299 results on '"Michael G. Richmond"'
Search Results
2. Reactions of diphosphine-stabilized Os3 clusters with triphenylantimony: syntheses and structures of new antimony-containing Os3 clusters via Sb–Ph bond cleavage
- Author
-
Fahmida Islam, Md. Sohag Hasan, Shishir Ghosh, Michael G. Richmond, Shariff E. Kabir, and Herbert W. Roesky
- Subjects
General Chemical Engineering ,General Chemistry - Abstract
The reactivity of the clusters [Os3(CO)10(μ-dppm)] and [HOs3(CO)8{μ3-Ph2PCH2PPh(C6H4-μ2,σ1)}] with SbPh3 has been examined.
- Published
- 2023
3. An investigation of steric influence on the reactivity of FeV(O)(OH) tautomers in stereospecific C–H hydroxylation
- Author
-
Mainak Mitra, Alexander Brinkmeier, Yong Li, Margarida Borrell, Arnau Call, Julio Lloret Fillol, Michael G. Richmond, Miquel Costas, and Ebbe Nordlander
- Subjects
Inorganic Chemistry - Abstract
The steric properties of the tetradentate ligand influences the reactivities of the Fe(v)O units in a number of tautomeric Fe(v)(O)(OH) complexes.
- Published
- 2023
4. Synthesis, Structure, and Reactivity of Triosmium Clusters Bearing a Metalated Dialkyl-substituted Pyrazine Ligand
- Author
-
Md. Mahbub Alam, Fahima Islam, Subas Rajbangshi, Kazi A. Azam, Shishir Ghosh, Vladimir N. Nesterov, Michael G. Richmond, and Shariff E. Kabir
- Subjects
General Chemistry ,Condensed Matter Physics - Published
- 2022
5. New Tri- and Pentanuclear Osmium Clusters from [Os3(CO)10(NCMe)2] and Pyrrolidine: Reactivity and X-ray Structure of [HOs3(CO)9(µ,κ2-N=C(CH3)C=NCH2CH2CH2)] and [H3Os5(CO)14(µ,η1,κ1-N=CCH2CH2CH2)]
- Author
-
Shahin A. Begum, Mst. Asha Aktar, Israt Jahan, Vladimir N. Nesterov, Michael G. Richmond, Shishir Ghosh, and Shariff E. Kabir
- Subjects
General Materials Science ,General Chemistry ,Condensed Matter Physics ,Biochemistry - Published
- 2022
6. Biomimics of [FeFe]-hydrogenases incorporating redox-active ligands: synthesis, redox properties and spectroelectrochemistry of diiron-dithiolate complexes with ferrocenyl-diphosphines as Fe4S4 surrogates
- Author
-
Georgia R. F. Orton, Shishir Ghosh, Lucy Alker, Jagodish C. Sarker, David Pugh, Michael G. Richmond, František Hartl, and Graeme Hogarth
- Subjects
Inorganic Chemistry - Abstract
Ferrocenyl diphosphine bridged diiron dithiolate complexes have been prepared and their redox properties investigated by CV and IR SEC supported by DFT calculations to give insight into their proton reduction and hydrogen oxidation activity.
- Published
- 2022
7. Syntheses of Group 5 Amide Amidinates and Their Reactions with Water: Different Reactivities of Nb(V) and Ta(V) Complexes
- Author
-
Adam T. Hand, Adam C. Lamb, Michael G. Richmond, Xiaoping Wang, Carlos A. Steren, and Zi-Ling Xue
- Subjects
Inorganic Chemistry ,Physical and Theoretical Chemistry - Abstract
Chemistries of Nb(V) and Ta(V) compounds are essentially identical as a result of lanthanide contraction. Hydrolysis of M(NMe
- Published
- 2022
8. Synthesis and characterization of µ,κ1- and µ,κ2-dithiolate di- and triruthenium complexes with a dppf ligand [dppf = 1,1'-bis(diphenylphosphino)ferrocene]
- Author
-
Abdullah Al Mamun, Jakir Hossain, Shishir Ghosh, Michael G. Richmond, and Shariff E. Kabir
- Subjects
Inorganic Chemistry ,Materials Chemistry ,Physical and Theoretical Chemistry - Published
- 2023
9. Asymmetric hydrogenation of an α-unsaturated carboxylic acid catalyzed by intact chiral transition metal carbonyl clusters – diastereomeric control of enantioselectivity
- Author
-
Ebbe Nordlander, Amrendra K. Singh, Isa Doverbratt, Michael G. Richmond, Yusuf Theibich, Arun K. Raha, Matti Haukka, Ahmed F. Abdel-Magied, and Ahibur Rahaman
- Subjects
chemistry.chemical_classification ,karboksyylihapot ,Carboxylic acid ,Asymmetric hydrogenation ,Diastereomer ,Tiglic acid ,asymmetric hydrogenation ,Medicinal chemistry ,Catalysis ,Inorganic Chemistry ,chemistry.chemical_compound ,chemistry ,Ferrocene ,Transition metal ,katalyysi ,carboxylic acid - Abstract
Twenty clusters of the general formula [(μ-H)2Ru3(μ3-S)(CO)7(μ-P–P*)] (P–P* = chiral diphosphine of the ferrocene-based Walphos or Josiphos families) have been synthesised and characterised. The clusters have been tested as catalysts for asymmetric hydrogenation of tiglic acid [trans-2-methyl-2-butenoic acid]. The observed enantioselectivities and conversion rates strongly support catalysis by intact Ru3 clusters. A catalytic mechanism involving an active Ru3 catalyst generated by CO loss from [(μ-H)2Ru3(μ3-S)(CO)7(μ-P–P*)] has been investigated by DFT calculations. peerReviewed
- Published
- 2020
10. Reaction of [Ru3{µ-Ph2PCH(Me)PPh2}(CO)10] with diphenylacetylene resulted in clusters [Ru3(µ3,η2-PhCCPh){µ,η2-Ph2PCH(Me)PPh2}(CO)7,8], [Ru3{κ2-Ph2PCH(Me)- PPh2}{µ3,η4-(CPh)4}(CO)4(µ-CO)2], and [Ru2{µ-Ph2PCH(Me)PPh2}{µ,η4-(CPh)4}(CO)4]
- Author
-
Michael G. Richmond, subas rajbangshi, shishir ghosh, graeme hogarth, volodymyr nesterov, vladimir nesterov, shariff Kabir, and herbert w. roesky
- Published
- 2022
11. Polynuclear ruthenium clusters containing stibine, stibene, and stibinidene ligands
- Author
-
Mihir L. Bhowmik, Md. Abdullah Al Mamun, Shishir Ghosh, Vladimir N. Nesterov, Michael G. Richmond, Shariff E. Kabir, and Herbert W. Roesky
- Subjects
Inorganic Chemistry ,Organic Chemistry ,Materials Chemistry ,Physical and Theoretical Chemistry ,Biochemistry - Published
- 2023
12. Polyhedral Flexibility in the Sulfido-Capped Cluster H2Ru3(CO)9(μ3-S) on Reaction with 2-(Diphenylphosphino)thioanisole (PS) and Reversible Tripodal Rotation of the Chelated PS Ligand in H2Ru3(CO)7(κ2-PS)(μ3-S)
- Author
-
Vladimir N. Nesterov, Darrell D. Mayberry, and Michael G. Richmond
- Subjects
Flexibility (anatomy) ,010405 organic chemistry ,Ligand ,Chemistry ,Organic Chemistry ,Thioanisole ,010402 general chemistry ,Rotation ,01 natural sciences ,0104 chemical sciences ,Inorganic Chemistry ,Crystallography ,medicine.anatomical_structure ,medicine ,Cluster (physics) ,Chelation ,Physical and Theoretical Chemistry - Abstract
Treatment of the tetrahedral cluster H2Ru3(CO)9(μ3-S) (1) with 2-(diphenylphosphino)thioanisole (PS) furnishes the cluster H2Ru3(CO)7(κ2-PS)(μ3-S) (2). Cluster 2, which exhibits a chelated thiophos...
- Published
- 2019
13. Reactivity of [Mo(CO)3(NCMe)3] towards pyrimidine-2-thiol (pymSH) and thiophenol (PhSH) in the presence of phosphine auxiliaries: Synthesis of mono- and dinuclear complexes bearing κ2 and µ,κ2-pymS coordination motifs
- Author
-
Roknuzzaman, Derek A. Tocher, Mohd. Rezaul Haque, S. M. Tareque Abedin, Shishir Ghosh, Shariff E. Kabir, and Michael G. Richmond
- Subjects
chemistry.chemical_classification ,Pyrimidine ,010405 organic chemistry ,Ligand ,Thiophenol ,Nuclear magnetic resonance spectroscopy ,010402 general chemistry ,01 natural sciences ,Medicinal chemistry ,Product distribution ,0104 chemical sciences ,Inorganic Chemistry ,chemistry.chemical_compound ,chemistry ,Materials Chemistry ,Thiol ,Reactivity (chemistry) ,Physical and Theoretical Chemistry ,Phosphine - Abstract
The reaction of [Mo(CO)3(NCMe)3] with added thiol in the presence of a phosphine auxiliary has been investigated. Treatment of [Mo(CO)3(NCMe)3] with pyrimidine-2-thiol (pymSH) and PPh3 at 60 °C in MeCN afforded the known mononuclear compounds cis-[Mo(CO)4(PPh3)2] (1) and [Mo(к2-pymS)4] (2), and the new binuclear compound [Mo2(CO)4(μ,к2-pymS)2(PPh3)2] (3), which possesses idealized C2 symmetry. A different product distribution was found when dppm was employed as the phosphine ligand. Of the five reaction products isolated, three consisted of mononuclear compounds, [Mo(CO)4(к2-dppm)] (4), [Mo(CO)3(κ2-dppm)(κ1-dppm)] (5) and [Mo(CO)(к2-pymS)2(к2-dppm)] (6), with the remaining two products corresponding to dinuclear compounds, [Mo2(CO)6(μ,к1-pymS)2(μ,к2-dppm)] (7) and [Mo2(CO)4(μ,к2-pymS)2(к2-dppm)] (8). Products 6–8 are new and have been fully characterized in solution by IR and NMR spectroscopy, and by X-ray crystallography in the case of 6 and 7. The reaction of [Mo(CO)3(NCMe)3] with PhSH and dppm at 60 °C in MeCN was also examined to assess the effect of thiol on the product distribution. The two principal products isolated were identified as the mononuclear compound [Mo(CO)2(κ1-PhS)2(κ2-dppm)] (9) and the dinuclear compound [Mo2(CO)6(μ,κ1-PhS)2(μ,κ2-dppm)] (10). The bonding in compounds 3, 6 and 7 was also examined by DFT, and highlights between the computational and experimental structures are discussed.
- Published
- 2019
14. Chalcogenide-capped triiron clusters [Fe3(CO)9(μ3-E)2], [Fe3(CO)7(μ3-CO)(μ3-E)(μ-dppm)] and [Fe3(CO)7(μ3-E)2(μ-dppm)] (E = S, Se) as proton-reduction catalysts
- Author
-
Sucharita Basak-Modi, Michael G. Richmond, George C. Lisensky, Shariff E. Kabir, Ebbe Nordlander, Matti Haukka, Shishir Ghosh, Ahibur Rahaman, Ahmed F. Abdel-Magied, and Graeme Hogarth
- Subjects
Sulfide ,Infrared spectroscopy ,Protonation ,organometalliyhdisteet ,Sulfonic acid ,010402 general chemistry ,Electrochemistry ,01 natural sciences ,Biochemistry ,Medicinal chemistry ,Catalysis ,Inorganic Chemistry ,chalcogenide ,chemistry.chemical_compound ,Selenide ,Materials Chemistry ,Physical and Theoretical Chemistry ,cluster ,ta116 ,proton-reduction ,chemistry.chemical_classification ,010405 organic chemistry ,Chalcogenide ,Organic Chemistry ,triiron ,sähkökemia ,0104 chemical sciences ,electrochemistry ,chemistry ,Cluster ,Triiron ,Proton-reduction ,Cyclic voltammetry - Abstract
Chalcogenide-capped triiron clusters [Fe3(CO)7(μ3-CO)(μ3-E)(μ-dppm)] and [Fe3(CO)7(μ3-E)2(μ-dppm)] (E = S, Se) have been examined as proton-reduction catalysts. Protonation studies show that [Fe3(CO)9(μ3-E)2] are unaffected by strong acids. Mono-capped [Fe3(CO)7(μ3-CO)(μ3-E)(μ-dppm)] react with HBF4.Et2O but changes in IR spectra are attributed to BF3 binding to the face-capping carbonyl, while bicapped [Fe3(CO)7(μ3-E)2(μ-dppm)] are protonated but in a process that is not catalytically important. DFT calculations are presented to support these protonation studies. Cyclic voltammetry shows that [Fe3(CO)9(μ3-Se)2] exhibits two reduction waves, and upon addition of strong acids, proton-reduction occurs at a range of potentials. Mono-chalcogenide clusters [Fe3(CO)7(μ3-CO)(μ3-E)(μ-dppm)] (E = S, Se) exhibit proton-reduction at ca.-1.85 (E = S) and -1.62 V (E = Se) in the presence of p-toluene sulfonic acid (p-TsOH). Bicapped [Fe3(CO)7(μ3-E)2(μ-dppm)] undergo quasi-reversible reductions at -1.55 (E = S) and -1.45 V (E = Se) and reduce p-TsOH to hydrogen but protonated species do not appear to be catalytically important. Current uptake is seen at the first reduction potential in each case, showing that [Fe3(CO)7(μ3-E)2(μ-dppm)]- are catalytically active but a far greater response is seen at ca.-1.9 V being tentatively associated with reduction of [H2Fe3(CO)7(μ3-E)2(μ-dppm)]+. In general, selenide clusters are reduced at slightly lower potentials than sulfide analogues and show slightly higher current uptake under comparable conditions. peerReviewed
- Published
- 2019
15. Highly efficient electrocatalytic proton-reduction by coordinatively and electronically unsaturated Fe(CO)(κ2-dppn)(κ2-tdt)
- Author
-
Shishir Ghosh, Graeme Hogarth, Shariff E. Kabir, Michael G. Richmond, and Shahed Rana
- Subjects
Proton ,010405 organic chemistry ,010402 general chemistry ,Electrochemistry ,DFT ,Square-pyramidal ,01 natural sciences ,Medicinal chemistry ,Square pyramidal molecular geometry ,0104 chemical sciences ,Catalysis ,Inorganic Chemistry ,Solvent ,Reduction (complexity) ,Proton reduction ,chemistry.chemical_compound ,Diphosphine ,chemistry ,Materials Chemistry ,Physical and Theoretical Chemistry ,Redox-active ligand ,Single crystal ,Naphthalene - Abstract
Coordinatively and electronically unsaturated square-pyramidal Fe(CO)(κ2-dppn)(κ2-tdt) (2) is shown to be amongst the most efficient proton-reduction catalysts reported to date. It is formed from the reaction of Fe2(CO)6(μ-tdt) (tdt = 3,4-toluenedithiolate) with 1,8-bis(diphenylphosphino)naphthalene (dppn) in presence of Me3NO·2H2O affording Fe2(CO)4(κ2-dppn)(μ-tdt) (1) as the major product, together with smaller but reproducible amounts of 2. Both have been characterized by single crystal X-ray diffraction. The electrochemistry of 2 is solvent dependent but in both CH2Cl2 and a 1:1 mixture of CH2Cl2/MeCN it shows a reversible reduction at E1/2 = –1.54 V and E1/2 = –1.68 V respectively. While 2 degrades in the presence of the strong acid HBF4·2H2O it is catalytically active for proton-reduction using CF3CO2H. Catalysis occurs at the first reduction potential and it displays an impressive icat/ip ratio of 33 after addition of 20 equivalents CF3CO2H. It is amongst the most efficient molecular proton-reduction catalysts reported to date.
- Published
- 2019
16. Activation of thiosaccharin at a polynuclear osmium cluster
- Author
-
Kazi A. Azam, Michael G. Richmond, Md. Jadu Mia, Md. Mahbub Alam, Matiar Rahman, Shariff E. Kabir, Shishir Ghosh, Derek A. Tocher, and Tamanna Pinky
- Subjects
010405 organic chemistry ,Hydride ,Organic Chemistry ,chemistry.chemical_element ,010402 general chemistry ,01 natural sciences ,Biochemistry ,Toluene ,Sulfur ,0104 chemical sciences ,Inorganic Chemistry ,chemistry.chemical_compound ,Crystallography ,chemistry ,Materials Chemistry ,Cluster (physics) ,Moiety ,Osmium ,Physical and Theoretical Chemistry ,Isomerization ,Bond cleavage - Abstract
The reaction of thiosaccharin (tsacH) with the triosmium cluster [Os3(CO)10(NCMe)2] furnishes the decacarbonyl isomers [HOs3(CO)10(μ-S-tsac)] (1) and [HOs3(CO)10(μ-N,S-1,3-tsac)] (2) in a 3:1 ratio at room temperature. These isomers differ by the coordination mode displayed by tsac ligand. The tsac moiety functions as an edge-bridging ligand via the sulfur atom in 1 while in 2 the bridging of adjacent osmium centers is achieved through the sulfur and nitrogen groups. The ancillary hydride in both products shares the Os Os edge that is bridged by the heterocyclic ligand. Heating 1 at 80 °C affords 2 and demonstrates that the former cluster is the product of kinetic control. The conversion of 1 → 2 has been investigated by DFT and the isomerization pathway elucidated. The DFT calculations confirm cluster 2 as the thermodynamically preferred isomer in this pair of products. Thermolysis of 2 in refluxing toluene affords the hexanuclear cluster [H2Os6(CO)17(μ-C,N-1,2-C6H4CNSO2)2(μ3-S)(μ4-S)] (3) via carbon-sulfur bond scission and subsequent capture of the extruded sulfur by the cluster core. The molecular structures for the three new clusters have been determined by single-crystal X-ray diffraction analyses.
- Published
- 2019
17. New molecular architectures containing low-valent cluster centres with di- and trimetalated 2-vinylpyrazine ligands: synthesis and molecular structures of Ru5(CO)15(μ5-C4H2N2CHCH)(μ-H)2 and Ru8(CO)24(μ7-C4H2N2CHC)(μ-H)3
- Author
-
Shishir Ghosh, Graeme Hogarth, Michael G. Richmond, Vladimir N. Nesterov, Nahid Akter, Shariff E. Kabir, and Md. Monir Hossain
- Subjects
Chemistry ,General Chemical Engineering ,Aromaticity ,02 engineering and technology ,General Chemistry ,Electronic structure ,010402 general chemistry ,021001 nanoscience & nanotechnology ,Ring (chemistry) ,01 natural sciences ,0104 chemical sciences ,Crystallography ,Cluster (physics) ,0210 nano-technology - Abstract
Reaction of 2-vinylpyrazine with Ru3(CO)12 results in multiple C–H bond activations to afford penta- and octa-ruthenium clusters, Ru5(CO)15(μ5-C4H2N2CHCH)(μ-H)2 (2) and Ru8(CO)24(μ7-C4H2N2CHC)(μ-H)3 (3), in which a Ru3 sub-unit is linked to Ru2 and Ru5 centres via di- and tri-metalated 2-vinylpyrazine ligands, exhibiting novel coordination modes including the loss of ring aromaticity in 2. The bonding of 2 and the mechanism for the fluxional behaviour of the hydrides have been examined by electronic structure calculations.
- Published
- 2019
18. Stereochemical control of the diphosphine and alkyne ligands in triruthenium clusters: The effect of reversible CO loss/addition on the ligand distribution in [Ru3(µ3,η2-PhCCPh){µ-Ph2PCH(Me)PPh2}(CO)7,8]
- Author
-
Subas Rajbangshi, Shishir Ghosh, Graeme Hogarth, Volodymyr V. Nesterov, Vladimir N. Nesterov, Michael G. Richmond, Shariff E. Kabir, and Herbert W. Roesky
- Subjects
Inorganic Chemistry ,Organic Chemistry ,Materials Chemistry ,Physical and Theoretical Chemistry ,Biochemistry - Published
- 2022
19. Proton reduction by phosphinidene-capped triiron clusters
- Author
-
Ahibur Rahaman, Michael G. Richmond, George C. Lisensky, Derek A. Tocher, Matti Haukka, Ebbe Nordlander, and Stephen B. Colbran
- Subjects
rauta ,phosphinidine ,010402 general chemistry ,Electrochemistry ,01 natural sciences ,Biochemistry ,Medicinal chemistry ,Redox ,proton reduction ,DFT ,Catalysis ,Inorganic Chemistry ,chemistry.chemical_compound ,katalyytit ,elektrokatalyysi ,Diphosphines ,Materials Chemistry ,Cluster (physics) ,electrocatalysis ,Physical and Theoretical Chemistry ,fosfori ,010405 organic chemistry ,Ligand ,Organic Chemistry ,tiheysfunktionaaliteoria ,kompleksiyhdisteet ,triiron ,Toluene ,0104 chemical sciences ,chemistry ,Phosphinidene - Abstract
Bis(phosphinidene)-capped triiron carbonyl clusters, including electron rich derivatives formed by substitution with chelating diphosphines, have been prepared and examined as proton reduction catalysts. Treatment of the known cluster [Fe3(CO)9(µ3-PPh)2] (1) with various diphosphines in refluxing THF (for 5, refluxing toluene) afforded the new clusters [Fe3(CO)7(µ3-PPh)2(κ2-dppb)] (2), [Fe3(CO)7(µ3-PPh)2(κ2-dppv)] (3), [Fe3(CO)7(µ3-PPh)2(κ2-dppe)] (4) and [Fe3(CO)7(µ3-PPh)2(µ-κ2-dppf)] (5) in moderate yields, together with small amounts of the corresponding [Fe3(CO)8(µ3-PPh)2(κ1-Ph2PxPPh2)] cluster (x = -C4H6-, -C2H2-, -C2H4-, -C3H6-, -C5H4FeC5H4-). The molecular structures of complexes 3 and 5 have been established by X-ray crystallography. Complexes 1–5 have been examined as proton reduction catalysts in the presence of p-toluenesulfonic acid (p-TsOH) in CH2Cl2. Cluster 1 exhibits two one-electron quasi-reversible reduction waves at –1.39 V (ΔE = 195 mV) and at –1.66 V (ΔE = 168 mV; potentials vs. Fc+/Fc). Upon addition of p-TsOH the unsubstituted cluster 1 shows a first catalytic wave at –1.57 V and two further proton reduction processes at –1.75 and –2.29 V, each with a good current response. The diphosphine-substituted derivatives of 1 are reduced at more negative potentials than the parent cluster 1. Clusters 2–4 each exhibit an oxidation at ca. +0.1 V and a reduction at ca. –1.6 V; for 4 conversion to a redox active successor species is seen upon both oxidation and reduction. Clusters 2–4 show catalytic waves in the presence of p-TsOH, with cluster 4 exhibiting the highest relative catalytic current (icat/i0 ≈ 57) in the presence of acid, albeit at a new third reduction process not observed for 2 and 3. Addition of the dppf ligand to the parent diphosphinidene cluster 1 gave cluster 5 which exhibited a single reduction process at –1.95 V and three oxidation processes, all at positive values as compared to 2–4. Cluster 5 showed only weak catalytic activity for proton reduction with p-TsOH. The bonding in 4 was investigated by DFT calculations, and the nature of the radical anion and dianion is discussed with respect to the electrochemical data. peerReviewed
- Published
- 2021
20. Reactions of triosmium and triruthenium clusters with 2-ethynylpyridine: new modes for alkyne C-C bond coupling and C-H bond activation
- Author
-
Md. Emdad Hossain, Shariff E. Kabir, Md. Tuhinur R. Joy, Roknuzzaman, Derek A. Tocher, Michael G. Richmond, and Shishir Ghosh
- Subjects
chemistry.chemical_classification ,Diene ,010405 organic chemistry ,General Chemical Engineering ,Substrate (chemistry) ,chemistry.chemical_element ,Alkyne ,General Chemistry ,010402 general chemistry ,Coupling (probability) ,01 natural sciences ,Medicinal chemistry ,0104 chemical sciences ,Ruthenium ,chemistry.chemical_compound ,chemistry ,Cluster (physics) ,Moiety ,Molecule - Abstract
The reaction of the trimetallic clusters [H2Os3(CO)10] and [Ru3(CO)10L2] (L = CO, MeCN) with 2-ethynylpyridine has been investigated. Treatment of [H2Os3(CO)10] with excess 2-ethynylpyridine affords [HOs3(CO)10(μ-C5H4NCH=CH)] (1), [HOs3(CO)9(μ3-C5H4NCCH2)] (2), [HOs3(CO)9(μ3-C5H4NCCCO2)] (3), and [HOs3(CO)10(μ-CHCHC5H4N)] (4) formed through either the direct addition of the Os–H bond across the CC bond or acetylenic C–H bond activation of the 2-ethynylpyridine substrate. In contrast, the dominant pathway for the reaction between [Ru3(CO)12] and 2-ethynylpyridine is C–C bond coupling of the alkyne moiety to furnish the triruthenium clusters [Ru3(CO)7(μ-CO){μ3-C5H4NCCHC(C5H4N)CH}] (5) and [Ru3(CO)7(μ-CO){μ3-C5H4NCCHC(C5H4N)CHCHC(C5H4N)}] (6). Cluster 5 contains a metalated 2-pyridyl-substituted diene while 6 exhibits a metalated 2-pyridyl-substituted triene moiety. The functionalized pyridyl ligands in 5 and 6 derive via the formal C–C bond coupling of two and three 2-ethynylpyridine molecules, respectively, and 5 and 6 provide evidence for facile alkyne insertion at ruthenium clusters. The solid-state structures of 1–3, 5, and 6 have been determined by single-crystal X-ray diffraction analyses, and the bonding in the product clusters has been investigated by DFT. In the case of 1, the computational results reveal a rare thermodynamic preference for a terminal hydride ligand as opposed to a hydride-bridged Os–Os bond (3c,2e Os–Os–H bond).
- Published
- 2020
21. Electrocatalytic proton-reduction behaviour of telluride-capped triiron clusters: tuning of overpotentials and stabilization of redox states relative to lighter chalcogenide analogues
- Author
-
Graeme Hogarth, Jess Browder-Long, Michael G. Richmond, Ebbe Nordlander, David A. Hrovat, Ahibur Rahaman, and George C. Lisensky
- Subjects
Inorganic Chemistry ,Crystallography ,chemistry.chemical_compound ,Chemistry ,Trifluoroacetic acid ,Singlet state ,Overpotential ,Cyclic voltammetry ,Electrochemistry ,Redox ,Phosphine ,Catalysis - Abstract
Reaction of [Fe3(CO)9(μ3-Te)2] (1) with the corresponding phosphine has been used to prepare the phosphine-substituted tellurium-capped triiron clusters [Fe3(CO)9(μ3-Te)2(PPh3)] (2), [Fe3(CO)8(μ3-Te)2(PPh3)] (3) and [Fe3(CO)7(μ3-Te)2(μ-R2PXPR2)] (X = CH2, R = Ph (4), Cy (5); X = NPri, R = Ph (6)). The directly related cluster [Fe3(CO)7(μ3-CO)(μ3-Te)(μ-dppm)] (7) was isolated from the reaction of [Fe3(CO)10(μ-Ph2PCH2PPh2)] with elemental tellurium. The electrochemistry of these new clusters has been probed by cyclic voltammetry, and selected complexes have been tested as proton reduction catalysts. Each 50-electron dicapped cluster exhibits two reductive processes; the first has good chemical reversibility in all cases but the reversibility of the second is dependent upon the nature of the supporting ligands. For the parent cluster 1 and the diphosphine derivatives 4–5 this second reduction is reversible, but for the PPh3 complex 3 it is irreversible, possibly as a result of CO or phosphine loss. The nature of the reduced products of 1 has been probed by DFT calculations. Upon addition of one electron, an elongation of one of the Fe–Te bonding interactions is found, while the addition of the second electron affords an open-shell triplet which is more stable by 8.8 kcal mol−1 than the closed-shell singlet dianion and has two elongated Fe–Te bonds. The phosphine-substituted clusters also exhibit oxidation chemistry but with poor reversibility in all cases. Since the reduction potentials for the tellurium-capped clusters occur at more positive potentials than for the sulfur and selenium analogues, and the redox processes also show better reversibility than for the S/Se analogues, the tellurium-capped clusters 1 and 3–5 have been examined as proton reduction catalysts. In the presence of p-toluenesulfonic acid (TsOH) or trifluoroacetic acid (TFA), these clusters reduce protons to H2 at both their first and second reduction potentials. Electron uptake at the second reduction potential is far greater than the first, suggesting that the open-shell triplet dianions are efficient catalysts. As expected, the catalytic overpotential increases upon successive phosphine substitution but so does the current response. A mechanistic scheme that takes the roles of the supporting ligands on the preferred route(s) to H2 production and release into account is presented.
- Published
- 2020
22. Electron transfer mediated by iron carbonyl clusters enhance light-driven hydrogen evolution in water by quantum dots
- Author
-
Jie Meng, Ebbe Nordlander, Mohamed Abdellah, Michael G. Richmond, Ahibur Rahaman, Kaibo Zheng, Chuanshuai Li, Hassan Mourad, Weihua Lin, Meiyuan Guo, and Alireza Honarfar
- Subjects
Materials science ,Hydrogen ,General Chemical Engineering ,chemistry.chemical_element ,Electron donor ,02 engineering and technology ,010402 general chemistry ,Photochemistry ,01 natural sciences ,proton reduction ,chemistry.chemical_compound ,Electron transfer ,Environmental Chemistry ,General Materials Science ,SDG 7 - Affordable and Clean Energy ,Hydrogen production ,Full Paper ,quantum dot ,Full Papers ,021001 nanoscience & nanotechnology ,Solar fuel ,Ascorbic acid ,electron transfer ,0104 chemical sciences ,iron carbonyl cluster ,General Energy ,chemistry ,photoluminescence spectroscopy ,Quantum dot ,0210 nano-technology ,Photocatalytic water splitting - Abstract
Photocatalytic water splitting has become a promising strategy for converting solar energy into clean and carbon‐neutral solar fuels in a low‐cost and environmentally benign way. Hydrogen gas is such a potential solar fuel/energy carrier. In a classical artificial photosynthetic system, a photosensitizer is generally associated with a co‐catalyst to convert photogenerated charge into (a) chemical bond(s). In the present study, assemblies consisting of CdSe quantum dots that are coupled with one of two molecular complexes/catalysts, that is, [Fe2S2(CO)6] or [Fe3Te2(CO)9], using an interface‐directed approach, have been tested as catalytic systems for hydrogen production in aqueous solution/organic solution. In the presence of ascorbic acid as a sacrificial electron donor and proton source, these assemblies exhibit enhanced activities for the rate of hydrogen production under visible light irradiation for 8 h in aqueous solution at pH 4.0 with up to 110 μmol of H2 per mg of assembly, almost 8.5 times that of pure CdSe quantum dots under the same conditions. Transient absorption and time‐resolved photoluminescence spectroscopies have been used to investigate the charge carrier transfer dynamics in the quantum dot/iron carbonyl cluster assemblies. The spectroscopic results indicate that effective electron transfer from the molecular iron complex to the valence band of the excited CdSe quantum dots significantly inhibits the recombination of photogenerated charge carriers, boosting the photocatalytic activity for hydrogen generation; that is, the iron clusters function as effective intermediaries for electron transfer from the sacrificial electron donor to the valence band of the quantum dots., Iron clusters for electron transfer: An assembly consisting of CdSe quantum dots and [Fe3Te2(CO)9] has been tested as a catalytic system for proton reduction in aqueous solution. Transient absorption and time‐resolved photoluminescence spectroscopies indicate that the iron cluster functions as an effective intermediary for electron transfer from a sacrificial electron donor to the valence band of the quantum dots.
- Published
- 2020
23. Cis- and trans molybdenum oxo complexes of a prochiral tetradentate aminophenolate ligand : Synthesis, characterization and oxotransfer activity
- Author
-
Kamal Hossain, Michael G. Richmond, Ebbe Nordlander, Anja Köhntopp, Matti Haukka, and Ari Lehtonen
- Subjects
010405 organic chemistry ,Chemistry ,computational modelling ,Stereoisomerism ,010402 general chemistry ,01 natural sciences ,Medicinal chemistry ,0104 chemical sciences ,Catalysis ,Inorganic Chemistry ,Solvent ,variable temperature NMR ,Tripodal ligand ,epoxidation ,Materials Chemistry ,Theoretical chemistry ,Proton NMR ,tripodal ligand ,Physical and Theoretical Chemistry ,Isomerization ,Oxo Atom Transfer ,Cis–trans isomerism - Abstract
Reaction of [MoO2Cl2(dmso)2] with the tetradentate O2N2 donor ligand papy [H2papy = N-(2-hydroxybenzyl)-N-(2-picolyl)glycine] leads to formation of the dioxomolybdenum(VI) complex [MoO2(papy)] (1) as a mixture of cis and trans isomers. Recrystallization from methanol furnishes solid cis-1, whereas the use of a dichloromethane-hexane mixture allows for the isolation of the trans-1 isomer. Both isomers have been structurally characterized by X-ray crystallography and the energy difference between the isomeric pair has been investigated by electronic structure calculations. Optimization of two configurational isomers in the gas phase predicts the trans isomer to lie 2.5 kcal/mol lower in energy (ΔG) than the cis isomer, which is inconsistent with the solution NMR data in d3-MeCN that exhibit a Keq of ca. 3 at 298 K for the trans ⇌ cis equilibrium. The DFT-computed energy difference is significantly improved (Keq = 5.4) by the inclusion of the MeCN solvent using the polarization continuum model (PCM). Density functional calculations reveal that the isomerization proceeds via a Ray-Dutt twist mechanism with a barrier of 14.5 kcal/mol, which is in accordance with the 1H NMR spectral data and the rapid equilibration of these isomers in solution. The catalytic reactivity of [MoO2(papy)] in the epoxidation of cis-cyclooctene is described, as well as its ability to effect oxo transfer from DMSO to PPh3.
- Published
- 2020
24. Synthesis of the labile rhenium(I) complexes fac-Re(CO)3(L)[κ2-O,O-FcC(O)CHC(O)Me] (where Fc = ferrocenyl; L = THF, H2O, alkyne) and alkyne addition to the diketonate ligand
- Author
-
Xiaoping Wang, Silvia Atim, Michael G. Richmond, Volodymyr V. Nesterov, and Li Yang
- Subjects
chemistry.chemical_classification ,Dimethyl acetylenedicarboxylate ,010405 organic chemistry ,Ligand ,Alkene ,Organic Chemistry ,Alkyne ,chemistry.chemical_element ,Rhenium ,010402 general chemistry ,01 natural sciences ,Biochemistry ,Medicinal chemistry ,0104 chemical sciences ,Inorganic Chemistry ,chemistry.chemical_compound ,Monomer ,chemistry ,Phenylacetylene ,Materials Chemistry ,Moiety ,Physical and Theoretical Chemistry - Abstract
Refluxing equimolar amounts of 1-ferrocenyl-1,3-butanedione (Fcacac) with BrRe(CO)5 in THF yielded the labile solvent complex fac-Re(CO)3 (THF)[κ2-O,O-FcC(O)CHC(O)Me] (1) in 80% isolated yield. 1 may also be prepared in quantitative yield using fac-BrRe(THF)2(CO)3 as the starting rhenium complex. The THF molecule in 1 is replaced by H2O during chromatographic purification to furnish the corresponding aqua complex fac-Re(CO)3(H2O)[κ2-O,O-FcC(O)CHC(O)Me] (2) while the reaction of 1 with PPh3 gives fac-Re(CO)3(PPh3)[κ2-O,O-FcC(O)CHC(O)Me] (3). Treatment of 1 with the terminal alkynes phenylacetylene and methyl propiolate proceeds rapidly at room temperature to give the π intermediates fac-Re(CO)3 (alkyne)[κ2-O,O-FcC(O)CHC(O)Me] [4 (PhC≡CH); 6 (MeO2CC CH)] that are not stable and undergo a 1,4-addition with regiospecific alkyne attack at the γ-methine site of the Fcacac moiety to afford fac-Re(CO)3 [κ3-C,O,O-FcC(O)CH(E-PhC = CH)C(O)Me] (5) and fac-Re(CO)3 [κ3-C,O,O-FcC(O)CH(Z-HC=CCO2Me)C(O)Me] (7). The solid-state structure for 5 and 7 was established by X-ray crystallography, and the computed mechanisms that account for the formation of these two products from 1 and alkyne are presented. 1 reacts with the internal alkyne dimethyl acetylenedicarboxylate (DMAD) to furnish the novel dimeric compound [fac-Re(CO)3{FcC(O)CH2C(CO2Me)C(CO2Me)}]2 (9). The structure of 9 confirms the linking of the monomeric fac-Re(CO)3{FcC(O)CH2C(CO2Me)C(CO2Me)} fragments through the intermolecular coordination of the ester oxygen atom associated with the α-carbomethoxy moiety of the E-metalated alkene; the latter moiety is traced to the initial DMAD insertion into the 6-membered metallocyclic ring in the π precursor fac-Re(CO)3 (MeO2CC CCO2Me)[κ2-O,O-FcC(O)CHC(O)Me].
- Published
- 2018
25. Synthesis and molecular structures of the 52-electron triiron telluride clusters [Fe3(CO)8(μ3-Te)2(κ2-diphosphine)] - Electrochemical properties and activity as proton reduction catalysts
- Author
-
George C. Lisensky, Derek A. Tocher, Ebbe Nordlander, Ahibur Rahaman, Michael G. Richmond, and Graeme Hogarth
- Subjects
010405 organic chemistry ,Chemistry ,Iron ,Organic Chemistry ,Infrared spectroscopy ,010402 general chemistry ,Electrochemistry ,01 natural sciences ,Biochemistry ,0104 chemical sciences ,Catalysis ,Inorganic Chemistry ,Proton reduction ,chemistry.chemical_compound ,Crystallography ,Diphosphine ,Cubane ,Diphosphines ,Materials Chemistry ,Tellurium ,Physical and Theoretical Chemistry ,Cyclic voltammetry ,HOMO/LUMO ,Phosphine - Abstract
Heating the 50-electron cluster [Fe3(CO)9 (μ3-Te)2] (1) with the diphosphines Ph2P-R-PPh2 [R = -CH2CH2- (dppe), Z-CH=CH- (dppv), 1,2-C6H4 (dppb), -CH2CH2CH2- (dpp), ferrocenyl (dppf), naphthalenyl (dppbn)] in benzene affords the 52-electron diphosphine-containing tellurium-capped triiron clusters [Fe3(CO)8 (μ3-Te)2 (κ2-diphosphine)] (diphosphine = dppe, dppv, dppb, dpp, dppf, dppnd) (2–7) in moderate yields, resulting from both phosphine addition and carbonyl loss. With 1,2-bis(diphenylphosphino)benzene (dppb) a second product is the cubane cluster [Fe4(CO)10(μ3-Te)4 (κ2-dppb)] (8). Cyclic voltammetry measurements on 2–7 reveal that all clusters show irreversible reductive behaviour at ca. −1.85 V with a series of associated small back oxidation waves, suggesting that reduction leads to significant structural change but that this can be reversed chemically. Oxidation occurs at relatively low potentials and is diphosphine-dependent. The first oxidation appears at ca. +0.35 V for 2–6 with a small degree of reversibility but is as low as +0.14 V for the bis(diphenylphosphino)naphthalene derivative 7 and in some cases is followed by further closely-spaced oxidation. Addition of [Cp2Fe][PF6] to 2–7 results in the formation of new clusters formulated as [Fe3(CO)8(μ3-Te)2(κ2-diphosphine)]+, with their IR spectra suggesting oxidation at the diiron centre. This is supported by computational studies (DFT) of the bis(diphenylphosphino)propane cluster 5 showing that the HOMO is the Fe Fe σ-bonding orbital, while the LUMO is centered on the diphosphine-substituted iron atom and has significant Fe Te σ∗-anti-bonding character consistent with the irreversible nature of the reduction. Complexes 2–7 have been examined as proton reduction catalysts in the presence of para-toluenesulfonic acid (TsOH). All are active at their first reduction potential, with a second catalytic process being observed at slightly higher potentials. While their overall electrocatalytic behaviour is similar to that noted for [Fe2(CO)6{μ-E(CH2)3E}] (E = S, Se, Te), the DFT results suggest that as the added electron is localised on the unique iron atom. The mechanistic aspects of hydrogen formation are likely to be quite different from the more widely studied diiron models.
- Published
- 2018
26. Synthesis and redox properties of fac-BrRe(CO)3[1,2-(PPh2)2-closo-1,2-C2B10H10]: The first structurally characterized rhenium carbonyl containing a carboranyl-based diphosphine ligand
- Author
-
Vladimir N. Nesterov, Chen-Hao Lin, and Michael G. Richmond
- Subjects
010405 organic chemistry ,Ligand ,Stereochemistry ,Organic Chemistry ,chemistry.chemical_element ,Rhenium ,010402 general chemistry ,Electrochemistry ,01 natural sciences ,Redox ,Medicinal chemistry ,0104 chemical sciences ,Analytical Chemistry ,Inorganic Chemistry ,chemistry ,X-ray crystallography ,Carborane ,Cyclic voltammetry ,HOMO/LUMO ,Spectroscopy - Abstract
The diphosphine 1,2-(PPh2)2-closo-1,2-C2B10H10 reacts with BrRe(CO)5 and fac-BrRe(CO)3(THF)2 to give fac-BrRe(CO)3[1,2-(PPh2)2-closo-1,2-C2B10H10] (1) in high yields (>80%). Compound 1 is the first structurally characterized rhenium carbonyl that contains an ancillary carborane-based diphosphine ligand. 1 has been characterized in solution by IR and NMR spectroscopies (1H and 31P), and the solid-state structure has been determined by X-ray diffraction analysis. The electrochemical properties of 1 have been investigated by cyclic voltammetry, and the composition of the DFT-computed HOMO and LUMO levels are discussed relative to the electrochemical data. The thermodynamics for the formation of 1 from the rhenium precursors BrRe(CO)5 and fac-BrRe(CO)3(THF)2 have been evaluated by DFT calculations.
- Published
- 2018
27. Mixed-valence dimolybdenum complexes containing hard oxo and soft carbonyl ligands: synthesis, structure, and electrochemistry of Mo2(O)(CO)2(μ-κ2-S(CH2)nS)2(κ2-diphosphine)
- Author
-
Tasneem A. Siddiquee, Mohd. Rezaul Haque, Shishir Ghosh, Graeme Hogarth, Vladimir N. Nesterov, Md. Matiar Rahman, Michael G. Richmond, and Shariff E. Kabir
- Subjects
Valence (chemistry) ,010405 organic chemistry ,Ligand ,Diastereomer ,chemistry.chemical_element ,010402 general chemistry ,Electrochemistry ,01 natural sciences ,Bond order ,0104 chemical sciences ,Inorganic Chemistry ,Crystallography ,chemistry ,Molybdenum ,Diphosphines ,Cis–trans isomerism - Abstract
Mixed-valence dimolybdenum complexes Mo2(O)(CO)2{μ-κ2-S(CH2)nS}2(κ2-Ph2P(CH2)mPPh2) (n = 2, 3; m = 1, 2) (1–4) have been synthesized from one-pot reactions of fac-Mo(CO)3(NCMe)3 and dithiols, HS(CH2)nSH, in the presence of diphosphines. The dimolybdenum framework is supported by two thiolate bridges, with one molybdenum carrying a terminal oxo ligand and the second two carbonyls. The dppm (m = 1) products exist as a pair of diastereomers differing in the relative orientation of the two carbonyls (cis and trans) at the Mo(CO)2(dppm) center, while dppe (m = 2) complexes are found solely as the trans isomers. Small amounts of Mo(CO){κ3-S(CH2CH2S)2}(κ2-dppe) (5) also result from the reaction using HS(CH2)2SH and dppe. The bonding in isomers of 1–4 has been computationally explored by DFT calculations, trans diastereomers being computed to be more stable than the corresponding pair of cis diastereomers for all. The calculations confirm the existence of MoO and Mo–Mo bond orders and suggest that the new dimeric compounds are best viewed as Mo(V)–Mo(I) mixed-valence systems. The electrochemical properties of 1 have been investigated by CV and show a reversible one-electron reduction associated with the Mo(V) centre, while two closely spaced irreversible oxidation waves are tentatively assigned to oxidation of the Mo(I) centre of the two isomers as supported by DFT calculations.
- Published
- 2018
28. Experimental and computational preference for phosphine regioselectivity and stereoselective tripodal rotation in HOs3(CO)8(PPh3)2(μ-1,2-N,C-η1,κ1-C7H4NS)
- Author
-
Shariff E. Kabir, Kenneth I. Hardcastle, Edward Rosenberg, Shishir Ghosh, Derek A. Tocher, Shahin A. Begum, Michael G. Richmond, Md. Arshad H. Chowdhury, and Li Yang
- Subjects
010405 organic chemistry ,Chemistry ,Stereochemistry ,Hydride ,General Chemical Engineering ,chemistry.chemical_element ,Regioselectivity ,General Chemistry ,Nuclear magnetic resonance spectroscopy ,010402 general chemistry ,01 natural sciences ,0104 chemical sciences ,chemistry.chemical_compound ,Molecule ,Moiety ,Stereoselectivity ,Osmium ,Phosphine - Abstract
The site preference for ligand substitution in the benzothiazolate-bridged cluster HOs3(CO)10(μ-1,2-N,C-η1,κ1-C7H4NS) (1) has been investigated using PPh3. 1 reacts with PPh3 in the presence of Me3NO to afford the mono- and bisphosphine substituted clusters HOs3(CO)9(PPh3)(μ-1,2-N,C-η1,κ1-C7H4NS) (2) and HOs3(CO)8(PPh3)2(μ-1,2-N,C-η1,κ1-C7H4NS) (3), respectively. 2 exists as a pair of non-interconverting isomers where the PPh3 ligand is situated at one of the equatorial sites syn to the edge-bridging hydride that shares a common Os–Os bond with the metalated heterocycle. The solid-state structure of the major isomer establishes the PPh3 regiochemistry at the N-substituted osmium center. DFT calculations confirm the thermodynamic preference for this particular isomer relative to the minor isomer whose phosphine ligand is located at the adjacent C-metalated osmium center. 2 also reacts with PPh3 to give 3. The locus of the second substitution occurs at one of the two equatorial sites at the Os(CO)4 moiety in 2 and gives rise to a pair of fluxional stereoisomers where the new phosphine ligand is scrambled between the two equatorial sites at the Os(CO)3P moiety. The molecular structure of the major isomer has been determined by X-ray diffraction analysis and found to represent the lowest energy structure of the different stereoisomers computed for HOs3(CO)8(PPh3)2(μ-1,2-N,C-η1,κ1-C7H4NS). The fluxional behavior displayed by 3 has been examined by VT NMR spectroscopy, and DFT calculations provide evidence for stereoselective tripodal rotation at the Os(CO)3P moiety that serves to equilibrate the second phosphine between the two available equatorial sites.
- Published
- 2018
29. Ligand coordination in [Re2(CO)9(NCMe)] and [H3Re3(CO)11(NCMe)] by triphenylantimony: Reactivity studies and Sb–Ph bond cleavage to give new antimony-containing di- and trirhenium complexes
- Author
-
Shishir Ghosh, Subas Rajbangshi, M. A. Al Mamun, Shariff E. Kabir, and Michael G. Richmond
- Subjects
Ligand ,Organic Chemistry ,Biochemistry ,Medicinal chemistry ,Toluene ,Inorganic Chemistry ,chemistry.chemical_compound ,chemistry ,Materials Chemistry ,Molecule ,Phenyl group ,Reactivity (chemistry) ,Physical and Theoretical Chemistry ,Acetonitrile ,Benzene ,Bond cleavage - Abstract
The reactions of the MeCN-substituted compounds [Re2(CO)9(NCMe)] and [H3Re3(CO)11(NCMe)] with SbPh3 are described. Treatment of [Re2(CO)9(NCMe)] with SbPh3 in refluxing benzene furnished [Re2(CO)9(SbPh3)] (1) in good yield. Complex 1 reacts with SbPh3 in the presence of Me3NO in acetonitrile to give [Re2(CO)8(σ-C6H5)(µ-SbPh2)(NCMe)] (2), which on further reaction with additional SbPh3 in refluxing benzene afforded the known complex [Re2(CO)8(σ-C6H5)(µ-SbPh2)(SbPh3)] (3). Heating 1 with SbPh3 at 110 °C yields 3, which on prolonged heating at 110 °C in the presence of SbPh3 furnished the known stibene-bridged compound [Re2(CO)7(µ-SbPh2)2(SbPh3)] (4). Complex 2 is a rare example of a dirhenium complex containing a sigma-bonded phenyl group and a labile acetonitrile ligand in the same molecule. The stibene-bridged dirhenium complexes (2−4) result from the cleavage of antimony-carbon and Re–Re bonds. Heating [H3Re3(CO)11(NCMe)] with SbPh3 in refluxing toluene afforded the new compounds [H3Re3(CO)11(SbPh3)] (5), [H2Re3(CO)11(µ-SbPh2)(SbPh3)] (6), and [HRe2(CO)6(µ-SbPh2)(SbPh3)2] (8) together with previously reported compound [HRe2(CO)7(µ-SbPh2)(SbPh3)] (7). Compound 5 is a simple monosubstituted product, while 6 contains a terminally coordinated SbPh3 ligand and a bridging SbPh2 group formed by Sb–Ph bond scission. Compound 8 results from cluster fragmentation and SbPh3 ligand activation by Sb–Ph bond cleavage. A series of separate thermolysis experiments have been performed in order to establish the relationship between the different products. Compounds 1, 2, 5, 6, and 8 have been structurally characterized by X-ray crystallography. The bonding in these new compounds has been investigated by electronic structure calculations.
- Published
- 2021
30. Oxygen atom transfer catalysis by dioxidomolybdenum(VI) complexes of pyridyl aminophenolate ligands
- Author
-
Michael G. Richmond, Matti Haukka, Ari Lehtonen, Ebbe Nordlander, Nadia C. Mösch-Zanetti, Jörg A. Schachner, and Kamal Hossain
- Subjects
chemistry.chemical_classification ,Aqueous solution ,010405 organic chemistry ,Chemistry ,Ligand ,Alkene ,Dimer ,Cationic polymerization ,010402 general chemistry ,01 natural sciences ,0104 chemical sciences ,Catalysis ,Inorganic Chemistry ,chemistry.chemical_compound ,Octahedral molecular geometry ,Polymer chemistry ,Materials Chemistry ,Reactivity (chemistry) ,Physical and Theoretical Chemistry - Abstract
A series of new cationic dioxidomolybdenum(VI) complexes [MoO2(Ln)]PF6 (2–5) with the tripodal tetradentate pyridyl aminophenolate ligands HL2-HL5 have been synthesized and characterized. Ligands HL2-HL4 carry substituents in the 4-position of the phenolate ring, viz. Cl, Br and NO2, respectively, whereas the ligand HL5, N-(2-hydroxy-3,5-di-tert-butylbenzyl)-N,N-bis(2-pyridylmethyl)amine, is a derivative of 3,5-di-tert-butylsalicylaldehyde. X-ray crystal structures of complexes 2, 3 and 5 reveal that they have a distorted octahedral geometry with the bonding parameters around the metal centres being practically similar. Stoichiometric oxygen atom transfer (OAT) properties of 5 with PPh3 were investigated using UV–Vis, 31P NMR and mass spectrometry. In CH2Cl2 solution, a dimeric Mo(V) complex [(µ-O){MoO(L5)}2](PF6)2 6 was formed while in methanol solution an air-sensitive Mo(IV) complex [MoO(OCH3)(L5)] 7 was obtained. The solid-state structure of the µ-oxo bridged dimer 6 was determined by X-ray diffraction. Complex 7 is unstable under ambient conditions and capable of reducing DMSO, thus showing reactivity analogous to that of DMSO reductases. Similarly, the OAT reactions of complexes 2–4 also resulted in the formation of dimeric Mo(V) and unsaturated monomeric Mo(IV) complexes that are analogous to complexes 6 and 7. Catalytic OAT at 25 °C could also be observed, using complexes 1–5 as catalysts for oxidation of PPh3 in deuterated dimethylsulfoxide (DMSO‑d6), which functioned both as a solvent and oxidant. All complexes were also tested as catalysts for sulfoxidation of methyl-p-tolylsulfide and epoxidation of various alkene substrates with tert-butyl hydroperoxide (TBHP) as an oxidant. Complex 1 did not exhibit any sulfoxidation activity under the conditions used, while 2–5 catalyzed the sulfoxidation of methyl-p-tolylsulfide. Only complexes 2 and 3, with ligands containing halide substituents, exhibited good to moderate activity for epoxidation of all alkene substrates studied, and, in general, good activity for all molybdenum(VI) catalysts was only exhibited when cis-cyclooctene was used as a substrate. No complex catalysed epoxidation of cis-cyclooctene when an aqueous solution of H2O2 was used as potential oxidant.
- Published
- 2021
31. The reaction of Os3(CO)12 with triphos {MeC(CH2PPh2)3}: A case of multiple C-P and C-H bond activations
- Author
-
Li Yang, Jade Y. Jung, Vladimir N. Nesterov, Soo Hun Yoon, Michael G. Richmond, David M. Marolf, and Gregory L. Powell
- Subjects
010405 organic chemistry ,Ligand ,Stereochemistry ,Aryl ,Organic Chemistry ,chemistry.chemical_element ,Nuclear magnetic resonance spectroscopy ,010402 general chemistry ,01 natural sciences ,Biochemistry ,Triphos ,0104 chemical sciences ,Inorganic Chemistry ,Metal ,chemistry.chemical_compound ,Crystallography ,chemistry ,visual_art ,Tripodal ligand ,Materials Chemistry ,visual_art.visual_art_medium ,Osmium ,Physical and Theoretical Chemistry ,Benzene - Abstract
The reaction of Os3(CO)12 with the tripodal ligand 1,1,1-tris(diphenylphosphinomethyl)ethane, triphos, results in the formation of the 48e trinuclear cluster complex Os3[μ-Ph2PCH2C(Me)(CH2PPh)CH2PC6H4](CO)7(η1-Ph). This new complex has been characterized by X-ray crystallography as well as by IR and NMR spectroscopy. The activation of the triphos ligand to yield the novel 9e donor ligand Ph2PCH2C(Me)(CH2PPh)CH2PC6H4 is unprecedented. The triphos ligand does not cap the triangular Os3 face; instead, the three phosphorus atoms are coordinated to only two of the metal atoms with the remaining osmium atom coordinated by an ortho-metalated aryl group. The triphos ligand is also transformed in several ways. Two C–P bonds are cleaved during the reaction, yielding an η1 phenyl ligand in the product and free benzene. The bonding in 1 has been examined by electronic structure calculations.
- Published
- 2017
32. Ambidentate Ligand Reactivity with the Rhenium(I) Compounds [BrRe(CO) 4 ] 2 and cis ‐BrRe(CO) 4 L: A Kinetic and Mechanistic Study
- Author
-
Vladimir N. Nesterov, Michael G. Richmond, and Darrell D. Mayberry
- Subjects
Substitution reaction ,Reaction mechanism ,010405 organic chemistry ,Ligand ,Dimer ,Regioselectivity ,010402 general chemistry ,Photochemistry ,01 natural sciences ,Medicinal chemistry ,0104 chemical sciences ,Inorganic Chemistry ,chemistry.chemical_compound ,chemistry ,Pyridine ,Moiety ,Phosphine - Abstract
The reaction of the dimer [BrRe(CO)4]2 with the ambidentate donor 2-(diphenylphosphino)pyridine (PN) has been investigated. The substitution reaction is rapid at room temperature and regioselective for phosphorus coordination to yield exclusively cis-BrRe(CO)4(κP-PN) (1). Thermolysis of 1 furnishes the PN-chelated product fac-BrRe(CO)3(κP,N-PN) (2) and CO. The regioselectivity in the ligand substitution reaction involving [BrRe(CO)4]2 and the tridentate donor 6-(diphenylphosphino)-2-formylpyridine (PON) was also studied, and consistent with the PN donor, only the κP-product, cis-BrRe(CO)4(κP-PON) (3) is formed. The reaction of the dimer [BrRe(CO)4]2 with 2-(diphenylphosphino)pyridine (PN) has been computationally modeled by DFT calculations. The energetics for the creation of an unsaturated intermediate through the cleavage of one of the bridging bromide ligands, followed by the addition of the donor to the unsaturated rhenium center, versus the direct attack of the pnictogen donor on the dimer have been evaluated. The latter process is computed as the preferred route for dimer activation, with an attack of the pyridyl moiety slightly favored compared to the phosphine moiety. Concerning the reaction of the PN ligand with [BrRe(CO)4]2, we predict the initial formation of the κN-isomer of 1 as the kinetic product of substitution, which in turn undergoes a rapid isomerization to furnish the thermodynamically more stable κP-isomer through a reversible ligand dissociation process.
- Published
- 2017
33. Reactions of Ru3(CO)10(μ-dppm) with Ph3GeH: Ge–H and Ge–C bond cleavage in Ph3GeH at triruthenium clusters
- Author
-
Shishir Ghosh, Mehedi Mahabub Khan, Shariff E. Kabir, Derek A. Tocher, Herbert W. Roesky, Mahbub Alam, Ahibur Rahaman, and Michael G. Richmond
- Subjects
010405 organic chemistry ,Hydride ,Organic Chemistry ,Inorganic chemistry ,010402 general chemistry ,01 natural sciences ,Biochemistry ,Oxidative addition ,Methane ,0104 chemical sciences ,3. Good health ,Inorganic Chemistry ,Crystallography ,chemistry.chemical_compound ,chemistry ,Materials Chemistry ,Cluster (physics) ,Physical and Theoretical Chemistry ,Bond cleavage - Abstract
The activation of Ph3GeH at the dppm-bridged cluster Ru3(CO)10(μ-dppm) [dppm = bis(diphenylphosphino)methane] has been investigated. Ru3(CO)10(μ-dppm) reacts with Ph3GeH at room temperature in the presence of Me3NO to give the new cluster products Ru3(CO)9(GePh3)(μ-dppm)(μ-H) (1) and Ru3(CO)8(GePh3)2(μ-dppm)(μ-H)2 (2) via successive oxidation-addition of two Ge–H bonds. Refluxing 1 in THF furnishes the diruthenium complex Ru2(CO)6(μ-GePh2)(μ-dppm) (3) as the major product (44%), in addition to Ru3(CO)7(μ-CO)(GePh3){μ3-PhPCH2P(Ph)C6H4}(μ-H) (4) and the known cluster Ru3(CO)9(μ-H)(μ3-Ph2PCH2PPh) (5) in 7 and 8% yields, respectively. Heating samples of cluster 2 also afforded 3 as the major product together with a small amount of Ru3(CO)7(GePh3)(μ-OH)(μ-dppm)(μ-H)2 (6). DFT calculations establish the stability of the different possible isomers for clusters 1, 2, and 6, in addition to providing insight into the mechanism for hydride fluxionality in 2. All new compounds have been characterized by analytical and spectroscopic methods, and the molecular structures of 1, 3, and 6 have been established by single-crystal X-ray diffraction analyses.
- Published
- 2017
34. Reversible C-H bond activation at a triosmium centre: A comparative study of the reactivity of unsaturated triosmium clusters Os 3 (CO) 8 (μ-dppm)(μ-H) 2 and Os 3 (CO) 8 (μ-dppf)(μ-H) 2 with activated alkynes
- Author
-
Shaikh M. Mobin, Graeme Hogarth, Md. Arshad H. Chowdhury, Herbert W. Roesky, Michael G. Richmond, Shishir Ghosh, Shariff E. Kabir, Mohd. Rezaul Haque, and Derek A. Tocher
- Subjects
Diphosphines ,Alkyne ,010402 general chemistry ,Photochemistry ,DFT ,01 natural sciences ,Biochemistry ,Medicinal chemistry ,Inorganic Chemistry ,chemistry.chemical_compound ,Reversible C-H bond activation ,Cyclopentadienyl complex ,Materials Chemistry ,C-C bond scission ,Physical and Theoretical Chemistry ,Bond cleavage ,chemistry.chemical_classification ,010405 organic chemistry ,Hydride ,Ligand ,Organic Chemistry ,Nuclear magnetic resonance spectroscopy ,0104 chemical sciences ,Ferrocene ,chemistry ,Unsaturated osmium clusters ,Activated alkynes - Abstract
Heating a benzene solution of the unsaturated cluster Os3(CO)8(μ-dppm)(μ-H)2 (1) [dppm = bis(diphenylphosphino)methane] with MeO2CC CCO2Me (DMAD) or EtO2CC CCO2Et (DEAD) at 80 °C furnished the dinuclear compounds Os2(CO)4(μ-dppm)(μ-η2;η1;к1-RO2CCCHCO2R)(μ-H) (3a, R = Me, 3b, R = Et) and the saturated trinuclear complexes Os3(CO)7(μ-dppm)(μ3-η2;η1;η1-RO2CCCCO2R)(μ-H)2 (4a, R = Me, 4b, R = Et). In contrast, similar reactions using unsaturated Os3(CO)8(μ-dppf)(μ-H)2 (2) [dppf = bis(diphenylphosphino)ferrocene] afforded only the trinuclear complexes Os3(CO)8(μ-dppf)(μ-η2;η1-RO2CCHCCO2R)(μ-H) (5a, R = Me; 5b, R = Et) and Os3(CO)7(μ-dppf)(μ3-η2;η1;η1-RO2CCCCO2R)(μ-H)2 (6a, R = Me; 6b, R = Et). Control experiments confirm that 5a and 5b decarbonylate at 80 °C to give 6a and 6b, respectively. Both 5a and 5b exist as a pair of isomers in solution, as demonstrated by 1H NMR and 31P{1H} NMR spectroscopy. DFT calculations on cluster 5a (as the dppf-Me4 derivative) indicate that the isomeric mixture derives from a torsional motion that promotes the conformational flipping of the cyclopentadienyl groups of the dppf-Me4 ligand relative to the metallic plane. VT NMR measurements on clusters 6a and 6b indicate that while the hydride ligand associated with the dppf-bridged Os-Os bond is nonfluxional at room temperature, the second hydride rapidly oscillates between the two non-dppf-bridged Os-Os edges. DFT examination of this hydride fluxionality confirms a “windshield wiper” motion for the labile hydride that gives rise to a time-average coupling of this hydride to both phosphorus centers of the dppf ligand. Thermolysis of 6a and 6b in refluxing toluene yielded Os3(CO)7(μ-dppf)(μ-η2;η1;к1-CCHCO2R) (7a, R=Me; 7b, R=Et). The vinylidene moieties in 7a and 7b derive from the carbon-carbon bond cleavage of coordinated alkyne ligands, and these two products exhibit high thermal stability in refluxing toluene.
- Published
- 2017
35. 5,7-Dihydroxy-2-(4-hydroxyphenyl)chroman-4-one (naringenin): X-ray diffraction structures of the naringenin enantiomers and DFT evaluation of the preferred ground-state structures and thermodynamics for racemization
- Author
-
Feroza K. Choudhury, Vladimir Shulaev, Khadiza Zaman, Vladimir N. Nesterov, Volodymyr V. Nesterov, Antonella Longo, Lev N. Zakharov, Michael G. Richmond, William Farrell, and Jose G. Calderon
- Subjects
010405 organic chemistry ,Stereochemistry ,Hydrogen bond ,Aryl ,010401 analytical chemistry ,Organic Chemistry ,Substituent ,Thermodynamics ,Ring (chemistry) ,01 natural sciences ,0104 chemical sciences ,Analytical Chemistry ,Inorganic Chemistry ,chemistry.chemical_compound ,chemistry ,Intramolecular force ,Molecule ,Enantiomer ,Racemization ,Spectroscopy - Abstract
The R- and S-enantiomers of naringenin were separated by chiral supercritical fluid (SCF) and the absolute configuration of each enantiomer was established by X-ray crystallography. The solid-state data is in agreement with the reported circular dichroism spectra. Both enantiomers crystallize in the monoclinic crystal system in the space group P21 with two independent molecules in the asymmetric unit. In all molecules, the pyrone ring adopts a flattened chair-like conformation in which the C1 atom deviates from the plane drawn through the remaining five atoms of this heterocycle. The 4-hydroxyphenyl substituent located at C1 of the pyrone ring occupies an equatorial position and lies in a plane that is almost perpendicular to the aromatic platform associated with the heterocyclic portion of the molecule. Strong intramolecular O-H⋯O hydrogen bonding exists between the carbonyl moiety and the aryl hydroxyl group at C5. In both enantiomers, a favorable mutual orientation of two independent molecules promotes the formation of intermolecular O-H⋯O hydrogen bonds that link them into dimers. There are additional long-range intermolecular O-H⋯O hydrogen bonds and weak C-H⋯O contacts within the unit cell of each enantiomer that connect dimers in an extended network. DFT calculations have been performed and the thermodynamics for naringenin racemization via an acyclic chalcone have been computed. Eight energetically accessible conformations have been verified for S-naringenin.
- Published
- 2017
36. Diphosphine-bridged digold(I) compounds: Structural and computational studies on the aurophilic interaction in Au2Cl2(μ-bpcd) and Au2Cl2(μ-bmi)
- Author
-
Michael G. Richmond, Rogers Nyamwihura, Li Yang, and Vladimir N. Nesterov
- Subjects
010405 organic chemistry ,Chemistry ,Organic Chemistry ,Electronic structure ,Nuclear magnetic resonance spectroscopy ,Dihedral angle ,Staggered conformation ,010402 general chemistry ,01 natural sciences ,0104 chemical sciences ,Analytical Chemistry ,Inorganic Chemistry ,Crystallography ,X-ray crystallography ,Potential energy surface ,Molecule ,HOMO/LUMO ,Spectroscopy - Abstract
The reaction of AuCl(tht) with the diphosphine donors 4,5-bis(diphenylphosphino)-4-cyclopenten-1,3-dione (bpcd) and 2,3-bis(diphenylphosphino)- N -phenylmaleimide (bmi) in a 2:1 stoichiometry affords the diphosphine-bridged digold(I) complexes Au 2 Cl 2 (bpcd) ( 1 ) and Au 2 Cl 2 (bmi) ( 2 ), respectively. 1 and 2 have been isolated and characterized in solution by IR and NMR spectroscopy ( 1 H and 31 P), and the solid-state structures established by X-ray crystallography. The X-Au-Au-X (X = Cl, P) atoms in both Au 2 dimers exhibit a gauche-type (staggered) interaction based on a torsion angle of −54° for 1 (X = Cl) and −70° for 2 (X = Cl, mean angle for the two independent molecules). Each Au 2 product displays a weak aurophilic interaction based on a Au-Au internuclear distance on the order of 2.9 A. The bonding in 1 and 2 has been investigated by electronic structure calculations and the composition of the HOMO and LUMO levels determined in the case of 2 . The potential energy surface for the interconversion of 2 to an alternative staggered conformation has been computed, and this transformation takes place through an eclipsed transition structure. The preference for a structure that contains a staggered orientation of Cl-Au-Au-Cl and P-Au-Au-Cl atoms is discussed.
- Published
- 2017
37. Catalytic C-H oxidations by nonheme mononuclear Fe(II) complexes of two pentadentate ligands: Evidence for an Fe(IV) oxo intermediate
- Author
-
Mainak Mitra, Ebbe Nordlander, Hassan Nimir, Miquel Costas, Albert A. Shteinman, Michael G. Richmond, and David A. Hrovat
- Subjects
chemistry.chemical_classification ,Ketone ,010405 organic chemistry ,Process Chemistry and Technology ,Radical ,Inorganic chemistry ,010402 general chemistry ,01 natural sciences ,Medicinal chemistry ,Redox ,Catalysis ,0104 chemical sciences ,chemistry.chemical_compound ,chemistry ,Kinetic isotope effect ,Hydroxyl radical ,Physical and Theoretical Chemistry ,Hydrogen peroxide ,Bond cleavage - Abstract
The oxidation reactions of alkanes with hydrogen peroxide and peracids (peracetic acid (PAA) and m-chloroperoxybenzoic acid (mCPBA)) catalysed by two Fe(II) complexes of pentadentate {N5}-donor ligands have been investigated. Kinetic isotope effect experiments and the use of other mechanistic probes have also been performed. While the total yields of oxidized products are similar regardless of oxidant (e.g. 30–39% for oxidation of cyclohexane), the observed alcohol/ketone ratios and kinetic isotope effects differ significantly with different oxidants. Catalytic reactions in H2O2 medium are consistent with the involvement of hydroxyl radicals in the Csingle bondH bond cleavage step, and resultant low kinetic isotope effect values. On the other hand, catalytic reactions performed using peracid media indicate the involvement of an oxidant different from the hydroxyl radical. For these reactions, the kinetic isotope effect values are relatively high (within a range of 4.2–5.1) and the C3/C2 selectivity parameters in adamantane oxidation are greater than 11, thereby excluding the presence of hydroxyl radicals in the Csingle bondH bond cleavage step. A low spin Fe(III)-OOH species has been detected in the H2O2-based catalytic system by UV/Vis, mass spectrometry and EPR spectroscopy, while an Fe(IV)-oxo species is postulated to be the active oxidant in the peracid-based catalytic systems. Computational studies on the Csingle bondH oxidation mechanism reveal that while the hydroxyl radical is mainly responsible for the H-atom abstraction in the H2O2-based catalytic system, it is the Fe(IV)-oxo species that abstracts the H-atom from the substrate in the peracid-based catalytic systems, in agreement with the experimental observations. (Less)
- Published
- 2017
38. New molecular architectures containing low-valent cluster centres with di- and trimetalated 2-vinylpyrazine ligands: synthesis and molecular structures of Ru
- Author
-
Md Monir, Hossain, Nahid, Akter, Shishir, Ghosh, Vladimir N, Nesterov, Michael G, Richmond, Graeme, Hogarth, and Shariff E, Kabir
- Abstract
Reaction of 2-vinylpyrazine with Ru
- Published
- 2019
39. Hydrogenase biomimics containing redox-active ligands: Fe2(CO)4(μ-edt)(κ2-bpcd) with electron-acceptor 4,5-bis(diphenylphosphino)-4-cyclopenten-1,3-dione (bpcd) as a potential [Fe4–S4]H surrogate
- Author
-
Mark R. Warren, David A. Hrovat, Nathan Hollingsworth, Graeme Hogarth, Michael G. Richmond, and Shishir Ghosh
- Subjects
chemistry.chemical_classification ,010405 organic chemistry ,Hydride ,Ligand ,Maleic anhydride ,Protonation ,Electron acceptor ,010402 general chemistry ,01 natural sciences ,Redox ,Medicinal chemistry ,0104 chemical sciences ,Inorganic Chemistry ,chemistry.chemical_compound ,chemistry ,Bond cleavage ,Coordination geometry - Abstract
[FeFe]-hydrogenases contain strongly electronically coupled diiron [2Fe]H and tetrairon [Fe4–S4]H clusters, and thus much recent effort has focused on the chemistry of diiron-dithiolate biomimics with appended redox-active ligands. Here we report on the synthesis and electrocatalytic activity of Fe2(CO)4(μ-edt)(κ2-bpcd) (2) in which the electron-acceptor 4,5-bis(diphenylphosphino)-4-cyclopenten-1,3-dione (bpcd) acts as a surrogate of the [Fe4–S4]H sub-cluster. The complex is prepared in low yield but has been fully characterised, including a crystallographic study which shows that the diphosphine adopts a basal-apical coordination geometry in the solid state. Cyclic voltammetry shows that 2 undergoes four reduction events with DFT studies confirming that the first reduction is localised on the low-lying π* system of the diphosphine ligand. The addition of the second electron furnishes a triplet dianion that exhibits spin density distributed over the diphosphine and diiron subunits. Protonation at the Fe–Fe bond of the triplet dianion furnishes the corresponding bridging hydride as the thermodynamically favoured species that contains a reduced bpcd ligand. Complex 2 functions as a catalyst for proton-reduction at its second reduction potential, in contrast to the related 2,3-bis(diphenylphosphino)maleic anhydride (bma) complex, Fe2(CO)4(μ-pdt)(κ2-bma) (1), which shows similar electrochemical behaviour but is not catalytically active. The difference in chemical behaviour is attributed to greater stability of the 4-cyclopenten-1,3-dione platform in 2 as compared to the maleic anhydride ring of the bma ligand in 1 following the uptake of the second electron. Thus protonation of the Fe–Fe bond in the 22− affords a species which is stable enough to undergo a further reduction–protonation event, unlike the bma ligand whose maleic anhydride ring undergoes deleterious C–O bond scission upon protonation or reaction with adventitious moisture. DFT studies, however, suggest that electron-transfer from the diphosphine to the diiron centre is not significant, probably due to their poor redox levelling. Thus, while the diphosphine is readily reduced, the added electron is apparently not utilised in proton-reduction and hence cannot truly be considered as an [Fe4–S4]H surrogate.
- Published
- 2019
40. Models of the iron-only hydrogenase enzyme: structure, electrochemistry and catalytic activity of Fe
- Author
-
David G, Unwin, Shishir, Ghosh, Faith, Ridley, Michael G, Richmond, Katherine B, Holt, and Graeme, Hogarth
- Abstract
A series of diiron bis(2-diphenylphosphinoethyl)phenylphosphine (triphos) complexes Fe2(CO)3(μ-dithiolate)(μ,κ1,κ2-triphos) (1-4) [dithiolate = 1 pdt; 2 edt; 3 adt (R = Bz), 4 (SMe)2] have been prepared and investigated as biomimics of the diiron site of [FeFe]-hydrogenases. The triphos ligand bridges the diiron vector whilst also chelating to one iron and 1-3 exist as a mixture of basal-basal-apical (bba) and basal-basal-basal (bbb) isomers which differ in the mode of chelation. In solution the bba and bbb forms do not interconvert on the NMR time scale, but the bba isomers are fluxional, and at low temperature four forms of 1bba are seen as the conformations for the pdt ring and triphos methylene groups are frozen. Crystallographic studies have established bba (pdt) and bbb (adt) ground state conformations and in both there is a significant deviation away from the expected eclipsed conformation (Lap-Fe-Fe-Lap torsion angle 0°) by 49.4 and 24.9° respectively, suggesting that introduction of triphos leads to significant strain and DFT calculations have been used to understand the relative energies of isomers. The electron rich nature of the diiron centre in 1-4 would suggest rapid protonation, but while bridging hydride complexes such as [Fe2(CO)3(μ-pdt)(μ,κ1,κ2-triphos)(μ-H)][BF4] (1H+) can be formed the process is slow. This behavior is likely a result of the high energy barrier in forming the initial (not observed) terminal hydride which requires a significant conformational change in triphos coordination. CV studies show that all starting compounds oxidize at low potentials and the addition of [Cp2Fe][PF6] to 1 affords [Fe2(CO)3(μ-pdt)(μ,κ1,κ2-triphos)][PF6] (1+) which has been characterised by IR spectroscopy. DFT studies suggest a ground state for 1+ with a partially rotated Fe(CO)2P moiety that yields a weak semi-bridging carbonyl with the adjacent Fe(CO)P2 group. No reduction peaks are seen for 1-4 within the solvent window but 1H+ undergoes reduction at -1.7 V. All complexes act as proton-reduction catalysts in the presence of HBF4·Et2O. For 1, three separate processes are observed and their dependence on acid concentration has been probed, and a mechanistic scheme is proposed based on formation via a CECE process of 1(μ-H)H which can either slowly release H2 or undergo further reduction. Relative contributions of the three processes to the total current were found to be highly dependent upon the background electrolyte, being attributed to their relative abilities to facilitate proton transfer processes. While 2 and 4 show similar proton reduction behaviour, the adt complex 3 is quite different being attributed to facile protonation of nitrogen which is followed by addition of a second proton at the diiron centre.
- Published
- 2019
41. Bimodal substitution behavior in the reaction of N,N’-diisopropylformamidine with [Os3(CO)10(NCMe)2]: Kinetics and molecular structures of the formamidinate-substituted clusters HOs3(CO)9[μ-C(O)NPr C(H)NPr ], HOs3(CO)10[μ-NPr C(H)NPr ], and HOs3(CO)9[μ3-NPr C(H)NPr ]
- Author
-
Michael G. Richmond, Li Yang, and Vladimir N. Nesterov
- Subjects
Chemistry ,Ligand ,Organic Chemistry ,Thermal decomposition ,Kinetics ,Decarbonylation ,Atmospheric temperature range ,Biochemistry ,Medicinal chemistry ,Molecularity ,Inorganic Chemistry ,Yield (chemistry) ,Materials Chemistry ,Cluster (physics) ,Physical and Theoretical Chemistry - Abstract
The dinitrogen donor N,N’-diisopropylformamidine [PriN=C(H)NHPri] reacts with the triosmium cluster Os3(CO)10(NCMe)2 at room temperature to yield the isomeric clusters HOs3(CO)9[μ-C(O)NPriC(H)NPri] (1) and HOs3(CO)10[μ-NPriC(H)NPri] (2) in a 1:2.8 ratio. 1 contains an edge-bridging iminocarbamoyl ligand, while 2 contains a bridging formamidinate ligand. Thermolysis of 1 yields 2 plus the face-capped cluster HOs3(CO)9[μ3-NPriC(H)NPri] (3). The decarbonylation of 2 to 3 + CO confirms the molecularity of the observed reaction steps. The three products have been fully characterized in solution by IR and NMR spectroscopies, and the solid-state structures for 1-3 have been determined by X-ray crystallography. The kinetics for the thermolysis reaction were investigated over the temperature range 342–383 K, and the concentration versus time profiles for the conversion of 1 → 2 → 3 + CO have been successfully modeled using two consecutive, irreversible first-order reactions. The bonding in clusters 1-3 have been examined by DFT, and these data support cluster 1 as the kinetic substitution product and cluster 2 as the thermodynamically favored isomer.
- Published
- 2021
42. Microwave-induced dppm ligand substitution in triosmium clusters: Structural and DFT evaluation of Os3 clusters containing multiply activated dppm ligands through cyclometalation, ortho metalation, and P C bond cleavage
- Author
-
Nigel Gwini, Michelle L. Parker, Gregory L. Powell, David M. Marolf, Jade Y. Jung, Vladimir N. Nesterov, Li Yang, James E. Johnstone, Soo Hun Yoon, Michael G. Richmond, David K. Kempe, and Audrey G. Fikes
- Subjects
010405 organic chemistry ,Chemistry ,Stereochemistry ,Metalation ,Organic Chemistry ,chemistry.chemical_element ,Electronic structure ,010402 general chemistry ,01 natural sciences ,Biochemistry ,Medicinal chemistry ,0104 chemical sciences ,Inorganic Chemistry ,Metal ,chemistry.chemical_compound ,visual_art ,Materials Chemistry ,visual_art.visual_art_medium ,Cluster (physics) ,Osmium ,Physical and Theoretical Chemistry ,Bond cleavage ,Phosphine ,Microwave - Abstract
Four new triosmium carbonyl complexes containing multiple dppm ligands were produced by microwave heating of solutions containing Os3(CO)12 and excess dppm. Three of these complexes, Os3(μ-H)2(CO)6(μ-dppm)[μ3-Ph2PCHP(C6H4)Ph] (2), Os3(CO)6[μ3-Ph2PCH2P(C6H4)Ph]2 (3), and Os3(μ-H)(CO)6[μ3-PhPCH2P(C6H4)Ph][μ3-PhPCH(C6H4)PPh] (4), contain two dppm ligands per Os3 unit, while the fourth, Os3(μ-H)(CO)5(dppm)[μ3-PhPCH2P(C6H4)Ph][μ3-PhPCH(C6H4)PPh] (5), is the first example of an Os3 cluster containing three dppm ligands. Microwave heating was also used to prepare the known complex Os3(μ-dppm)2(CO)8 (1) more efficiently than previously reported. The new complexes 2–5 have been characterized by IR, NMR, mass spectrometry, and X-ray crystallography. In addition, the bonding in these complexes has been examined by electronic structure calculations. All of the new complexes contain at least one dppm ligand that has undergone C H and/or C P bond activation. Complex 2 is a 48e cluster that contains one intact dppm ligand and one face-capping dppm ligand that coordinates all three osmium sites through the phosphine moieties and cyclometalated and ortho-metalated carbon atoms. Complex 3 is a 48e cluster with two ortho-metalated dppm ligands while complexes 4 and 5 are 50e clusters that possess a common metallic framework with one cyclometalated dppm ligand and one ortho-metalated dppm ligand.
- Published
- 2016
43. Thermal transformations of tris(2-thienyl)phosphine (PTh3) at low-valent ruthenium cluster centers: Part I. Carbon–hydrogen, carbon–phosphorus and carbon–sulfur bond activation yielding Ru3(CO)8L{μ-Th2P(C4H2S)}(μ-H) (L = CO, PTh3), Ru3(CO)7(μ-PTh2)2(μ3-η2-C4H2S), Ru4(CO)9(μ-CO)2(μ4-η2-C4H2S)(μ4-PTh) and Ru5(CO)11(μ-PTh2)(μ4-η4-C4H3)(μ4-S)
- Author
-
Ebbe Nordlander, Jagodish C. Sarker, Noorjahan Begum, Shishir Ghosh, Derek A. Tocher, Michael G. Richmond, Graeme Hogarth, Miaz Uddin, and Shariff E. Kabir
- Subjects
010405 organic chemistry ,Ligand ,Organic Chemistry ,chemistry.chemical_element ,Nuclear magnetic resonance spectroscopy ,010402 general chemistry ,Photochemistry ,01 natural sciences ,Biochemistry ,Medicinal chemistry ,0104 chemical sciences ,Ruthenium ,Inorganic Chemistry ,chemistry.chemical_compound ,chemistry ,Phosphinidene ,Materials Chemistry ,Thiophene ,Physical and Theoretical Chemistry ,Benzene ,Phosphine ,Bond cleavage - Abstract
Reaction of Ru 3 (CO) 12 with tris(2-thienyl)phosphine (PTh 3 ) in CH 2 Cl 2 at room temperature or in THF in the presence of a catalytic amount of Na[Ph 2 CO] furnishes the carbonyl substitution products Ru 3 (CO) 11 (PTh 3 ) ( 1 ), Ru 3 (CO) 10 (PTh 3 ) 2 ( 2 ), and Ru 3 (CO) 9 (PTh 3 ) 3 ( 3 ). Heating 1 in toluene affords the cyclometalated cluster Ru 3 (CO) 9 {μ-Th 2 P(C 4 H 2 S)}(μ-H) ( 4 ) resulting from carbonyl loss and carbon–hydrogen bond activation, and both 4 and the substituted derivative Ru 3 (CO) 8 {μ-Th 2 P(C 4 H 2 S)}(PTh 3 )(μ-H) ( 5 ) resulted from the direct reaction of Ru 3 (CO) 12 and PTh 3 at 110 °C in toluene. Interestingly, thermolysis of 2 in benzene at 80 °C affords 5 together with phosphido-bridged Ru 3 (CO) 7 (μ-PTh 2 ) 2 (μ 3 -η 2 -C 4 H 2 S) ( 6 ) resulting from both phosphorus–carbon and carbon–hydrogen bond activation of coordinated PTh 3 ligand(s). Cluster 6 is the only product of the thermolysis of 2 in toluene. Heating cyclometalated 4 with Ru 3 (CO) 12 in toluene at 110 °C yielded the tetranuclear phosphinidine cluster, Ru 4 (CO) 9 (μ-CO) 2 (μ 4 -η 2 -C 4 H 2 S)(μ 4 -PTh) ( 7 ), resulting from carbon–phosphorus bond scission, together with the pentaruthenium sulfide cluster, Ru 5 (CO) 11 (μ-PTh 2 )(μ 4 -η 4 -C 4 H 3 )(μ 4 -S) ( 8 ), in which sulfur is extruded from a thiophene ring. All the new compounds were characterized by elemental analysis, mass spectrometry, IR and NMR spectroscopy, and by single crystal X-ray diffraction analysis in case of clusters 4 , 6 , 7 , and 8 . Cluster 4 consists of a triangular ruthenium framework containing a μ 3 -Th 2 P(C 4 H 2 S) ligand, while 6 consists of a ruthenium triangle containing η 2 -μ 3 -thiophyne ligand and two edge-bridging PTh 2 ligands. Cluster 7 exhibits a distorted square arrangement of ruthenium atoms that are capped on one side by a μ 4 -phosphinidene ligand and on the other by a 4e donating μ 4 -η 2 -C 4 H 2 S ligand. The structure of 8 represents a rare example of a pentaruthenium wing-tip bridged-butterfly skeleton capped by μ 4 -S and μ 4 -η 4 -C 4 H 3 ligands. The compounds 4 , 6 , 7 , and 8 have been examined by density functional theory (DFT), and the lowest energy structure computed coincides with the X-ray diffraction structure. The hemilabile nature of the activated thienyl ligand in 4 and 6 has also been computationally investigated.
- Published
- 2016
44. Syntheses and Characterization of Tantalum Alkyl Imides and Amide Imides. DFT Studies of Unusual α-SiMe3 Abstraction by an Amide Ligand
- Author
-
Carlos A. Steren, Seth C. Hunter, Zi-Ling Xue, Michael G. Richmond, and Shu-Jian Chen
- Subjects
chemistry.chemical_classification ,Ligand ,Stereochemistry ,Dimer ,Organic Chemistry ,Medicinal chemistry ,Inorganic Chemistry ,chemistry.chemical_compound ,Reaction rate constant ,chemistry ,Reagent ,Amide ,Physical and Theoretical Chemistry ,Imide ,Cis–trans isomerism ,Alkyl - Abstract
Reaction of TaCl2(═NSiMe3)[N(SiMe3)2] (1) with alkylating reagents form the alkyl amide imide complexes TaR2(═NSiMe3)[N(SiMe3)2] (R = Me (2), CH2Ph (3)) and mixed amide imide compounds Ta(NR′2)2(═NSiMe3)[N(SiMe3)2] (R′ = Me (4), Et (5)). The reaction of 2 and 0.5 equiv of O2 leads to preferential oxygen insertion into one Ta–Me bond, yielding the alkoxy-bridged alkyl dimer Ta2(μ-OMe)2Me2(═NSiMe3)2[N(SiMe3)2]2 (6) as cis and trans isomers. Crystallization of the cis-6 and trans-6 mixture gave only crystals of trans-6. When the crystals of trans-6 were dissolved in benzene-d6, conversion of trans-6 to cis-6 occurred until the trans-6 ⇌ cis-6 equilibrium was reached with Keq = 0.79(0.02) at 25.0(0.1) °C. Kinetic studies of the exchange gave the rate constants k = 0.018(0.002) min–1 for the trans-6 → cis-6 conversion and k′ = 0.022(0.002) min–1 for the reverse cis-6 → trans-6 conversion at 25.0(0.1) °C. Complex 6 reacts with additional O2, forming the dialkoxy dimer Ta2(μ-OMe)2(OMe)2(═NSiMe3)2[N(SiMe3)2]2 (7)...
- Published
- 2015
45. Thermolysis of [HOs3(CO)8{µ3-Ph2PCH2P(Ph)C6H4}]: New Os2- and Os3- cluster products based on multiple C H bond activation of the bis(diphenylphosphino)methane ligand
- Author
-
Nikhil Chandra Bhoumik, Shishir Ghosh, Tuhinur R. Joy, Michael G. Richmond, and Shariff E. Kabir
- Subjects
010405 organic chemistry ,Ligand ,Metalation ,Aryl ,chemistry.chemical_element ,010402 general chemistry ,01 natural sciences ,Oxidative addition ,Medicinal chemistry ,0104 chemical sciences ,Inorganic Chemistry ,chemistry.chemical_compound ,chemistry ,Materials Chemistry ,Molecule ,Osmium ,Osmium Compounds ,Physical and Theoretical Chemistry ,Methylene - Abstract
Thermolysis of [HOs3(CO)8{µ3-Ph2PCH2P(Ph)C6H4}] in refluxing xylene has been investigated, and three new polynuclear osmium compounds containing an activated dppm ligand(s) as a result of multiple C H bond cleavages have been isolated. The new complexes are [H2Os3(CO)8{µ3-Ph2PCHP(Ph)C6H4}] (1), [H2Os3(CO)6(µ3-Ph2PCHPPh2){µ3-Ph2PCH2P(Ph)C6H4}] (2) and [Os2(CO)6{µ-H4C6(Ph)PCH2P(Ph)C6H4}] (3), each of which exhibits a different coordination mode involving the activated diphosphine ligand. Cluster 1 is formed by C H bond oxidative addition of the backbone methylene spacer and an ortho site at an ancillary aryl ring. Control experiments confirm that cluster 2 is formed from cluster 1 and dppm. The dinuclear compound 3 is formed via metalation of two phenyl rings on the dppm ligand and is accompanied by the release of an osmium atom. The molecular structure of each new complex has been established by single-crystal X-ray diffraction analysis, and the bonding in compounds 1–3 has been examined by DFT calculations.
- Published
- 2020
46. Facile Os-Os bond cleavage in the reactions of [Os3(CO)10(NCMe)2] and [Os3(CO)10(μ-H)2] with tetramethylthiuram disulfide (tmtd): Syntheses and crystal structures of new polynuclear osmium carbonyl complexes containing a dimethyldithiocarbamate ligand(s)
- Author
-
Shariff E. Kabir, Tapan Kumar Saha, Michael G. Richmond, Nikhil Chandra Bhoumik, Vladimir N. Nesterov, and Shishir Ghosh
- Subjects
010405 organic chemistry ,Chemistry ,Ligand ,Organic Chemistry ,Center (category theory) ,chemistry.chemical_element ,Crystal structure ,010402 general chemistry ,01 natural sciences ,Biochemistry ,Medicinal chemistry ,0104 chemical sciences ,Inorganic Chemistry ,Yield (chemistry) ,Materials Chemistry ,Cluster (physics) ,Chelation ,Osmium ,Physical and Theoretical Chemistry ,Bond cleavage - Abstract
The reaction of the labile triosmium cluster [Os3(CO)10(NCMe)2] with tetramethylthiuram disulfide [(Me2NCS2)2, tmtd] has been investigated at room temperature. Conducting the reaction under mild conditions has permitted the isolation of the three new polynuclear osmium complexes [Os3(CO)10{κ2(S,S)-S2CNMe2}2] (1), [Os4(CO)12{κ2(S,S)-S2CNMe2}2{μ3-η1(C),κ2(O,O)-CO2)}(μ3-S)] (2) and [Os3(CO)9{μ3-η1(C),κ1(S)-SCNMe2}{μ-κ1(S)-SC(O)NMe2}] (3) in 20, 16 and 10% yield, respectively, in addition to the known mononuclear complex [Os(CO)2{κ2(S,S)-S2CNMe2}2] (4) in 10% yield. The triosmium complex 1 has a linear arrangement of osmium atoms with each terminal osmium center containing a chelating Me2NCS2 ligand. The tetraosmium complex 2 also possesses a μ3-CO2 ligand in addition to capping sulfido and chelating Me2NCS2 ligands. Complex 3 contains an open triosmium core with the open Os···Os edge bridged by a Me2NC(O)S ligand and a capping Me2NCS ligand. A similar reaction between the unsaturated cluster [Os3(CO)10(μ-H)2] and [(Me2NCS2)2] yielded the triosmium complexes [Os3(CO)10{μ-κ1(S)-S2CNMe2}(μ-H)] (5) and [Os3(CO)9{μ3-κ2(S,S)-S2CNMe2}(μ-H)] (6) in addition to compounds 1 and 4 and the known hydroxyl cluster [Os3(CO)10(μ-OH)(μ-H)] (7) in 7, 14, 14, 10, and 6% yield, respectively. Both 5 and 6 possess a triosmium core, and the Me2NCS2 ligand acts in edge-bridging capacity in the former while it serves as a face-capping ligand in the latter. All new complexes have been characterized by combustion analyses and IR and NMR spectroscopies, and the solid-state structures of 1–3, and 6 have been established by X-ray crystallography. The bonding in clusters 1-3 and 6 has been examined by electronic structure calculations, and the thermodynamics for the formation of 1 and MeCN (two equiv) from [Os3(CO)10(MeCN)2] and tmtd, relative to other chelated and bridged isomers of [Os3(CO)10(tmtd)] are discussed.
- Published
- 2020
47. Reactions of [Ru3(CO)12] with thiosaccharin: Synthesis and structure of di-, tri-, tetra- and penta-ruthenium complexes containing a thiosaccharinate ligand(s)
- Author
-
Shariff E. Kabir, Shishir Ghosh, Vladimir N. Nesterov, Md. Jadu Mia, Michael G. Richmond, Nikhil Chandra Bhoumik, and Md. Selim Reza
- Subjects
biology ,010405 organic chemistry ,Chemistry ,Ligand ,Organic Chemistry ,chemistry.chemical_element ,010402 general chemistry ,biology.organism_classification ,01 natural sciences ,Biochemistry ,Medicinal chemistry ,0104 chemical sciences ,Ruthenium ,Inorganic Chemistry ,Materials Chemistry ,Tetra ,Physical and Theoretical Chemistry - Abstract
Reactions of [Ru3(CO)12] with thiosaccharin (tsacH) at different temperatures have been investigated. At 40 °C, the diruthenium complex [Ru2(CO)6(μ-N,S-tsac)2] (1) is produced and whose ruthenium atoms are bridged by two tsac ligands that are oriented in a head-to-tail fashion. When this reaction is carried out at 66 °C, the tri-, tetra- and penta-ruthenium complexes [H2Ru3(CO)7(μ-N,S-tsac)(μ-C,N–C6H4CNSO2)(μ3-S)] (2), [Ru4(CO)12(μ-N,S-tsac)2(μ4-S)] (3) and [H2Ru5(CO)13(μ-N,S-tsac)(μ-C,N–C6H4CNSO2)(μ3-S)(μ4-S)] (4), respectively, are also isolated in addition to 1. The triruthenium complex 2 exhibits an arachno SRu3 polyhedron containing edge-bridging tsac and C6H4CNSO2 ligands. The tetraruthenium complex 3 consists of two [Ru2(CO)6(μ-N,S-tsac)] fragments linked via a μ4-S ligand, while the pentaruthenium complex 4 is composed of individual Ru3 and Ru2 units linked via a μ4-S ligand. At 81 °C, the same reaction furnishes the pentaruthenium complex [HRu5(CO)15(μ-N,S-tsac)(μ5-S)] (5) containing tsac and μ5-S bridging ligands. The molecular structures of the new complexes have been determined by single-crystal X-ray diffraction analyses, and the bonding in each product has been examined by DFT.
- Published
- 2020
48. Hydrogenase biomimetics with redox-active ligands:Synthesis, structure, and electrocatalytic studies on [Fe2(CO)4(κ2-dppn)(μ-edt)] (edt = ethanedithiolate; dppn = 1,8-bis(diphenylphosphino)naphthalene)
- Author
-
Shariff E. Kabir, Graeme Hogarth, Shishir Ghosh, Michael G. Richmond, Nathan Hollingsworth, and Shahed Rana
- Subjects
Dppn ,Hydrogenase ,hydrogenase biomimetics ,dithiolate ,Protonation ,010402 general chemistry ,01 natural sciences ,Medicinal chemistry ,Redox ,Catalysis ,Inorganic Chemistry ,dppn ,lcsh:Inorganic chemistry ,HOMO/LUMO ,Bond cleavage ,proton-reduction ,010405 organic chemistry ,Chemistry ,Redox-active ,Antibonding molecular orbital ,lcsh:QD146-197 ,0104 chemical sciences ,Hydrogenase biomimetics ,Dithiolate ,Cyclic voltammetry ,redox-active ,Proton-reduction - Abstract
Addition of the bulky redox-active diphosphine 1,8-bis(diphenylphosphino)naphthalene (dppn) to [Fe2(CO)6(µ-edt)] (1) (edt = 1,2-ethanedithiolate) affords [Fe2(CO)4(κ2-dppn)(µ-edt)] (3) as the major product, together with small amounts of a P⁻C bond cleavage product [Fe2(CO)5{κ1-PPh2(1-C10H7)}(µ-edt)] (2). The redox properties of 3 have been examined by cyclic voltammetry and it has been tested as a proton-reduction catalyst. It undergoes a reversible reduction at E1/2 = −2.18 V and exhibits two overlapping reversible oxidations at E1/2 = −0.08 V and E1/2 = 0.04 V. DFT calculations show that while the Highest Occupied Molecular Orbital (HOMO) is metal-centred (Fe⁻Fe σ-bonding), the Lowest Unoccupied Molecular Orbital (LUMO) is primarily ligand-based, but also contains an antibonding Fe⁻Fe contribution, highlighting the redox-active nature of the diphosphine. It is readily protonated upon addition of strong acids and catalyzes the electrochemical reduction of protons at Ep = −2.00 V in the presence of CF3CO2H. The catalytic current indicates that it is one of the most efficient diiron electrocatalysts for the reduction of protons, albeit operating at quite a negative potential.
- Published
- 2018
49. Diphosphine-induced thiolate-bridge scission of [Re(CO)3(μ,κ2-S,N-thpymS)]2 (thpymS = 1,4,5,6-tetrahydropyrimidine-2-thiolate):Structural and computational studies of configurational isomers of [Re(CO)3(κ2-S,N-thpymS)]2(μ,κ1,κ1-dppe)
- Author
-
Graeme Hogarth, Md. Rassel Moni, Shishir Ghosh, Derek A. Tocher, Shaikh M. Mobin, Michael G. Richmond, and Shariff E. Kabir
- Subjects
Denticity ,Diphosphines ,010405 organic chemistry ,Ligand ,Newman projection ,Organic Chemistry ,chemistry.chemical_element ,Stereoisomerism ,Rhenium ,010402 general chemistry ,01 natural sciences ,Biochemistry ,DFT ,0104 chemical sciences ,Inorganic Chemistry ,Crystallography ,chemistry.chemical_compound ,chemistry ,Materials Chemistry ,Physical and Theoretical Chemistry ,Tetrahydropyrimidine-2-thiolate ,Phosphine ,Bond cleavage - Abstract
Reactions of binuclear [Re(CO)3(μ,κ2-S,N-thpymS)]2 (1) with diphosphines have been investigated. At 298 K, dppm reacts to give mononuclear [Re(CO)3(κ1-dppm)(κ2-S,N-thpymS)] (2) through a phosphine-promoted scission of the dithiolate bridges that leaves one of the phosphine moieties free (dangling). Refluxing 2 in toluene leads to CO loss and formation of dinuclear [Re2(CO)4(μ-dppm)(μ,κ2-S,N-thpymS)2] (3) whose rhenium centers are bridged by two thiolate groups and the dppm ligand. Treatment of 1 with dppe at room temperature furnishes [Re(CO)3(κ2-S,N-thpymS)]2(μ,κ1,κ1-dppe) (4) where each phosphine center ligates the respective d6-ML5 rhenium fragment. Complex 4 exists as two distinct configurational isomers (4a and 4b) that have been isolated and the solid-state structures characterized crystallographically. The principal difference in the stereoisomeric products is the orientation of the two [Re(CO)3(κ2-S,N-thpymS)] moieties at the anti-staggered Newman projection involving the P-C-C-P backbone of the dppe ligand. Both stereoisomers retain their identity in solution at ambient temperatures but equilibrate to a 1:1 mixture upon heating at 363 K for 1 h. The reaction of 1 with dppe in toluene at 383 K affords [Re(CO)2(κ1-dppe)2(κ2-S,N-thpymS)] (5) containing two monodentate (dangling) diphosphine ligands. Thus, these seemingly simple reactions afford a range of different products whose composition is highly dependent upon the experimental conditions employed and the nature of the diphosphine backbone. The reaction of 1 with dppe and the process responsible for the equilibration of the two configurational isomers of 4 have been investigated by electronic structure calculations.
- Published
- 2018
50. Experimental and computational preference for phosphine regioselectivity and stereoselective tripodal rotation in HOs
- Author
-
Shahin A, Begum, Md Arshad H, Chowdhury, Shishir, Ghosh, Derek A, Tocher, Michael G, Richmond, Li, Yang, K I, Hardcastle, Edward, Rosenberg, and Shariff E, Kabir
- Abstract
The site preference for ligand substitution in the benzothiazolate-bridged cluster HOs
- Published
- 2018
Catalog
Discovery Service for Jio Institute Digital Library
For full access to our library's resources, please sign in.