1
|
Balcells D, Clot E, Eisenstein O. C—H Bond Activation in Transition Metal Species from a Computational Perspective. Chem Rev 2010; 110:749-823. [PMID: 20067255 DOI: 10.1021/cr900315k] [Citation(s) in RCA: 861] [Impact Index Per Article: 57.4] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/11/2022]
|
|
15 |
861 |
2
|
Hull JF, Balcells D, Blakemore JD, Incarvito CD, Eisenstein O, Brudvig GW, Crabtree RH. Highly active and robust Cp* iridium complexes for catalytic water oxidation. J Am Chem Soc 2009; 131:8730-1. [PMID: 19496565 DOI: 10.1021/ja901270f] [Citation(s) in RCA: 472] [Impact Index Per Article: 29.5] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/29/2022]
Abstract
A series of Cp*Ir catalysts are the most active known by over an order of magnitude for water oxidation with Ce(IV). DFT calculations support a Cp*Ir=O complex as an active species.
Collapse
|
Research Support, U.S. Gov't, Non-P.H.S. |
16 |
472 |
3
|
Blakemore JD, Schley ND, Balcells D, Hull JF, Olack GW, Incarvito CD, Eisenstein O, Brudvig GW, Crabtree RH. Half-sandwich iridium complexes for homogeneous water-oxidation catalysis. J Am Chem Soc 2010; 132:16017-29. [PMID: 20964386 DOI: 10.1021/ja104775j] [Citation(s) in RCA: 417] [Impact Index Per Article: 27.8] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/12/2022]
Abstract
Iridium half-sandwich complexes of the types Cp*Ir(N-C)X, [Cp*Ir(N-N)X]X, and [CpIr(N-N)X]X are catalyst precursors for the homogeneous oxidation of water to dioxygen. Kinetic studies with cerium(IV) ammonium nitrate as primary oxidant show that oxygen evolution is rapid and continues over many hours. In addition, [Cp*Ir(H(2)O)(3)]SO(4) and [(Cp*Ir)(2)(μ-OH)(3)]OH can show even higher turnover frequencies (up to 20 min(-1) at pH 0.89). Aqueous electrochemical studies on the cationic complexes having chelate ligands show catalytic oxidation at pH > 7; conversely, at low pH, there are no oxidation waves up to 1.5 V vs NHE for the complexes. H(2)(18)O isotope incorporation studies demonstrate that water is the source of oxygen atoms during cerium(IV)-driven catalysis. DFT calculations and kinetic experiments, including kinetic-isotope-effect studies, suggest a mechanism for homogeneous iridium-catalyzed water oxidation and contribute to the determination of the rate-determining step. The kinetic experiments also help distinguish the active homogeneous catalyst from heterogeneous nanoparticulate iridium dioxide.
Collapse
|
Research Support, U.S. Gov't, Non-P.H.S. |
15 |
417 |
4
|
Balcells D, Nova A, Clot E, Gnanamgari D, Crabtree RH, Eisenstein O. Mechanism of Homogeneous Iridium-Catalyzed Alkylation of Amines with Alcohols from a DFT Study. Organometallics 2008. [DOI: 10.1021/om800134d] [Citation(s) in RCA: 140] [Impact Index Per Article: 8.2] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/30/2022]
|
|
17 |
140 |
5
|
Beromi MM, Nova A, Balcells D, Brasacchio AM, Brudvig GW, Guard LM, Hazari N, Vinyard DJ. Mechanistic Study of an Improved Ni Precatalyst for Suzuki-Miyaura Reactions of Aryl Sulfamates: Understanding the Role of Ni(I) Species. J Am Chem Soc 2017; 139:922-936. [PMID: 28009513 PMCID: PMC5360380 DOI: 10.1021/jacs.6b11412] [Citation(s) in RCA: 119] [Impact Index Per Article: 14.9] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/21/2022]
Abstract
Nickel precatalysts are potentially a more sustainable alternative to traditional palladium precatalysts for the Suzuki-Miyaura coupling reaction. Currently, there is significant interest in Suzuki-Miyaura coupling reactions involving readily accessible phenolic derivatives such as aryl sulfamates, as the sulfamate moiety can act as a directing group for the prefunctionalization of the aromatic backbone of the electrophile prior to cross-coupling. By evaluating complexes in the Ni(0), (I), and (II) oxidation states we report a precatalyst, (dppf)Ni(o-tolyl)(Cl) (dppf = 1,1'-bis(diphenylphosphino)ferrocene), for Suzuki-Miyaura coupling reactions involving aryl sulfamates and boronic acids, which operates at a significantly lower catalyst loading and at milder reaction conditions than other reported systems. In some cases it can even function at room temperature. Mechanistic studies on precatalyst activation and the speciation of nickel during catalysis reveal that Ni(I) species are formed in the catalytic reaction via two different pathways: (i) the precatalyst (dppf)Ni(o-tolyl)(Cl) undergoes comproportionation with the active Ni(0) species; and (ii) the catalytic intermediate (dppf)Ni(Ar)(sulfamate) (Ar = aryl) undergoes comproportionation with the active Ni(0) species. In both cases the formation of Ni(I) is detrimental to catalysis, which is proposed to proceed via a Ni(0)/Ni(II) cycle. DFT calculations are used to support experimental observations and provide insight about the elementary steps involved in reactions directly on the catalytic cycle, as well as off-cycle processes. Our mechanistic investigation provides guidelines for designing even more active nickel catalysts.
Collapse
|
Research Support, N.I.H., Extramural |
8 |
119 |
6
|
Melvin PR, Nova A, Balcells D, Dai W, Hazari N, Hruszkewycz DP, Shah HP, Tudge MT. Design of a Versatile and Improved Precatalyst Scaffold for Palladium-Catalyzed Cross-Coupling: (η3-1-tBu-indenyl)2(μ-Cl)2Pd2. ACS Catal 2015. [DOI: 10.1021/acscatal.5b00878] [Citation(s) in RCA: 107] [Impact Index Per Article: 10.7] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/30/2022]
|
|
10 |
107 |
7
|
Hruszkewycz DP, Balcells D, Guard LM, Hazari N, Tilset M. Insight into the efficiency of cinnamyl-supported precatalysts for the Suzuki-Miyaura reaction: observation of Pd(I) dimers with bridging allyl ligands during catalysis. J Am Chem Soc 2014; 136:7300-16. [PMID: 24824779 DOI: 10.1021/ja412565c] [Citation(s) in RCA: 102] [Impact Index Per Article: 9.3] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/19/2022]
Abstract
Despite widespread use of complexes of the type Pd(L)(η(3)-allyl)Cl as precatalysts for cross-coupling, the chemistry of related Pd(I) dimers of the form (μ-allyl)(μ-Cl)Pd2(L)2 has been underexplored. Here, the relationship between the monomeric and the dimeric compounds is investigated using both experiment and theory. We report an efficient synthesis of the Pd(I) dimers (μ-allyl)(μ-Cl)Pd2(IPr)2 (allyl = allyl, crotyl, cinnamyl; IPr = 1,3-bis(2,6-diisopropylphenyl)imidazol-2-ylidene) through activation of Pd(IPr)(η(3)-allyl)Cl type monomers under mildly basic reaction conditions. The catalytic performance of the Pd(II) monomers and their Pd(I) μ-allyl dimer congeners for the Suzuki-Miyaura reaction is compared. We propose that the (μ-allyl)(μ-Cl)Pd2(IPr)2-type dimers are activated for catalysis through disproportionation to Pd(IPr)(η(3)-allyl)Cl and monoligated IPr-Pd(0). The microscopic reverse comproportionation reaction of monomers of the type Pd(IPr)(η(3)-allyl)Cl with IPr-Pd(0) to form Pd(I) dimers is also studied. It is demonstrated that this is a facile process, and Pd(I) dimers are directly observed during catalysis in reactions using Pd(II) precatalysts. In these catalytic reactions, Pd(I) μ-allyl dimer formation is a deleterious process which removes the IPr-Pd(0) active species from the reaction mixture. However, increased sterics at the 1-position of the allyl ligand in the Pd(IPr)(η(3)-crotyl)Cl and Pd(IPr)(η(3)-cinnamyl)Cl precatalysts results in a larger kinetic barrier to comproportionation, which allows more of the active IPr-Pd(0) catalyst to enter the catalytic cycle when these substituted precatalysts are used. Furthermore, we have developed reaction conditions for the Suzuki-Miyaura reaction using Pd(IPr)(η(3)-cinnamyl)Cl which are compatible with mild bases.
Collapse
|
Journal Article |
11 |
102 |
8
|
Nova A, Balcells D, Schley ND, Dobereiner GE, Crabtree RH, Eisenstein O. An Experimental−Theoretical Study of the Factors That Affect the Switch between Ruthenium-Catalyzed Dehydrogenative Amide Formation versus Amine Alkylation. Organometallics 2010. [DOI: 10.1021/om101015u] [Citation(s) in RCA: 98] [Impact Index Per Article: 6.5] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/29/2022]
|
|
15 |
98 |
9
|
|
|
18 |
95 |
10
|
Hull JF, Balcells D, Sauer ELO, Raynaud C, Brudvig GW, Crabtree RH, Eisenstein O. Manganese catalysts for C-H activation: an experimental/theoretical study identifies the stereoelectronic factor that controls the switch between hydroxylation and desaturation pathways. J Am Chem Soc 2010; 132:7605-16. [PMID: 20481432 DOI: 10.1021/ja908744w] [Citation(s) in RCA: 93] [Impact Index Per Article: 6.2] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/29/2022]
Abstract
We describe competitive C-H bond activation chemistry of two types, desaturation and hydroxylation, using synthetic manganese catalysts with several substrates. 9,10-Dihydrophenanthrene (DHP) gives the highest desaturation activity, the final products being phenanthrene (P1) and phenanthrene 9,10-oxide (P3), the latter being thought to arise from epoxidation of some of the phenanthrene. The hydroxylase pathway also occurs as suggested by the presence of the dione product, phenanthrene-9,10-dione (P2), thought to arise from further oxidation of hydroxylation intermediate 9-hydroxy-9,10-dihydrophenanthrene. The experimental work together with the density functional theory (DFT) calculations shows that the postulated Mn oxo active species, [Mn(O)(tpp)(Cl)] (tpp = tetraphenylporphyrin), can promote the oxidation of dihydrophenanthrene by either desaturation or hydroxylation pathways. The calculations show that these two competing reactions have a common initial step, radical H abstraction from one of the DHP sp(3) C-H bonds. The resulting Mn hydroxo intermediate is capable of promoting not only OH rebound (hydroxylation) but also a second H abstraction adjacent to the first (desaturation). Like the active Mn(V)=O species, this Mn(IV)-OH species also has radical character on oxygen and can thus give H abstraction. Both steps have very low and therefore very similar energy barriers, leading to a product mixture. Since the radical character of the catalyst is located on the oxygen p orbital perpendicular to the Mn(IV)-OH plane, the orientation of the organic radical with respect to this plane determines which reaction, desaturation or hydroxylation, will occur. Stereoelectronic factors such as the rotational orientation of the OH group in the enzyme active site are thus likely to constitute the switch between hydroxylase and desaturase behavior.
Collapse
|
Research Support, U.S. Gov't, Non-P.H.S. |
15 |
93 |
11
|
Balcells D, Raynaud C, Crabtree RH, Eisenstein O. A rational basis for the axial ligand effect in C-H oxidation by [MnO(porphyrin)(X)]+ (X = H2O, OH-, O2-) from a DFT study. Inorg Chem 2008; 47:10090-9. [PMID: 18788735 DOI: 10.1021/ic8013706] [Citation(s) in RCA: 83] [Impact Index Per Article: 4.9] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/28/2023]
Abstract
Oxyl radical character in the MnO group of the title system is shown from a density functional theory study to be essential for efficient C-H cleavage, which is a key step in C-H oxidation. Since oxyl species have elongated Mn-O bonds relative to the more usual oxo species of type MnO, the normal expectation would be that high trans-influence ligands X should facilitate oxyl character by elongating the Mn-O bond and thus enhance both oxyl character and reactivity. Contrary to this expectation, but in line with the experimental data (Jin, N.; Ibrahim, M.; Spiro, T. G.; Groves, J. T. J. Am. Chem. Soc. 2007, 129, 12416), we find that reactivity increases along the series X = O(2-) < OH(-) < H2O for the following reasons. The ground-state singlet (S) is unreactive for all X, and only the higher-energy triplet (T) and quintet (Q) states have the oxyl character needed for reactivity, but the higher trans-influence X ligands are also shown to increase the S/T and S/Q gaps, thus making attainment of the needed T and Q states harder. The latter effect is dominant, and high trans-influence X ligands thus disfavor reaction. The higher reactivity in the presence of acid noted by Groves and co-workers is thus rationalized by the preference for having X = H2O over OH(-) or O(2-).
Collapse
|
Journal Article |
17 |
83 |
12
|
Zhou M, Balcells D, Parent AR, Crabtree RH, Eisenstein O. Cp* Iridium Precatalysts for Selective C–H Oxidation via Direct Oxygen Insertion: A Joint Experimental/Computational Study. ACS Catal 2012. [DOI: 10.1021/cs2005899] [Citation(s) in RCA: 79] [Impact Index Per Article: 6.1] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/04/2023]
|
|
13 |
79 |
13
|
Melvin PR, Balcells D, Hazari N, Nova A. Understanding Precatalyst Activation in Cross-Coupling Reactions: Alcohol Facilitated Reduction from Pd(II) to Pd(0) in Precatalysts of the Type (η3-allyl)Pd(L)(Cl) and (η3-indenyl)Pd(L)(Cl). ACS Catal 2015. [DOI: 10.1021/acscatal.5b01291] [Citation(s) in RCA: 79] [Impact Index Per Article: 7.9] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/28/2022]
|
|
10 |
79 |
14
|
Friederich P, Dos Passos Gomes G, De Bin R, Aspuru-Guzik A, Balcells D. Machine learning dihydrogen activation in the chemical space surrounding Vaska's complex. Chem Sci 2020; 11:4584-4601. [PMID: 33224459 PMCID: PMC7659707 DOI: 10.1039/d0sc00445f] [Citation(s) in RCA: 78] [Impact Index Per Article: 15.6] [Reference Citation Analysis] [Abstract] [Track Full Text] [Download PDF] [Figures] [Journal Information] [Subscribe] [Scholar Register] [Received: 01/23/2020] [Accepted: 04/06/2020] [Indexed: 12/15/2022] Open
Abstract
Homogeneous catalysis using transition metal complexes is ubiquitously used for organic synthesis, as well as technologically relevant in applications such as water splitting and CO2 reduction. The key steps underlying homogeneous catalysis require a specific combination of electronic and steric effects from the ligands bound to the metal center. Finding the optimal combination of ligands is a challenging task due to the exceedingly large number of possibilities and the non-trivial ligand-ligand interactions. The classic example of Vaska's complex, trans-[Ir(PPh3)2(CO)(Cl)], illustrates this scenario. The ligands of this species activate iridium for the oxidative addition of hydrogen, yielding the dihydride cis-[Ir(H)2(PPh3)2(CO)(Cl)] complex. Despite the simplicity of this system, thousands of derivatives can be formulated for the activation of H2, with a limited number of ligands belonging to the same general categories found in the original complex. In this work, we show how DFT and machine learning (ML) methods can be combined to enable the prediction of reactivity within large chemical spaces containing thousands of complexes. In a space of 2574 species derived from Vaska's complex, data from DFT calculations are used to train and test ML models that predict the H2-activation barrier. In contrast to experiments and calculations requiring several days to be completed, the ML models were trained and used on a laptop on a time-scale of minutes. As a first approach, we combined Bayesian-optimized artificial neural networks (ANN) with features derived from autocorrelation and deltametric functions. The resulting ANNs achieved high accuracies, with mean absolute errors (MAE) between 1 and 2 kcal mol-1, depending on the size of the training set. By using a Gaussian process (GP) model trained with a set of selected features, including fingerprints, accuracy was further enhanced. Remarkably, this GP model minimized the MAE below 1 kcal mol-1, by using only 20% or less of the data available for training. The gradient boosting (GB) method was also used to assess the relevance of the features, which was used for both feature selection and model interpretation purposes. Features accounting for chemical composition, atom size and electronegativity were found to be the most determinant in the predictions. Further, the ligand fragments with the strongest influence on the H2-activation barrier were identified.
Collapse
|
research-article |
5 |
78 |
15
|
Conde A, Vilella L, Balcells D, Díaz-Requejo MM, Lledós A, Pérez PJ. Introducing Copper as Catalyst for Oxidative Alkane Dehydrogenation. J Am Chem Soc 2013; 135:3887-96. [DOI: 10.1021/ja310866k] [Citation(s) in RCA: 74] [Impact Index Per Article: 6.2] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/20/2023]
|
|
12 |
74 |
16
|
Balcells D, Maseras F, Ujaque G. Computational Rationalization of the Dependence of the Enantioselectivity on the Nature of the Catalyst in the Vanadium-Catalyzed Oxidation of Sulfides by Hydrogen Peroxide. J Am Chem Soc 2005; 127:3624-34. [PMID: 15755184 DOI: 10.1021/ja0435384] [Citation(s) in RCA: 65] [Impact Index Per Article: 3.3] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/29/2022]
Abstract
A computational study with the IMOMM(Becke3LYP:MM3) method is carried out on the mechanism of the enantioselective reaction of complex V(O)(L)(OOH), L= bulky tridentate Schiff base, and bis(tert-butyl) disulfide. The reaction with a given L ligand A is first systematically studied: different conformers of the catalyst are optimized, and the large number of associated transition states are systematically searched. The study is then extended to the geometry optimization of selected transition states associated to other ligands B, C, and D, similar to A but differing in the nature of certain substituents R1, R2, R3. The experimental trends in selectivity for catalysts based on ligands A to D are faithfully reproduced by the calculations. Analysis of the computational results leads finally to the formulation of a simple model that can explain one of the most remarkable aspect of this reaction, namely the large effect on enantioselectivity of ligands seemingly far from each other in the catalyst.
Collapse
|
|
20 |
65 |
17
|
Balcells D, Raynaud C, Crabtree RH, Eisenstein O. The rebound mechanism in catalytic C–H oxidation by MnO(tpp)Cl from DFT studies: electronic nature of the active species. Chem Commun (Camb) 2008:744-6. [DOI: 10.1039/b715939k] [Citation(s) in RCA: 64] [Impact Index Per Article: 3.8] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/31/2022]
|
|
17 |
64 |
18
|
Balcells D, Nova A. Designing Pd and Ni Catalysts for Cross-Coupling Reactions by Minimizing Off-Cycle Species. ACS Catal 2018. [DOI: 10.1021/acscatal.8b00230] [Citation(s) in RCA: 61] [Impact Index Per Article: 8.7] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/20/2022]
|
|
7 |
61 |
19
|
Wu J, Nova A, Balcells D, Brudvig GW, Dai W, Guard LM, Hazari N, Lin PH, Pokhrel R, Takase MK. Nickel(I) Monomers and Dimers with Cyclopentadienyl and Indenyl Ligands. Chemistry 2014; 20:5327-37. [DOI: 10.1002/chem.201305021] [Citation(s) in RCA: 60] [Impact Index Per Article: 5.5] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 12/24/2013] [Revised: 01/28/2014] [Indexed: 11/10/2022]
|
|
11 |
60 |
20
|
Langseth E, Scheuermann ML, Balcells D, Kaminsky W, Goldberg KI, Eisenstein O, Heyn RH, Tilset M. Generation and structural characterization of a gold(III) alkene complex. Angew Chem Int Ed Engl 2013; 52:1660-3. [PMID: 23283802 DOI: 10.1002/anie.201209140] [Citation(s) in RCA: 53] [Impact Index Per Article: 4.4] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 11/14/2012] [Indexed: 01/15/2023]
|
Journal Article |
12 |
53 |
21
|
Balcells D, Skjelstad BB. tmQM Dataset-Quantum Geometries and Properties of 86k Transition Metal Complexes. J Chem Inf Model 2020; 60:6135-6146. [PMID: 33166143 PMCID: PMC7768608 DOI: 10.1021/acs.jcim.0c01041] [Citation(s) in RCA: 51] [Impact Index Per Article: 10.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Download PDF] [Figures] [Journal Information] [Subscribe] [Scholar Register] [Received: 09/03/2020] [Indexed: 12/19/2022]
Abstract
We report the transition metal quantum mechanics (tmQM) data set, which contains the geometries and properties of a large transition metal-organic compound space. tmQM comprises 86,665 mononuclear complexes extracted from the Cambridge Structural Database, including Werner, bioinorganic, and organometallic complexes based on a large variety of organic ligands and 30 transition metals (the 3d, 4d, and 5d from groups 3 to 12). All complexes are closed-shell, with a formal charge in the range {+1, 0, -1}e. The tmQM data set provides the Cartesian coordinates of all metal complexes optimized at the GFN2-xTB level, and their molecular size, stoichiometry, and metal node degree. The quantum properties were computed at the DFT(TPSSh-D3BJ/def2-SVP) level and include the electronic and dispersion energies, highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) energies, HOMO/LUMO gap, dipole moment, and natural charge of the metal center; GFN2-xTB polarizabilities are also provided. Pairwise representations showed the low correlation between these properties, providing nearly continuous maps with unusual regions of the chemical space, for example, complexes combining large polarizabilities with wide HOMO/LUMO gaps and complexes combining low-energy HOMO orbitals with electron-rich metal centers. The tmQM data set can be exploited in the data-driven discovery of new metal complexes, including predictive models based on machine learning. These models may have a strong impact on the fields in which transition metal chemistry plays a key role, for example, catalysis, organic synthesis, and materials science. tmQM is an open data set that can be downloaded free of charge from https://github.com/bbskjelstad/tmqm.
Collapse
|
research-article |
5 |
51 |
22
|
Balcells D, Maseras F, Lledós A. Density functional study on the mechanism of the vanadium-catalyzed oxidation of sulfides by hydrogen peroxide. J Org Chem 2003; 68:4265-74. [PMID: 12762725 DOI: 10.1021/jo034052t] [Citation(s) in RCA: 50] [Impact Index Per Article: 2.3] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/30/2022]
Abstract
A computational study with the Becke3LYP method is carried out on the mechanism of the reaction of complexes V(O)(L)(OOH) and V(O)(LH)(OO) (L = O(CH)(3)N(CH(2))(2)O) with CH(3)S-SCH(3), a system that stands as a model for experimental systems where the metal complex contains larger chelating Schiff bases and the substrate is bis(tert-butyl) disulfide. The different possible isomers of both the hydroperoxo V(O)(L)(OOH) and the peroxo V(O)(LH)(OO) forms of the catalyst are explored, and the reactivity of the most stable among them with the dimethyl disulfide substrate is studied through location of the corresponding transition states. A large variety of reactive paths happen to exist, though in all cases the reaction takes place through a direct transfer process, with the simultaneous formation of the S-O bond and breaking of the O-O bond being the rate-limiting step.
Collapse
|
|
22 |
50 |
23
|
Balcells D, Raynaud C, Crabtree RH, Eisenstein O. C–H oxidation by hydroxo manganese(v) porphyrins: a DFT study. Chem Commun (Camb) 2009:1772-4. [DOI: 10.1039/b821029b] [Citation(s) in RCA: 43] [Impact Index Per Article: 2.7] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/21/2022]
|
|
16 |
43 |
24
|
Vilella L, Vidossich P, Balcells D, Lledós A. Basic ancillary ligands promote O–O bond formation in iridium-catalyzed water oxidation: A DFT study. Dalton Trans 2011; 40:11241-7. [DOI: 10.1039/c1dt10660k] [Citation(s) in RCA: 41] [Impact Index Per Article: 2.9] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/21/2022]
|
|
14 |
41 |
25
|
Artús Suàrez L, Culakova Z, Balcells D, Bernskoetter WH, Eisenstein O, Goldberg KI, Hazari N, Tilset M, Nova A. The Key Role of the Hemiaminal Intermediate in the Iron-Catalyzed Deaminative Hydrogenation of Amides. ACS Catal 2018. [DOI: 10.1021/acscatal.8b02184] [Citation(s) in RCA: 40] [Impact Index Per Article: 5.7] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/28/2022]
|
|
7 |
40 |