauer oxidation.5, 10-19 The hydrocarbon oxidation catalysis that has been reported tends to be unselective; the documented olefin epoxidation, for instance, is accompanied by substantial allylic oxidation.12-17, 20 Reported here is the selective epoxidation of olefins by the previously described [Ga(phen)2Cl2]Cl (phen = 1,10-phenanthroline).14, 17, 20-22 To the best of our knowledge, this is the first instance of a gallium(III) compound serving as a homogeneous catalyst in alkene epoxidation (Scheme 3.1), although this metal ion has previously been a component in heterogeneous catalysts for such reactions.14, 17 Scheme 3.1 Illustration of epoxidation catalyzed by [Ga(phen)2Cl2]Cl 3.2 Experimental Materials Unless stated otherwise, all chemicals were purchased from Sigma-Aldrich and used as received. Anhydrous ethanol and diethyl ether (ether) were bought from Pharmco-Aaper and 41 Fisher Scientific, respectively. Deuterated acetonitrile (CD3CN) was purchased from Cambridge Isotopes. [Ga(phen)2C12]Cl was prepared and characterized as previously described.20-22 The Ga(III) complex crystallized upon adding ether to a saturated solution of the compound in acetonitrile (MeCN). The identity and purity of the Ga(III) catalyst were confirmed by nuclear magnetic resonance (NMR). Preparation of Custom-Made Peracetic Acid (PAAR) The more basic grade of peracetic acid (PAAR) was prepared through a slightly modified version of a reported procedure.4 17 g of 50% H2O2 (0.25 mol) was slowly added to glacial acetic acid (150 g, 2.5 mol), which was stirred in a polyethylene bottle at room temperature. After 5 min, 5.0 g of Amberlite IR-120 was added. The resultant mixture was allowed to stir behind a blast shield for 24 h. After this duration, the solution was filtered, and the molar concentration of PAA was determined by its integrated 13C NMR resonance relative to that of acetic acid. The solution was stored in a standard -20 oC freezer when not in use. The content of PAA was determined to be 7.1% (molar) by 13C NMR analysis; this molar concentration was replicated in multiple batches. CAUTION: Peracids and mixtures of peroxides and organic solvents are potentially explosive and should be handled with care. The dangers can be minimized by using minimal amounts of these materials, using proper protective equipment including blast shields, and working at lower temperatures. Instrumentation Proton and carbon-13 nuclear magnetic resonance (1H NMR, 13C NMR) spectra were acquired on a 400 MHz AV Bruker NMR spectrometer at 294 K. All resonances were referenced to internal standards. Gas chromatography (GC) was obtained on a ThermoScientific Trace GC Ultra spectrometer with a flame ionization detector (FID). Mass spectrometry was performed on an Ultra Performance LC System (ACQUITY, Waters Corporation, Milford, MA) coupled to a quadrupole time-of-flight mass spectrometer (QS2 TOF Premier, Waters Corporation, Milford, MA) via direct probe analysis operated in the positive ion mode. All samples were run as MeCN solutions. Inductively coupled plasma optical emission spectrometry (ICP-OES) was performed on a 42 SPECTRO CIROS ICP-OES spectrometer in Auburn University?s Department of Agronomy and Soils. All samples were prepared and run as solutions in 10% (by volume) aqueous HCl. X-Ray Crystallography The crystal was mounted in paratone oil on a glass fiber and aligned on a Bruker SMART APEX CCD X-ray diffractometer. Intensity measurements were performed using graphite monochromated Mo K? radiation (? = 0.71073 ?) from a sealed tube and monocapillary collimator. SMART (v 5.624) was used to determine the preliminary cell constants and control the data acquisition. The intensities of reflections of a sphere were collected through the compilation of three sets of exposures (frames). Each set had a different ? angle for the crystal, with each exposure spanning 0.3o in ?. A total of 1800 frames were collected with exposure times of 40 s per frame. After the data were corrected for Lorentz and polarization effects, the structure was solved using direct methods and expanded using Fourier techniques. All non-hydrogen atoms were refined anisotropically. Hydrogen atoms were included at idealized positions 0.95 ? from their parent atoms prior to the final refinement. Details regarding the data acquisition and analysis are included in Table 3.1. 43 Table 3.1. Selected crystallographic data for [Ga(phen)2Cl2]Cl. Parameter [Ga(phen)2Cl2]Cl.0.5MeCN.H2O Formula C50H39Cl6Ga2N9O2 Mw 1150.04 Crystal System Orthorhombic Space group Pbcn (#60) a (?) 12.7377 (11) b (?) 23.275 (2) c (?) 16.5439 (15) ? (deg) 90 ? (deg) 90 ? (deg) 90 V (?3) 4904.8 (8) Z 4 Crystal color Yellow T (K) 183 Reflns collected 47811 Unique reflns 4661 R1 (F, I > 2?(I)) 0.0476 wR2 (F2, all data) 0.121 R1 = ???Fo?- ?Fc??/??Fo?; wR2 = [?w(Fo2-Fc2)2/?w(Fo2)2]1/2 Reactivity studies All reactions were run in acetonitrile (MeCN) at 0 oC under N2. The reaction mixtures were stirred for 1 h, at which point excess ether was added to precipitate the gallium compounds. With all substrates, the quenched reaction mixtures were sequentially filtered through plugs of neutral alumina, basic alumina, and silica gel to remove the remaining PAA and metal salts. The workup did not remove organic starting materials or products from the reaction mixture, as assessed by analyzing control mixtures of alkenes and epoxides that were subjected to the 44 same protocol. All products were identified by their GC retention times and NMR. The yields of the products were quantified relative to an internal standard of 1,2-dichlorobenzene, which was found to be unreactive under these conditions. All reactions were run at least four times to ensure reproducibility. The results were also found to be reproducible with different batches of both [Ga(phen)2Cl2]Cl and PAAR. Isolated Yields Isolated yields were obtained for cyclohexene, cyclooctene, and 1-octene. Cyclohexene oxide. [Ga(phen)2C12]Cl (0.054 g, 0.10 mmol) and cyclohexene (0.842 g, 10.0 mmol) were dissolved in 2 mL of MeCN. After the solution was put under N2 and cooled to 0 oC, PAAR solution (11.4 g, 20.0 mmol) was added dropwise. After 1 h, the reaction solution was filtered through plugs of silica gel and K2CO3. Cyclohexene oxide was isolated from the filtrate through distillation, yielding 0.38 g of product (39% vs. 46% GC yield). Cyclooctene oxide. [Ga(phen)2C12]Cl (0.0540 g, 0.10 mmol) and cyclooctene (1.18 g, 10.0 mmol) were dissolved in 2 mL of MeCN. The solution was put under N2 through nitrogen bubbling and cooled to 0 oC. Subsequently, PAAR solution (11.4 g, 20.0 mmol) was added slowly. After 1 h, the reaction mixture was filtered through silica gel and K2CO3. Cyclooctene oxide was distilled from the filtrate (0.97 g, 77% vs. 87% GC yield). 1-Octene oxide. [Ga(phen)2C12]Cl (0.054 g, 0.10 mmol) and 1-octene (1.14 g, 10.0 mmol) were dissolved in 2 mL of MeCN. The reaction mixture was put under N2 and cooled to 0 oC. Once the temperature equilibrated, 11.4 g of PAAR solution (20.0 mmol) was added dropwise. 1 h after the addition was completed, the reaction mixture was filtered through plugs of silica gel and K2CO3. 1-Octene oxide was distilled from the filtrate (0.47 g, 37% vs. 41% GC yield). 3.3 Results and Discussion The [Ga(phen)2Cl2]Cl complex was crystallized from ethanol (Figure 3.1). The structural data confirm the compositions of the inner and outer spheres, with the outer-sphere chloride 45 located 5.57 ? away from the GaIII ion. The halides are terminal rather than bridging, analogous to what is observed in the structure of [Ga(phen)2Br2]Br.23 The structure is a rare example of a hexacoordinate dihalide gallium(III) complex.23-25 The chlorides are cis to each other, as has been observed in other hexacoordinate dihalide gallium(III) complexes with bidentate ligands.23-25 Figure 3.1 ORTEP representation of the cation [Ga(phen)2Cl2]+. The counterion, H atoms, and outer-sphere solvent molecules have been removed for clarity. All thermal ellipsoids are drawn at 50% probability. A fuller description of the structure, including a table of metrical parameters, is included in Table 3.2. 46 Table 3.2 Bond lengths (?) and bond angles (?) for [Ga(phen)2Cl2]+. Bond Length Angle Ga-N (1) 2.085 (2) N(1)-Ga-Cl(1) 94.19 (6) Ga-N (12) 2.127 (2) N(1)-Ga-Cl(2) 95.09 (7) Ga-N (15) 2.134 (2) N(12)-Ga-Cl(1) 88.16 (6) Ga-N (26) 2.084 (2) N(12)-Ga-Cl(2) 171.58 (6) Ga-Cl (1) 2.2849 (7) N(15)-Ga-N(26) 78.57 (9) Ga-Cl (2) 2.2642 (8) N(15)-Ga-Cl(1) 169.26 (6) Angle N(15)-Ga-Cl(2) 90.47 (6) N(1)-Ga-N(12) 78.73 (8) N(26)-Ga-Cl(1) 94.32 (7) N(1)-Ga-N(15) 91.51 (8) N(26)-Ga-Cl(2) 93.61 (7) N(1)-Ga-N(26) 166.86 (9) Cl(1)-Ga-Cl(1) 98.07 (3) The Ga(III)-N bond distances and the L-Ga(III)-L bond angles are similar to those reported for [Ga(phen)2Br2]+.23 The major difference between the two structures is that the bonds between the Ga(III) and the nitrogen atoms on the 10-positions of the phenanthroline ligands are slightly longer in the chloride structure, averaging 2.13 ? as opposed to the 2.10 ? average for the bromide structure. The ability of [Ga(phen)2Cl2]Cl to catalyze the epoxidation of alkenes in acetonitrile (MeCN) was initially tested using H2O2. Although H2O2 served as a competent terminal oxidant in aluminum(III)-catalyzed alkene epoxidation,12, 15 it did not lead to any epoxidation with the gallium(III) catalyst. Speculating that the poor binding affinity of H2O2 to GaIII was responsible for the lack of activity,26 we investigated two grades of peracetic acid (PAA). In addition to commercially available PAA (PAAC, pH ? 1), we prepared and investigated a less acidic, custom-made grade (PAAR, pH ? 4)27 that previously led to excellent activity in manganese-catalyzed epoxidation.28-29 PAAR was prepared by reacting acetic acid and H2O2 in the presence of a highly acidic resin (Amberlite IR 120); PAAR lacks the ?1% H2SO4 additive that is used as the catalyst for the commercial preparation of PAA. The peracids should ligate metal ions more readily than H2O2, particularly under more basic conditions. Using PAAC as the terminal oxidant led to modest epoxidation at 0 ?C, but PAAR was 47 much more effective, particularly with electron-deficient substrates (Table 3.3). As is commonly observed, the more electron-deficient olefins tend to react less readily. The two substrates with two alkene groups, ethyl sorbate and 4-vinylcyclohexene, are oxidized exclusively at the more electron-rich site. These results parallel those found for the aforementioned manganese(II)-catalyzed alkene epoxidation.28-29 Although the yields of the epoxide do increase when more PAAR is added, the reactions become noticeably less efficient. Adding one additional equiv of PAAR to the cyclooctene reaction, for instance, only raises the yield of the epoxide from 64% to 87%, with a concomitant drop in the oxidative efficiency from 64% to 44%. 48 Table 3.3 Yields/Turnover Numbers of Gallium(III) Catalyzed Alkene Epoxidation Reactions by Various Grades and Amounts of PAAa a Sandard reaction conditions: MeCN, 0 ?C, initial concentrations of 5.0 mM [Ga(phen)2Cl2]Cl and 500 mM alkene. The yields, defined as the percentage of alkene converted to the epoxide, were measured at 1 h via gas chromatography. The yields also serve as the turnover numbers because the substrate is present in a 100-fold excess relative to the gallium(III). b The ratio of [Ga(phen)2Cl2]Cl/alkene/PAA was 1:100:200 when 2 equiv of PAA was used and 1:100:100 when 1 equiv of terminal oxidant was added. c Isolated yields for cyclohexene oxide (39%), cyclooctene oxide (77%), and 1-octene oxide (37%). The control studies confirm that both the phen ligands and the GaIII ions are essential for the observed catalysis. GaCl3 by itself does not catalyze epoxidation to an appreciably greater extent than the metal-free controls. The summary of control experiments is in Table 3.4. We hypothesize that the electron-withdrawing nature of the phen ligands30 amplifies the Lewis acidity of the GaIII metal center relative to that in the chloride salt. 49 Table 3.4 Yields of Epoxide Products (%) in Control Reactions. The conditions are identical to those described for the reactions in Table 3.3, with the exceptions of the identity and concentration of the catalyst. For the above reactions, the ratio of alkene/PAA was 1:2 when 2 equiv. of PAA were used and 1:1 when 1 equiv. of terminal oxidant was added, with the concentration of alkene set at 500 mM. All reactions were run in MeCN at 0 ?C under N2. The yields, defined here as the percentage of alkene converted to the epoxide, were measured at 1 h through a comparison of the GC integration of the epoxide peak to that of an internal standard, 1,2-dichlorobenzene. No other organic products were observed in the above reactions. 50 In order to exclude the mechanistic possibility of a highly active transition-metal contaminant, we assessed the catalytic activity of 0.1 mM [Mn(phen)2Cl2] and found it to be inferior to that of 5.0 mM [Ga(phen)2Cl2]+. The loading of the manganese(II) complex, which is known to catalyze olefin epoxidation,29 is meant to mimic an active 2% molar impurity. The 2% value is a conservative upper limit, given the high purities of the available gallium(III) salts and the crystallinity of the [Ga(phen)2Cl2]+ catalyst (Figure 3.1). The purity of the gallium(III) complex is further confirmed by inductively coupled plasma optical emission spectrometry (ICP-OES), which found that the iron and manganese contents were equal within error to those of a blank sample. The results of ICP-OES are summarized in Table 3.5. That the reactivity cannot be fully accounted for by a highly active transition-metal contaminant demonstrates that a gallium(III) species is indeed catalyzing olefin epoxidation. 51 Table 3.5 Summary of ICP-OES Analysis a The ? 2?(I)) 0.0302 0.0327 wR2 (F2, all data) 0.0828 0.0728 R1 = ???Fo?- ?Fc??/??Fo?; wR2 = [?w(Fo2-Fc2)2/?w(Fo2)2]1/2. 72 Crystals were also grown from saturated solutions of 2 in mixtures of MeOH and MeCN (Table 4.1). The structural data, however, correspond to the formula [Ga(trispicen)Cl](GaCl4)(Cl) (7, Figure 4.2), as opposed to [Ga(trispicen)Cl2]Cl. Given that only the latter composition is consistent with the NMR, IR, and elemental analysis data for freshly isolated solid samples of 2, we believe that 2 converts to 7 over the prolonged period of time required for crystal growth. In the structure of 7, the bound trispicen is ?-5 rather than ?-4. Two different counteranions are in the outer sphere: [GaCl4]- and Cl-. The presence of the tetrachlorogallate(III) anion indicates that some of the Ga(III) in the sample had dissociated from the trispicen ligand over the time required for crystallization. Crystals that grew from saturated solutions of 3 were likewise consistent with decomposition products rather than the formula anticipated from the NMR and elemental analysis of the freshly precipitated product. In some cases, the Ga(III) centers in the crystals were coordinated by bispicen, with unit cell parameters that were identical to those of the aforementioned crystals of 1. In other cases, Ga(III) complexes with trispicen were isolated from the solutions. Both degradation products contain ligands that are missing picolyl arms. Water molecules are present in these crystal structures, which may suggest that these portions are being lost through hydrolytic processes. Attempts to grow the crystals in the strict absence of air and moisture were unsuccessful. Short-Term Reactivity The six Ga(III) compounds were tested for their ability to catalyze the epoxidation of alkenes by PAAR (Table 4.2). A 1% loading of catalyst and two equiv of PAAR per equiv of substrate were used, primarily to facilitate comparison to our previous results with [Ga(phen)2Cl2]Cl.20 For each catalyzed oxidation reaction, the epoxide is the sole organic product observed at 1 h; no allylic oxidation or dihydroxylation is observed with the 1% mol catalytic loading. In all cases, the percentage of the alkene substrate that has been consumed is equal within error to the yield of the epoxide product, suggesting that the GC analysis has accounted for all organic products. As anticipated, the more electron-deficient terminal alkenes react less readily, as evidenced by the lower yields of epoxides. 73 Table 4.2. Alkene Conversions/Yields of Epoxides (%) of Reactions Catalyzed by 1-6a Substrate Product [Ga(phen)2Cl2]Clb 1 2 3 4 5 6 46 (?5) 46 (?8) 31 (?4) 32 (?4) 27 (?2) 25 (?2) 14 (?1) 15 (?1) 56 (?4) 56 (?7) 43 (?2) 40 (?3) 41 (?2) 40 (?3) 90 (?3) 87 (?4) 80 (?3) 82 (?2) 66 (?4) 68 (?5) 32 (?2) 31 (?3) 93 (?3) 94 (?2) 87 (?3) 84 (?2) 85 (?5) 83 (?5) 10 (?1) 11 (?2) 7 (?0.4) 7 (?0.6) 3 (?0.2) 3 (?0.2) 0 0 13 (?1) 13 (?1) 10 (?1) 9 (?1) 10 (?1) 9 (?1) 38 (?4) 41 (?3) 2 (?0.2) 2 (?0.3) 2 (?0.2) 2 (?0.2) 0 0 51 (?3) 47 (?3) 34 (?3) 35 (?3) 2 (?0.2) 2 (?0.2) 59 (?8) 61 (?6) 40 (?4) 37 (?3) 26 (?4) 28 (?4) 17 (?2) 17 (?2) 70 (?6) 72 (?4) 52 (?5) 50 (?5) 38 (?2) 39 (?3) aStandard reaction conditions: MeCN, 273 K, N 2, [Ga(III)] = 5.0 mM, [alkene]o = 500 mM, [PAAR]o = 1.0 M. The alkene conversion, defined as the percentage of alkene that has been consumed by the reaction, was measured at 1 h via GC. The alkene conversions are italicized in the above table. The yield of epoxide, defined as the percentage of alkene converted to the epoxide, was measured at 1 h via GC. Since the substrate is in a 100-fold excess relative to the catalyst, the above yield of epoxides also serve as turnover numbers. bFrom reference 20. Three trends can be observed in the epoxidation activities. First, the Ga(III) complexes with phen derivatives promote alkene epoxidation to much greater extents. In particular, compounds 4 and 5 promote the oxidation of 1-octene to 1-octene oxide to much higher degrees than any of the other four newly reported Ga(III) complexes. The increases in the yields of the other epoxide products are much less pronounced. Second, the use of more electron-deficient ligands tends to result in more epoxidation. Comparison of the activities of [Ga(phen)2Cl2]Cl, 4, and 5 reveals that the installation of a more electron-withdrawing NO2 group on the 5-position of the phen ligands increases the turnover rate; the addition of an electron-donating NH2 group reduces the yield of the epoxide at 1 h. Third, as the maximum denticity of the ligand increases past ?-4, the Ga(III) complexes become less able to quickly catalyze olefin epoxidation. Within the series of complexes with pyridylamine ligands, the catalytic activity associated with the tetradentate ligand bispicen is fastest; conversely, the complex with the potentially hexadentate tpen, 3, is the worst at promoting olefin epoxidation. 74 In the previously studied alkene oxidation catalyzed by [Ga(phen)2Cl2]Cl, the selectivity for the epoxide product is lost upon switching to a lower catalyst loading. With a 0.1% catalyst loading of the phen complex, the most heavily represented products result from allylic C-H oxidation and bishydroxylation.20 A loss of selectivity for the epoxide is also observed for the oxidation of cyclohexene catalyzed by 1, but with a substantially different product distribution. With a 0.1% loading of 1, 3.0 M cyclohexene is oxidized by 6.0 M PAAR to predominantly trans-1,2-cyclohexanediol (49%), with cyclohexene oxide as a minor product (6%) and the remainder of the substrate failing to react (45%). Unexpectedly, 2-cyclohexenol is not observed. Another potential product of allylic C-H oxidation, 2-cyclohexenone, is not formed in the olefin oxidations catalyzed by low concentrations of either 1 or [Ga(phen)2Cl2]Cl.20 Time-Scale of Reactivity The ability of the Ga(III) complexes to catalyze the epoxidation of olefins by peracetic acid was monitored beyond 1 h in order to assess whether a more highly chelating ligand could extend the lifetime of the reactivity. Figure 4.3 shows the yields as a function of time for the epoxidations of five olefins catalyzed by the bispicen complex 1. There is no noticeable initiation period, although this may occur on a much shorter timescale. The data show that the catalytic activity of 1 is maintained past 1 h; the observed increases in the yields beyond this time point are too large to be attributed to the uncatalyzed reactions between the olefins and terminal oxidant.20 The most reactive substrate, cyclooctene, is fully converted to cyclooctene oxide by 2 h. Beyond 5 h, diol products consistent with the subsequent opening of the epoxide are observed. With cyclohexene as the substrate, cyclohexene oxide is the sole organic product until 5 h, forming in 75% yield from the alkene. By 6 h, noticeable amounts of trans-1,2-cyclohexanediol are observed in addition to the epoxide. 75 Figure 4.3. Yields of alkene epoxidations by PAAR catalyzed by the bispicen complex 1 as a function of time. The reaction conditions are identical to those listed for Table 4.2. Mass spectrometry (MS) of the cyclohexene epoxidation at 3 h reveals that the bispicen-Ga(III) adduct is still intact. At both 5 min and 3 h, the most intense feature has a m/z of 369.0817, consistent with [Ga(bispicen-H)(CH3CO2)]+ (predicted m/z = 369.0842). This suggests that the two chlorides in 1 are displaced by acetate shortly after the beginning of the reaction. Under the ionization conditions, one of the amine protons is likely lost, explaining both the reduced mass and the +1 charge of the observed ion. The trispicen and tpen complexes 2 and 3 likewise appear to retain their catalytic activity for longer periods of time than [Ga(phen)2Cl2]Cl, as assessed by their abilities to catalyze the oxidation of cyclohexene by PAAR. Figure 4.4A compares the longer term reactivities of compounds 1-6 and [Ga(phen)2Cl2]Cl, using the oxidation of cyclohexene to cyclohexene oxide as a standard.20 The plot shows that the activities of the complexes with the bidentate ligands have decreased substantially by 3 h. Catalysts 1-3, though slower, appear to retain more of their activity, with the overall catalytic activity of 1 becoming approximately equivalent to those of 5 and [Ga(phen)2Cl2]Cl by 3 h. As anticipated, compounds 2 and 3 are much less active than 1 throughout the 3 h. MS analyses of 2 and 3 at 3 h do not reveal any 76 peaks consistent with the ligand hydrolysis observed in the crystals grown from solutions of 3. Although 3, for instance, eventually decomposes to 1 and 2 (Figure 4.2), we did not observe m/z features for either of these compounds in the 3 h data for reactions using 3 as a catalyst. As with 1, the peaks observed at 3 h are consistent with acetate adducts. The plots of the cyclohexene oxide yields versus time for catalysts 4-6 are more curved than those observed for compounds 1-3, suggesting that the former set is less durable under the reaction conditions. Despite this, catalyst 4 remains the best catalyst of the six at the 3 h timepoint. Losses of activity for 4 and 5 are also observed over time when 1-octene is used as the standard (Figure 4.4B), but with this particular substrate, the durabilities of 1-3 do not come close to compensating for their slower activity. A. B. Figure 4.4. Yields of alkene epoxidation by PAAR catalyzed by the 1-6 and [Ga(phen)2Cl2]Cl as a function of time. Panel A shows the oxidation of cyclohexene; whereas, B shows the oxidation of 1-octene. The reaction conditions are identical to those listed for Table 4.2. The data for [Ga(phen)2Cl2]Cl are from Reference 20. 4.4 Discussion The syntheses of the Ga(III) complexes are generally straightforward, in that no special measures are necessary to form adducts between the N-donor ligands and the metal ion and that pure products can be isolated without chromatography. The obtained yields range from 77 moderate (51%) to high (85%). One complication with the syntheses is that the compounds are hygroscopic, and prolonged exposure to moisture appears to degrade the pyridylamine ligands, as evidenced by the crystallized decomposition products of 3. The complexes with the phen derivatives are likely isolated as mixtures of cis isomers (Scheme 4.4). Three pyridylamine ligands were investigated: bispicen, trispicen, and tpen (Scheme 4.1). The bispicen molecule is a well-established tetradentate ligand.32, 46-48 The tpen ligand typically coordinates metal ions in a hexadentate fashion, although tetra- and pentadentate modes of coordination have been observed.49-53 With respect to previously reported Group 13 chemistry, there exists a heptacoordinate complex of tpen with In(III), with the tpen itself providing six of the donor atoms in the inner sphere.54 The coordination chemistry of trispicen is much less established, but a singly methylated derivative, N?-methyl-N,N,N?-tris(2-pyridylmethyl)-1,2-ethanediamine, is pentadentate in its structurally characterized complexes with Mn(II) and Fe(II).34, 55 NMR analysis of 1, 2, and 3 suggests that all three ligands predominantly coordinate Ga(III) through four N-donors, with one or two pyridine rings failing to ligate the metal in 2 and 3, respectively. This is supported by IR analysis of freshly precipitated powder samples of these compounds. Although the crystal structure of 1 (Figure 4.1) is consistent with its NMR spectra and elemental analysis, crystals grown from solutions of 2 are not consistent with the initially obtained [Ga(trispicen)Cl2]Cl. The crystal structure of the isolated 7 (Figure 4.2) demonstrates that the coordination of trispicen to Ga(III) is not limited to the hypodentate mode supported by the NMR data. In one of the complex ions within the asymmetric unit, [Ga(trispicen)Cl]2+, the ligand is pentacoordinate; whereas, in the other complex ion, [GaCl4]-, the trispicen is entirely absent from the gallium. The [Ga(trispicen)Cl]2+ dication represents only the second example of a structurally characterized Ga(III) ion with a N5X coordination sphere (X = F, Cl, Br, I).56 The capability of the trispicen and tpen ligands to more fully coordinate the Ga(III) may explain the decreased epoxidation activities exhibited by 2 and 3 (Table 4.2, Figure 4.4). Focusing the analysis on the complexes with the three pyridylamine ligands, the bispicen complex 1 promotes the most extensive epoxidation of all substrates except 1-octene, which 78 essentially doesn?t react with either 1, 2, or 3 present as the catalyst. The tpen-containing 3 is the worst catalyst of the three compounds, with 1 h yields that are less than half of those measured for 1. Based on the results, we speculate that the pendant pyridine rings in 2 and 3 may compete with the terminal oxidant for the coordination sites vacated by the initially bound chloride anions. Previous work suggests that the affinities of pyridine rings to Ga(III) are comparable to those of carboxylate ligands.28 Reducing the number of pendant pyridine donors would alleviate the competition between intramolecular and intermolecular binding and facilitate the coordination of PAA to the Ga(III) center necessary for catalysis. Alternatively, the steric bulk provided by the additional binding arms in 2 and 3 may hinder coordination of the terminal oxidant and/or limit the accessibility of the alkene substrate to the reactive portion of the active oxidant. The results may suggest that two available coordination sites are needed to activate the PAA most efficiently, which is reminiscent of the Sharpless type mechanism proposed for the epoxidation of alkenes by mixtures of [Al(H2O)6]3+ and H2O2.27, 57 In the mechanistic scheme proposed by Shul?pin et al for this Al(III) catalysis, coordination of both oxygen atoms of the H2O2 to the metal center precedes a concerted oxygen atom transfer from the bound oxidant to an uncoordinated alkene.27 If two coordination sites are likewise needed for the activation of PAA by the Ga(III) centers of the current catalysts, 2 would be anticipated to be an inferior catalyst to 1, in accordance with our results (Table 4.2). Scheme 4.5 illustrates a possible mechanism for the epoxidation reactions that highlights the potential benefit of a second coordination site for the PAA. The illustrated mechanism assumes that the N-donor ligands remain fully bound during the reaction; this assumption has not been experimentally confirmed. We are currently assessing the feasibility of various mechanistic pathways through computational analysis and experiments involving Ga(III) complexes with sterically encumbered bispicen derivatives. 79 Scheme 4.5 Illustration of possible mechanism for the epoxidation reactions The higher catalytic activities of the complexes with the phen derivatives suggest that the Ga(III) centers enable catalysis through their abilities to act as Lewis acids. That more electron-rich alkenes tend to react more readily would suggest an electrophilic oxidant, which would be stabilized and rendered less active by more electron-rich ligands. The electrophilicity of the oxidant in the Ga(III) catalysis is also consistent with the inability of H2O2 and t-BuOOH to serve as terminal oxidants for the epoxidation reactions. Carbonyl groups are widely considered to be electron-withdrawing functional groups. When PAA is used as the terminal oxidant, its carbonyl group may reduce the electron density on the transferred oxygen atom sufficiently to allow oxygen atom transfer to the alkene substrate. The acetate produced upon oxygen atom transfer would also be a better leaving group than the hydroxide or t-butyloxide resulting from H2O2 or t-BuOOH. 1,10-Phenanthroline is widely viewed as being electron-poor relative to amine N-donor ligands and should render chelated metal ions more electron-deficient. Complexes 4 and 5 are notable for their ability to catalyze the oxidation of 1-octene; the other four Ga(III) compounds reported here, conversely, cannot promote this reaction to a significant degree. The addition of electron-withdrawing nitro groups onto the phen ligands improves the yields of the epoxidation reactions; conversely, the installation of electron-donating amino groups decreases the yields (Table 4.1). Although the complexes with the phen ligands are more active over shorter durations, the compounds with the pyridylamine ligands can potentially achieve superior turnover numbers over longer periods of time due to the greater stability provided by chelate effects. One noted shortcoming with the [Ga(phen)2Cl2]Cl catalyst is that it decomposed to unreactive Ga(III) 80 salts and free phen within approximately the first hour of the epoxidation reactions.20 The phen appears to slowly dissociate from the Ga(III) center under acidic conditions. The stability of the Ga(III) catalyst can be increased by using a single tetradentate ligand in the place of two bidentate ligands. The ability of bispicen, trispicen, and tpen to sustain the catalysis over longer periods of time confirms that this is a viable strategy for Ga(III)-catalyzed olefin epoxidations (Figures 4.3 and 4.4). Compound 1 is an equivalent catalyst to [Ga(phen)2Cl2]Cl when the yields of cyclohexene epoxidation are measured at 3 h (Figure 4.4). Unlike the phen compound, the Ga(III)-bispicen adduct remains intact at 3 h. The observed decomposition of 2 and 3 (Figure 4.2, Figure A1), however, suggest that the pyridylamine-containing complexes do have finite catalytic lifetimes. Compounds 1-3 appear to retain their intact polydentate ligands through the 3 h necessary for the alkene epoxidation, however, suggesting that the ligand degradation is slow relative to the alkene epoxidation. The reactivity with the 1% catalyst loading is selective for the epoxide product for short durations, with the epoxide accounting for the entirety of the observed organic product. The reactions using 1 as the catalyst do start to produce trans-1,2-cyclohexanediol between 5 and 6 h after the start of the reaction. As with the olefin oxidation catalyzed by [Ga(phen)2Cl2]Cl,20 the selectivity for the epoxide product is lost when the catalyst loading of 1 is reduced to 0.1%. Oddly, the chemistry associated with 1 differs from that promoted by the phen compound in that the lower concentration of 1 does not promote allylic C-H activation. Instead, bishydroxylation is overwhelmingly favored under such circumstances. When cyclohexene oxide is used as a substrate with the standard reaction conditions and a 1% mol loading of gallium (3.0 mM 1, 6.0 M PAAR, 0 ?C, 7 h), no diol is observed, suggesting that cyclohexene may be converted directly to trans-1,2-cyclohexanediol as opposed to a two-step reaction involving epoxidation followed by ring-opening. We speculate that another Ga(III)-based oxidant is responsible and are continuing to investigate this side-reactivity through experimental and computational methods. The mechanism displayed on Scheme 4.5 certainly doesn?t explain all of the observed oxidative activity of the Ga(III) complexes. 81 4.5 Conclusions Six new gallium(III) complexes with N-donor ligands have been prepared and tested as catalysts for the epoxidation of alkenes by peracetic acid. Comparison of the compounds? activities provides three major insights into Ga(III)-catalyzed alkene epoxidation. First, more electron-deficient ligands are found to support more rapid catalytic turnover, with phen derivatives leading to superior initial activity over ligands with amine and pyridine chelating groups. The Ga(III) complexes with phen derivatives are markedly better at activating the electron-deficient substrate 1-octene, and [Ga(NO2-phen)2Cl2]Cl is the best catalyst of the six under all characterized reaction conditions. Second, more highly coordinating ligands, such as the tetradentate bispicen, can prolong the catalytic activity by stabilizing the ligand-metal adduct. Third, it is found that the more highly chelating N-donor ligands, trispicen and tpen, decrease the catalytic activity. The additional binding arms may compete with terminal oxidant for the coordination sites, or they may impede the reactivity through steric effects. 82 Appendix Figure A1. ORTEP representation of [Ga(trispicen)Cl]Cl2?MeCN?3H2O grown from a solution of [Ga(tpen)Cl2]Cl in MeCN. All hydrogen atoms are omitted for clarity. All thermal ellipsoids are drawn at 50% probability. 83 Table A1. Selected crystallographic data for [Ga(trispicen)Cl]Cl2. Parameter [Ga(trispicen)Cl]Cl2?MeCN? Formula C22H32Cl3GaN6O3 MW 602.61 Crystal system Triclinic Space group P1(#2) a (?) 9.1540(3) b (?) 11.9245(4) c (?) 12.8632(4) ? (deg) 80.3060(10) ? (deg) 81.8800(10) ? (deg)? 88.0630(10) V (?3) 1370.11(8) ?? 2 Cryst color Colorless ?????? 198 Reflns 36681 Unique reflns 3770 R1 (F, I > 2?(I)) 0.0364 wR2 (F2, all 0.0943 R1 = ???Fo?- ?Fc??/??Fo?; wR2 = [?w(Fo2-Fc2)2/?w(Fo2)2]1/2. Figure A2. Mass spectrum (ESI) of [Ga(tpen)Cl2]Cl. The m/z features at 563.1000 and 565.0953 are assigned to [Ga(tpen)Cl2]+ (calcd m/z = 563.1008 and 565.0979). The m/z feature at 264.0689 is assigned to [Ga(tpen)Cl]2+ (calcd m/z = 264.0650). 84 Figure A3. 1H NMR spectrum of 1 in CD3OD at 294 K. The data were acquired on a 400 MHz instrument. Solvent peaks are present at 4.87 (water) and 3.31 (methanol) ppm; these have been truncated for clarity. The doublet centered at 4.88 ppm overlaps with the water peak. Figure A4. 13C NMR spectrum of 1 in CD3OD at 294 K. The solvent peak at 49.86 ppm (methanol) has been truncated for clarity. 85 Figure A5. IR spectrum of 1. The sample was prepared as a KBr pellet. The full listing of peak frequencies in cm-1 is as follows: 3416 (s), 3389 (s), 3100 (s), 3040 (s), 2991 (s), 2914 (s), 2850 (s), 2482 (w), 2395 (w), 2270 (w), 2166 (w), 1665 (w), 1612 (s), 1578 (m), 1481 (s), 1440 (s), 1363 (s), 1300 (s) 1262 (m), 1226 (s), 1160 (m), 1099 (s), 1048 (s), 1033 (s), 980 (m), 906 (w), 849 (m), 821 (w), 777 (s), 723 (m), 649 (m), 567 (m), 506 (m), 480 (m), 425 (m). Figure A6. 1H NMR spectrum of 2 in CD3OD at 294 K. The data were acquired on a 400 MHz instrument. The solvent peaks present at 4.87 (water) and 3.31 (methanol) ppm have been truncated for clarity. 86 Figure A7. 13C NMR spectrum of 2 in CD3OD at 294 K. The solvent peak at 49.86 ppm (methanol) has been truncated for clarity. Figure A8. IR spectrum of a powder sample of 2. The sample was prepared as a KBr pellet. The full listing of peak frequencies in cm-1 is as follows: 3487 (s), 3387 (s), 3334 (s), 3115 (s), 2850 (s), 2657 (s), 2476 (m), 2392 (m), 2271 (m), 2157 (m), 2034 (w), 1936 (w), 1882 (w), 1815 (w), 1747 (w), 1611 (s), 1478 (s), 1438 (s), 1363 (s), 1298 (s), 1230 (s), 1159 (s), 1107 (s), 1040 (s), 1031 (s), 977 (s), 910 (w), 849 (m), 777 (s), 720 (m), 646 (m), 565 (m), 492 (m), 424 (m). Note the shoulder on the feature at 1611 cm-1. 87 Figure A9. IR spectrum of a crystalline sample of 7, which was grown from a saturated solution of 2. The sample was prepared as a KBr pellet.Note that the peak at 1613 cm-1 lacks the shoulder found for the powder sample of 2, which would be consistent with the lack of non-coordinated pyridine rings in the solid. Figure A10.1H NMR spectrum of 3 in CDCl3 at 294 K. The data were acquired on a 250 MHz instrument. The solvent peak present at 7.26 (chloroform) has been truncated for clarity. 88 Figure A11.1H NMR spectrum of 3 in CD3CN at 294 K. The data were acquired on a 400 MHz instrument. The solvent peak present at 1.94 (acetonitrile) has been truncated for clarity. Figure A12. 13C NMR spectrum of 3 in CDCl3 at 294 K. The solvent peak at 77.16 (chloroform) has been truncated for clarity. 89 Figure A13. IR spectrum of a powder sample of 3. The sample was prepared as a KBr pellet. The full listing of peak frequencies in cm-1 is as follows: 3045 (m), 3008 (s), 2945 (s), 2907 (s), 2863 (s), 2815 (s), 2800 (s), 2764 (s), 2695 (m), 2550 (w), 2300 (w), 1905 (w), 1859 (w), 1796 (w), 1662 (w), 1588 (s), 1572 (s), 1472 (s), 1433 (s), 1364 (s), 1345 (s), 1298 (s), 1245 (s), 1209 (m), 1170 (s), 1126 (s), 1085 (m), 1045 (m), 1016 (w), 989 (s), 902 (m), 864 (m), 839 (m), 786 (s), 765 (s), 732 (m), 687 (w), 662 (w), 617 (m), 593 (w), 546 (w), 470 (w).Note the intense feature at 1588 cm-1, which is consistent with the presence of two non-coordinated pyridine rings in the solid. Figure A14. 1H NMR spectrum of 4 in CD3OD at 294 K. The data were acquired on a 600 MHz instrument. Only solvent resonances were observed outside the shown 7.0-10.5 ppm range. 90 Figure A15. 13C NMR spectrum of 4 in CD3OD at 294 K. The feature at 49.00 ppm corresponds to solvent and has been truncated for clarity. Figure A16. IR spectrum of a powder sample of 4. The sample was prepared as a KBr pellet. The full listing of peak frequencies in cm-1 is as follows: 3422 (m), 3113 (w), 3071 (m), 3004 (w), 2970 (w), 2934 (w), 2871 (w), 2337 (w), 1970 (w), 1946 (w), 1841 (w), 1627 (m), 1586 (m), 1540 (s), 1522 (s), 1507 (s), 1490 (m), 1453 (m), 1421 (s), 1388 (m), 1349 (s), 1335 (s), 1261 (w), 1207 (m), 1181 (m), 1153 (m), 1117 (m), 1104 (m), 1038 (w), 992 (w), 974 (w), 920 (m), 838 (s), 823 (s), 810 (m), 751 (m), 734 (s), 721 (s), 652 (m), 619 (w), 541 (w), 507 (w), 429 (w). 91 Figure A17. 1H NMR spectrum of 5 in DMSO-d6 at 294 K. The data were acquired on a 250 MHz instrument. The solvent contributes a resonance feature at 2.50 ppm. Figure A18. 1H NMR spectrum of 5 in CD3OD at 294 K. The data were acquired on a 600 MHz instrument. Only solvent resonances were observed outside the shown 7.0-10.5 ppm range. 92 Figure A19. 13C NMR spectrum of 5 in DMSO-d6 at 294 K. The feature at 39.51 ppm corresponds to solvent and has been truncated for clarity. Figure A20. IR spectrum of a powder sample of 5. The sample was prepared as a KBr pellet. The full listing of peak frequencies in cm-1 is as follows: 3399 (s), 3339 (s), 3208 (s), 3080 (m), 1640 (s), 1620 (s), 1604 (s), 1586 (m), 1522 (w), 1494 (s), 1465 (m), 1435 (s), 1350 (w), 1323 (w), 1287 (w), 1226 (w), 1164 (w), 1146 (w), 1122 (w), 1087 (w), 913 (w), 851 (w), 826 (w), 802 (w), 725 (s), 665 (w), 653 (w), 427 (w). 93 Figure A21. 1H NMR spectrum of 6 in DMSO-d6 at 294 K. The data were acquired on a 400 MHz instrument. A solvent peak is present at 2.50 ppm. Figure A22. 13C NMR spectrum of 6 in DMSO-d6 at 294 K. The solvent peak at 39.51 ppm has been truncated for clarity. 94 Figure A23. IR spectrum of a powder sample of 6. The sample was prepared as a KBr pellet. The full listing of peak frequencies in cm-1 is as follows: 3420 (s), 3225 (s), 2978 (s), 2912 (s), 2804 (s), 2747 (s), 2711 (s), 2677 (s), 2634 (s), 2575 (m), 2521 (m), 2420 (m), 2280 (w), 2055 (m), 1624 (m), 1602 (m), 1504 (s), 1342 (w), 1182 (w), 1085 (m), 1034 (s), 1008 (w), 973 (w), 917 (w), 786 (w), 568 (m), 467 (m), 445 (w), 431 (w). Figure A24.The yields of the organic products of [Ga(bispicen)Cl2]Cl-catalyzed cyclohexene oxidation by PAAR over time. The reaction was run at 0 ?C in MeCN with initial concentrations of 5.0 mM Ga(III) catalyst, 500 mM cyclohexene, and 1000 mM PAAR.The points on the plot are the averages of three separate reactions, with errors of ?2% for the yields of cyclohexene oxide and ?1% for the yields of 1,2-cyclohexanediol. The diol is not observed until the 6 h time point. 95 Figure A25. Mass spectrum (ESI) of the gallium species present in solution 3 h after the beginning of the cyclohexene epoxidation reaction catalyzed by 1. 500 mM cyclohexene, 1000 mM peracetic acid, and 5.0 mM [Ga(bispicen)Cl2]Cl were mixed in MeCN and allowed to stir at 0 ?C for 3 h. The m/z features at 369.0817 and 371.0823 are assigned to [Ga(bispicen-H)(CH3CO2)]+ (calcd m/z = 369.0842 and 371.0833). 96 Figure A26. Mass spectrum (ESI) of the gallium species present in solution 3 h after the beginning of the cyclohexene epoxidation reaction catalyzed by 2. 500 mM cyclohexene, 1000 mM peracetic acid, and 5.0 mM [Ga(trispicen)Cl2]Cl were mixed in MeCN and allowed to stir at 0 ?C for 3 h. The m/z features at 460.1083 and 462.1095 are assigned to [Ga(trispicen-H)(CH3CO2)]+ (calcd m/z = 460.1264 and 462.1255). The peaks at 281.5377 and 219.5382 are assigned to [Ga(trispicen)Cl]2+ (calcd m/z = 218.5445 and 219.5445). The peaks at 200.5546 and 201.5544 are assigned to [Ga(trispicen-H)]2+ (calcd m/z = 200.5566 and 201.5561). 97 Figure A27. Mass spectrum (ESI) of the gallium species present in solution 3 h after the beginning of the cyclohexene epoxidation reaction catalyzed by 3. 500 mM cyclohexene, 1000 mM peracetic acid, and 5.0 mM [Ga(tpen)Cl2]Cl were mixed in MeCN and allowed to stir at 0 ?C for 3 h. The m/z features at 551.1714 and 553.1232 are assigned to [Ga(tpen-H)(CH3CO2)]+ (calcd m/z = 551.1686 and 553.1677). 98 References (1) Chi, Y.; Chou, T.-Y.; Wang, Y.-J.; Huang, S.-F.; Carty, A. J.; Scoles, L.; Udachin, K. A.; Peng, S.-M.; Lee, G.-H. Organometallics 2003, 23, 95-103. (2) Park, J.-H.; Afzaal, M.; Helliwell, M.; Malik, M. A.; O'Brien, P.; Raftery, J. Chem. Mater. 2003, 15, 4205-4210. (3) Valet, M.; Hoffman, D. M. Chem. Mater. 2001, 13, 2135-2143. (4) Suh, S.; Hoffman, D. M. Chem. Mater. 2000, 12, 2794-2797. (5) MacInnes, A. N.; Power, M. B.; Barron, A. R. Chem. Mater. 1993, 5, 1344-1351. (6) Kouvetakis, J.; Beach, D. B. Chem. Mater. 1989, 1, 476-478. (7) Cowley, A. H. J. Organomet. Chem. 2000, 600, 168-173. (8) Bartholom?, M. D. Inorg. Chim. Acta 2012, 389, 36-51. (9) Wadas, T. J.; Wong, E. H.; Weisman, G. R.; Anderson, C. J. Chem. Rev. 2010, 110, 2858-2902. (10) Bartholom?, M. D.; Louie, A. S.; Valliant, J. F.; Zubieta, J. Chem. Rev. 2010, 110, 2903-2920. (11) Boros, E.; Ferreira, C. L.; Cawthray, J. F.; Price, E. W.; Patrick, B. O.; Wester, D. W.; Adam, M. J.; Orvig, C. J. Am. Chem. Soc. 2010, 132, 15726-15733. (12) Dagorne, S.; Atwood, D. A. Chem. Rev. 2008, 108, 4037-4071. (13) Yamaguchi, M.; Nishimura, Y. Chem. Commun. 2008, 35-48. (14) Wehmschulte, R. J.; Steele, J. M.; Young, J. D.; Khan, M. A. J. Am. Chem. Soc. 2003, 125, 1470-1471. (15) Lichtenberg, C.; Spaniol, T. P.; Okuda, J. Inorg. Chem. 2012, 51, 2254-2262. (16) Tang, S.; Monot, J.; El-Hellani, A.; Michelet, B.; Guillot, R.; Bour, C.; Gandon, V. Chem. Eur. J. 2012, 18, 10239-10243. (17) Carty, A. J.; Dymock, K. R.; Boorman, P. M. Can. J. Chem. 1970, 48, 3524-3529. (18) Ivanov-Emin, B. N.; Nisel'son, L. A.; Rabovik, Y. I.; Larionova, L. E. Russ. J. Inorg. Chem. 1961, 6, 1142-1146. 99 (19) McPhail, A. T.; Miller, R. W.; Pitt, C. G.; Gupta, G.; Srivastava, S. C. J. Chem. Soc., Dalton Trans. 1976, 1657-1661. (20) Jiang, W.; Gorden, J. D.; Goldsmith, C. R. Inorg. Chem. 2012, 51, 2725-2727. (21) Fricke, R.; Kosslick, H.; Lischke, G.; Richter, M. Chem. Rev. 2000, 100, 2303-2405. (22) Pescarmona, P. P.; Janssen, K. P. F.; Jacobs, P. A. Chem. Eur. J. 2007, 13, 6562-6572. (23) Stoica, G.; Santiago, M.; Jacobs, P. A.; P?rez-Ram?rez, J.; Pescarmona, P. P. Appl. Catal. A 2009, 371, 43-53. (24) Cates, C. D.; Myers, T. W.; Berben, L. A. Inorg. Chem. 2012, 51, 11891-11897. (25) Nam, W.; Valentine, J. S. J. Am. Chem. Soc. 1990, 112, 4977-4979. (26) Rinaldi, R.; de Oliveira, H. F. N.; Schumann, H.; Schuchardt, U. J. Mol. Catal. A 2009, 307, 1-8. (27) Kuznetsov, M. L.; Kozlov, Y. N.; Mandelli, D.; Pombeiro, A. J. L.; Shul'pin, G. B. Inorg. Chem. 2011, 50, 3996-4005. (28) Caravan, P.; Rettig, S. J.; Orvig, C. Inorg. Chem. 1997, 36, 1306-1315. (29) Murphy, A.; Pace, A.; Stack, T. D. P. Org. Lett. 2004, 6, 3119-3122. (30) Sheldrick, G. M., 6.12 ed.; Siemens Analytical X-ray Instruments, Inc.: Madison, WI, 2001. (31) Sheldrick, G. M.; Bruker Analytical X-ray Systems: Madison, WI, 1996. (32) Toftlund, H.; Pedersen, E.; Yde-Andersen, S. Acta. Chem. Scand. A 1984, 38, 693-697. (33) Anderegg, G.; Wenk, F. Helv. Chim. Acta 1967, 50, 2330-2332. (34) Mialane, P.; Nivorojkine, A.; Pratviel, G.; Az?ma, L.; Slany, M.; Godde, F.; Simaan, A.; Banse, F.; Kargar-Grisel, T.; Bouchoux, G.; Sainton, J.; Horner, O.; Guilhem, J.; Tchertanova, L.; Meunier, B.; Girerd, J.-J. Inorg. Chem. 1999, 38, 1085-1092. (35) Goldsmith, C. R.; Jonas, R. T.; Cole, A. P.; Stack, T. D. P. Inorg. Chem. 2002, 41, 4642-4652. (36) Junk, P. C.; Skelton, B. W.; White, A. H. Aust. J. Chem. 2006, 59, 147-154. (37) Gill, N. S.; Nuttall, R. H.; Scaife, D. E.; Sharp, D. W. A. J. Inorg. Nucl. Chem. 1961, 18, 79-87. 100 (38) Willett, R. D.; G?mez-Garc?a, C. J.; Twamley, B.; G?mez-Coca, S.; Ruiz, E. Inorg. Chem. 2012, 51, 5487-5493. (39) Goodson, P. A.; Glerup, J.; Hodgson, D. J.; Michelsen, K.; Pedersen, E. Inorg. Chem. 1990, 29, 503-508. (40) Coates, C. M.; Fiedler, S. R.; McCullough, T. L.; Albrecht-Schmitt, T. E.; Shores, M. P.; Goldsmith, C. R. Inorg. Chem. 2010, 49, 1481-1486. (41) Collins, M. A.; Hodgson, D. J.; Michelsen, K.; Towle, D. K. J. Chem. Soc., Chem. Commun. 1987, 1659-1660. (42) Arulsamy, N.; Hodgson, D. J.; Glerup, J. Inorg. Chim. Acta 1993, 209, 61-69. (43) Ardon, M.; Bino, A.; Michelsen, K.; Pedersen, E.; Thompson, R. C. Inorg. Chem. 1997, 36, 4147-4150. (44) Restivo, R.; Palenik, G. J. J. Chem. Soc., Dalton Trans. 1972, 341-344. (45) Sofetis, A.; Raptopoulou, C. P.; Terzis, A.; Zafiropoulos, T. F. Inorg. Chim. Acta 2006, 359, 3389-3395. (46) Glerup, J.; Goodson, P. A.; Hazell, A.; Hazell, R.; Hodgson, D. J.; McKenzie, C. J.; Michelsen, K.; Rychlewska, U.; Toftlund, H. Inorg. Chem. 1994, 33, 4105-4111. (47) Arulsamy, N.; Goodson, P. A.; Hodgson, D. J.; Glerup, J.; Michelsen, K. Inorg. Chim. Acta 1994, 216, 21-29. (48) Heinrichs, M. A.; Hodgson, D. J.; Michelsen, K.; Pedersen, E. Inorg. Chem. 1984, 23, 3174-3180. (49) Hureau, C.; Blanchard, S.; Nierlich, M.; Blain, G.; Rivi?re, E.; Girerd, J.-J.; Anxolab?h?re-Mallart, E.; Blondin, G. Inorg. Chem. 2004, 43, 4415-4426. (50) Musso, S.; Anderegg, G.; Ruegger, H.; Schl?pfer, C. W.; Gramlich, V. Inorg. Chem. 1995, 34, 3329-3338. (51) Chang, H. R.; McCusker, J. K.; Toftlund, H.; Wilson, S. R.; Trautwein, A. X.; Winkler, H.; Hendrickson, D. N. J. Am. Chem. Soc. 1990, 112, 6814-6827. (52) Yoon, D. C.; Lee, U.; Oh, C. E. Bull. Korean Chem. Soc. 2004, 25, 796-800. (53) Dueland, L.; Hazell, R.; McKenzie, C. J.; Nielsen, L. P.; Toftlund, H. J. Chem. Soc., Dalton Trans. 2001, 152-156. 101 (54) Harrington, J. M.; Oscarson, K. A.; Jones, S. B.; Reibenspies, J. H.; Bartolotti, L. J.; Hancock, R. D. Z. Naturforsch. 2007, 62b, 386-396. (55) Hureau, C.; Groni, S.; Guillot, R.; Blondin, G.; Duboc, C.; Anxolab?h?re-Mallart, E. Inorg. Chem. 2008, 47, 9238-9247. (56) Radish, K. M.; Cornillon, J. L.; Korp, J. D.; Guilard, R. J. Heterocycl. Chem. 1989, 26, 1101-1104. (57) Sharpless, K. B.; Townsend, J. M.; Williams, D. R. J. Am. Chem. Soc. 1972, 94, 295-296.