The Grignard Reaction – Unraveling a Chemical Puzzle
- Raphael Mathias PeltzerRaphael Mathias PeltzerDepartment of Chemistry and Hylleraas Centre for Quantum Molecular Sciences, University of Oslo, P.O. Box 1033 Blindern, Oslo 0315, NorwayMore by Raphael Mathias Peltzer
- ,
- Jürgen GaussJürgen GaussDepartment Chemie, Johannes Gutenberg-Universität Mainz, Duesbergweg 10-14, Mainz 55128, GermanyMore by Jürgen Gauss
- ,
- Odile Eisenstein*Odile Eisenstein*[email protected]Department of Chemistry and Hylleraas Centre for Quantum Molecular Sciences, University of Oslo, P.O. Box 1033 Blindern, Oslo 0315, NorwayICGM, Université de Montpellier, CNRS, ENSCM, Montpellier 34095 Cedex 5, FranceMore by Odile Eisenstein
- , and
- Michele Cascella*Michele Cascella*[email protected]Department of Chemistry and Hylleraas Centre for Quantum Molecular Sciences, University of Oslo, P.O. Box 1033 Blindern, Oslo 0315, NorwayMore by Michele Cascella
Abstract

More than 100 years since its discovery, the mechanism of the Grignard reaction remains unresolved. Ambiguities arise from the concomitant presence of multiple organomagnesium species and the competing mechanisms involving either nucleophilic addition or the formation of radical intermediates. To shed light on this topic, quantum-chemical calculations and ab initio molecular dynamics simulations are used to study the reaction of CH3MgCl in tetrahydrofuran with acetaldehyde and fluorenone as prototypical reagents. All organomagnesium species coexisting in solution due to the Schlenk equilibrium are found to be competent reagents for the nucleophilic pathway. The range of activation energies displayed by all of these compounds is relatively small. The most reactive species are a dinuclear Mg complex in which the substrate and the nucleophile initially bind to different Mg centers and the mononuclear dimethyl magnesium. The radical reaction, which requires the homolytic cleavage of the Mg–CH3 bond, cannot occur unless a substrate with a low-lying π*(CO) orbital coordinates the Mg center. This rationalizes why a radical mechanism is detected only in the presence of substrates with a low reduction potential. This feature, in turn, does not necessarily favor the nucleophilic addition, as shown for the reaction with fluorenone. The solvent needs to be considered as a reactant for both the nucleophilic and the radical reactions, and its dynamics is essential for representing the energy profile. The similar reactivity of several species in fast equilibrium implies that the reaction does not occur via a single process but by an ensemble of parallel reactions.
Introduction


Figure 1

Figure 1. Schematic representation of possible mechanisms for the Grignard reaction: (a) polar mechanism, heterolytic Mg–C bond breaking, with subsequent formation of a nucleophilic carbon that adds to the electrophilic carbonyl carbon; or (b) radical mechanism, homolytic Mg–C bond breaking, with subsequent recombination of the species with unpaired electrons.
Computational Methods
Ab Initio Molecular Dynamics Simulations
Simulation Protocol
Polar Mechanism – Activation Energies
Radical Formation Energy
Results and Discussion
Polar Mechanism
Figure 2

Figure 2. Organomagnesium complexes considered as reactants for the polar mechanism, classified as a function of the relative positions of the substrate (ACA) and nucleophile (methyl group). The labels geminal, vicinal, and bridging describe the initial position of the reactive groups with respect to the Mg center(s). The activation free energies, ΔA⧧, defined as the difference in free energy between the TS and the related reactant species, are given in kcal mol–1.
Reaction with Geminal and Bridging Groups
Structural Features of the Geminal Reaction
Figure 3

Figure 3. Geminal reaction for compound Bgem. Representative geometries for the reactant, transition state, and product. The Mg atom and its ligands are represented as balls-and-sticks. Other solvating THF molecules are drawn as lines. The color codes are mauve for magnesium, red for oxygen, gray for carbon, and white for hydrogen.
Figure 4

Figure 4. Transition state structures of the geminal and vicinal reactions (Bgem, Fvic). Left: Evidence of the structural similarity of the two TSs. Balls-and-sticks are used to represent the reacting moiety, while the rest of the system is drawn as transparent cylinders. THF molecules not bound to Mg are not shown. The green line marks the incipient C–CH3 bond and its associated distance. Top right: Schematic definition of the angles formed by the C3 axis of the nucleophile methyl group. Bottom right: Distances and angles formed by the atoms involved in the four-center TS. Distances are reported in angstroms, and angles are in degrees. ϕX is the angle at the vertex X in the four-member ring (top right panel). The color code is as in Figure 3.
Reaction with Vicinal Groups
Figure 5

Figure 5. Vicinal reaction for Fvic. Representative geometries for the reactant, transition state, and product. The Mg atoms (left Mg1, right Mg2) and first coordination sphere ligands are represented as balls-and-sticks. Other solvating THF molecules are drawn as lines. The color code is cyan for chlorine and green for the nucleophilic methyl, while the other elements are colored as in Figure 3.
Role of the Schlenk Equilibrium in the Nucleophilic Grignard Reaction
Figure 6

Figure 6. Red: Activation energies of all Grignard species derived from CH3MgCl in THF solution when reacting with acetaldehyde. Blue: Relative free energies of the species involved in the Schlenk equilibrium (data from ref (31)). The green box highlights the intrinsically most reactive species identified in this work.
Radical Mechanism
Figure 7

Figure 7. Mg–CH3 BDE and spin-density localization (in fractions of e) for MgCl(Sub)(THF)nCH3 (Sub = substrate). The green wire-frame shows the isosurface of the spin density map at a value of 0.0065 au.
Nucleophilic Attack versus Homolytic Cleavage: The Case of Fluorenone
Figure 8

Figure 8. Energetics of the Grignard reaction mechanisms for fluorenone. Top panel: Activation free energy of the nucleophilic addition with Mg(CH3)2. Bottom panel: Bond dissociation energy. All energy values are in kcal mol–1. On the right is the distribution of the spin density in Mg(CH3)(fluorenone)THF2: 33% of the unpaired electron localizes on the carbonyl carbon, 16% on the carbonyl oxygen, and 45% on the remaining π system.
Figure 9

Figure 9. Structural reorganization from (left) reactant to (right) transition state in the nucleophilic pathway for (top) ACA and (bottom) fluorenone. The CH3–Mg–O–C dihedral angle (corresponding to the blue arrow) is 14 ± 15° for ACA and 114 ± 30° for fluorenone.
Conclusions
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.9b11829.
Reference coupled-cluster calculations and benchmark BDE calculations, and structures of Mg(CH3)2, Mg(CH3)2(CH3CHO), and CH3Mg(OCH(CH3)2) in THF (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.
Acknowledgments
This work was supported by the Research Council of Norway (RCN) through the CoE Hylleraas Center for Quantum Molecular Sciences (Grant number 262695) and by the Norwegian Supercomputing Program (NOTUR) (Grant number NN4654K). M.C. and J.G. acknowledge funding by the Deutsche Forschungsgemeinschaft (DFG) within the project B5 of the TRR 146 (project number 233630050). We thank Mats Tilset for being at the right place at the right time to ask the right question. We also thank Filippo Lipparini for useful discussions.
References
This article references 65 other publications.
- 1Vollhardt, K.; Schore, N. Organic Chemistry: Structure and Function; W. H. Freeman: New York, 2014.Google ScholarThere is no corresponding record for this reference.
- 2Grignard, V. C. Sur quelques nouvelles combinaisons organométalliques du magnésium et leur application à des synthèses daalcools et d’hydrocarbures. Compt. Rend. Hebd. Séances Acad. Sci. 1900, 130, 1322– 24Google Scholar2https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaD28XitV2ltb0%253D&md5=0883715fab5d100533088de258b7c324About some new organo-metallic compounds of the magnesium and their application to the synthesis of alcohols and hydrocarbons. [machine translation]Grignard, V.Comptes Rendus Hebdomadaires des Seances de l'Academie des Sciences (1900), 130 (), 1322-24CODEN: COREAF; ISSN:0001-4036.[Machine Translation of Descriptors]. During effect of CH3J on magnesium powders in presence of anhydrous ether, an organo-metallic compound forms under a very lively reaction. The Mg dissolves gradually, and finally receives a clear liquid, which leaves a gray, crystalline, hygroscopic mass with the evaporation of the ether. One sets to that more ether solution which on 1 atom Mg of 1 mole CH3J contains 1 mole an aldehyde or a ketone, then takes place under a very lively reaction for formation of an organo-metallic compound, which supplies the appropriate secondary or tertiary alcohol with decomposing with acidified water with an yield of about 70%. The reaction takes place after the subsequent equations. 1. CH3I + Mg = CH3MgI. 2. CH3MgI + RCHO = R. CH (CH3) OMgI. 3. R. CH (CH3) OMgI + H2O = R. CH (OH) . CH3 + MgIOH. The bromine and iodine compounds of the saturated and unsaturated alkyls in the same way benzyl bromide give appropriate alcohol in same way. Author represented the subsequent alcohols in this way: Phenylisobutylcarbinol, C6H5. CH (OH): C4H9, from benzaldehyde, Isobutylbromide and Mg are colorless and weakly smelling liquid, Kp9, 122°. Dimethylphenylcarbinol, C6H5. C (OH) (CH3) 2, from acetophenone, Mg and CH3I, are pleasantly smelling and colorless liquid, Kp10, 93-95°. Dimethylbenzylcarbinol, C6H5. CH2. C (OH) (CH3) 2, from acetone, benzyl bromide and Mg are weakly smelling and colorless liquid, Kp10, 103-105°. With application of unsaturated aldehydes and ketones, yields with the double bond of the cobalt group and the developing alcohol is sometimes changeable. It splits off, then with the distillation H2O, and an unsaturated hydrocarbon develops. Thus the Dimethyl-2.4-pentadien-2.4 becomes, CH3. during effect of Mesityloxide on magnesium methyl iodide; C (CH3): CH. C (CH3): CH3, Kp750, 92-93°.
- 3Rappoport, Z., Marek, I., Eds. The Chemistry of Organomagnesium Compounds; Wiley-VCH: Weinheim, Germany, 2008.Google ScholarThere is no corresponding record for this reference.
- 4Schlenk, W.; Schlenk, W. Über die Konstitution der Grignardschen Magnesiumverbindungen. Ber. Dtsch. Chem. Ges. B 1929, 62, 920– 924, DOI: 10.1002/cber.19290620422Google Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaB1MXksFSitg%253D%253D&md5=30432803972ea7f1a470a7245fbf7813The constitution of the Grignard magnesium derivativesSchlenk, W.; Schlenk, Wilh., Jr.Berichte der Deutschen Chemischen Gesellschaft [Abteilung] B: Abhandlungen (1929), 62B (), 920-4CODEN: BDCBAD; ISSN:0365-9488.Several Grignard compds. have been prepd., then fractionally pptd. from their Et2O soln. by means of O(CH2CH2)2O. The Mg:X ratios of the fractions have been examd. Grignard compds. must be represented by 2RMgX .dblharw. MgR2 + MgX2. For EtI, the compn. of the Grignard deriv. would be: 6EtMgI + 4MgEt2 + 4MgI2. For PhBr: PhMgBr + 0.115MgPh2 + 0.115MgBr2.
- 5Seyferth, D. The Grignard Reagents. Organometallics 2009, 28, 1598– 1605, DOI: 10.1021/om900088zGoogle Scholar5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXivVeis70%253D&md5=e0c332b6494bdecec6bd4bcfce564ee0The Grignard ReagentsSeyferth, DietmarOrganometallics (2009), 28 (6), 1598-1605CODEN: ORGND7; ISSN:0276-7333. (American Chemical Society)A review of prepn. and reactions of Grignard reagents.
- 6Silverman, G. S.; Rakita, P. E. Handbook of Grignard Reagents; CRC Press: New York, 1996.Google ScholarThere is no corresponding record for this reference.
- 7Robertson, S. D.; Uzelac, M.; Mulvey, R. E. Alkali-metal-mediated synergistic effects in polar main group organometallic chemistry. Chem. Rev. 2019, 119, 8332– 8405, DOI: 10.1021/acs.chemrev.9b00047Google Scholar7https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXltVWrtro%253D&md5=f40eeff96aacc4152b568c0b79d6fa78Alkali-Metal-Mediated Synergistic Effects in Polar Main Group Organometallic ChemistryRobertson, Stuart D.; Uzelac, Marina; Mulvey, Robert E.Chemical Reviews (Washington, DC, United States) (2019), 119 (14), 8332-8405CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. The development of synthetic chem. since the early 1900s owes much to the service of organolithium reagents. Brilliant bases (e.g., deprotonating C-H bonds), nucleophiles (e.g., adding to unsatd. mols.), and transfer agents (e.g., delivering ligands to other metals), these versatile virtuosi and to a lesser extent the org. derivs. of the other common alkali metals sodium and potassium have proved indispensable in both academia and technol. Today these monometallic compds. are still utilized widely in synthetic campaigns, but in recent years they have been joined by an assortment of bimetallic formulations that also contain an alkali metal but in company with another metal. These bimetallic formulations often exhibit unique chem. that can be interpreted in terms of synergistic effects, for which the alkali metal is essential, though it is often the second metal that performs the synthetic transformation. Here, this "alkali-metal-mediated" chem. is surveyed focusing mainly on bimetallic formulations contg. two alkali metals or an alkali metal paired with magnesium, calcium, zinc, aluminum, or gallium. In this International Year of the Periodic Table (IYPT), we ponder whether a Periodic Table of Element Pairs will emerge in the future.
- 8Fauvarque, J.; Rouget, E. Compt. Rend. Hebd. Séances Acad. Sci., Ser. C 1968, 257, 1355Google ScholarThere is no corresponding record for this reference.
- 9Blomberg, C.; Mosher, H. A radical process in a reaction of a Grignard compound. J. Organomet. Chem. 1968, 13, 519– 522, DOI: 10.1016/S0022-328X(00)82781-3Google Scholar9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaF1cXkvVKjs7s%253D&md5=79ed18eeed10bd80dccd480a57115d24A radical process in a reaction of a Grignard compoundBlomberg, Cornelis; Mosher, Harry S.Journal of Organometallic Chemistry (1968), 13 (2), 519-22CODEN: JORCAI; ISSN:0022-328X.Chem. and E.S.R. evidence is reported for the occurrence of radicals during the reaction of the Grignard reagent from neopentyl chloride with benzophenone in tetrahydrofuran; 1,1-diphenyl-3,3-dimethylbutanol and benzopinacol were isolated and the formation of neopentane was observed. The steric bulk of the neopentyl group (R') so retards the normal addn. reaction [Ph2CO:Mg(R') X → Rh2R'COMgX] that the radicals, formed in the process [Ph2CO:Mg(R')X → Ph2C·OMgX + R·'], are able to escape from the solvent cage. The neopentyl radical can then react with the solvent to give R' and the ketyl can dimerize to give the Mg halide salt of benzopinacol.
- 10Blomberg, C.; Grootveld, H.; Gerner, T.; Bickelhaupt, F. Radical formation during reactions of Grignard reagents with quinones. J. Organomet. Chem. 1970, 24, 549– 553, DOI: 10.1016/S0022-328X(00)84483-6Google Scholar10https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE3cXltV2rs7s%253D&md5=b149433224f258111263fa1ea8aa0dafRadical formation during reactions of Grignard reagents with quinonesBlomberg, Cornelis; Grootveld, H. H.; Gerner, T. H.; Bickelhaupt, F.Journal of Organometallic Chemistry (1970), 24 (3), 549-53CODEN: JORCAI; ISSN:0022-328X.In dil. soln. the addn. of PhMgBr to acenaphthenequinone and phenanthrenequinone (I) leads by single electron transfer to the formation of the corresponding semiquinones, which can be identified by ESR. I yields ∼20% 9-hydroxy-10-phenoxyphenanthrene, the formation of which is explained by combination of the initially formed semiquinone and Ph radicals. The analogy with photochem. reactions and the possible occurrence of similar reactions in organomagnesium and organozinc chemistry are briefly discussed.
- 11Ashby, E. C.; Nackashi, J.; Parris, G. E. Composition of Grignard compounds. X. NMR, IR, and molecular association studies of some methylmagnesium alkoxides in diethyl ether, tetrahydrofuran, and benzene. J. Am. Chem. Soc. 1975, 97, 3162– 3171, DOI: 10.1021/ja00844a040Google Scholar11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE2MXksFOrurs%253D&md5=92a29e90c3ba2d15dc364ae2a5347ffcComposition of Grignard compounds. X. NMR, ir, and molecular association studies of some methylmagnesium alkoxides in diethyl ether, tetrahydrofuran, and benzeneAshby, E. C.; Nackashi, J.; Parris, G. E.Journal of the American Chemical Society (1975), 97 (11), 3162-71CODEN: JACSAT; ISSN:0002-7863.Mol. assocn. of MeMgOR (R = OCPh2Me, OCMe3, OCHMe2, OPr) in Et2O, THF, and C6H6 was examd. using ir and NMR spectral data. The steric bulk of the alkoxy group and the coordinating ability of the solvent determine the thermodynamically preferred soln. compn. In THF, solvated dimers are preferred. In Et2O, linear oligomers and cubane tetramers are preferred provided the alkoxy group is not bulkier than the tert-butoxy group. In C6H6, cubane tetramers are obsd. for alkoxy groups of intermediate bulk such as tert-butoxy and isopropoxy, but the less bulky n-propoxy group permits the formation of an oligomer contg. seven to nine monomer units. For the reagents with alkoxy groups less bulky than tert-butoxy, the equilibria involving various structures are established very rapidly. However, the dimer-linear oligomer ↹ cubane tetramer equilibrium is established very slowly for methylmagnesium tert-butoxide compds. The cubane form is very inert and does not exchange or otherwise interact with Me2Mg in Et2O. The dimer-linear oligomer form is quite labile and readily exchanges with Me2Mg forming mixed-bridged compds. However, in Et2O, the mixed bridge is not sufficiently strong to prevent slow conversion of methylmagnesium tert-butoxide to the cubane form thus releasing Me2Mg.
- 12Ashby, E. C.; Lopp, I. G.; Buhler, J. D. Mechanisms of Grignard reactions with ketones. Polar vs. single electron transfer pathways. J. Am. Chem. Soc. 1975, 97, 1964– 1966, DOI: 10.1021/ja00840a066Google Scholar12https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE2MXhsFCrs78%253D&md5=9c20322ef87fe6aa0b10e5a3826f32b5Mechanisms of Grignard reactions with ketones. Polar vs. single electron transfer pathwaysAshby, E. C.; Lopp, Irene G.; Buhler, Jerry D.Journal of the American Chemical Society (1975), 97 (7), 1964-6CODEN: JACSAT; ISSN:0002-7863.The addn. of MeMgBr to 2-methylbenzophenone in Et2O proceeds via a normal polar mechanism, whereas in the presence of small amts. of transition metal catalysts (e.g. 0.05 mole % FeCl3) the reaction proceeds via a single electron transfer pathway. The role of an intermediate (I) in by-product formation is examd.
- 13Ashby, E. C. A detailed description of the mechanism of reaction of Grignard reagents with ketones. Pure Appl. Chem. 1980, 52, 545– 569, DOI: 10.1351/pac198052030545Google Scholar13https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3cXksVGgs7k%253D&md5=6b83e73f4ac9654be92428627d98a4b4A detailed description of the mechanism of reaction of Grignard reagents with ketonesAshby, E. C.Pure and Applied Chemistry (1980), 52 (3), 545-69CODEN: PACHAS; ISSN:0033-4545.A review, chiefly of the work of the author, with 29 refs.
- 14Ashby, E. C.; Bowers, J. R. Organometallic reaction mechanisms. 17. Nature of alkyl transfer in reactions of Grignard reagents with ketones. Evidence for radical intermediates in the formation of 1,2-addition product involving tertiary and primary Grignard reagents. J. Am. Chem. Soc. 1981, 103, 2242– 2250, DOI: 10.1021/ja00399a018Google Scholar14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3MXitVyjtbY%253D&md5=894ebdd380e00f11030a6aee19f48826Organometallic reaction mechanisms. 17. Nature of alkyl transfer in reactions of Grignard reagents with ketones. Evidence for radical intermediates in the formation of 1,2-addition product involving tertiary and primary Grignard reagentsAshby, E. C.; Bowers, Joseph R., Jr.Journal of the American Chemical Society (1981), 103 (9), 2242-50CODEN: JACSAT; ISSN:0002-7863.The title 1,2-addn. products (formed after dissocn. of the radical anion-radical cation pair) show free-radical character, as indicated by the cyclized 1,2-addn. products formed from the reaction of a tertiary Grignard reagent probe with Ph2CO in THF and from the reaction of a primary Grignard reagent probe (neooctenyl Grignard reagent) with Ph2CO in Et2O. The 1,6-addn. products show free-radical character, as evidenced by the cyclized 1,6-addn. products formed in all of the reactions which involve the tertiary Grignard reagent probe (in all solvents studied) with Ph2CO and 2-MeC6H4COPh (I) and also in the reaction of the neooctenyl Grignard reagent probe with I.
- 15Lund, T.; Pedersen, M. L.; Frandsen, L. A. Does the reaction between fluorenone and grignard reagents involve free fluorenone anion radicals?. Tetrahedron Lett. 1994, 35, 9225– 9226, DOI: 10.1016/0040-4039(94)88472-2Google Scholar15https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXislagsrw%253D&md5=7a16badf7a3077c3a004c11eab03afeaDoes the reaction between fluorenone and Grignard reagents involve free fluorenone anion radicals?Lund, Torben; Pedersen, Morten L.; Frandsen, Lars A.Tetrahedron Letters (1994), 35 (49), 9225-6CODEN: TELEAY; ISSN:0040-4039. (Elsevier)The ratio between 1,6- and 1,2-addn. in the reactions of electrogenerated fluorenone anion radicals with RX in THF were similar to the ratio obtained in the Grignard reaction of fluorenone with RMgX in THF. This indicates that the addn. products in the Grignard reaction may be obtained via the coupling of freely diffusing fluorenone anion radicals with R radicals.
- 16Blomberg, C.; Salinger, R. M.; Mosher, H. S. Reaction of Grignard reagent from neopentyl chloride with benzophenone. A nuclear magnetic resonance study. J. Org. Chem. 1969, 34, 2385– 2388, DOI: 10.1021/jo01260a028Google Scholar16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaF1MXks12qsbw%253D&md5=ff6e7144106079007095c9c65674f978Reaction of the Grignard reagent from neopentyl chloride with benzophenone. Nuclear magnetic resonance study.Blomberg, Cornelis; Salinger, Rudolf M.; Mosher, Harry S.Journal of Organic Chemistry (1969), 34 (8), 2385-8CODEN: JOCEAH; ISSN:0022-3263.The rate of change in the N.M.R. spectrum of a mixt. of the Grignard reagent from neopentyl chloride and benzophenone in tetrahydrofuran was studied. The expected addn. reaction was complicated by the simultaneous occurrence of a radical reaction to produce neopentane and benzopinacol. These spectra are interpretable in terms of a reaction to give an initial product which in turn undergoes further reaction with the Grignard reagent to give a new reactive species. This reactive intermediate presumably is either the alkylmagnesium alkoxide (RMgOR') or a complex of the initial product with the Grignard reagent (RMgCl.R'OMgCl). This constitutes a direct observation of a process often postulated in the reaction of a Grignard compd. Qual. generalizations could be made but because of these complications it was not possible to make a quant. kinetic anal. of the data.
- 17Hoffmann, R. W.; Hölzer, B. Concerted and stepwise Grignard additions, probed with a chiral Grignard reagent. Chem. Commun. 2001, 491– 492, DOI: 10.1039/b009678oGoogle Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXhtl2ntLw%253D&md5=44546c2d93ae9b98c42f9d2b65f6033dConcerted and stepwise Grignard additions, probed with a chiral Grignard reagentHoffmann, Reinhard W.; Holzer, BettinaChemical Communications (Cambridge, United Kingdom) (2001), (5), 491-492CODEN: CHCOFS; ISSN:1359-7345. (Royal Society of Chemistry)The Grignard reagent (S)-PhCH2CH(MgCl)CH2CH3 6, in which the magnesium-bearing carbon atom is the sole stereogenic center has been added to CO2, PhNCO, PhNCS and certain aldehydes with full retention of configuration. Reaction with benzophenone, electron-deficient aldehydes and several allyl halides proceeded with partial or complete racemization. The findings are discussed with respect to a dichotomy between concerted polar and stepwise SET reaction pathways.
- 18Hoffmann, R. W. The quest for chiral Grignard reagents. Chem. Soc. Rev. 2003, 32, 225– 230, DOI: 10.1039/b300840cGoogle Scholar18https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXkt1emtL8%253D&md5=58237aeb3e5721d0f3d00c65d99db0b0The quest for chiral Grignard reagentsHoffmann, Reinhard W.Chemical Society Reviews (2003), 32 (4), 225-230CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review on the prepn. of chiral Grignard reagents. The involvement of single electron transfer, i.e. of free radicals in the reactions of organomagnesium reagents could be detected with the aid of a chiral secondary Grignard reagent, in which the magnesium-bearing carbon atom is the sole stereogenic center. So far, however, such reagents have not been accessible, because the std. prepn. of Grignard reagents proceeds via free radicals. The authors review and summarize here their efforts to generate (S)-1-Benzylpropylmagnesium chloride 36 by asym. synthesis starting from an enantiomerically pure compd. ArSOCH(Cl)CH2Ph (Ar = p-cl-C6H5) 30b using a sulfoxide/magnesium exchange reaction to give trans-β-chloro-α-phenylbenzenepropanol 33 followed by a carbenoid homologation reaction. Grignard reagent 36 turned out to be an ideal probe to learn about the extent to which SET is involved in reactions of organomagnesium reagents.
- 19Gajewski, J. J.; Bocian, W.; Harris, N. J.; Olson, L. P.; Gajewski, J. P. Secondary deuterium kinetic isotope effects in irreversible additions of hydride and carbon nucleophiles to aldehydes: a spectrum of transition states from complete bond formation to single electron transfer. J. Am. Chem. Soc. 1999, 121, 326– 334, DOI: 10.1021/ja982504rGoogle Scholar19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXjsFyi&md5=57e98328aac6eabcdfae331f64ccb1e7Secondary deuterium kinetic isotope effects in irreversible additions of hydride and carbon nucleophiles to aldehydes: a spectrum of transition states from complete bond formation to single electron transferGajewski, Joseph J.; Bocian, Wojciech; Harris, Nathan J.; Olson, Leif P.; Gajewski, John P.Journal of the American Chemical Society (1999), 121 (2), 326-334CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The competitive kinetics of hydride and organometallic addns. to PhCHO and PhCDO were detd. at -78° using LiAlH4, LiBHEt3, NaBH4, LiBH4, LiAlH(OCMe3)3, NaBH(OMe)3, NaBH(OAc)3 (at 20°), RMgBr (R = Me, Ph, allyl), and RLi (R = Me, Ph, Bu, Me3C, allyl). The hydride addns. had an inverse secondary D kinetic isotope effects in all cases, but the magnitude of the effect varied inversely with the apparent reactivity of the hydride. In the addns. of MeMgBr and of MeLi and PhLi, inverse secondary D isotope effects were obsd.; little if any isotope effect was obsd. with PhMgBr, BuLi or Me3CLi. With CH2:CHCH2M (M = MgBr, Li), a normal secondary D kinetic isotope effect was obsd. Rate-detg. single-electron transfer occurs with allyl reagents, but direct nucleophilic reaction occurs with all of the other reagents, with the extent of bond formation depending on the reactivity of the reagent. In the addn. of MeLi to cyclohexanecarboxaldehyde (I), a less inverse secondary D kinetic isotope effect was obsd. than that obsd. in the addn. of MeLi to PhCHO, and allyllithium addn. to I had a kinetic isotope effect near unity. The data with organometallic addns., which are not incompatible with observations of carbonyl C isotope effects, suggest that electrochem. detd. redox potentials which indicate endoergonic electron transfer with energies .ltorsim.13 kcal/mol allow electron-transfer mechanisms to compete well with direct polar addns. to aldehydes, provided that the reagent is highly stabilized, like allyl species. MeLi, PhLi, MeMgBr and PhMgBr are estd. to undergo electron transfer with endoergonicities >30 kcal/mol with PhCHO, so these react by direct polar addns. A working hypothesis is that BuLi reagents undergo polar addns., despite redox potentials which indicate ≤13 kcal/mol endoergonic electron transfer, because of the great exoergonicity assocd. with the 2-electron addn., which is responsible for a low barrier for polar reactions.
- 20Otte, D. A. L.; Woerpel, K. A. Evidence that Additions of Grignard Reagents to Aliphatic Aldehydes do Not Involve Single-Electron-Transfer Processes. Org. Lett. 2015, 17, 3906– 3909, DOI: 10.1021/acs.orglett.5b01893Google Scholar20https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXht1GnurrO&md5=ae186bb421050b8bc1d639a7ccb83cb6Evidence that Additions of Grignard Reagents to Aliphatic Aldehydes Do Not Involve Single-Electron-Transfer ProcessesOtte, Douglas A. L.; Woerpel, K. A.Organic Letters (2015), 17 (15), 3906-3909CODEN: ORLEF7; ISSN:1523-7052. (American Chemical Society)Addn. of allylmagnesium reagents to an aliph. aldehyde bearing a radical clock gave only addn. products and no evidence of ring-opened products that would suggest single-electron-transfer reactions. The analogous Barbier reaction also did not provide evidence for a single-electron-transfer mechanism in the addn. step. Other Grignard reagents (methyl-, vinyl-, t-Bu-, and triphenylmethylmagnesium halides) also do not appear to add to an alkyl aldehyde by a single-electron-transfer mechanism.
- 21Garst, J. F.; Soriaga, M. P. Grignard reagent formation. Coord. Chem. Rev. 2004, 248, 623– 652, DOI: 10.1016/j.ccr.2004.02.018Google Scholar21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXksF2ku7s%253D&md5=ccb7be559b5068df60420274187c985aGrignard reagent formationGarst, John F.; Soriaga, Manuel P.Coordination Chemistry Reviews (2004), 248 (7-8), 623-652CODEN: CCHRAM; ISSN:0010-8545. (Elsevier Science B.V.)A review with 74 refs. Probably reactions of Mg metal with org. halides RX in ether solvents are typical metallic corrosions in which the stabilization of Mg2+, substantially through its coordination by the solvent, drives its loss from the metal and consequently the redns. of RX and reaction intermediates such as R· at the metal surface. Although alkyl halides form Grignard reagents through nonchain mechanisms in which intermediate radicals diffuse in soln., very small amts. of radical isomerization occur in Grignard reactions of certain vinyl and aryl halides, even when intermediate radicals R· would isomerize very rapidly. This suggests a dominant nonradical mechanism for these vinyl and aryl halides or a mechanism in which intermediate radicals R· have extremely short lifetimes. Since the former seems more likely, a dianion mechanism, through a transition state [RX2-]‡, is proposed. Surface studies of polycryst. Mg show that the oxide layer is mostly Mg(OH)2 and that it is mech. passivating. In the absence of promoters, Grignard reactions occur very slowly until enough RX has seeped to the Mg surface and reacted there to undercut and cause the Mg(OH)2 layer to flake off.
- 22Chen, Z.-N.; Fu, G.; Xu, X. Theoretical studies on Grignard reagent formation: radical mechanism versus non-radical mechanism. Org. Biomol. Chem. 2012, 10, 9491– 9500, DOI: 10.1039/c2ob26658jGoogle Scholar22https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xhs1yks7fF&md5=3a16a90835cdf9ce691ba0aba4f9e968Theoretical studies on Grignard reagent formation: radical mechanism versus non-radical mechanismChen, Zhe-Ning; Fu, Gang; Xu, XinOrganic & Biomolecular Chemistry (2012), 10 (47), 9491-9500CODEN: OBCRAK; ISSN:1477-0520. (Royal Society of Chemistry)Here the systematic theor. study on the mechanisms of Grignard reagent formation is presented (GRF) for CH3Cl reacting with Mg atom, Mg2 and Mg clusters (Mg4-Mg20). The calcns. reveal that the ground state Mg atom is inactive under matrix condition, whereas it is active under metal vapor synthesis (MVS) conditions. However, the excited state Mg (3P) atom, as produced by laser-ablation, can react with CH3Cl without barriers, and hence is active under matrix condition. The prediction is that the bimagnesium Grignard reagent, though often proposed, can barely be obsd. exptl., due to its high reactivity towards addnl. CH3Cl to produce more stable Grignard reagent dimer, and that the cluster Grignard reagent RMg4X possesses a flat Mg4 unit rather than a tetrahedral geometry. The calcns. further reveal that the radical pathway (T4) is prevalent on Mg, Mg2 and Mgn clusters of small size, while the no-radical pathway (T2), which starts at Mg4, becomes competitive with T4 as the cluster size increases. A structure-reactivity relation between barrier heights and ionization potentials of Mgn is established. These findings not only resolve controversy in expt. and theory, but also provide insights which can be used in the design of effective synthesis approaches for the prepn. of chiral Grignard reagents.
- 23Shao, Y.; Liu, Z.; Huang, P.; Liu, B. A unified model of Grignard reagent formation. Phys. Chem. Chem. Phys. 2018, 20, 11100– 11108, DOI: 10.1039/C8CP01031EGoogle Scholar23https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXltFWqurw%253D&md5=0468205a9e6ba31fd7aeff16043280ccA unified model of Grignard reagent formationShao, Yunqi; Liu, Zhen; Huang, Pan; Liu, BopingPhysical Chemistry Chemical Physics (2018), 20 (16), 11100-11108CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)Grignard reagents are among the most fundamental reagents in org. synthesis, yet studies have hitherto failed to fully explain the selectivity and kinetics of Grignard reagent formation (GRF). The present study provides new insights into the intermediates and pathways of GRF using d. functional theory (DFT) calcns. Potential energy surfaces of RX dissocn. along different directions reveal the origin of configuration retention of alkenyl and arom. halides. Radical intermediates participate solely in the dissocn. stage, and depend on the geometry of the reactant halide. Dissocn. of org. halides yields stabilized surface anions, and the rest of the reaction is ionic in nature. MgX+/RMg+ are proposed as the key intermediates of Mg leaving from the surface in the self-activation of GRF, which explains the accelerated kinetics upon addn. of RMgX or MgX2. The intermediacy of the cations was supported by a simple electrochem. expt. To the best of the authors' knowledge, this is the 1st unified ionic model (I-model) developed for resolving the controversial issues of GRF.
- 24Axten, J.; Troy, J.; Jiang, P.; Trachtman, M.; Bock, C. W. An ab initio molecular orbital study of the Grignard reagents CH3MgCl and [CH3MgCl]2: the Schlenk equilibrium. Struct. Chem. 1994, 5, 99– 108, DOI: 10.1007/BF02265351Google Scholar24https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2cXmtlSrs7s%253D&md5=62a601c4a2da0528a295f40433e172e7An ab initio molecular orbital study of the Grignard reagents CH3MgCl and [CH3MgCl]2: The Schlenk equilibriumAxten, Jeffrey; Troy, Jennifer; Jiang, Peter; Trachtman, Mendel; Bock, Charles W.Structural Chemistry (1994), 5 (2), 99-108CODEN: STCHES; ISSN:1040-0400.Ab initio MO calcns. are used to study the modified Schlenk equil.: 2RMgCl .dblharw. MgR2 + MgCl2 .dblharw. Mg(Cl2)MgR2 with R = H and CH3. In the absence of any solvents, calcns. indicate that the formation of the various possible bridged dimers (RMgCl)2 is substantially exothermic. However, using di-Me ether as a model solvent, we show that the formation of the dimer (Me2O)(CH3)Mg(μ-Cl2)Mg(CH3)(OMe2) is exothermic only when entropic effects are included.
- 25Ehlers, A. W.; van Klink, G. P. M.; van Eis, M. J.; Bickelhaupt, F.; Nederkoorn, P. H. J.; Lammertsma, K. Density-Functional study of (solvated) Grignard complexes. J. Mol. Model. 2000, 6, 186– 194, DOI: 10.1007/s0089400060186Google Scholar25https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXkt1KhsrY%253D&md5=0ead5657fb9ba5cf003b7c95813d7f0eDensity-functional study of (solvated) Grignard complexesEhlers, Andreas W.; Van Klink, Gerard P. M.; Van Eis, Maurice J.; Bickelhaupt, Friedrich; Nederkoorn, Paul H. J.; Lammertsma, KoopJournal of Molecular Modeling (2000), 6 (2), 186-194CODEN: JMMOFK; ISSN:0948-5023. (Springer-Verlag)D. functional calcns. have been used to study the solvent effect of di-Et ether on the Schlenk equil. and the aggregation of Grignard reagents RMgX with R = Me, Et, Ph. Solvent stabilization of the Mg complexes of the first solvent is larger than that of the second one. The solvation energy decreases on going from the dihalides MgX2 to the monohalides RMgX to the diorgano compds. MgR2. The calcns. indicate that the energetic preference of the unsym. species reduces upon solvation. The strong tendency to dimerization of the un- and partly solvated compd. vanishes for the higher solvated cases.
- 26Tammiku, J.; Burk, P.; Tuulmets, A. 1,10-phenanthroline and its complexes with magnesium compounds. Disproportionation equilibria. J. Phys. Chem. A 2001, 105, 8554– 8561, DOI: 10.1021/jp011476nGoogle Scholar26https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXlvFyms70%253D&md5=8e9f052d75c330fadf679e58939df0e41,10-Phenanthroline and Its Complexes with Magnesium Compounds. Disproportionation EquilibriaTammiku, Jaana; Burk, Peeter; Tuulmets, AntsJournal of Physical Chemistry A (2001), 105 (37), 8554-8561CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The solvation, complexation, and disproportionation equil., which might be important during titrn. of a Grignard reagent RMgX with an alc. in the presence of 1,10-phenanthroline (phen), have been studied both in the gas phase and soln. using the d. functional theory (DFT) B3LYP/6-31+G* method. Solvation was modeled using the supermol. approach. NBO at. charge analyses were performed at the B3LYP/6-31G* level. The absorption spectra of the complexes were calcd. by the DFT TD/MPW1PW91/6-311+G** method. According to our calcns. the complexation of magnesium halide MgX2 with 1,10-phenanthroline is the reason for the disappearance of the red color of the complex RMgX(phen) near the titrn. end point.
- 27Tammiku-Taul, J.; Burk, P.; Tuulmets, A. Theoretical study of magnesium compounds: The Schlenk equilibrium in the gas phase and in the presence of Et2O and THF molecules. J. Phys. Chem. A 2004, 108, 133– 139, DOI: 10.1021/jp035653rGoogle Scholar27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXps1ymsLc%253D&md5=a004a2954b11ec82106aec89d2425056Theoretical Study of Magnesium Compounds: The Schlenk Equilibrium in the Gas Phase and in the Presence of Et2O and THF MoleculesTammiku-Taul, Jaana; Burk, Peeter; Tuulmets, AntsJournal of Physical Chemistry A (2004), 108 (1), 133-139CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The Schlenk equil. involving RMgX, R2Mg, and MgX2 (R = Me, Et, Ph and X = Cl, Br) has been studied both in the gas phase and in Et2O and THF solns. by the d. functional theory (DFT) B3LYP/6-31+G* method. Solvation was modeled using the supermol. approach. The stabilization due to interaction with solvent mols. decreases in the order MgX2 > RMgX > R2Mg and among the groups (R and X) Ph > Me > Et and Cl > Br. Studied magnesium compds. are more strongly solvated by THF compared to Et2O. The magnesium halide is solvated with up to four solvent mols. in THF soln., assuming that trans-dihalotetrakis(tetrahydrofurano)magnesium(II) complex forms. The formation of cis-dihalotetrakis(tetrahydrofurano)magnesium(II) is energetically less favorable than the formation of corresponding disolvated complexes. The predominant species in the Schlenk equil. are RMgX in Et2O and R2Mg + MgX2 in THF, which is consistent with exptl. data.
- 28Yamabe, S.; Yamazaki, S. In The Chemistry of Organomagnesium Compounds; Rappoport, Z., Marek, I., Eds.; Wiley-VCH: Weinheim, Germany, 2008; pp 369– 402.Google ScholarThere is no corresponding record for this reference.
- 29Yamazaki, S.; Yamabe, S. A computational study on addition of grignard reagents to carbonyl compounds. J. Org. Chem. 2002, 67, 9346– 9353, DOI: 10.1021/jo026017cGoogle Scholar29https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD38XovVChtL8%253D&md5=21d1d2fbbe9559d21f36ca0e767b850cA Computational Study on Addition of Grignard Reagents to Carbonyl CompoundsYamazaki, Shoko; Yamabe, ShinichiJournal of Organic Chemistry (2002), 67 (26), 9346-9353CODEN: JOCEAH; ISSN:0022-3263. (American Chemical Society)The mechanism of stereoselective addn. of Grignard reagents to carbonyl compds. was studied using B3LYP d. functional theory calcns. The study of the reaction of methylmagnesium chloride and formaldehyde in di-Me ether revealed a new reaction path involving carbonyl compd. coordination to Mg atoms in a dimeric Grignard reagent. The structure of the transition state for the addn. step shows that an interaction between a vicinal-Mg bonding alkyl group and C:O causes the C-C bond formation. The simplified mechanism shown by this model is in accord with the aggregation nature of Grignard reagents and their high reactivities toward carbonyl compds. Concerted and four-centered formation of strong O-Mg and C-C bonds was suggested as a polar mechanism. When the alkyl group is bulky, C-C bond formation is blocked and the Mg-O bond formation takes precedence. A diradical is formed with the odd spins localized on the alkyl group and carbonyl moiety. Diradical formation and its recombination probably are a single electron transfer (SET) process. The criteria for the concerted polar and stepwise SET processes were discussed in terms of precursor geometries and relative energies.
- 30Mori, T.; Kato, S. Grignard reagents in solution: Theoretical study of the equilibria and the reaction with a carbonyl compound in diethyl ether solvent. J. Phys. Chem. A 2009, 113, 6158– 6165, DOI: 10.1021/jp9009788Google Scholar30https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXlsFWjt7c%253D&md5=d774f6bc6bc37600ab68e8bf77155a04Grignard Reagents in Solution: Theoretical Study of the Equilibria and the Reaction with a Carbonyl Compound in Diethyl Ether SolventMori, Toshifumi; Kato, ShigekiJournal of Physical Chemistry A (2009), 113 (21), 6158-6165CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The equil. of Grignard reagents, CH3MgCl and CH3MgBr, in di-Et ether (Et2O) solvent as well as the reaction of the reagents with acetone are studied theor. To describe the equil. and reactions in Et2O solvent, the authors employ the ref. interaction site model SCF method with the second-order Moller-Plesset perturbation (RISM-MP2) free energy gradient method. Since the solvent mols. strongly coordinate to the Grignard reagents, the authors construct a cluster model by including several Et2O mols. into the quantum mech. region and embed it into the bulk solvent. Probably instead of the traditionally accepted cyclic dimer, the linear form of dimer is as stable as the monomer pair and participates in the equil. For the reaction with acetone, two important reaction paths (i.e., monomeric and linear dimeric paths) are studied. The barrier height for the monomeric path is much higher than that for the linear dimeric path, indicating that the reaction of the Grignard reagent with acetone proceeds through the linear dimeric reaction path. The change of solvation structure during the reaction is examd. From the calcd. free energy profiles, the entire reaction mechanisms of the Grignard reagents with aliph. ketones in Et2O solvent are discussed.
- 31Peltzer, R. M.; Eisenstein, O.; Nova, A.; Cascella, M. How solvent dynamics controls the Schlenk equilibrium of Grignard reagents: A computational study of CH3MgCl in tetrahydrofuran. J. Phys. Chem. B 2017, 121, 4226– 4237, DOI: 10.1021/acs.jpcb.7b02716Google Scholar31https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXltlWhtbg%253D&md5=731a756b355e2ecc9f28c13ae641c66bHow Solvent Dynamics Controls the Schlenk Equilibrium of Grignard Reagents: A Computational Study of CH3MgCl in TetrahydrofuranPeltzer, Raphael M.; Eisenstein, Odile; Nova, Ainara; Cascella, MicheleJournal of Physical Chemistry B (2017), 121 (16), 4226-4237CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)The Schlenk equil. is a complex reaction governing the presence of multiple chem. species in soln. of Grignard reagents. The full characterization at the mol. level of the transformation of CH3MgCl into MgCl2 and Mg(CH3)2 in THF by means of ab initio mol. dynamics simulations with enhanced-sampling metadynamics is presented. The reaction occurs via formation of dinuclear species bridged by chlorine atoms. At room temp., the different chem. species involved in the reaction accept multiple solvation structures, with two to four THF mols. that can coordinate the Mg atoms. The energy difference between all dinuclear solvated structures is lower than 5 kcal mol-1. The solvent is shown to be a direct key player driving the Schlenk mechanism. In particular, this study illustrates how the most stable sym. solvated dinuclear species, (THF)CH3Mg(μ-Cl)2MgCH3(THF) and (THF)CH3Mg(μ-Cl)(μ-CH3)MgCl(THF), need to evolve to less stable asym. solvated species, (THF)CH3Mg(μ-Cl)2MgCH3(THF)2 and (THF)CH3Mg(μ-Cl)(μ-CH3)MgCl(THF)2, in order to yield ligand exchange or product dissocn. In addn., the transferred ligands are always departing from an axial position of a pentacoordinated Mg atom. Thus, solvent dynamics is key to successive Mg-Cl and Mg-CH3 bond cleavages because bond breaking occurs at the most solvated Mg atom and the formation of bonds takes place at the least solvated one. The dynamics of the solvent also contributes to keep relatively flat the free energy profile of the Schlenk equil. These results shed light on one of the most used organometallic reagents whose structure in solvent remains exptl. unresolved. These results may also help to develop a more efficient catalyst for reactions involving these species.
- 32Hohenberg, P.; Kohn, W. Inhomogeneous electron gas. Phys. Rev. 1964, 136, B864– B871, DOI: 10.1103/PhysRev.136.B864Google ScholarThere is no corresponding record for this reference.
- 33Kohn, W.; Sham, L. J. Self-consistent equations including exchange and correlation effects. Phys. Rev. 1965, 140, 1133, DOI: 10.1103/PhysRev.140.A1133Google ScholarThere is no corresponding record for this reference.
- 34Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865– 3868, DOI: 10.1103/PhysRevLett.77.3865Google Scholar34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XmsVCgsbs%253D&md5=55943538406ee74f93aabdf882cd4630Generalized gradient approximation made simplePerdew, John P.; Burke, Kieron; Ernzerhof, MatthiasPhysical Review Letters (1996), 77 (18), 3865-3868CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)Generalized gradient approxns. (GGA's) for the exchange-correlation energy improve upon the local spin d. (LSD) description of atoms, mols., and solids. We present a simple derivation of a simple GGA, in which all parameters (other than those in LSD) are fundamental consts. Only general features of the detailed construction underlying the Perdew-Wang 1991 (PW91) GGA are invoked. Improvements over PW91 include an accurate description of the linear response of the uniform electron gas, correct behavior under uniform scaling, and a smoother potential.
- 35VandeVondele, J.; Hutter, J. Gaussian basis sets for accurate calculations on molecular systems in gas and condensed phases. J. Chem. Phys. 2007, 127, 114105, DOI: 10.1063/1.2770708Google Scholar35https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXhtFSrsLvM&md5=d7fdb937efb88cf3fca85792bb49ec27Gaussian basis sets for accurate calculations on molecular systems in gas and condensed phasesVandeVondele, Joost; Hutter, JurgJournal of Chemical Physics (2007), 127 (11), 114105/1-114105/9CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)We present a library of Gaussian basis sets that has been specifically optimized to perform accurate mol. calcns. based on d. functional theory. It targets a wide range of chem. environments, including the gas phase, interfaces, and the condensed phase. These generally contracted basis sets, which include diffuse primitives, are obtained minimizing a linear combination of the total energy and the condition no. of the overlap matrix for a set of mols. with respect to the exponents and contraction coeffs. of the full basis. Typically, for a given accuracy in the total energy, significantly fewer basis functions are needed in this scheme than in the usual split valence scheme, leading to a speedup for systems where the computational cost is dominated by diagonalization. More importantly, binding energies of hydrogen bonded complexes are of similar quality as the ones obtained with augmented basis sets, i.e., have a small (down to 0.2 kcal/mol) basis set superposition error, and the monomers have dipoles within 0.1 D of the basis set limit. However, contrary to typical augmented basis sets, there are no near linear dependencies in the basis, so that the overlap matrix is always well conditioned, also, in the condensed phase. The basis can therefore be used in first principles mol. dynamics simulations and is well suited for linear scaling calcns.
- 36Goedecker, S.; Teter, M.; Hutter, J. Separable dual-space Gaussian pseudopotentials. Phys. Rev. B 1996, 54, 1703– 1710, DOI: 10.1103/PhysRevB.54.1703Google Scholar36https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XksFOht78%253D&md5=de0d078249d924ff884f32cb1e02595cSeparable dual-space Gaussian pseudopotentialsGoedecker, S.; Teter, M.; Hutter, J.Physical Review B: Condensed Matter (1996), 54 (3), 1703-1710CODEN: PRBMDO; ISSN:0163-1829. (American Physical Society)We present pseudopotential coeffs. for the first two rows of the Periodic Table. The pseudopotential is of an analytic form that gives optimal efficiency in numerical calculations using plane waves as a basis set. At most, even coeffs. are necessary to specify its analytic form. It is separable and has optimal decay properties in both real and Fourier space. Because of this property, the application of the nonlocal part of the pseudopotential to a wave function can be done efficiently on a grid in real space. Real space integration is much faster for large systems than ordinary multiplication in Fourier space, since it shows only quadratic scaling with respect to the size of the system. We systematically verify the high accuracy of these pseudopotentials by extensive at. and mol. test calcns.
- 37Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104, DOI: 10.1063/1.3382344Google Scholar37https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXkvVyks7o%253D&md5=2bca89d904579d5565537a0820dc2ae8A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-PuGrimme, Stefan; Antony, Jens; Ehrlich, Stephan; Krieg, HelgeJournal of Chemical Physics (2010), 132 (15), 154104/1-154104/19CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The method of dispersion correction as an add-on to std. Kohn-Sham d. functional theory (DFT-D) has been refined regarding higher accuracy, broader range of applicability, and less empiricism. The main new ingredients are atom-pairwise specific dispersion coeffs. and cutoff radii that are both computed from first principles. The coeffs. for new eighth-order dispersion terms are computed using established recursion relations. System (geometry) dependent information is used for the first time in a DFT-D type approach by employing the new concept of fractional coordination nos. (CN). They are used to interpolate between dispersion coeffs. of atoms in different chem. environments. The method only requires adjustment of two global parameters for each d. functional, is asymptotically exact for a gas of weakly interacting neutral atoms, and easily allows the computation of at. forces. Three-body nonadditivity terms are considered. The method has been assessed on std. benchmark sets for inter- and intramol. noncovalent interactions with a particular emphasis on a consistent description of light and heavy element systems. The mean abs. deviations for the S22 benchmark set of noncovalent interactions for 11 std. d. functionals decrease by 15%-40% compared to the previous (already accurate) DFT-D version. Spectacular improvements are found for a tripeptide-folding model and all tested metallic systems. The rectification of the long-range behavior and the use of more accurate C6 coeffs. also lead to a much better description of large (infinite) systems as shown for graphene sheets and the adsorption of benzene on an Ag(111) surface. For graphene it is found that the inclusion of three-body terms substantially (by about 10%) weakens the interlayer binding. We propose the revised DFT-D method as a general tool for the computation of the dispersion energy in mols. and solids of any kind with DFT and related (low-cost) electronic structure methods for large systems. (c) 2010 American Institute of Physics.
- 38Swope, W. C.; Andersen, H. C.; Berens, P. H.; Wilson, K. R. A computer simulation method for the calculation of equilibrium constants for the formation of physical clusters of molecules: Application to small water clusters. J. Chem. Phys. 1982, 76, 648, DOI: 10.1063/1.442716Google ScholarThere is no corresponding record for this reference.
- 39Bussi, G.; Donadio, D.; Parrinello, M. Canonical sampling through velocity rescaling. J. Chem. Phys. 2007, 126, 014101, DOI: 10.1063/1.2408420Google Scholar39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXosVCltg%253D%253D&md5=9c182b57bfc65bca6be23c8c76b4be77Canonical sampling through velocity rescalingBussi, Giovanni; Donadio, Davide; Parrinello, MicheleJournal of Chemical Physics (2007), 126 (1), 014101/1-014101/7CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The authors present a new mol. dynamics algorithm for sampling the canonical distribution. In this approach the velocities of all the particles are rescaled by a properly chosen random factor. The algorithm is formally justified and it is shown that, in spite of its stochastic nature, a quantity can still be defined that remains const. during the evolution. In numerical applications this quantity can be used to measure the accuracy of the sampling. The authors illustrate the properties of this new method on Lennard-Jones and TIP4P water models in the solid and liq. phases. Its performance is excellent and largely independent of the thermostat parameter also with regard to the dynamic properties.
- 40Nosé, S. A unified formulation of the constant temperature molecular dynamics methods. J. Chem. Phys. 1984, 81, 511– 519, DOI: 10.1063/1.447334Google Scholar40https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2cXkvFOrs7k%253D&md5=2974515ec89e5601868e35871c0f19c2A unified formulation of the constant-temperature molecular-dynamics methodsNose, ShuichiJournal of Chemical Physics (1984), 81 (1), 511-19CODEN: JCPSA6; ISSN:0021-9606.Three recently proposed const. temp. mol. dynamics methods [N., (1984) (1); W. G. Hoover et al., (1982) (2); D. J. Evans and G. P. Morris, (1983) (2); and J. M. Haile and S. Gupta, 1983) (3)] are examd. anal. via calcg. the equil. distribution functions and comparing them with that of the canonical ensemble. Except for effects due to momentum and angular momentum conservation, method (1) yields the rigorous canonical distribution in both momentum and coordinate space. Method (2) can be made rigorous in coordinate space, and can be derived from method (1) by imposing a specific constraint. Method (3) is not rigorous and gives a deviation of order N-1/2 from the canonical distribution (N the no. of particles). The results for the const. temp.-const. pressure ensemble are similar to the canonical ensemble case.
- 41Hoover, W. G. Canonical dynamics: Equilibrium phase-space distributions. Phys. Rev. A 1985, 31, 1695– 1697, DOI: 10.1103/PhysRevA.31.1695Google Scholar41https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC2sjotlWltA%253D%253D&md5=99a2477835b37592226a5d18a760685cCanonical dynamics: Equilibrium phase-space distributionsHooverPhysical review. A, General physics (1985), 31 (3), 1695-1697 ISSN:0556-2791.There is no expanded citation for this reference.
- 42Martyna, G. J.; Klein, M. L.; Tuckerman, M. Nosé—Hoover chains: The canonical ensemble via continuous dynamics. J. Chem. Phys. 1992, 97, 2635– 2643, DOI: 10.1063/1.463940Google ScholarThere is no corresponding record for this reference.
- 43Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual molecular dynamics. J. Mol. Graphics 1996, 14, 33– 38, DOI: 10.1016/0263-7855(96)00018-5Google Scholar43https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28Xis12nsrg%253D&md5=1e3094ec3151fb85c5ff05f8505c78d5VDM: visual molecular dynamicsHumphrey, William; Dalke, Andrew; Schulten, KlausJournal of Molecular Graphics (1996), 14 (1), 33-8, plates, 27-28CODEN: JMGRDV; ISSN:0263-7855. (Elsevier)VMD is a mol. graphics program designed for the display and anal. of mol. assemblies, in particular, biopolymers such as proteins and nucleic acids. VMD can simultaneously display any no. of structures using a wide variety of rendering styles and coloring methods. Mols. are displayed as one or more "representations," in which each representation embodies a particular rendering method and coloring scheme for a selected subset of atoms. The atoms displayed in each representation are chosen using an extensive atom selection syntax, which includes Boolean operators and regular expressions. VMD provides a complete graphical user interface for program control, as well as a text interface using the Tcl embeddable parser to allow for complex scripts with variable substitution, control loops, and function calls. Full session logging is supported, which produces a VMD command script for later playback. High-resoln. raster images of displayed mols. may be produced by generating input scripts for use by a no. of photorealistic image-rendering applications. VMD has also been expressly designed with the ability to animate mol. dynamics (MD) simulation trajectories, imported either from files or from a direct connection to a running MD simulation. VMD is the visualization component of MDScope, a set of tools for interactive problem solving in structural biol., which also includes the parallel MD program NAMD, and the MDCOMM software used to connect the visualization and simulation programs, VMD is written in C++, using an object-oriented design; the program, including source code and extensive documentation, is freely available via anonymous ftp and through the World Wide Web.
- 44Carter, E. A.; Ciccotti, G.; Heynes, J. T.; Kapral, R. Constrained reaction coordinate dynamics for the simulation of rare events. Chem. Phys. Lett. 1989, 156, 472– 477, DOI: 10.1016/S0009-2614(89)87314-2Google Scholar44https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1MXks1CitrY%253D&md5=bff35139b4bb8f1c06d29386dfa786dfConstrained reaction coordinate dynamics for the simulation of rare eventsCarter, E. A.; Ciccotti, Giovanni; Hynes, James T.; Kapral, RaymondChemical Physics Letters (1989), 156 (5), 472-7CODEN: CHPLBC; ISSN:0009-2614.A computationally efficient mol. dynamics method for estg. the rates of rare events that occur by activated processes is described. The system is constrained at "bottleneck" regions on a general many-body reaction coordinate in order to generate a biased configurational distribution. Suitable reweighting of this biased distribution, along with correct momentum distribution sampling, provides a new ensemble, the constrained-reaction-coordinate-dynamics ensemble, with which to study rare events of this type. Applications to chem. reaction rates are made.
- 45Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Universal solvation model based on solute electron density and on a continuum model of the solvent defined by the bulk dielectric constant and atomic surface tensions. J. Phys. Chem. B 2009, 113, 6378– 6396, DOI: 10.1021/jp810292nGoogle Scholar45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXksV2is74%253D&md5=54931a64c70d28445ee53876a8b1a4b9Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface TensionsMarenich, Aleksandr V.; Cramer, Christopher J.; Truhlar, Donald G.Journal of Physical Chemistry B (2009), 113 (18), 6378-6396CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)We present a new continuum solvation model based on the quantum mech. charge d. of a solute mol. interacting with a continuum description of the solvent. The model is called SMD, where the "D" stands for "d." to denote that the full solute electron d. is used without defining partial at. charges. "Continuum" denotes that the solvent is not represented explicitly but rather as a dielec. medium with surface tension at the solute-solvent boundary. SMD is a universal solvation model, where "universal" denotes its applicability to any charged or uncharged solute in any solvent or liq. medium for which a few key descriptors are known (in particular, dielec. const., refractive index, bulk surface tension, and acidity and basicity parameters). The model separates the observable solvation free energy into two main components. The first component is the bulk electrostatic contribution arising from a self-consistent reaction field treatment that involves the soln. of the nonhomogeneous Poisson equation for electrostatics in terms of the integral-equation-formalism polarizable continuum model (IEF-PCM). The cavities for the bulk electrostatic calcn. are defined by superpositions of nuclear-centered spheres. The second component is called the cavity-dispersion-solvent-structure term and is the contribution arising from short-range interactions between the solute and solvent mols. in the first solvation shell. This contribution is a sum of terms that are proportional (with geometry-dependent proportionality consts. called at. surface tensions) to the solvent-accessible surface areas of the individual atoms of the solute. The SMD model has been parametrized with a training set of 2821 solvation data including 112 aq. ionic solvation free energies, 220 solvation free energies for 166 ions in acetonitrile, methanol, and DMSO, 2346 solvation free energies for 318 neutral solutes in 91 solvents (90 nonaq. org. solvents and water), and 143 transfer free energies for 93 neutral solutes between water and 15 org. solvents. The elements present in the solutes are H, C, N, O, F, Si, P, S, Cl, and Br. The SMD model employs a single set of parameters (intrinsic at. Coulomb radii and at. surface tension coeffs.) optimized over six electronic structure methods: M05-2X/MIDI!6D, M05-2X/6-31G*, M05-2X/6-31+G**, M05-2X/cc-pVTZ, B3LYP/6-31G*, and HF/6-31G*. Although the SMD model has been parametrized using the IEF-PCM protocol for bulk electrostatics, it may also be employed with other algorithms for solving the nonhomogeneous Poisson equation for continuum solvation calcns. in which the solute is represented by its electron d. in real space. This includes, for example, the conductor-like screening algorithm. With the 6-31G* basis set, the SMD model achieves mean unsigned errors of 0.6-1.0 kcal/mol in the solvation free energies of tested neutrals and mean unsigned errors of 4 kcal/mol on av. for ions with either Gaussian03 or GAMESS.
- 46Zhao, Y.; Truhlar, D. G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215– 241, DOI: 10.1007/s00214-007-0310-xGoogle Scholar46https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXltFyltbY%253D&md5=c31d6f319d7c7a45aa9b716220e4a422The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other functionalsZhao, Yan; Truhlar, Donald G.Theoretical Chemistry Accounts (2008), 120 (1-3), 215-241CODEN: TCACFW; ISSN:1432-881X. (Springer GmbH)We present two new hybrid meta exchange-correlation functionals, called M06 and M06-2X. The M06 functional is parametrized including both transition metals and nonmetals, whereas the M06-2X functional is a high-nonlocality functional with double the amt. of nonlocal exchange (2X), and it is parametrized only for nonmetals. The functionals, along with the previously published M06-L local functional and the M06-HF full-Hartree-Fock functionals, constitute the M06 suite of complementary functionals. We assess these four functionals by comparing their performance to that of 12 other functionals and Hartree-Fock theory for 403 energetic data in 29 diverse databases, including ten databases for thermochem., four databases for kinetics, eight databases for noncovalent interactions, three databases for transition metal bonding, one database for metal atom excitation energies, and three databases for mol. excitation energies. We also illustrate the performance of these 17 methods for three databases contg. 40 bond lengths and for databases contg. 38 vibrational frequencies and 15 vibrational zero point energies. We recommend the M06-2X functional for applications involving main-group thermochem., kinetics, noncovalent interactions, and electronic excitation energies to valence and Rydberg states. We recommend the M06 functional for application in organometallic and inorganometallic chem. and for noncovalent interactions.
- 47Ditchfield, R.; Hehre, W. J.; Pople, J. A. Self-consistent molecular-orbital methods. IX. An extended Gaussian-type basis for molecular-orbital studies of organic molecules. J. Chem. Phys. 1971, 54, 724– 728, DOI: 10.1063/1.1674902Google Scholar47https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE3MXksFOiuw%253D%253D&md5=abce439752b61fad3aa748515ec78c71Self-consistent molecular-orbital methods. IX. Extended Gaussian-type basis for molecular-orbital studies of organic moleculesDitchfield, R.; Hehre, Warren J.; Pople, John A.Journal of Chemical Physics (1971), 54 (2), 724-8CODEN: JCPSA6; ISSN:0021-9606.An extended basis set of at. functions expressed as fixed linear combinations of Gaussian functions is presented for H and the first-row atoms C to F. In this set. described as 4-31 G, each inner shell is represented by a single basis function taken as a sum of 4 Gaussians, and each valence orbital is split into inner and outer parts described by 3 and 1 Gaussian function, resp. The expansion coeffs. and Gaussian exponents are detd. by minimizing the total calcd. energy of the at. ground state. This basis set is then used in single-determinant MO studies of a group of small polyat. mols. Optimization of valence-shell scaling factors shows that considerable rescaling of at. functions occurs in mols., the largest effects being obsd. for H and C. However, the range of optimum scale factors for each atom is small enough to allow the selection of a std. mol. set. The use of this std. basis gives theoretical equil. geometries in reasonable agreement with expt.
- 48Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Petersson, G. A.; Nakatsuji, H.; Li, X.; Caricato, M.; Marenich, A. V.; Bloino, J.; Janesko, B. G.; Gomperts, R.; Mennucci, B.; Hratchian, H. P.; Ortiz, J. V.; Izmaylov, A. F.; Sonnenberg, J. L.; Williams-Young, D.; Ding, F.; Lipparini, F.; Egidi, F.; Goings, J.; Peng, B.; Petrone, A.; Henderson, T.; Ranasinghe, D.; Zakrzewski, V. G.; Gao, J.; Rega, N.; Zheng, G.; Liang, W.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Throssell, K.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M. J.; Heyd, J. J.; Brothers, E. N.; Kudin, K. N.; Staroverov, V. N.; Keith, T. A.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A. P.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Millam, J. M.; Klene, M.; Adamo, C.; Cammi, R.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Farkas, O.; Foresman, J. B.; Fox, D. J. Gaussian 09, revision B.01; Gaussian, Inc.: Wallingford, CT, 2016.Google ScholarThere is no corresponding record for this reference.
- 49Raghavachari, K.; Trucks, G. W.; Pople, J. A.; Head-Gordon, M. A fifth-order perturbation comparison of electron correlation theories. Chem. Phys. Lett. 1969, 157, 479– 483, DOI: 10.1016/S0009-2614(89)87395-6Google ScholarThere is no corresponding record for this reference.
- 50Dunning, T. H. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 1007– 1023, DOI: 10.1063/1.456153Google Scholar50https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1MXksVGmtrk%253D&md5=c6cd67a3748dc61692a9cb622d2694a0Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogenDunning, Thom H., Jr.Journal of Chemical Physics (1989), 90 (2), 1007-23CODEN: JCPSA6; ISSN:0021-9606.Guided by the calcns. on oxygen in the literature, basis sets for use in correlated at. and mol. calcns. were developed for all of the first row atoms from boron through neon, and for hydrogen. As in the oxygen atom calcns., the incremental energy lowerings, due to the addn. of correlating functions, fall into distinct groups. This leads to the concept of correlation-consistent basis sets, i.e., sets which include all functions in a given group as well as all functions in any higher groups. Correlation-consistent sets are given for all of the atoms considered. The most accurate sets detd. in this way, [5s4p3d2f1g], consistently yield 99% of the correlation energy obtained with the corresponding at.-natural-orbital sets, even though the latter contains 50% more primitive functions and twice as many primitive polarization functions. It is estd. that this set yields 94-97% of the total (HF + 1 + 2) correlation energy for the atoms neon through boron.
- 51CFOUR, a quantum-chemical program package written by Stanton, J. F.; Gauss, J.; Cheng, L.; Harding, M. E.; Matthews, D. A.; Szalay, P. G. with contributions from Auer, A. A.; Bartlett, R. J.; Benedikt, U.; Berger, C.; Bernholdt, D. E.; Bomble, Y. J.; Christiansen, O.; Engel, F.; Faber, R.; Heckert, M.; Heun, O.; Hilgenberg, M.; Huber, C.; Jagau, T.-C.; Jonsson, D.; Jusélius, J.; Kirsch, T.; Klein, K.; Lauderdale, W. J.; Lipparini, F.; Metzroth, T.; Mück, L.A.; O’Neill, D. P.; Price, D. R.; Prochnow, E.; Puzzarini, C.; Ruud, K.; Schiffmann, F.; Schwalbach, W.; Simmons, C.; Stopkowicz, S.; Tajti, A.; Vázquez, J.; Wang, F.; Watts, J. D. and the integral packages MOLECULE (Almlöf, J.; Taylor, P. R.), PROPS (Taylor, P. R.), ABACUS (Helgaker, T.; Jensen, H. J. Aa.; Jørgensen, P.; Olsen, J.), and ECP routines by Mitin, A. V. and van Wüllen, C.; website: http://www.cfour.de.Google ScholarThere is no corresponding record for this reference.
- 52Buergi, H. B.; Dunitz, J. D.; Shefter, E. Geometrical reaction coordinates. II. Nucleophilic addition to a carbonyl group. J. Am. Chem. Soc. 1973, 95, 5065– 5067, DOI: 10.1021/ja00796a058Google ScholarThere is no corresponding record for this reference.
- 53Buergi, H. B.; Lehn, J. M.; Wipff, G. Ab initio study of nucleophilic addition to a carbonyl group. J. Am. Chem. Soc. 1974, 96, 1956– 1957, DOI: 10.1021/ja00813a062Google Scholar53https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE2cXhtlalsL0%253D&md5=9646a20668e917e66b8876136bba887cAb initio study of nucleophilic addition to a carbonyl groupBuergi, H. B.; Lehn, J. M.; Wipff, G.Journal of the American Chemical Society (1974), 96 (6), 1956-7CODEN: JACSAT; ISSN:0002-7863.Ab initio SCF-LCGO (linear combination of Gaussian orbitals)-MO computations were performed on the reaction of hydride ion with HCHO, considered as model for nucleophilic addns. to the carbonyl group. Geometrical changes and electronic rearrangements were obtained as a function of the reaction coordinate. Orientational constraints in the course of the reaction bear relation to orbital steering, togetherness, and proximity effects considered in the literature. The calcd. geometrical changes correlate with the changes obsd. in crystal structures for the approach of an amino site toward a carbonyl group. A computation of the NH3-HCHO system at one fixed sepn. was also performed for comparison purposes.
- 54Walker, F. W.; Ashby, E. C. Composition of Grignard compounds. VI. Nature of association in tetrahydrofuran and diethyl ether solutions. J. Am. Chem. Soc. 1969, 91, 3845– 3850, DOI: 10.1021/ja01042a027Google Scholar54https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaF1MXktlerur0%253D&md5=98ef0f28cd3bdbeaec2ff3c655aee557Composition of Grignard compounds. VI. Nature of association in tetrahydrofuran and diethyl ether solutionsWalker, Frank W.; Ashby, E. C.Journal of the American Chemical Society (1969), 91 (14), 3845-50CODEN: JACSAT; ISSN:0002-7863.Ebullioscopic data are presented for tetrahydrofuran (I) and Et2O solns. of several Grignard and related Mg compounds over a wide concn. range. Anal. of the data is accomplished by observing the change in assocn. with concn. and by consideration of the constancy of the equil. consts. calcd. for several possible descriptions of the assocd. system. The expected nonideality of the solns. studied was considered in the interpretation of the data. While all the compds. studied were monomeric in I, the alkyl- and arylmagnesium bromides and iodides were monomeric in Et2O only at low concn. (<0.1 m), exhibiting in general an increase in assocn. with concn. These compds. are assocd. in a polymeric fashion. In contrast, the alkylmagnesium chlorides assoc. in Et2O to form stable dimers with the assocn. insensitive to concn. changes. Comparison of the data for Mg halides and dialkylmagnesium compds. in Et2O indicates that, except for the Me compd., assocn. is considerably stronger for the Mg halides than for the dialkylmagnesium compds. Thus, except for methylmagnesium halides, Grignard compds. assoc. with bridging mainly through the halogen atom. The methylmagnesium halides are exceptional since Me bridging is strong enough in Et2O to permit assocn. by bridging through either the Me group or the halogen atom. Although the steric and electronic nature of the alkyl group has some effect on the assocn. of Grignard compds., the effect is generally small compared to to the effect of halogen or solvent.
- 55Holm, T. Thermochemical bond dissociation energies of carbon-magnesium bonds. J. Chem. Soc., Perkin Trans. 2 1981, 2, 464– 467, DOI: 10.1039/P29810000464Google ScholarThere is no corresponding record for this reference.
- 56We consider here thermodynamically equilibrated solutions of the Grignard reagent. In fact, the formation of the Grignard reagent may occur via a radical mechanism; see, for instance, the recent study:Henriques, A. M.; Barbosa, A. G. H. Chemical bonding and the equilibrium composition of Grignard reagents in ethereal solution. J. Phys. Chem. A 2011, 115, 12259– 12270, DOI: 10.1021/jp202762pGoogle Scholar56https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhtlWjt7bJ&md5=dd04ac72cb7f1d42b5ed2e58265d14e7Chemical bonding and the equilibrium composition of Grignard reagents in ethereal solutionsHenriques, Andre M.; Barbosa, Andre G. H.Journal of Physical Chemistry A (2011), 115 (44), 12259-12270CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)A thorough anal. of the electronic structure and thermodn. aspects of Grignard reagents and its assocd. equil. compn. in ethereal solns. is performed. Considering methylmagnesium halides contg. fluorine, chlorine, and bromine, we studied the neutral, charged, and radical species assocd. with their chem. equil. in soln. The ethereal solvents considered, THF and di-Et ether, were modeled using the polarizable continuum model (PCM) and also by explicit coordination to the Mg atoms in a cluster. The chem. bonding of the species that constitute the Grignard reagent is analyzed in detail with generalized valence bond (GVB) wave functions. Equil. consts. were calcd. with the DFT/M06 functional and GVB wave functions, yielding similar results. According to our calcns. and existing kinetic and electrochem. evidence, the species R·, R-, ·MgX, and RMgX2- must be present in low concn. in the equil. We conclude that depending on the halogen, a different route must be followed to produce the relevant equil. species in each case. Chloride and bromide must preferably follow a "radical-based" pathway, and fluoride must follow a "carbanionic-based" pathway. These different mechanisms are contrasted against the available exptl. results and are proven to be consistent with the existing thermodn. data on the Grignard reagent equil.
- 57Ziegler, D. S.; Wei, B.; Knochel, P. Improving the halogen–magnesium exchange by using new turbo-Grignard reagents. Chem. - Eur. J. 2019, 25, 2695– 2703, DOI: 10.1002/chem.201803904Google Scholar57https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXisFektbnE&md5=7f13713cfdbc0ca7df641474933c6bb6Improving the Halogen-Magnesium Exchange by using New Turbo-Grignard ReagentsZiegler, Dorothee S.; Wei, Baosheng; Knochel, PaulChemistry - A European Journal (2019), 25 (11), 2695-2703CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. This Minireview describes the scope of the halogen-magnesium exchange. It shows that the use of the turbo-Grignard reagent (iPrMgCl·LiCl) greatly enhances the rate of the Br- and I-Mg exchange. Furthermore, this magnesium reagent allows the performance of a fast sulfoxide-magnesium exchange. Also, the use of s-BuMgOR·LiOR (R=2-ethylhexyl) enables a Br-Mg exchange in toluene leading to various Grignard reagents in toluene. Addnl., the new exchange reagent s-Bu2Mg·2LiOR (R = 2-ethylhexyl) further increases the rate of the halogen-magnesium exchange allowing one to perform a chlorine-magnesium exchange for arom. chlorides bearing an ortho-methoxy substituent in toluene.
- 58Harutyunyan, S. R.; Lopez, F.; Browne, W. R.; Correa, A.; Pena, D.; Badorrey, R.; Meetsma, A.; Minaard, A. J.; Feringa, B. L. On the mechanism of the copper-catalyzed enantioselective 1,4-addition of Grignard reagents to α, β-unsaturated carbonyl compounds. J. Am. Chem. Soc. 2006, 128, 9103– 9118, DOI: 10.1021/ja0585634Google Scholar58https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28Xmt12msbw%253D&md5=5c3b262d1e881545199120f63476f146On the Mechanism of the Copper-Catalyzed Enantioselective 1,4-Addition of Grignard Reagents to α,β-Unsaturated Carbonyl CompoundsHarutyunyan, Syuzanna R.; Lopez, Fernando; Browne, Wesley R.; Correa, Arkaitz; Pena, Diego; Badorrey, Ramon; Meetsma, Auke; Minnaard, Adriaan J.; Feringa, Ben L.Journal of the American Chemical Society (2006), 128 (28), 9103-9118CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The mechanism of the enantioselective 1,4-addn. of Grignard reagents to α,β-unsatd. carbonyl compds. promoted by copper complexes of chiral ferrocenyl diphosphines is explored through kinetic, spectroscopic, and electrochem. anal. On the basis of these studies, a structure of the active catalyst is proposed. The roles of the solvent, copper halide, and the Grignard reagent have been examd. Kinetic studies support a reductive elimination as the rate-limiting step in which the chiral catalyst, the substrate, and the Grignard reagent are involved. The thermodn. activation parameters were detd. from the temp. dependence of the reaction rate. The putative active species and the catalytic cycle of the reaction are discussed.
- 59Lopez, F.; Minaard, A. J.; Feringa, B. L. Catalytic enantioselective conjugate addition with Grignard reagents. Acc. Chem. Res. 2007, 40, 179– 188, DOI: 10.1021/ar0501976Google Scholar59https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XhtlaksrrL&md5=418c3907b4d4aed60135f06abc29c389Catalytic enantioselective conjugate addition with Grignard reagentsLopez, Fernando; Minnaard, Adriaan J.; Feringa, Ben L.Accounts of Chemical Research (2007), 40 (3), 179-188CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)A review on recent advances in Cu-catalyzed asym. conjugate addn. of Grignard reagents, asym. SN2' substitution reactions of allylic bromides with Grignard reagents, and application to synthesis of natural products.
- 60Seitz, T. A.; Seitz, J. A. A general two-metal-ion mechanism for catalytic RNA. Proc. Natl. Acad. Sci. U. S. A. 1993, 90, 6498– 6502, DOI: 10.1073/pnas.90.14.6498Google ScholarThere is no corresponding record for this reference.
- 61De Vivo, M.; Dal Peraro, M.; Klein, M. L. Phosphodiester cleavage in ribonuclease H occurs via an associative two-metal-aided catalytic mechanism. J. Am. Chem. Soc. 2008, 130, 10955– 10962, DOI: 10.1021/ja8005786Google Scholar61https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXptVGisr8%253D&md5=db9ad0480dccbdb588f0f86ecc2d710dPhosphodiester Cleavage in Ribonuclease H Occurs via an Associative Two-Metal-Aided Catalytic MechanismDe Vivo, Marco; Dal Peraro, Matteo; Klein, Michael L.Journal of the American Chemical Society (2008), 130 (33), 10955-10962CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)RNase H belongs to the nucleotidyl-transferase (NT) superfamily and hydrolyzes the phosphodiester linkages that form the backbone of the RNA strand in RNA•DNA hybrids. This enzyme is implicated in replication initiation and DNA topol. restoration and represents a very promising target for anti-HIV drug design. Structural information has been provided by high-resoln. crystal structures of the complex RNase H/RNA•DNA from Bacillus halodurans (Bh), which reveals that two metal ions are required for formation of a catalytic active complex. Here, we use classical force field-based and quantum mechanics/mol. mechanics calcns. for modeling the nucleotidyl transfer reaction in RNase H, clarifying the role of the metal ions and the nature of the nucleophile (water vs. hydroxide ion). During the catalysis, the two metal ions act cooperatively, facilitating nucleophile formation and stabilizing both transition state and leaving group. Importantly, the two Mg2+ metals also support the formation of a meta-stable phosphorane intermediate along the reaction, which resembles the phosphorane intermediate structure obtained only in the debated β-phosphoglucomutase crystal (Lahiri, S. D.; et al. Science 2003, 299 (5615), 2067-2071). The nucleophile formation (i.e., water deprotonation) can be achieved in situ, after migration of one proton from the water to the scissile phosphate in the transition state. This proton transfer is actually mediated by solvation water mols. Due to the highly conserved nature of the enzymic bimetal motif, these results might also be relevant for structurally similar enzymes belonging to the NT superfamily.
- 62Rosta, E.; Nowotny, M.; Yang, W.; Hummer, G. Catalytic mechanism of RNA backbone cleavage by ribonuclease H from quantum mechanics/molecular mechanics simulations. J. Am. Chem. Soc. 2011, 133, 8934– 8941, DOI: 10.1021/ja200173aGoogle Scholar62https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXmsFymt7Y%253D&md5=0819abe0e0cedd6f352f09dee5c1c6f2Catalytic Mechanism of RNA Backbone Cleavage by Ribonuclease H from Quantum Mechanics/Molecular Mechanics SimulationsRosta, Edina; Nowotny, Marcin; Yang, Wei; Hummer, GerhardJournal of the American Chemical Society (2011), 133 (23), 8934-8941CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)We use quantum mechanics/mol. mechanics simulations to study the cleavage of the RNA (RNA) backbone catalyzed by RNase H. This protein is a prototypical member of a large family of enzymes that use two-metal catalysis to process nucleic acids. By combining Hamiltonian replica exchange with a finite-temp. string method, we calc. the free energy surface underlying the RNA-cleavage reaction and characterize its mechanism. We find that the reaction proceeds in two steps. In a first step, catalyzed primarily by magnesium ion A and its ligands, a water mol. attacks the scissile phosphate. Consistent with thiol-substitution expts., a water proton is transferred to the downstream phosphate group. The transient phosphorane formed as a result of this nucleophilic attack decays by breaking the bond between the phosphate and the ribose oxygen. In the resulting intermediate, the dissocd. but unprotonated leaving group forms an alkoxide coordinated to magnesium ion B. In a second step, the reaction is completed by protonation of the leaving group, with a neutral Asp132 as a likely proton donor. The overall reaction barrier of ∼15 kcal mol-1, encountered in the first step, together with the cost of protonating Asp132, is consistent with the slow measured rate of ∼1-100/min. The two-step mechanism is also consistent with the bell-shaped pH dependence of the reaction rate. The nonmonotonic relative motion of the magnesium ions along the reaction pathway agrees with X-ray crystal structures. Proton-transfer reactions and changes in the metal ion coordination emerge as central factors in the RNA-cleavage reaction.
- 63Casalino, L.; Palermo, G.; Rothlisberger, U.; Magistrato, A. Who activates the nucleophile in ribozyme catalysis? An answer from the splicing mechanism of group II introns. J. Am. Chem. Soc. 2016, 138, 10374– 10377, DOI: 10.1021/jacs.6b01363Google Scholar63https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XpvFylsb0%253D&md5=e01f4d5f916c9ea688b90d8baeeb0a69Who Activates the Nucleophile in Ribozyme Catalysis? An Answer from the Splicing Mechanism of Group II IntronsCasalino, Lorenzo; Palermo, Giulia; Rothlisberger, Ursula; Magistrato, AlessandraJournal of the American Chemical Society (2016), 138 (33), 10374-10377CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Group II introns are Mg2+-dependent ribozymes that are considered to be the evolutionary ancestors of the eukaryotic spliceosome, thus representing an ideal model system to understand the mechanism of conversion of premature mRNA (mRNA) into mature mRNA. Neither in splicing nor for self-cleaving ribozymes has the role of the two Mg2+ ions been established, and even the way the nucleophile is activated is still controversial. Here we employed hybrid quantum-classical QM(Car-Parrinello)/MM mol. dynamics simulations in combination with thermodn. integration to characterize the mol. mechanism of the first and rate-detg. step of the splicing process (i.e., the cleavage of the 5'-exon) catalyzed by group II intron ribozymes. Remarkably, our results show a new RNA-specific dissociative mechanism in which the bulk water accepts the nucleophile's proton during its attack on the scissile phosphate. The process occurs in a single step with no Mg2+ ion activating the nucleophile, at odds with nucleases enzymes. We suggest that the novel reaction path elucidated here might be an evolutionary ancestor of the more efficient two-metal-ion mechanism found in enzymes.
- 64Genna, V.; Vidossich, P.; Ippoliti, E.; Carloni, P.; De Vivo, M. A self-activated mechanism for nucleic acid polymerization catalyzed by DNA/RNA polymerases. J. Am. Chem. Soc. 2016, 138, 14592– 14598, DOI: 10.1021/jacs.6b05475Google Scholar64https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XhtlCrsLzE&md5=f8ee07a32f8b12a2978ab9c974a701b9A Self-Activated Mechanism for Nucleic Acid Polymerization Catalyzed by DNA/RNA PolymerasesGenna, Vito; Vidossich, Pietro; Ippoliti, Emiliano; Carloni, Paolo; Vivo, Marco DeJournal of the American Chemical Society (2016), 138 (44), 14592-14598CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The enzymic polymn. of DNA and RNA is at the basis of genetic inheritance for all living organisms. It is catalyzed by the DNA/RNA polymerase (Pol) superfamily. Here, bioinformatics anal. revealed that the incoming nucleotide substrate always forms an H-bond between its 3'-OH and β-phosphate moieties upon formation of the Michaelis complex. This previously unrecognized H-bond implies a novel self-activated mechanism (SAM), which synergistically connects the in situ nucleophile formation with subsequent nucleotide addn. and, importantly, nucleic acid translocation. Thus, SAM allows an elegant and efficient closed-loop sequence of chem. and phys. steps for Pol catalysis. This is markedly different from previous mechanistic hypotheses. This proposed mechanism was corroborated via ab initio QM/MM simulations on a specific Pol, human DNA polymerase-η, an enzyme involved in repairing damaged DNA. The structural conservation of DNA and RNA Pols supports the possible extension of SAM to Pol enzymes from the 3 domains of life.
- 65Genna, V.; Donati, E.; De Vivo, M. The catalytic mechanism of DNA and RNA polymerases. ACS Catal. 2018, 8, 11103– 11118, DOI: 10.1021/acscatal.8b03363Google Scholar65https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXhvFCks7nI&md5=79eb79ea021ae74916df15ef0bbc0c17The Catalytic Mechanism of DNA and RNA PolymerasesGenna, Vito; Donati, Elisa; De Vivo, MarcoACS Catalysis (2018), 8 (12), 11103-11118CODEN: ACCACS; ISSN:2155-5435. (American Chemical Society)A review. DNA and RNA polymerases (Pols) catalyze nucleic acid biosynthesis in all domains of life, with implications for human diseases and health. Pols carry out nucleic acid extension through the addn. of one incoming nucleotide trisphosphate at the 3'-OH terminus of the growing primer strand, at every catalytic cycle. Thus, Pol catalysis involves chem. reactions for nucleophile 3'-OH deprotonation and nucleotide addn., as well as major protein conformational motions and structural rearrangements for nucleotide selection, binding, and nucleic acid translocation to complete the overall catalytic cycle. In this respect, quantum and mol. mechanics simulations, integrated with exptl. data, have advanced our mechanistic understanding of how Pols operate at the at. level. This Perspective outlines how modern mol. simulations can further deepen our understanding of Pol catalytic reactions and fidelity, which may help in devising strategies for designing drugs and artificial Pols for biotechnol. and clin. purposes.
Cited By
This article is cited by 78 publications.
- Anh D. H. Pham, Johnny Bui, Kenneth W. Foreman. Minimal Theoretical Description of Magnesium Halogen Exchanges. Organometallics 2023, 42
(22)
, 3266-3274. https://doi.org/10.1021/acs.organomet.3c00382
- Junu Kim, Yusuke Hayashi, Sara Badr, Kazuya Okamoto, Toshikazu Hakogi, Haruo Furukawa, Satoshi Yoshikawa, Hayao Nakanishi, Hirokazu Sugiyama. Hybrid Modeling of an Active Pharmaceutical Ingredient Flow Synthesis in a Ring-Opening Reaction of an Epoxide with a Grignard Reagent. Industrial & Engineering Chemistry Research 2023, 62
(43)
, 17824-17834. https://doi.org/10.1021/acs.iecr.3c02137
- Hamed Zarei, Sara Sobhani, José Miguel Sansano. First Reusable Catalyst for the Reductive Coupling Reaction of Organohalides with Aldehydes. ACS Omega 2023, 8
(40)
, 36801-36814. https://doi.org/10.1021/acsomega.3c03414
- Marinella de Giovanetti, Sondre H. Hopen Eliasson, Abril C. Castro, Odile Eisenstein, Michele Cascella. Morphological Plasticity of LiCl Clusters Interacting with Grignard Reagent in Tetrahydrofuran. Journal of the American Chemical Society 2023, 145
(30)
, 16305-16309. https://doi.org/10.1021/jacs.3c04238
- Gustavo G. Flores-Bernal, Ma. Elena Vargas-Díaz, Hugo A. Jiménez-Vázquez, Marcos Hernández-Rodríguez, L. Gerardo Zepeda-Vallejo. Structural Features of Diacyldodecaheterocycles with Pseudo-C2-Symmetry: Promising Stereoinductors for Divergent Synthesis of Chiral Alcohols. ACS Omega 2023, 8
(23)
, 20611-20620. https://doi.org/10.1021/acsomega.3c01161
- Gerald J. Tanoury, Satish Kumar Iyemperumal, Elaine C. Lee. Toward a Combined Molecular Dynamics and Quantum Mechanical Approach to Understanding Solvent Effects on Chemical Processes in the Pharmaceutical Industry: The Case of a Lewis Acid-Mediated SNAr Reaction. Organic Process Research & Development 2023, 27
(4)
, 742-754. https://doi.org/10.1021/acs.oprd.3c00010
- Tingting Sun, Jintong Zhang, Yijun Fang, Yu Zhou, Haiqun Cao, Gen Luo, Zhi-Chao Cao. Enantioselective Alkylation of Unactivated C–O Bond: Solvent Molecule Affects Competing β-H Elimination and Reductive Elimination Dynamics. ACS Catalysis 2023, 13
(6)
, 3702-3709. https://doi.org/10.1021/acscatal.2c06054
- Min Su, Meng-Qin Pu, Hang Xiao, Yu-Jiao Chen, Tao Li, Quan-Xi Shi, Yu-Jing Sheng, Hongli Bao, Wen-Ming Wan. Turbo-Grignard Reagent Mediated Polymerization of Styrene under Mild Conditions Capable of Low Đ and Reactive Hydrogen Compatibility. Macromolecules 2022, 55
(23)
, 10422-10429. https://doi.org/10.1021/acs.macromol.2c01904
- Joseph E. Reynolds, III, Austin C. Acosta, ShinYoung Kang, Sichi Li, Andrew S. Lipton, Mark E. Bowden, Nicholas R. Myllenbeck, Andreas Schneemann, Noemi Leick, Austin Bhandarkar, Christopher Reed, Robert D. Horton, Thomas Gennett, Brandon C. Wood, Mark D. Allendorf, Vitalie Stavila. Teaching an Old Reagent New Tricks: Synthesis, Unusual Reactivity, and Solution Dynamics of Borohydride Grignard Compounds. Organometallics 2022, 41
(20)
, 2823-2832. https://doi.org/10.1021/acs.organomet.2c00284
- Przemysław J. Boratyński. Stereoselective Domino Rearrangement peri-Annulation of Cinchona Alkaloid Derivatives with 8-Bromo-1-naphthyl Grignard. The Journal of Organic Chemistry 2022, 87
(17)
, 11602-11607. https://doi.org/10.1021/acs.joc.2c01249
- Yusuke Kuroda. 1,1-Carboamination of Terminal Alkenes via a Reaction of Azo-Ene Adducts with Grignard Reagents. Organic Letters 2022, 24
(33)
, 6224-6229. https://doi.org/10.1021/acs.orglett.2c02585
- Nicole D. Bartolo, Krystyna M. Demkiw, Jacquelyne A. Read, Elizabeth M. Valentín, Yingying Yang, Alexandra M. Dillon, Chunhua T. Hu, Michael D. Ward, K. A. Woerpel. Conformationally Biased Ketones React Diastereoselectively with Allylmagnesium Halides. The Journal of Organic Chemistry 2022, 87
(5)
, 3042-3065. https://doi.org/10.1021/acs.joc.1c02844
- Christoph Taeschler, Eva Kirchner, Emilia Păunescu, Ulrich Mayerhöffer. Copper-Free Alternatives to Access Ketone Building Blocks from Grignard Reagents. ACS Omega 2022, 7
(4)
, 3613-3617. https://doi.org/10.1021/acsomega.1c06202
- Eliot F. Woods, Alexandra J. Berl, Leanna P. Kantt, Christopher T. Eckdahl, Michael R. Wasielewski, Brandon E. Haines, Julia A. Kalow. Light Directs Monomer Coordination in Catalyst-Free Grignard Photopolymerization. Journal of the American Chemical Society 2021, 143
(44)
, 18755-18765. https://doi.org/10.1021/jacs.1c09595
- Jennifer A. Clark, Pranav J. Thacker, Charles J. McGill, Jason R. Miles, Phillip R. Westmoreland, Kirill Efimenko, Jan Genzer, Erik E. Santiso. DFT Analysis of Organotin Catalytic Mechanisms in Dehydration Esterification Reactions for Terephthalic Acid and 2,2,4,4-Tetramethyl-1,3-cyclobutanediol. The Journal of Physical Chemistry A 2021, 125
(23)
, 4943-4956. https://doi.org/10.1021/acs.jpca.1c00850
- Nicole D. Bartolo, Krystyna M. Demkiw, Elizabeth M. Valentín, Chunhua T. Hu, Alya A. Arabi, K. A. Woerpel. Diastereoselective Additions of Allylmagnesium Reagents to α-Substituted Ketones When Stereochemical Models Cannot Be Used. The Journal of Organic Chemistry 2021, 86
(10)
, 7203-7217. https://doi.org/10.1021/acs.joc.1c00553
- Philipp Rinke, Helmar Görls, Robert Kretschmer. Calcium and Magnesium Bis(β-diketiminate) Complexes: Impact of the Alkylene Bridge on Schlenk-Type Rearrangements. Inorganic Chemistry 2021, 60
(7)
, 5310-5321. https://doi.org/10.1021/acs.inorgchem.1c00301
- Yue Fu, Leonardo Bernasconi, Peng Liu. Ab Initio Molecular Dynamics Simulations of the SN1/SN2 Mechanistic Continuum in Glycosylation Reactions. Journal of the American Chemical Society 2021, 143
(3)
, 1577-1589. https://doi.org/10.1021/jacs.0c12096
- Akachukwu D. Obi, Jacob E. Walley, Nathan C. Frey, Yuen Onn Wong, Diane A. Dickie, Charles Edwin Webster, Robert J. Gilliard, Jr.. Tris(carbene) Stabilization of Monomeric Magnesium Cations: A Neutral, Nontethered Ligand Approach. Organometallics 2020, 39
(23)
, 4329-4339. https://doi.org/10.1021/acs.organomet.0c00462
- Zeng Hong, Jiancheng Ruan, Xinzhi Chen, Chao Qian, Xin Ge, Shaodong Zhou. On the Origin of the Promoting Effect Exerted by Magnesium in the ZnCl2-Catalyzed Synthesis of N,N-Diisopropylethylamine. ACS Omega 2020, 5
(46)
, 29903-29912. https://doi.org/10.1021/acsomega.0c04188
- Thomas Q. Davies, Michael J. Tilby, Jack Ren, Nicholas A. Parker, David Skolc, Adrian Hall, Fernanda Duarte, Michael C. Willis. Harnessing Sulfinyl Nitrenes: A Unified One-Pot Synthesis of Sulfoximines and Sulfonimidamides. Journal of the American Chemical Society 2020, 142
(36)
, 15445-15453. https://doi.org/10.1021/jacs.0c06986
- Jarvis Hill, Asiri A. Hettikankanamalage, David Crich. Diversity-Oriented Synthesis of N,N,O-Trisubstituted Hydroxylamines from Alcohols and Amines by N–O Bond Formation. Journal of the American Chemical Society 2020, 142
(35)
, 14820-14825. https://doi.org/10.1021/jacs.0c05991
- Nicole D. Bartolo, K. A. Woerpel. Evidence against Single-Electron Transfer in the Additions of Most Organomagnesium Reagents to Carbonyl Compounds. The Journal of Organic Chemistry 2020, 85
(12)
, 7848-7862. https://doi.org/10.1021/acs.joc.0c00481
- Jerik Mathew Valera Lauridsen, Sung Yeon Cho, Han Yong Bae, Ji-Woong Lee. CO2 (De)Activation in Carboxylation Reactions: A Case Study Using Grignard Reagents and Nucleophilic Bases. Organometallics 2020, 39
(9)
, 1652-1657. https://doi.org/10.1021/acs.organomet.9b00838
- Sabrina G. Sobel . Active Learning through Discussions of Current Research in Inorganic Chemistry Classes. 2020, 21-30. https://doi.org/10.1021/bk-2020-1370.ch003
- Anjali Gupta, Joydev K. Laha. Growing Utilization of Radical Chemistry in the Synthesis of Pharmaceuticals. The Chemical Record 2023, 23
(11)
https://doi.org/10.1002/tcr.202300207
- Christoph Helling, Cameron Jones. Schlenk‐Type Equilibria of Grignard‐Analogous Arylberyllium Complexes: Steric Effects**. Chemistry – A European Journal 2023, 29
(60)
https://doi.org/10.1002/chem.202302222
- Magnus R. Buchner, Lewis R. Thomas‐Hargreaves, Chantsalmaa Berthold, Deniz F. Bekiş, Sergei I. Ivlev. A Preference for Heterolepticity ‐
Schlenk
Type Equilibria in Organometallic Beryllium Systems. Chemistry – A European Journal 2023, 29
(60)
https://doi.org/10.1002/chem.202302495
- Erin M. Hanada, Patrick J. McShea, Suzanne A. Blum. Trimethylsilyl Chloride Aids in Solubilization of Oxidative Addition Intermediates from Zinc Metal. Angewandte Chemie 2023, 135
(43)
https://doi.org/10.1002/ange.202307787
- Erin M. Hanada, Patrick J. McShea, Suzanne A. Blum. Trimethylsilyl Chloride Aids in Solubilization of Oxidative Addition Intermediates from Zinc Metal. Angewandte Chemie International Edition 2023, 62
(43)
https://doi.org/10.1002/anie.202307787
- Humberto A. Rodríguez, F. Matthias Bickelhaupt, Israel Fernández. Origin of the Bürgi‐Dunitz Angle. ChemPhysChem 2023, 24
(17)
https://doi.org/10.1002/cphc.202300379
- Sambasivarao Kotha, Usha Nandan Chaurasia, Kunkumita Jena. Ring‐Rearrangement Metathesis Approach to Fused [5/5/6/5/5] and [6/5/5/5/5/6] Carbocyclic Derivatives.. ChemistrySelect 2023, 8
(30)
https://doi.org/10.1002/slct.202302094
- Khalifah A. Salmeia, Akef T. Afaneh, Reem R. Habash, Antonia Neels. Trivinylphosphine Oxide: Synthesis, Characterization, and Polymerization Reactivity Investigated Using Single-Crystal Analysis and Density Functional Theory. Molecules 2023, 28
(16)
, 6097. https://doi.org/10.3390/molecules28166097
- Philipp Schüler, Simon Sengupta, Sven Krieck, Matthias Westerhausen. In Situ Generation of Magnesium‐ and Calcium‐Based Grignard Reagents for Amide Synthesis. Chemistry – A European Journal 2023, 29
(40)
https://doi.org/10.1002/chem.202300833
- Hans-Jürgen Federsel. What enables and blocks synthetic chemistry methods in becoming industrially significant?. Cell Reports Physical Science 2023, 4
(7)
, 101493. https://doi.org/10.1016/j.xcrp.2023.101493
- Shunya Ono, Aya Sugiyama, Nao Sakamoto, Kazuma Kuwabara, Mao Minoura, Toshiaki Murai. Two-Step Substitution Reaction of Phosphonates Carrying a Binaphthyl Group with Grignard Reagents Leading to the Formation of P-Chirogenic Phosphine Oxides. Synlett 2023, 34
(12)
, 1502-1506. https://doi.org/10.1055/a-1979-6245
- Andreas Hermann, Rana Seymen, Lukas Brieger, Johannes Kleinheider, Bastian Grabe, Wolf Hiller, Carsten Strohmann. Umfassende Studie der Gesteigerten Reaktivität von Turbo‐Grignard‐Reagenzien**. Angewandte Chemie 2023, 135
(25)
https://doi.org/10.1002/ange.202302489
- Andreas Hermann, Rana Seymen, Lukas Brieger, Johannes Kleinheider, Bastian Grabe, Wolf Hiller, Carsten Strohmann. Comprehensive Study of the Enhanced Reactivity of Turbo‐Grignard Reagents**. Angewandte Chemie International Edition 2023, 62
(25)
https://doi.org/10.1002/anie.202302489
- Venkata D. B. C. Dasireddy, Gurwinder Singh, Stalin Joseph, Yoshihiro Sugi, Ajayan Vinu. Homogeneous
F
riedel–
C
rafts Alkylation. 2023, 555-594. https://doi.org/10.1002/9783527827992.ch20
- Simon de Graaff, Marc Schmidtmann, Rüdiger Beckhaus. Elemental Magnesium as an Early Transition Metal Substitute: From Pentafulvene Coordination to Deprotonation of Amines. Chemistry – A European Journal 2023, 29
(30)
https://doi.org/10.1002/chem.202300221
- Haoran Ding, Marat Orazov. Anodically‐Generated Alkyl Radicals Derived from Carboxylic Acids as Reactive Intermediates for Addition to Alkenes. ChemElectroChem 2023, 10
(10)
https://doi.org/10.1002/celc.202201099
- Min Zhou, Jet Tsien, Ryan Dykstra, Jonathan M. E. Hughes, Byron K. Peters, Rohan R. Merchant, Osvaldo Gutierrez, Tian Qin. Alkyl sulfinates as cross-coupling partners for programmable and stereospecific installation of C(sp3) bioisosteres. Nature Chemistry 2023, 15
(4)
, 550-559. https://doi.org/10.1038/s41557-023-01150-z
- Maurice Metzler, Michael Bolte, Matthias Wagner, Hans-Wolfram Lerner. Crystal structure of [
t
BuMgCl]
2
[MgCl
2
(Et
2
O)
2
]
2. Acta Crystallographica Section E Crystallographic Communications 2023, 79
(4)
, 341-344. https://doi.org/10.1107/S2056989023002190
- Xinmin Hu, Xia Zhao, Xiangying Lv, Yan‐Bo Wu, Yuxiang Bu, Gang Lu. Ab Initio Metadynamics Simulations of Hexafluoroisopropanol Solvent Effects: Synergistic Role of Solvent H‐Bonding Networks and Solvent‐Solute C−H/π Interactions. Chemistry – A European Journal 2023, 29
(17)
https://doi.org/10.1002/chem.202203879
- Hayoung Song, Eunsung Lee. Revisiting the Reaction of IPr with Tritylium: An Alternative Mechanistic Pathway**. Chemistry – A European Journal 2023, 29
(12)
https://doi.org/10.1002/chem.202203364
- Alexander Düfert. Carbonylchemie. 2023, 39-186. https://doi.org/10.1007/978-3-662-65244-2_2
- Christopher E. Reimann, Kelly E. Kim, Alexander W. Rand, Farbod A. Moghadam, Brian M. Stoltz. What is a cross-coupling? An argument for a universal definition. Tetrahedron 2023, 130 , 133176. https://doi.org/10.1016/j.tet.2022.133176
- Jordan Rio, Lionel Perrin, Pierre‐Adrien Payard. Structure–Reactivity Relationship of Organozinc and Organozincate Reagents: Key Elements towards Molecular Understanding. European Journal of Organic Chemistry 2022, 2022
(44)
https://doi.org/10.1002/ejoc.202200906
- Simon Sengupta, Philipp Schüler, Helmar Görls, Phil Liebing, Sven Krieck, Matthias Westerhausen. In Situ Grignard Metalation Method for the Synthesis of Hauser Bases. Chemistry – A European Journal 2022, 28
(50)
https://doi.org/10.1002/chem.202201359
- Jennifer R. Lynch, Alan R. Kennedy, Jim Barker, Jacqueline Reid, Robert E. Mulvey. Crystallographic Characterisation of Organolithium and Organomagnesium Intermediates in Reactions of Aldehydes and Ketones. Helvetica Chimica Acta 2022, 105
(9)
https://doi.org/10.1002/hlca.202200082
- Lewis R. Thomas‐Hargreaves, Chantsalmaa Berthold, William Augustinov, Matthias Müller, Sergei I. Ivlev, Magnus R. Buchner. Reactivity of Diphenylberyllium as a Brønsted Base and Its Synthetic Application. Chemistry – A European Journal 2022, 28
(35)
https://doi.org/10.1002/chem.202200851
- Claudio Monasterolo, Ryan O'Gara, Saranna E. Kavanagh, Sadbh E. Byrne, Bartosz Bieszczad, Orla Murray, Michael Wiesinger, Rebecca A. Lynch, Kirill Nikitin, Declan G. Gilheany. Asymmetric addition of Grignard reagents to ketones: culmination of the ligand-mediated methodology allows modular construction of chiral tertiary alcohols. Chemical Science 2022, 13
(21)
, 6262-6269. https://doi.org/10.1039/D1SC06350B
- Daiki Kato, Tomoya Murase, Jalindar Talode, Haruki Nagae, Hayato Tsurugi, Masahiko Seki, Kazushi Mashima. Diarylcuprates for Selective Syntheses of Multifunctionalized Ketones from Thioesters under Mild Conditions. Chemistry – A European Journal 2022, 28
(26)
https://doi.org/10.1002/chem.202200474
- Rajeev Kumar, K.K. Bhasin, Jaspreet S. Dhau, Avtar Singh. Synthesis and characterization of 3-pyridylchalcogen compounds. Inorganic Chemistry Communications 2022, 139 , 109344. https://doi.org/10.1016/j.inoche.2022.109344
- Alisa S. Sunagatullina, Ferdinand H. Lutter, Paul Knochel. Herstellung von primären und sekundären Dialkylmagnesiumverbindungen durch eine radikalische I/Mg‐Austauschreaktion mit
s
Bu
2
Mg in Toluol. Angewandte Chemie 2022, 134
(13)
https://doi.org/10.1002/ange.202116625
- Alisa S. Sunagatullina, Ferdinand H. Lutter, Paul Knochel. Preparation of Primary and Secondary Dialkylmagnesiums by a Radical I/Mg‐Exchange Reaction Using
s
Bu
2
Mg in Toluene. Angewandte Chemie International Edition 2022, 61
(13)
https://doi.org/10.1002/anie.202116625
- Ewa Pietrasiak, Eunsung Lee. Grignard reagent formation
via
C–F bond activation: a centenary perspective. Chemical Communications 2022, 58
(17)
, 2799-2813. https://doi.org/10.1039/D1CC06753B
- Victoria S. Pfennig, Romina C. Villella, Julia Nikodemus, Carsten Bolm. Mechanochemical Grignard Reactions with Gaseous CO
2
and Sodium Methyl Carbonate**. Angewandte Chemie 2022, 134
(9)
https://doi.org/10.1002/ange.202116514
- Victoria S. Pfennig, Romina C. Villella, Julia Nikodemus, Carsten Bolm. Mechanochemical Grignard Reactions with Gaseous CO
2
and Sodium Methyl Carbonate**. Angewandte Chemie International Edition 2022, 61
(9)
https://doi.org/10.1002/anie.202116514
- Giovanni M. Fusi, Zelong Lim, Stephen D. Lindell, Enrique Gomez‐Bengoa, Malcolm R. Gordon, Silvia Gazzola. 2‐ and 6‐Purinylmagnesium Halides in Dichloromethane: Scope and New Insights into the Solvent Influence on the C−Mg Bond. European Journal of Organic Chemistry 2022, 2022
(7)
https://doi.org/10.1002/ejoc.202101009
- Odile Eisenstein. From the Felkin‐Anh Rule to the Grignard Reaction: an Almost Circular 50 Year Adventure in the World of Molecular Structures and Reaction Mechanisms with Computational Chemistry**. Israel Journal of Chemistry 2022, 62
(1-2)
https://doi.org/10.1002/ijch.202100138
- Gantulga Norjmaa, Gregori Ujaque, Agustí Lledós. Beyond Continuum Solvent Models in Computational Homogeneous Catalysis. Topics in Catalysis 2022, 65
(1-4)
, 118-140. https://doi.org/10.1007/s11244-021-01520-2
- Sudip Baguli, Sumana Mondal, Chhotan Mandal, Santu Goswami, Debabrata Mukherjee. Cyclopentadienyl Complexes of the Alkaline Earths in Light of the Periodic Trends. Chemistry – An Asian Journal 2022, 17
(1)
https://doi.org/10.1002/asia.202100962
- S. Chantal E. Stieber. Computational Methods in Organometallic Chemistry. 2022, 176-210. https://doi.org/10.1016/B978-0-12-820206-7.00099-8
- Maren Podewitz. Trendbericht Theoretische Chemie 2/2: Mit dem Computer zu effizienteren Katalysatoren. Nachrichten aus der Chemie 2021, 69
(11)
, 60-62. https://doi.org/10.1002/nadc.20214119408
- George R. M. Dowson, Joshua Cooper, Peter Styring. Reactive capture using metal looping: the effect of oxygen. Faraday Discussions 2021, 230 , 292-307. https://doi.org/10.1039/D1FD00001B
- Agustí Lledós. Computational Organometallic Catalysis: Where We Are, Where We Are Going. European Journal of Inorganic Chemistry 2021, 2021
(26)
, 2547-2555. https://doi.org/10.1002/ejic.202100330
- Pietro Vidossich, Marco De Vivo. The role of first principles simulations in studying (bio)catalytic processes. Chem Catalysis 2021, 1
(1)
, 69-87. https://doi.org/10.1016/j.checat.2021.04.009
- Eamonn F. Healy. Organic chemistry as representation. Foundations of Chemistry 2021, 23
(1)
, 59-68. https://doi.org/10.1007/s10698-020-09379-z
- Friederike Ratsch, Joss Pepe Strache, Waldemar Schlundt, Jörg‐Martin Neudörfl, Andreas Adler, Sarwar Aziz, Bernd Goldfuss, Hans‐Günther Schmalz. Enantioselective Cleavage of Cyclobutanols Through Ir‐Catalyzed C−C Bond Activation: Mechanistic and Synthetic Aspects. Chemistry – A European Journal 2021, 27
(14)
, 4640-4652. https://doi.org/10.1002/chem.202004843
- Jan Paradies, Jennifer Andexer, Uwe Beifuss, Florian Beuerle, Malte Brasholz, Rolf Breinbauer, Martin Ernst, Ruth Ganardi, Tobias A. M. Gulder, Wolfgang Hüttel, Stephanie Kath‐Schorr, Karsten Körber, Markus Kordes, Matthias Lehmann, Thomas Lindel, Burkhard Luy, Christian Mück‐Lichtenfeld, Claudia Muhle‐Goll, Jochen Niemeyer, Roland Pfau, Jörg Pietruszka, Johannes L. Röckl, Norbert Schaschke, Mathias O. Senge, Bernd F. Straub, Siegfried R. Waldvogel, Thomas Werner, Daniel B. Werz, Christian Winter. Organische Chemie. Nachrichten aus der Chemie 2021, 69
(3)
, 38-68. https://doi.org/10.1002/nadc.20214105947
- Ferran Planas, Stefanie V. Kohlhepp, Genping Huang, Abraham Mendoza, Fahmi Himo. Computational and Experimental Study of Turbo‐Organomagnesium Amide Reagents: Cubane Aggregates as Reactive Intermediates in Pummerer Coupling. Chemistry – A European Journal 2021, 27
(8)
, 2767-2773. https://doi.org/10.1002/chem.202004164
- Jorge Espina-Casado, Alfonso Fernández-González, Marta E. Díaz-García, Rosana Badía-Laíño. Smart carbon dots as chemosensor for control of water contamination in organic media. Sensors and Actuators B: Chemical 2021, 329 , 129262. https://doi.org/10.1016/j.snb.2020.129262
- Katharina Dilchert, Michelle Schmidt, Angela Großjohann, Kai‐Stephan Feichtner, Robert E. Mulvey, Viktoria H. Gessner. Lösungsmitteleinflüsse auf die Struktur und Stabilität von Alkalimetallcarbenoiden. Angewandte Chemie 2021, 133
(1)
, 498-504. https://doi.org/10.1002/ange.202011278
- Katharina Dilchert, Michelle Schmidt, Angela Großjohann, Kai‐Stephan Feichtner, Robert E. Mulvey, Viktoria H. Gessner. Solvation Effects on the Structure and Stability of Alkali Metal Carbenoids. Angewandte Chemie International Edition 2021, 60
(1)
, 493-498. https://doi.org/10.1002/anie.202011278
- Janus J. Eriksen, Stella Stopkowicz, Thomas-C. Jagau, Trygve Helgaker. Foreword: Prof. Gauss Festschrift. Molecular Physics 2020, 118
(19-20)
, e1817247. https://doi.org/10.1080/00268976.2020.1817247
- Philipp C. Stegner, Christian A. Fischer, D. Thao Nguyen, Andreas Rösch, Johanne Penafiel, Jens Langer, Michael Wiesinger, Sjoerd Harder. Intramolecular Alkene Hydroamination with Hybrid Catalysts Consisting of a Metal Salt and a Neutral Organic Base. European Journal of Inorganic Chemistry 2020, 2020
(35)
, 3387-3394. https://doi.org/10.1002/ejic.202000671
- Odile Eisenstein, Gregori Ujaque, Agustí Lledós. What Makes a Good (Computed) Energy Profile?. 2020, 1-38. https://doi.org/10.1007/3418_2020_57
Abstract
Figure 1
Figure 1. Schematic representation of possible mechanisms for the Grignard reaction: (a) polar mechanism, heterolytic Mg–C bond breaking, with subsequent formation of a nucleophilic carbon that adds to the electrophilic carbonyl carbon; or (b) radical mechanism, homolytic Mg–C bond breaking, with subsequent recombination of the species with unpaired electrons.
Figure 2
Figure 2. Organomagnesium complexes considered as reactants for the polar mechanism, classified as a function of the relative positions of the substrate (ACA) and nucleophile (methyl group). The labels geminal, vicinal, and bridging describe the initial position of the reactive groups with respect to the Mg center(s). The activation free energies, ΔA⧧, defined as the difference in free energy between the TS and the related reactant species, are given in kcal mol–1.
Figure 3
Figure 3. Geminal reaction for compound Bgem. Representative geometries for the reactant, transition state, and product. The Mg atom and its ligands are represented as balls-and-sticks. Other solvating THF molecules are drawn as lines. The color codes are mauve for magnesium, red for oxygen, gray for carbon, and white for hydrogen.
Figure 4
Figure 4. Transition state structures of the geminal and vicinal reactions (Bgem, Fvic). Left: Evidence of the structural similarity of the two TSs. Balls-and-sticks are used to represent the reacting moiety, while the rest of the system is drawn as transparent cylinders. THF molecules not bound to Mg are not shown. The green line marks the incipient C–CH3 bond and its associated distance. Top right: Schematic definition of the angles formed by the C3 axis of the nucleophile methyl group. Bottom right: Distances and angles formed by the atoms involved in the four-center TS. Distances are reported in angstroms, and angles are in degrees. ϕX is the angle at the vertex X in the four-member ring (top right panel). The color code is as in Figure 3.
Figure 5
Figure 5. Vicinal reaction for Fvic. Representative geometries for the reactant, transition state, and product. The Mg atoms (left Mg1, right Mg2) and first coordination sphere ligands are represented as balls-and-sticks. Other solvating THF molecules are drawn as lines. The color code is cyan for chlorine and green for the nucleophilic methyl, while the other elements are colored as in Figure 3.
Figure 6
Figure 6. Red: Activation energies of all Grignard species derived from CH3MgCl in THF solution when reacting with acetaldehyde. Blue: Relative free energies of the species involved in the Schlenk equilibrium (data from ref (31)). The green box highlights the intrinsically most reactive species identified in this work.
Figure 7
Figure 7. Mg–CH3 BDE and spin-density localization (in fractions of e) for MgCl(Sub)(THF)nCH3 (Sub = substrate). The green wire-frame shows the isosurface of the spin density map at a value of 0.0065 au.
Figure 8
Figure 8. Energetics of the Grignard reaction mechanisms for fluorenone. Top panel: Activation free energy of the nucleophilic addition with Mg(CH3)2. Bottom panel: Bond dissociation energy. All energy values are in kcal mol–1. On the right is the distribution of the spin density in Mg(CH3)(fluorenone)THF2: 33% of the unpaired electron localizes on the carbonyl carbon, 16% on the carbonyl oxygen, and 45% on the remaining π system.
Figure 9
Figure 9. Structural reorganization from (left) reactant to (right) transition state in the nucleophilic pathway for (top) ACA and (bottom) fluorenone. The CH3–Mg–O–C dihedral angle (corresponding to the blue arrow) is 14 ± 15° for ACA and 114 ± 30° for fluorenone.
References
ARTICLE SECTIONSThis article references 65 other publications.
- 1Vollhardt, K.; Schore, N. Organic Chemistry: Structure and Function; W. H. Freeman: New York, 2014.Google ScholarThere is no corresponding record for this reference.
- 2Grignard, V. C. Sur quelques nouvelles combinaisons organométalliques du magnésium et leur application à des synthèses daalcools et d’hydrocarbures. Compt. Rend. Hebd. Séances Acad. Sci. 1900, 130, 1322– 24Google Scholar2https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaD28XitV2ltb0%253D&md5=0883715fab5d100533088de258b7c324About some new organo-metallic compounds of the magnesium and their application to the synthesis of alcohols and hydrocarbons. [machine translation]Grignard, V.Comptes Rendus Hebdomadaires des Seances de l'Academie des Sciences (1900), 130 (), 1322-24CODEN: COREAF; ISSN:0001-4036.[Machine Translation of Descriptors]. During effect of CH3J on magnesium powders in presence of anhydrous ether, an organo-metallic compound forms under a very lively reaction. The Mg dissolves gradually, and finally receives a clear liquid, which leaves a gray, crystalline, hygroscopic mass with the evaporation of the ether. One sets to that more ether solution which on 1 atom Mg of 1 mole CH3J contains 1 mole an aldehyde or a ketone, then takes place under a very lively reaction for formation of an organo-metallic compound, which supplies the appropriate secondary or tertiary alcohol with decomposing with acidified water with an yield of about 70%. The reaction takes place after the subsequent equations. 1. CH3I + Mg = CH3MgI. 2. CH3MgI + RCHO = R. CH (CH3) OMgI. 3. R. CH (CH3) OMgI + H2O = R. CH (OH) . CH3 + MgIOH. The bromine and iodine compounds of the saturated and unsaturated alkyls in the same way benzyl bromide give appropriate alcohol in same way. Author represented the subsequent alcohols in this way: Phenylisobutylcarbinol, C6H5. CH (OH): C4H9, from benzaldehyde, Isobutylbromide and Mg are colorless and weakly smelling liquid, Kp9, 122°. Dimethylphenylcarbinol, C6H5. C (OH) (CH3) 2, from acetophenone, Mg and CH3I, are pleasantly smelling and colorless liquid, Kp10, 93-95°. Dimethylbenzylcarbinol, C6H5. CH2. C (OH) (CH3) 2, from acetone, benzyl bromide and Mg are weakly smelling and colorless liquid, Kp10, 103-105°. With application of unsaturated aldehydes and ketones, yields with the double bond of the cobalt group and the developing alcohol is sometimes changeable. It splits off, then with the distillation H2O, and an unsaturated hydrocarbon develops. Thus the Dimethyl-2.4-pentadien-2.4 becomes, CH3. during effect of Mesityloxide on magnesium methyl iodide; C (CH3): CH. C (CH3): CH3, Kp750, 92-93°.
- 3Rappoport, Z., Marek, I., Eds. The Chemistry of Organomagnesium Compounds; Wiley-VCH: Weinheim, Germany, 2008.Google ScholarThere is no corresponding record for this reference.
- 4Schlenk, W.; Schlenk, W. Über die Konstitution der Grignardschen Magnesiumverbindungen. Ber. Dtsch. Chem. Ges. B 1929, 62, 920– 924, DOI: 10.1002/cber.19290620422Google Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaB1MXksFSitg%253D%253D&md5=30432803972ea7f1a470a7245fbf7813The constitution of the Grignard magnesium derivativesSchlenk, W.; Schlenk, Wilh., Jr.Berichte der Deutschen Chemischen Gesellschaft [Abteilung] B: Abhandlungen (1929), 62B (), 920-4CODEN: BDCBAD; ISSN:0365-9488.Several Grignard compds. have been prepd., then fractionally pptd. from their Et2O soln. by means of O(CH2CH2)2O. The Mg:X ratios of the fractions have been examd. Grignard compds. must be represented by 2RMgX .dblharw. MgR2 + MgX2. For EtI, the compn. of the Grignard deriv. would be: 6EtMgI + 4MgEt2 + 4MgI2. For PhBr: PhMgBr + 0.115MgPh2 + 0.115MgBr2.
- 5Seyferth, D. The Grignard Reagents. Organometallics 2009, 28, 1598– 1605, DOI: 10.1021/om900088zGoogle Scholar5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXivVeis70%253D&md5=e0c332b6494bdecec6bd4bcfce564ee0The Grignard ReagentsSeyferth, DietmarOrganometallics (2009), 28 (6), 1598-1605CODEN: ORGND7; ISSN:0276-7333. (American Chemical Society)A review of prepn. and reactions of Grignard reagents.
- 6Silverman, G. S.; Rakita, P. E. Handbook of Grignard Reagents; CRC Press: New York, 1996.Google ScholarThere is no corresponding record for this reference.
- 7Robertson, S. D.; Uzelac, M.; Mulvey, R. E. Alkali-metal-mediated synergistic effects in polar main group organometallic chemistry. Chem. Rev. 2019, 119, 8332– 8405, DOI: 10.1021/acs.chemrev.9b00047Google Scholar7https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXltVWrtro%253D&md5=f40eeff96aacc4152b568c0b79d6fa78Alkali-Metal-Mediated Synergistic Effects in Polar Main Group Organometallic ChemistryRobertson, Stuart D.; Uzelac, Marina; Mulvey, Robert E.Chemical Reviews (Washington, DC, United States) (2019), 119 (14), 8332-8405CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review. The development of synthetic chem. since the early 1900s owes much to the service of organolithium reagents. Brilliant bases (e.g., deprotonating C-H bonds), nucleophiles (e.g., adding to unsatd. mols.), and transfer agents (e.g., delivering ligands to other metals), these versatile virtuosi and to a lesser extent the org. derivs. of the other common alkali metals sodium and potassium have proved indispensable in both academia and technol. Today these monometallic compds. are still utilized widely in synthetic campaigns, but in recent years they have been joined by an assortment of bimetallic formulations that also contain an alkali metal but in company with another metal. These bimetallic formulations often exhibit unique chem. that can be interpreted in terms of synergistic effects, for which the alkali metal is essential, though it is often the second metal that performs the synthetic transformation. Here, this "alkali-metal-mediated" chem. is surveyed focusing mainly on bimetallic formulations contg. two alkali metals or an alkali metal paired with magnesium, calcium, zinc, aluminum, or gallium. In this International Year of the Periodic Table (IYPT), we ponder whether a Periodic Table of Element Pairs will emerge in the future.
- 8Fauvarque, J.; Rouget, E. Compt. Rend. Hebd. Séances Acad. Sci., Ser. C 1968, 257, 1355Google ScholarThere is no corresponding record for this reference.
- 9Blomberg, C.; Mosher, H. A radical process in a reaction of a Grignard compound. J. Organomet. Chem. 1968, 13, 519– 522, DOI: 10.1016/S0022-328X(00)82781-3Google Scholar9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaF1cXkvVKjs7s%253D&md5=79ed18eeed10bd80dccd480a57115d24A radical process in a reaction of a Grignard compoundBlomberg, Cornelis; Mosher, Harry S.Journal of Organometallic Chemistry (1968), 13 (2), 519-22CODEN: JORCAI; ISSN:0022-328X.Chem. and E.S.R. evidence is reported for the occurrence of radicals during the reaction of the Grignard reagent from neopentyl chloride with benzophenone in tetrahydrofuran; 1,1-diphenyl-3,3-dimethylbutanol and benzopinacol were isolated and the formation of neopentane was observed. The steric bulk of the neopentyl group (R') so retards the normal addn. reaction [Ph2CO:Mg(R') X → Rh2R'COMgX] that the radicals, formed in the process [Ph2CO:Mg(R')X → Ph2C·OMgX + R·'], are able to escape from the solvent cage. The neopentyl radical can then react with the solvent to give R' and the ketyl can dimerize to give the Mg halide salt of benzopinacol.
- 10Blomberg, C.; Grootveld, H.; Gerner, T.; Bickelhaupt, F. Radical formation during reactions of Grignard reagents with quinones. J. Organomet. Chem. 1970, 24, 549– 553, DOI: 10.1016/S0022-328X(00)84483-6Google Scholar10https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE3cXltV2rs7s%253D&md5=b149433224f258111263fa1ea8aa0dafRadical formation during reactions of Grignard reagents with quinonesBlomberg, Cornelis; Grootveld, H. H.; Gerner, T. H.; Bickelhaupt, F.Journal of Organometallic Chemistry (1970), 24 (3), 549-53CODEN: JORCAI; ISSN:0022-328X.In dil. soln. the addn. of PhMgBr to acenaphthenequinone and phenanthrenequinone (I) leads by single electron transfer to the formation of the corresponding semiquinones, which can be identified by ESR. I yields ∼20% 9-hydroxy-10-phenoxyphenanthrene, the formation of which is explained by combination of the initially formed semiquinone and Ph radicals. The analogy with photochem. reactions and the possible occurrence of similar reactions in organomagnesium and organozinc chemistry are briefly discussed.
- 11Ashby, E. C.; Nackashi, J.; Parris, G. E. Composition of Grignard compounds. X. NMR, IR, and molecular association studies of some methylmagnesium alkoxides in diethyl ether, tetrahydrofuran, and benzene. J. Am. Chem. Soc. 1975, 97, 3162– 3171, DOI: 10.1021/ja00844a040Google Scholar11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE2MXksFOrurs%253D&md5=92a29e90c3ba2d15dc364ae2a5347ffcComposition of Grignard compounds. X. NMR, ir, and molecular association studies of some methylmagnesium alkoxides in diethyl ether, tetrahydrofuran, and benzeneAshby, E. C.; Nackashi, J.; Parris, G. E.Journal of the American Chemical Society (1975), 97 (11), 3162-71CODEN: JACSAT; ISSN:0002-7863.Mol. assocn. of MeMgOR (R = OCPh2Me, OCMe3, OCHMe2, OPr) in Et2O, THF, and C6H6 was examd. using ir and NMR spectral data. The steric bulk of the alkoxy group and the coordinating ability of the solvent determine the thermodynamically preferred soln. compn. In THF, solvated dimers are preferred. In Et2O, linear oligomers and cubane tetramers are preferred provided the alkoxy group is not bulkier than the tert-butoxy group. In C6H6, cubane tetramers are obsd. for alkoxy groups of intermediate bulk such as tert-butoxy and isopropoxy, but the less bulky n-propoxy group permits the formation of an oligomer contg. seven to nine monomer units. For the reagents with alkoxy groups less bulky than tert-butoxy, the equilibria involving various structures are established very rapidly. However, the dimer-linear oligomer ↹ cubane tetramer equilibrium is established very slowly for methylmagnesium tert-butoxide compds. The cubane form is very inert and does not exchange or otherwise interact with Me2Mg in Et2O. The dimer-linear oligomer form is quite labile and readily exchanges with Me2Mg forming mixed-bridged compds. However, in Et2O, the mixed bridge is not sufficiently strong to prevent slow conversion of methylmagnesium tert-butoxide to the cubane form thus releasing Me2Mg.
- 12Ashby, E. C.; Lopp, I. G.; Buhler, J. D. Mechanisms of Grignard reactions with ketones. Polar vs. single electron transfer pathways. J. Am. Chem. Soc. 1975, 97, 1964– 1966, DOI: 10.1021/ja00840a066Google Scholar12https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE2MXhsFCrs78%253D&md5=9c20322ef87fe6aa0b10e5a3826f32b5Mechanisms of Grignard reactions with ketones. Polar vs. single electron transfer pathwaysAshby, E. C.; Lopp, Irene G.; Buhler, Jerry D.Journal of the American Chemical Society (1975), 97 (7), 1964-6CODEN: JACSAT; ISSN:0002-7863.The addn. of MeMgBr to 2-methylbenzophenone in Et2O proceeds via a normal polar mechanism, whereas in the presence of small amts. of transition metal catalysts (e.g. 0.05 mole % FeCl3) the reaction proceeds via a single electron transfer pathway. The role of an intermediate (I) in by-product formation is examd.
- 13Ashby, E. C. A detailed description of the mechanism of reaction of Grignard reagents with ketones. Pure Appl. Chem. 1980, 52, 545– 569, DOI: 10.1351/pac198052030545Google Scholar13https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3cXksVGgs7k%253D&md5=6b83e73f4ac9654be92428627d98a4b4A detailed description of the mechanism of reaction of Grignard reagents with ketonesAshby, E. C.Pure and Applied Chemistry (1980), 52 (3), 545-69CODEN: PACHAS; ISSN:0033-4545.A review, chiefly of the work of the author, with 29 refs.
- 14Ashby, E. C.; Bowers, J. R. Organometallic reaction mechanisms. 17. Nature of alkyl transfer in reactions of Grignard reagents with ketones. Evidence for radical intermediates in the formation of 1,2-addition product involving tertiary and primary Grignard reagents. J. Am. Chem. Soc. 1981, 103, 2242– 2250, DOI: 10.1021/ja00399a018Google Scholar14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL3MXitVyjtbY%253D&md5=894ebdd380e00f11030a6aee19f48826Organometallic reaction mechanisms. 17. Nature of alkyl transfer in reactions of Grignard reagents with ketones. Evidence for radical intermediates in the formation of 1,2-addition product involving tertiary and primary Grignard reagentsAshby, E. C.; Bowers, Joseph R., Jr.Journal of the American Chemical Society (1981), 103 (9), 2242-50CODEN: JACSAT; ISSN:0002-7863.The title 1,2-addn. products (formed after dissocn. of the radical anion-radical cation pair) show free-radical character, as indicated by the cyclized 1,2-addn. products formed from the reaction of a tertiary Grignard reagent probe with Ph2CO in THF and from the reaction of a primary Grignard reagent probe (neooctenyl Grignard reagent) with Ph2CO in Et2O. The 1,6-addn. products show free-radical character, as evidenced by the cyclized 1,6-addn. products formed in all of the reactions which involve the tertiary Grignard reagent probe (in all solvents studied) with Ph2CO and 2-MeC6H4COPh (I) and also in the reaction of the neooctenyl Grignard reagent probe with I.
- 15Lund, T.; Pedersen, M. L.; Frandsen, L. A. Does the reaction between fluorenone and grignard reagents involve free fluorenone anion radicals?. Tetrahedron Lett. 1994, 35, 9225– 9226, DOI: 10.1016/0040-4039(94)88472-2Google Scholar15https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXislagsrw%253D&md5=7a16badf7a3077c3a004c11eab03afeaDoes the reaction between fluorenone and Grignard reagents involve free fluorenone anion radicals?Lund, Torben; Pedersen, Morten L.; Frandsen, Lars A.Tetrahedron Letters (1994), 35 (49), 9225-6CODEN: TELEAY; ISSN:0040-4039. (Elsevier)The ratio between 1,6- and 1,2-addn. in the reactions of electrogenerated fluorenone anion radicals with RX in THF were similar to the ratio obtained in the Grignard reaction of fluorenone with RMgX in THF. This indicates that the addn. products in the Grignard reaction may be obtained via the coupling of freely diffusing fluorenone anion radicals with R radicals.
- 16Blomberg, C.; Salinger, R. M.; Mosher, H. S. Reaction of Grignard reagent from neopentyl chloride with benzophenone. A nuclear magnetic resonance study. J. Org. Chem. 1969, 34, 2385– 2388, DOI: 10.1021/jo01260a028Google Scholar16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaF1MXks12qsbw%253D&md5=ff6e7144106079007095c9c65674f978Reaction of the Grignard reagent from neopentyl chloride with benzophenone. Nuclear magnetic resonance study.Blomberg, Cornelis; Salinger, Rudolf M.; Mosher, Harry S.Journal of Organic Chemistry (1969), 34 (8), 2385-8CODEN: JOCEAH; ISSN:0022-3263.The rate of change in the N.M.R. spectrum of a mixt. of the Grignard reagent from neopentyl chloride and benzophenone in tetrahydrofuran was studied. The expected addn. reaction was complicated by the simultaneous occurrence of a radical reaction to produce neopentane and benzopinacol. These spectra are interpretable in terms of a reaction to give an initial product which in turn undergoes further reaction with the Grignard reagent to give a new reactive species. This reactive intermediate presumably is either the alkylmagnesium alkoxide (RMgOR') or a complex of the initial product with the Grignard reagent (RMgCl.R'OMgCl). This constitutes a direct observation of a process often postulated in the reaction of a Grignard compd. Qual. generalizations could be made but because of these complications it was not possible to make a quant. kinetic anal. of the data.
- 17Hoffmann, R. W.; Hölzer, B. Concerted and stepwise Grignard additions, probed with a chiral Grignard reagent. Chem. Commun. 2001, 491– 492, DOI: 10.1039/b009678oGoogle Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXhtl2ntLw%253D&md5=44546c2d93ae9b98c42f9d2b65f6033dConcerted and stepwise Grignard additions, probed with a chiral Grignard reagentHoffmann, Reinhard W.; Holzer, BettinaChemical Communications (Cambridge, United Kingdom) (2001), (5), 491-492CODEN: CHCOFS; ISSN:1359-7345. (Royal Society of Chemistry)The Grignard reagent (S)-PhCH2CH(MgCl)CH2CH3 6, in which the magnesium-bearing carbon atom is the sole stereogenic center has been added to CO2, PhNCO, PhNCS and certain aldehydes with full retention of configuration. Reaction with benzophenone, electron-deficient aldehydes and several allyl halides proceeded with partial or complete racemization. The findings are discussed with respect to a dichotomy between concerted polar and stepwise SET reaction pathways.
- 18Hoffmann, R. W. The quest for chiral Grignard reagents. Chem. Soc. Rev. 2003, 32, 225– 230, DOI: 10.1039/b300840cGoogle Scholar18https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXkt1emtL8%253D&md5=58237aeb3e5721d0f3d00c65d99db0b0The quest for chiral Grignard reagentsHoffmann, Reinhard W.Chemical Society Reviews (2003), 32 (4), 225-230CODEN: CSRVBR; ISSN:0306-0012. (Royal Society of Chemistry)A review on the prepn. of chiral Grignard reagents. The involvement of single electron transfer, i.e. of free radicals in the reactions of organomagnesium reagents could be detected with the aid of a chiral secondary Grignard reagent, in which the magnesium-bearing carbon atom is the sole stereogenic center. So far, however, such reagents have not been accessible, because the std. prepn. of Grignard reagents proceeds via free radicals. The authors review and summarize here their efforts to generate (S)-1-Benzylpropylmagnesium chloride 36 by asym. synthesis starting from an enantiomerically pure compd. ArSOCH(Cl)CH2Ph (Ar = p-cl-C6H5) 30b using a sulfoxide/magnesium exchange reaction to give trans-β-chloro-α-phenylbenzenepropanol 33 followed by a carbenoid homologation reaction. Grignard reagent 36 turned out to be an ideal probe to learn about the extent to which SET is involved in reactions of organomagnesium reagents.
- 19Gajewski, J. J.; Bocian, W.; Harris, N. J.; Olson, L. P.; Gajewski, J. P. Secondary deuterium kinetic isotope effects in irreversible additions of hydride and carbon nucleophiles to aldehydes: a spectrum of transition states from complete bond formation to single electron transfer. J. Am. Chem. Soc. 1999, 121, 326– 334, DOI: 10.1021/ja982504rGoogle Scholar19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXjsFyi&md5=57e98328aac6eabcdfae331f64ccb1e7Secondary deuterium kinetic isotope effects in irreversible additions of hydride and carbon nucleophiles to aldehydes: a spectrum of transition states from complete bond formation to single electron transferGajewski, Joseph J.; Bocian, Wojciech; Harris, Nathan J.; Olson, Leif P.; Gajewski, John P.Journal of the American Chemical Society (1999), 121 (2), 326-334CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The competitive kinetics of hydride and organometallic addns. to PhCHO and PhCDO were detd. at -78° using LiAlH4, LiBHEt3, NaBH4, LiBH4, LiAlH(OCMe3)3, NaBH(OMe)3, NaBH(OAc)3 (at 20°), RMgBr (R = Me, Ph, allyl), and RLi (R = Me, Ph, Bu, Me3C, allyl). The hydride addns. had an inverse secondary D kinetic isotope effects in all cases, but the magnitude of the effect varied inversely with the apparent reactivity of the hydride. In the addns. of MeMgBr and of MeLi and PhLi, inverse secondary D isotope effects were obsd.; little if any isotope effect was obsd. with PhMgBr, BuLi or Me3CLi. With CH2:CHCH2M (M = MgBr, Li), a normal secondary D kinetic isotope effect was obsd. Rate-detg. single-electron transfer occurs with allyl reagents, but direct nucleophilic reaction occurs with all of the other reagents, with the extent of bond formation depending on the reactivity of the reagent. In the addn. of MeLi to cyclohexanecarboxaldehyde (I), a less inverse secondary D kinetic isotope effect was obsd. than that obsd. in the addn. of MeLi to PhCHO, and allyllithium addn. to I had a kinetic isotope effect near unity. The data with organometallic addns., which are not incompatible with observations of carbonyl C isotope effects, suggest that electrochem. detd. redox potentials which indicate endoergonic electron transfer with energies .ltorsim.13 kcal/mol allow electron-transfer mechanisms to compete well with direct polar addns. to aldehydes, provided that the reagent is highly stabilized, like allyl species. MeLi, PhLi, MeMgBr and PhMgBr are estd. to undergo electron transfer with endoergonicities >30 kcal/mol with PhCHO, so these react by direct polar addns. A working hypothesis is that BuLi reagents undergo polar addns., despite redox potentials which indicate ≤13 kcal/mol endoergonic electron transfer, because of the great exoergonicity assocd. with the 2-electron addn., which is responsible for a low barrier for polar reactions.
- 20Otte, D. A. L.; Woerpel, K. A. Evidence that Additions of Grignard Reagents to Aliphatic Aldehydes do Not Involve Single-Electron-Transfer Processes. Org. Lett. 2015, 17, 3906– 3909, DOI: 10.1021/acs.orglett.5b01893Google Scholar20https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXht1GnurrO&md5=ae186bb421050b8bc1d639a7ccb83cb6Evidence that Additions of Grignard Reagents to Aliphatic Aldehydes Do Not Involve Single-Electron-Transfer ProcessesOtte, Douglas A. L.; Woerpel, K. A.Organic Letters (2015), 17 (15), 3906-3909CODEN: ORLEF7; ISSN:1523-7052. (American Chemical Society)Addn. of allylmagnesium reagents to an aliph. aldehyde bearing a radical clock gave only addn. products and no evidence of ring-opened products that would suggest single-electron-transfer reactions. The analogous Barbier reaction also did not provide evidence for a single-electron-transfer mechanism in the addn. step. Other Grignard reagents (methyl-, vinyl-, t-Bu-, and triphenylmethylmagnesium halides) also do not appear to add to an alkyl aldehyde by a single-electron-transfer mechanism.
- 21Garst, J. F.; Soriaga, M. P. Grignard reagent formation. Coord. Chem. Rev. 2004, 248, 623– 652, DOI: 10.1016/j.ccr.2004.02.018Google Scholar21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXksF2ku7s%253D&md5=ccb7be559b5068df60420274187c985aGrignard reagent formationGarst, John F.; Soriaga, Manuel P.Coordination Chemistry Reviews (2004), 248 (7-8), 623-652CODEN: CCHRAM; ISSN:0010-8545. (Elsevier Science B.V.)A review with 74 refs. Probably reactions of Mg metal with org. halides RX in ether solvents are typical metallic corrosions in which the stabilization of Mg2+, substantially through its coordination by the solvent, drives its loss from the metal and consequently the redns. of RX and reaction intermediates such as R· at the metal surface. Although alkyl halides form Grignard reagents through nonchain mechanisms in which intermediate radicals diffuse in soln., very small amts. of radical isomerization occur in Grignard reactions of certain vinyl and aryl halides, even when intermediate radicals R· would isomerize very rapidly. This suggests a dominant nonradical mechanism for these vinyl and aryl halides or a mechanism in which intermediate radicals R· have extremely short lifetimes. Since the former seems more likely, a dianion mechanism, through a transition state [RX2-]‡, is proposed. Surface studies of polycryst. Mg show that the oxide layer is mostly Mg(OH)2 and that it is mech. passivating. In the absence of promoters, Grignard reactions occur very slowly until enough RX has seeped to the Mg surface and reacted there to undercut and cause the Mg(OH)2 layer to flake off.
- 22Chen, Z.-N.; Fu, G.; Xu, X. Theoretical studies on Grignard reagent formation: radical mechanism versus non-radical mechanism. Org. Biomol. Chem. 2012, 10, 9491– 9500, DOI: 10.1039/c2ob26658jGoogle Scholar22https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xhs1yks7fF&md5=3a16a90835cdf9ce691ba0aba4f9e968Theoretical studies on Grignard reagent formation: radical mechanism versus non-radical mechanismChen, Zhe-Ning; Fu, Gang; Xu, XinOrganic & Biomolecular Chemistry (2012), 10 (47), 9491-9500CODEN: OBCRAK; ISSN:1477-0520. (Royal Society of Chemistry)Here the systematic theor. study on the mechanisms of Grignard reagent formation is presented (GRF) for CH3Cl reacting with Mg atom, Mg2 and Mg clusters (Mg4-Mg20). The calcns. reveal that the ground state Mg atom is inactive under matrix condition, whereas it is active under metal vapor synthesis (MVS) conditions. However, the excited state Mg (3P) atom, as produced by laser-ablation, can react with CH3Cl without barriers, and hence is active under matrix condition. The prediction is that the bimagnesium Grignard reagent, though often proposed, can barely be obsd. exptl., due to its high reactivity towards addnl. CH3Cl to produce more stable Grignard reagent dimer, and that the cluster Grignard reagent RMg4X possesses a flat Mg4 unit rather than a tetrahedral geometry. The calcns. further reveal that the radical pathway (T4) is prevalent on Mg, Mg2 and Mgn clusters of small size, while the no-radical pathway (T2), which starts at Mg4, becomes competitive with T4 as the cluster size increases. A structure-reactivity relation between barrier heights and ionization potentials of Mgn is established. These findings not only resolve controversy in expt. and theory, but also provide insights which can be used in the design of effective synthesis approaches for the prepn. of chiral Grignard reagents.
- 23Shao, Y.; Liu, Z.; Huang, P.; Liu, B. A unified model of Grignard reagent formation. Phys. Chem. Chem. Phys. 2018, 20, 11100– 11108, DOI: 10.1039/C8CP01031EGoogle Scholar23https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXltFWqurw%253D&md5=0468205a9e6ba31fd7aeff16043280ccA unified model of Grignard reagent formationShao, Yunqi; Liu, Zhen; Huang, Pan; Liu, BopingPhysical Chemistry Chemical Physics (2018), 20 (16), 11100-11108CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)Grignard reagents are among the most fundamental reagents in org. synthesis, yet studies have hitherto failed to fully explain the selectivity and kinetics of Grignard reagent formation (GRF). The present study provides new insights into the intermediates and pathways of GRF using d. functional theory (DFT) calcns. Potential energy surfaces of RX dissocn. along different directions reveal the origin of configuration retention of alkenyl and arom. halides. Radical intermediates participate solely in the dissocn. stage, and depend on the geometry of the reactant halide. Dissocn. of org. halides yields stabilized surface anions, and the rest of the reaction is ionic in nature. MgX+/RMg+ are proposed as the key intermediates of Mg leaving from the surface in the self-activation of GRF, which explains the accelerated kinetics upon addn. of RMgX or MgX2. The intermediacy of the cations was supported by a simple electrochem. expt. To the best of the authors' knowledge, this is the 1st unified ionic model (I-model) developed for resolving the controversial issues of GRF.
- 24Axten, J.; Troy, J.; Jiang, P.; Trachtman, M.; Bock, C. W. An ab initio molecular orbital study of the Grignard reagents CH3MgCl and [CH3MgCl]2: the Schlenk equilibrium. Struct. Chem. 1994, 5, 99– 108, DOI: 10.1007/BF02265351Google Scholar24https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2cXmtlSrs7s%253D&md5=62a601c4a2da0528a295f40433e172e7An ab initio molecular orbital study of the Grignard reagents CH3MgCl and [CH3MgCl]2: The Schlenk equilibriumAxten, Jeffrey; Troy, Jennifer; Jiang, Peter; Trachtman, Mendel; Bock, Charles W.Structural Chemistry (1994), 5 (2), 99-108CODEN: STCHES; ISSN:1040-0400.Ab initio MO calcns. are used to study the modified Schlenk equil.: 2RMgCl .dblharw. MgR2 + MgCl2 .dblharw. Mg(Cl2)MgR2 with R = H and CH3. In the absence of any solvents, calcns. indicate that the formation of the various possible bridged dimers (RMgCl)2 is substantially exothermic. However, using di-Me ether as a model solvent, we show that the formation of the dimer (Me2O)(CH3)Mg(μ-Cl2)Mg(CH3)(OMe2) is exothermic only when entropic effects are included.
- 25Ehlers, A. W.; van Klink, G. P. M.; van Eis, M. J.; Bickelhaupt, F.; Nederkoorn, P. H. J.; Lammertsma, K. Density-Functional study of (solvated) Grignard complexes. J. Mol. Model. 2000, 6, 186– 194, DOI: 10.1007/s0089400060186Google Scholar25https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXkt1KhsrY%253D&md5=0ead5657fb9ba5cf003b7c95813d7f0eDensity-functional study of (solvated) Grignard complexesEhlers, Andreas W.; Van Klink, Gerard P. M.; Van Eis, Maurice J.; Bickelhaupt, Friedrich; Nederkoorn, Paul H. J.; Lammertsma, KoopJournal of Molecular Modeling (2000), 6 (2), 186-194CODEN: JMMOFK; ISSN:0948-5023. (Springer-Verlag)D. functional calcns. have been used to study the solvent effect of di-Et ether on the Schlenk equil. and the aggregation of Grignard reagents RMgX with R = Me, Et, Ph. Solvent stabilization of the Mg complexes of the first solvent is larger than that of the second one. The solvation energy decreases on going from the dihalides MgX2 to the monohalides RMgX to the diorgano compds. MgR2. The calcns. indicate that the energetic preference of the unsym. species reduces upon solvation. The strong tendency to dimerization of the un- and partly solvated compd. vanishes for the higher solvated cases.
- 26Tammiku, J.; Burk, P.; Tuulmets, A. 1,10-phenanthroline and its complexes with magnesium compounds. Disproportionation equilibria. J. Phys. Chem. A 2001, 105, 8554– 8561, DOI: 10.1021/jp011476nGoogle Scholar26https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXlvFyms70%253D&md5=8e9f052d75c330fadf679e58939df0e41,10-Phenanthroline and Its Complexes with Magnesium Compounds. Disproportionation EquilibriaTammiku, Jaana; Burk, Peeter; Tuulmets, AntsJournal of Physical Chemistry A (2001), 105 (37), 8554-8561CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The solvation, complexation, and disproportionation equil., which might be important during titrn. of a Grignard reagent RMgX with an alc. in the presence of 1,10-phenanthroline (phen), have been studied both in the gas phase and soln. using the d. functional theory (DFT) B3LYP/6-31+G* method. Solvation was modeled using the supermol. approach. NBO at. charge analyses were performed at the B3LYP/6-31G* level. The absorption spectra of the complexes were calcd. by the DFT TD/MPW1PW91/6-311+G** method. According to our calcns. the complexation of magnesium halide MgX2 with 1,10-phenanthroline is the reason for the disappearance of the red color of the complex RMgX(phen) near the titrn. end point.
- 27Tammiku-Taul, J.; Burk, P.; Tuulmets, A. Theoretical study of magnesium compounds: The Schlenk equilibrium in the gas phase and in the presence of Et2O and THF molecules. J. Phys. Chem. A 2004, 108, 133– 139, DOI: 10.1021/jp035653rGoogle Scholar27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXps1ymsLc%253D&md5=a004a2954b11ec82106aec89d2425056Theoretical Study of Magnesium Compounds: The Schlenk Equilibrium in the Gas Phase and in the Presence of Et2O and THF MoleculesTammiku-Taul, Jaana; Burk, Peeter; Tuulmets, AntsJournal of Physical Chemistry A (2004), 108 (1), 133-139CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The Schlenk equil. involving RMgX, R2Mg, and MgX2 (R = Me, Et, Ph and X = Cl, Br) has been studied both in the gas phase and in Et2O and THF solns. by the d. functional theory (DFT) B3LYP/6-31+G* method. Solvation was modeled using the supermol. approach. The stabilization due to interaction with solvent mols. decreases in the order MgX2 > RMgX > R2Mg and among the groups (R and X) Ph > Me > Et and Cl > Br. Studied magnesium compds. are more strongly solvated by THF compared to Et2O. The magnesium halide is solvated with up to four solvent mols. in THF soln., assuming that trans-dihalotetrakis(tetrahydrofurano)magnesium(II) complex forms. The formation of cis-dihalotetrakis(tetrahydrofurano)magnesium(II) is energetically less favorable than the formation of corresponding disolvated complexes. The predominant species in the Schlenk equil. are RMgX in Et2O and R2Mg + MgX2 in THF, which is consistent with exptl. data.
- 28Yamabe, S.; Yamazaki, S. In The Chemistry of Organomagnesium Compounds; Rappoport, Z., Marek, I., Eds.; Wiley-VCH: Weinheim, Germany, 2008; pp 369– 402.Google ScholarThere is no corresponding record for this reference.
- 29Yamazaki, S.; Yamabe, S. A computational study on addition of grignard reagents to carbonyl compounds. J. Org. Chem. 2002, 67, 9346– 9353, DOI: 10.1021/jo026017cGoogle Scholar29https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD38XovVChtL8%253D&md5=21d1d2fbbe9559d21f36ca0e767b850cA Computational Study on Addition of Grignard Reagents to Carbonyl CompoundsYamazaki, Shoko; Yamabe, ShinichiJournal of Organic Chemistry (2002), 67 (26), 9346-9353CODEN: JOCEAH; ISSN:0022-3263. (American Chemical Society)The mechanism of stereoselective addn. of Grignard reagents to carbonyl compds. was studied using B3LYP d. functional theory calcns. The study of the reaction of methylmagnesium chloride and formaldehyde in di-Me ether revealed a new reaction path involving carbonyl compd. coordination to Mg atoms in a dimeric Grignard reagent. The structure of the transition state for the addn. step shows that an interaction between a vicinal-Mg bonding alkyl group and C:O causes the C-C bond formation. The simplified mechanism shown by this model is in accord with the aggregation nature of Grignard reagents and their high reactivities toward carbonyl compds. Concerted and four-centered formation of strong O-Mg and C-C bonds was suggested as a polar mechanism. When the alkyl group is bulky, C-C bond formation is blocked and the Mg-O bond formation takes precedence. A diradical is formed with the odd spins localized on the alkyl group and carbonyl moiety. Diradical formation and its recombination probably are a single electron transfer (SET) process. The criteria for the concerted polar and stepwise SET processes were discussed in terms of precursor geometries and relative energies.
- 30Mori, T.; Kato, S. Grignard reagents in solution: Theoretical study of the equilibria and the reaction with a carbonyl compound in diethyl ether solvent. J. Phys. Chem. A 2009, 113, 6158– 6165, DOI: 10.1021/jp9009788Google Scholar30https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXlsFWjt7c%253D&md5=d774f6bc6bc37600ab68e8bf77155a04Grignard Reagents in Solution: Theoretical Study of the Equilibria and the Reaction with a Carbonyl Compound in Diethyl Ether SolventMori, Toshifumi; Kato, ShigekiJournal of Physical Chemistry A (2009), 113 (21), 6158-6165CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The equil. of Grignard reagents, CH3MgCl and CH3MgBr, in di-Et ether (Et2O) solvent as well as the reaction of the reagents with acetone are studied theor. To describe the equil. and reactions in Et2O solvent, the authors employ the ref. interaction site model SCF method with the second-order Moller-Plesset perturbation (RISM-MP2) free energy gradient method. Since the solvent mols. strongly coordinate to the Grignard reagents, the authors construct a cluster model by including several Et2O mols. into the quantum mech. region and embed it into the bulk solvent. Probably instead of the traditionally accepted cyclic dimer, the linear form of dimer is as stable as the monomer pair and participates in the equil. For the reaction with acetone, two important reaction paths (i.e., monomeric and linear dimeric paths) are studied. The barrier height for the monomeric path is much higher than that for the linear dimeric path, indicating that the reaction of the Grignard reagent with acetone proceeds through the linear dimeric reaction path. The change of solvation structure during the reaction is examd. From the calcd. free energy profiles, the entire reaction mechanisms of the Grignard reagents with aliph. ketones in Et2O solvent are discussed.
- 31Peltzer, R. M.; Eisenstein, O.; Nova, A.; Cascella, M. How solvent dynamics controls the Schlenk equilibrium of Grignard reagents: A computational study of CH3MgCl in tetrahydrofuran. J. Phys. Chem. B 2017, 121, 4226– 4237, DOI: 10.1021/acs.jpcb.7b02716Google Scholar31https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXltlWhtbg%253D&md5=731a756b355e2ecc9f28c13ae641c66bHow Solvent Dynamics Controls the Schlenk Equilibrium of Grignard Reagents: A Computational Study of CH3MgCl in TetrahydrofuranPeltzer, Raphael M.; Eisenstein, Odile; Nova, Ainara; Cascella, MicheleJournal of Physical Chemistry B (2017), 121 (16), 4226-4237CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)The Schlenk equil. is a complex reaction governing the presence of multiple chem. species in soln. of Grignard reagents. The full characterization at the mol. level of the transformation of CH3MgCl into MgCl2 and Mg(CH3)2 in THF by means of ab initio mol. dynamics simulations with enhanced-sampling metadynamics is presented. The reaction occurs via formation of dinuclear species bridged by chlorine atoms. At room temp., the different chem. species involved in the reaction accept multiple solvation structures, with two to four THF mols. that can coordinate the Mg atoms. The energy difference between all dinuclear solvated structures is lower than 5 kcal mol-1. The solvent is shown to be a direct key player driving the Schlenk mechanism. In particular, this study illustrates how the most stable sym. solvated dinuclear species, (THF)CH3Mg(μ-Cl)2MgCH3(THF) and (THF)CH3Mg(μ-Cl)(μ-CH3)MgCl(THF), need to evolve to less stable asym. solvated species, (THF)CH3Mg(μ-Cl)2MgCH3(THF)2 and (THF)CH3Mg(μ-Cl)(μ-CH3)MgCl(THF)2, in order to yield ligand exchange or product dissocn. In addn., the transferred ligands are always departing from an axial position of a pentacoordinated Mg atom. Thus, solvent dynamics is key to successive Mg-Cl and Mg-CH3 bond cleavages because bond breaking occurs at the most solvated Mg atom and the formation of bonds takes place at the least solvated one. The dynamics of the solvent also contributes to keep relatively flat the free energy profile of the Schlenk equil. These results shed light on one of the most used organometallic reagents whose structure in solvent remains exptl. unresolved. These results may also help to develop a more efficient catalyst for reactions involving these species.
- 32Hohenberg, P.; Kohn, W. Inhomogeneous electron gas. Phys. Rev. 1964, 136, B864– B871, DOI: 10.1103/PhysRev.136.B864Google ScholarThere is no corresponding record for this reference.
- 33Kohn, W.; Sham, L. J. Self-consistent equations including exchange and correlation effects. Phys. Rev. 1965, 140, 1133, DOI: 10.1103/PhysRev.140.A1133Google ScholarThere is no corresponding record for this reference.
- 34Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865– 3868, DOI: 10.1103/PhysRevLett.77.3865Google Scholar34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XmsVCgsbs%253D&md5=55943538406ee74f93aabdf882cd4630Generalized gradient approximation made simplePerdew, John P.; Burke, Kieron; Ernzerhof, MatthiasPhysical Review Letters (1996), 77 (18), 3865-3868CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)Generalized gradient approxns. (GGA's) for the exchange-correlation energy improve upon the local spin d. (LSD) description of atoms, mols., and solids. We present a simple derivation of a simple GGA, in which all parameters (other than those in LSD) are fundamental consts. Only general features of the detailed construction underlying the Perdew-Wang 1991 (PW91) GGA are invoked. Improvements over PW91 include an accurate description of the linear response of the uniform electron gas, correct behavior under uniform scaling, and a smoother potential.
- 35VandeVondele, J.; Hutter, J. Gaussian basis sets for accurate calculations on molecular systems in gas and condensed phases. J. Chem. Phys. 2007, 127, 114105, DOI: 10.1063/1.2770708Google Scholar35https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXhtFSrsLvM&md5=d7fdb937efb88cf3fca85792bb49ec27Gaussian basis sets for accurate calculations on molecular systems in gas and condensed phasesVandeVondele, Joost; Hutter, JurgJournal of Chemical Physics (2007), 127 (11), 114105/1-114105/9CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)We present a library of Gaussian basis sets that has been specifically optimized to perform accurate mol. calcns. based on d. functional theory. It targets a wide range of chem. environments, including the gas phase, interfaces, and the condensed phase. These generally contracted basis sets, which include diffuse primitives, are obtained minimizing a linear combination of the total energy and the condition no. of the overlap matrix for a set of mols. with respect to the exponents and contraction coeffs. of the full basis. Typically, for a given accuracy in the total energy, significantly fewer basis functions are needed in this scheme than in the usual split valence scheme, leading to a speedup for systems where the computational cost is dominated by diagonalization. More importantly, binding energies of hydrogen bonded complexes are of similar quality as the ones obtained with augmented basis sets, i.e., have a small (down to 0.2 kcal/mol) basis set superposition error, and the monomers have dipoles within 0.1 D of the basis set limit. However, contrary to typical augmented basis sets, there are no near linear dependencies in the basis, so that the overlap matrix is always well conditioned, also, in the condensed phase. The basis can therefore be used in first principles mol. dynamics simulations and is well suited for linear scaling calcns.
- 36Goedecker, S.; Teter, M.; Hutter, J. Separable dual-space Gaussian pseudopotentials. Phys. Rev. B 1996, 54, 1703– 1710, DOI: 10.1103/PhysRevB.54.1703Google Scholar36https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28XksFOht78%253D&md5=de0d078249d924ff884f32cb1e02595cSeparable dual-space Gaussian pseudopotentialsGoedecker, S.; Teter, M.; Hutter, J.Physical Review B: Condensed Matter (1996), 54 (3), 1703-1710CODEN: PRBMDO; ISSN:0163-1829. (American Physical Society)We present pseudopotential coeffs. for the first two rows of the Periodic Table. The pseudopotential is of an analytic form that gives optimal efficiency in numerical calculations using plane waves as a basis set. At most, even coeffs. are necessary to specify its analytic form. It is separable and has optimal decay properties in both real and Fourier space. Because of this property, the application of the nonlocal part of the pseudopotential to a wave function can be done efficiently on a grid in real space. Real space integration is much faster for large systems than ordinary multiplication in Fourier space, since it shows only quadratic scaling with respect to the size of the system. We systematically verify the high accuracy of these pseudopotentials by extensive at. and mol. test calcns.
- 37Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104, DOI: 10.1063/1.3382344Google Scholar37https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXkvVyks7o%253D&md5=2bca89d904579d5565537a0820dc2ae8A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-PuGrimme, Stefan; Antony, Jens; Ehrlich, Stephan; Krieg, HelgeJournal of Chemical Physics (2010), 132 (15), 154104/1-154104/19CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The method of dispersion correction as an add-on to std. Kohn-Sham d. functional theory (DFT-D) has been refined regarding higher accuracy, broader range of applicability, and less empiricism. The main new ingredients are atom-pairwise specific dispersion coeffs. and cutoff radii that are both computed from first principles. The coeffs. for new eighth-order dispersion terms are computed using established recursion relations. System (geometry) dependent information is used for the first time in a DFT-D type approach by employing the new concept of fractional coordination nos. (CN). They are used to interpolate between dispersion coeffs. of atoms in different chem. environments. The method only requires adjustment of two global parameters for each d. functional, is asymptotically exact for a gas of weakly interacting neutral atoms, and easily allows the computation of at. forces. Three-body nonadditivity terms are considered. The method has been assessed on std. benchmark sets for inter- and intramol. noncovalent interactions with a particular emphasis on a consistent description of light and heavy element systems. The mean abs. deviations for the S22 benchmark set of noncovalent interactions for 11 std. d. functionals decrease by 15%-40% compared to the previous (already accurate) DFT-D version. Spectacular improvements are found for a tripeptide-folding model and all tested metallic systems. The rectification of the long-range behavior and the use of more accurate C6 coeffs. also lead to a much better description of large (infinite) systems as shown for graphene sheets and the adsorption of benzene on an Ag(111) surface. For graphene it is found that the inclusion of three-body terms substantially (by about 10%) weakens the interlayer binding. We propose the revised DFT-D method as a general tool for the computation of the dispersion energy in mols. and solids of any kind with DFT and related (low-cost) electronic structure methods for large systems. (c) 2010 American Institute of Physics.
- 38Swope, W. C.; Andersen, H. C.; Berens, P. H.; Wilson, K. R. A computer simulation method for the calculation of equilibrium constants for the formation of physical clusters of molecules: Application to small water clusters. J. Chem. Phys. 1982, 76, 648, DOI: 10.1063/1.442716Google ScholarThere is no corresponding record for this reference.
- 39Bussi, G.; Donadio, D.; Parrinello, M. Canonical sampling through velocity rescaling. J. Chem. Phys. 2007, 126, 014101, DOI: 10.1063/1.2408420Google Scholar39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXosVCltg%253D%253D&md5=9c182b57bfc65bca6be23c8c76b4be77Canonical sampling through velocity rescalingBussi, Giovanni; Donadio, Davide; Parrinello, MicheleJournal of Chemical Physics (2007), 126 (1), 014101/1-014101/7CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The authors present a new mol. dynamics algorithm for sampling the canonical distribution. In this approach the velocities of all the particles are rescaled by a properly chosen random factor. The algorithm is formally justified and it is shown that, in spite of its stochastic nature, a quantity can still be defined that remains const. during the evolution. In numerical applications this quantity can be used to measure the accuracy of the sampling. The authors illustrate the properties of this new method on Lennard-Jones and TIP4P water models in the solid and liq. phases. Its performance is excellent and largely independent of the thermostat parameter also with regard to the dynamic properties.
- 40Nosé, S. A unified formulation of the constant temperature molecular dynamics methods. J. Chem. Phys. 1984, 81, 511– 519, DOI: 10.1063/1.447334Google Scholar40https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2cXkvFOrs7k%253D&md5=2974515ec89e5601868e35871c0f19c2A unified formulation of the constant-temperature molecular-dynamics methodsNose, ShuichiJournal of Chemical Physics (1984), 81 (1), 511-19CODEN: JCPSA6; ISSN:0021-9606.Three recently proposed const. temp. mol. dynamics methods [N., (1984) (1); W. G. Hoover et al., (1982) (2); D. J. Evans and G. P. Morris, (1983) (2); and J. M. Haile and S. Gupta, 1983) (3)] are examd. anal. via calcg. the equil. distribution functions and comparing them with that of the canonical ensemble. Except for effects due to momentum and angular momentum conservation, method (1) yields the rigorous canonical distribution in both momentum and coordinate space. Method (2) can be made rigorous in coordinate space, and can be derived from method (1) by imposing a specific constraint. Method (3) is not rigorous and gives a deviation of order N-1/2 from the canonical distribution (N the no. of particles). The results for the const. temp.-const. pressure ensemble are similar to the canonical ensemble case.
- 41Hoover, W. G. Canonical dynamics: Equilibrium phase-space distributions. Phys. Rev. A 1985, 31, 1695– 1697, DOI: 10.1103/PhysRevA.31.1695Google Scholar41https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC2sjotlWltA%253D%253D&md5=99a2477835b37592226a5d18a760685cCanonical dynamics: Equilibrium phase-space distributionsHooverPhysical review. A, General physics (1985), 31 (3), 1695-1697 ISSN:0556-2791.There is no expanded citation for this reference.
- 42Martyna, G. J.; Klein, M. L.; Tuckerman, M. Nosé—Hoover chains: The canonical ensemble via continuous dynamics. J. Chem. Phys. 1992, 97, 2635– 2643, DOI: 10.1063/1.463940Google ScholarThere is no corresponding record for this reference.
- 43Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual molecular dynamics. J. Mol. Graphics 1996, 14, 33– 38, DOI: 10.1016/0263-7855(96)00018-5Google Scholar43https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK28Xis12nsrg%253D&md5=1e3094ec3151fb85c5ff05f8505c78d5VDM: visual molecular dynamicsHumphrey, William; Dalke, Andrew; Schulten, KlausJournal of Molecular Graphics (1996), 14 (1), 33-8, plates, 27-28CODEN: JMGRDV; ISSN:0263-7855. (Elsevier)VMD is a mol. graphics program designed for the display and anal. of mol. assemblies, in particular, biopolymers such as proteins and nucleic acids. VMD can simultaneously display any no. of structures using a wide variety of rendering styles and coloring methods. Mols. are displayed as one or more "representations," in which each representation embodies a particular rendering method and coloring scheme for a selected subset of atoms. The atoms displayed in each representation are chosen using an extensive atom selection syntax, which includes Boolean operators and regular expressions. VMD provides a complete graphical user interface for program control, as well as a text interface using the Tcl embeddable parser to allow for complex scripts with variable substitution, control loops, and function calls. Full session logging is supported, which produces a VMD command script for later playback. High-resoln. raster images of displayed mols. may be produced by generating input scripts for use by a no. of photorealistic image-rendering applications. VMD has also been expressly designed with the ability to animate mol. dynamics (MD) simulation trajectories, imported either from files or from a direct connection to a running MD simulation. VMD is the visualization component of MDScope, a set of tools for interactive problem solving in structural biol., which also includes the parallel MD program NAMD, and the MDCOMM software used to connect the visualization and simulation programs, VMD is written in C++, using an object-oriented design; the program, including source code and extensive documentation, is freely available via anonymous ftp and through the World Wide Web.
- 44Carter, E. A.; Ciccotti, G.; Heynes, J. T.; Kapral, R. Constrained reaction coordinate dynamics for the simulation of rare events. Chem. Phys. Lett. 1989, 156, 472– 477, DOI: 10.1016/S0009-2614(89)87314-2Google Scholar44https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1MXks1CitrY%253D&md5=bff35139b4bb8f1c06d29386dfa786dfConstrained reaction coordinate dynamics for the simulation of rare eventsCarter, E. A.; Ciccotti, Giovanni; Hynes, James T.; Kapral, RaymondChemical Physics Letters (1989), 156 (5), 472-7CODEN: CHPLBC; ISSN:0009-2614.A computationally efficient mol. dynamics method for estg. the rates of rare events that occur by activated processes is described. The system is constrained at "bottleneck" regions on a general many-body reaction coordinate in order to generate a biased configurational distribution. Suitable reweighting of this biased distribution, along with correct momentum distribution sampling, provides a new ensemble, the constrained-reaction-coordinate-dynamics ensemble, with which to study rare events of this type. Applications to chem. reaction rates are made.
- 45Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. Universal solvation model based on solute electron density and on a continuum model of the solvent defined by the bulk dielectric constant and atomic surface tensions. J. Phys. Chem. B 2009, 113, 6378– 6396, DOI: 10.1021/jp810292nGoogle Scholar45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXksV2is74%253D&md5=54931a64c70d28445ee53876a8b1a4b9Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface TensionsMarenich, Aleksandr V.; Cramer, Christopher J.; Truhlar, Donald G.Journal of Physical Chemistry B (2009), 113 (18), 6378-6396CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)We present a new continuum solvation model based on the quantum mech. charge d. of a solute mol. interacting with a continuum description of the solvent. The model is called SMD, where the "D" stands for "d." to denote that the full solute electron d. is used without defining partial at. charges. "Continuum" denotes that the solvent is not represented explicitly but rather as a dielec. medium with surface tension at the solute-solvent boundary. SMD is a universal solvation model, where "universal" denotes its applicability to any charged or uncharged solute in any solvent or liq. medium for which a few key descriptors are known (in particular, dielec. const., refractive index, bulk surface tension, and acidity and basicity parameters). The model separates the observable solvation free energy into two main components. The first component is the bulk electrostatic contribution arising from a self-consistent reaction field treatment that involves the soln. of the nonhomogeneous Poisson equation for electrostatics in terms of the integral-equation-formalism polarizable continuum model (IEF-PCM). The cavities for the bulk electrostatic calcn. are defined by superpositions of nuclear-centered spheres. The second component is called the cavity-dispersion-solvent-structure term and is the contribution arising from short-range interactions between the solute and solvent mols. in the first solvation shell. This contribution is a sum of terms that are proportional (with geometry-dependent proportionality consts. called at. surface tensions) to the solvent-accessible surface areas of the individual atoms of the solute. The SMD model has been parametrized with a training set of 2821 solvation data including 112 aq. ionic solvation free energies, 220 solvation free energies for 166 ions in acetonitrile, methanol, and DMSO, 2346 solvation free energies for 318 neutral solutes in 91 solvents (90 nonaq. org. solvents and water), and 143 transfer free energies for 93 neutral solutes between water and 15 org. solvents. The elements present in the solutes are H, C, N, O, F, Si, P, S, Cl, and Br. The SMD model employs a single set of parameters (intrinsic at. Coulomb radii and at. surface tension coeffs.) optimized over six electronic structure methods: M05-2X/MIDI!6D, M05-2X/6-31G*, M05-2X/6-31+G**, M05-2X/cc-pVTZ, B3LYP/6-31G*, and HF/6-31G*. Although the SMD model has been parametrized using the IEF-PCM protocol for bulk electrostatics, it may also be employed with other algorithms for solving the nonhomogeneous Poisson equation for continuum solvation calcns. in which the solute is represented by its electron d. in real space. This includes, for example, the conductor-like screening algorithm. With the 6-31G* basis set, the SMD model achieves mean unsigned errors of 0.6-1.0 kcal/mol in the solvation free energies of tested neutrals and mean unsigned errors of 4 kcal/mol on av. for ions with either Gaussian03 or GAMESS.
- 46Zhao, Y.; Truhlar, D. G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215– 241, DOI: 10.1007/s00214-007-0310-xGoogle Scholar46https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXltFyltbY%253D&md5=c31d6f319d7c7a45aa9b716220e4a422The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other functionalsZhao, Yan; Truhlar, Donald G.Theoretical Chemistry Accounts (2008), 120 (1-3), 215-241CODEN: TCACFW; ISSN:1432-881X. (Springer GmbH)We present two new hybrid meta exchange-correlation functionals, called M06 and M06-2X. The M06 functional is parametrized including both transition metals and nonmetals, whereas the M06-2X functional is a high-nonlocality functional with double the amt. of nonlocal exchange (2X), and it is parametrized only for nonmetals. The functionals, along with the previously published M06-L local functional and the M06-HF full-Hartree-Fock functionals, constitute the M06 suite of complementary functionals. We assess these four functionals by comparing their performance to that of 12 other functionals and Hartree-Fock theory for 403 energetic data in 29 diverse databases, including ten databases for thermochem., four databases for kinetics, eight databases for noncovalent interactions, three databases for transition metal bonding, one database for metal atom excitation energies, and three databases for mol. excitation energies. We also illustrate the performance of these 17 methods for three databases contg. 40 bond lengths and for databases contg. 38 vibrational frequencies and 15 vibrational zero point energies. We recommend the M06-2X functional for applications involving main-group thermochem., kinetics, noncovalent interactions, and electronic excitation energies to valence and Rydberg states. We recommend the M06 functional for application in organometallic and inorganometallic chem. and for noncovalent interactions.
- 47Ditchfield, R.; Hehre, W. J.; Pople, J. A. Self-consistent molecular-orbital methods. IX. An extended Gaussian-type basis for molecular-orbital studies of organic molecules. J. Chem. Phys. 1971, 54, 724– 728, DOI: 10.1063/1.1674902Google Scholar47https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE3MXksFOiuw%253D%253D&md5=abce439752b61fad3aa748515ec78c71Self-consistent molecular-orbital methods. IX. Extended Gaussian-type basis for molecular-orbital studies of organic moleculesDitchfield, R.; Hehre, Warren J.; Pople, John A.Journal of Chemical Physics (1971), 54 (2), 724-8CODEN: JCPSA6; ISSN:0021-9606.An extended basis set of at. functions expressed as fixed linear combinations of Gaussian functions is presented for H and the first-row atoms C to F. In this set. described as 4-31 G, each inner shell is represented by a single basis function taken as a sum of 4 Gaussians, and each valence orbital is split into inner and outer parts described by 3 and 1 Gaussian function, resp. The expansion coeffs. and Gaussian exponents are detd. by minimizing the total calcd. energy of the at. ground state. This basis set is then used in single-determinant MO studies of a group of small polyat. mols. Optimization of valence-shell scaling factors shows that considerable rescaling of at. functions occurs in mols., the largest effects being obsd. for H and C. However, the range of optimum scale factors for each atom is small enough to allow the selection of a std. mol. set. The use of this std. basis gives theoretical equil. geometries in reasonable agreement with expt.
- 48Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Petersson, G. A.; Nakatsuji, H.; Li, X.; Caricato, M.; Marenich, A. V.; Bloino, J.; Janesko, B. G.; Gomperts, R.; Mennucci, B.; Hratchian, H. P.; Ortiz, J. V.; Izmaylov, A. F.; Sonnenberg, J. L.; Williams-Young, D.; Ding, F.; Lipparini, F.; Egidi, F.; Goings, J.; Peng, B.; Petrone, A.; Henderson, T.; Ranasinghe, D.; Zakrzewski, V. G.; Gao, J.; Rega, N.; Zheng, G.; Liang, W.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Throssell, K.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M. J.; Heyd, J. J.; Brothers, E. N.; Kudin, K. N.; Staroverov, V. N.; Keith, T. A.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A. P.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Millam, J. M.; Klene, M.; Adamo, C.; Cammi, R.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Farkas, O.; Foresman, J. B.; Fox, D. J. Gaussian 09, revision B.01; Gaussian, Inc.: Wallingford, CT, 2016.Google ScholarThere is no corresponding record for this reference.
- 49Raghavachari, K.; Trucks, G. W.; Pople, J. A.; Head-Gordon, M. A fifth-order perturbation comparison of electron correlation theories. Chem. Phys. Lett. 1969, 157, 479– 483, DOI: 10.1016/S0009-2614(89)87395-6Google ScholarThere is no corresponding record for this reference.
- 50Dunning, T. H. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 1007– 1023, DOI: 10.1063/1.456153Google Scholar50https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1MXksVGmtrk%253D&md5=c6cd67a3748dc61692a9cb622d2694a0Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogenDunning, Thom H., Jr.Journal of Chemical Physics (1989), 90 (2), 1007-23CODEN: JCPSA6; ISSN:0021-9606.Guided by the calcns. on oxygen in the literature, basis sets for use in correlated at. and mol. calcns. were developed for all of the first row atoms from boron through neon, and for hydrogen. As in the oxygen atom calcns., the incremental energy lowerings, due to the addn. of correlating functions, fall into distinct groups. This leads to the concept of correlation-consistent basis sets, i.e., sets which include all functions in a given group as well as all functions in any higher groups. Correlation-consistent sets are given for all of the atoms considered. The most accurate sets detd. in this way, [5s4p3d2f1g], consistently yield 99% of the correlation energy obtained with the corresponding at.-natural-orbital sets, even though the latter contains 50% more primitive functions and twice as many primitive polarization functions. It is estd. that this set yields 94-97% of the total (HF + 1 + 2) correlation energy for the atoms neon through boron.
- 51CFOUR, a quantum-chemical program package written by Stanton, J. F.; Gauss, J.; Cheng, L.; Harding, M. E.; Matthews, D. A.; Szalay, P. G. with contributions from Auer, A. A.; Bartlett, R. J.; Benedikt, U.; Berger, C.; Bernholdt, D. E.; Bomble, Y. J.; Christiansen, O.; Engel, F.; Faber, R.; Heckert, M.; Heun, O.; Hilgenberg, M.; Huber, C.; Jagau, T.-C.; Jonsson, D.; Jusélius, J.; Kirsch, T.; Klein, K.; Lauderdale, W. J.; Lipparini, F.; Metzroth, T.; Mück, L.A.; O’Neill, D. P.; Price, D. R.; Prochnow, E.; Puzzarini, C.; Ruud, K.; Schiffmann, F.; Schwalbach, W.; Simmons, C.; Stopkowicz, S.; Tajti, A.; Vázquez, J.; Wang, F.; Watts, J. D. and the integral packages MOLECULE (Almlöf, J.; Taylor, P. R.), PROPS (Taylor, P. R.), ABACUS (Helgaker, T.; Jensen, H. J. Aa.; Jørgensen, P.; Olsen, J.), and ECP routines by Mitin, A. V. and van Wüllen, C.; website: http://www.cfour.de.Google ScholarThere is no corresponding record for this reference.
- 52Buergi, H. B.; Dunitz, J. D.; Shefter, E. Geometrical reaction coordinates. II. Nucleophilic addition to a carbonyl group. J. Am. Chem. Soc. 1973, 95, 5065– 5067, DOI: 10.1021/ja00796a058Google ScholarThere is no corresponding record for this reference.
- 53Buergi, H. B.; Lehn, J. M.; Wipff, G. Ab initio study of nucleophilic addition to a carbonyl group. J. Am. Chem. Soc. 1974, 96, 1956– 1957, DOI: 10.1021/ja00813a062Google Scholar53https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE2cXhtlalsL0%253D&md5=9646a20668e917e66b8876136bba887cAb initio study of nucleophilic addition to a carbonyl groupBuergi, H. B.; Lehn, J. M.; Wipff, G.Journal of the American Chemical Society (1974), 96 (6), 1956-7CODEN: JACSAT; ISSN:0002-7863.Ab initio SCF-LCGO (linear combination of Gaussian orbitals)-MO computations were performed on the reaction of hydride ion with HCHO, considered as model for nucleophilic addns. to the carbonyl group. Geometrical changes and electronic rearrangements were obtained as a function of the reaction coordinate. Orientational constraints in the course of the reaction bear relation to orbital steering, togetherness, and proximity effects considered in the literature. The calcd. geometrical changes correlate with the changes obsd. in crystal structures for the approach of an amino site toward a carbonyl group. A computation of the NH3-HCHO system at one fixed sepn. was also performed for comparison purposes.
- 54Walker, F. W.; Ashby, E. C. Composition of Grignard compounds. VI. Nature of association in tetrahydrofuran and diethyl ether solutions. J. Am. Chem. Soc. 1969, 91, 3845– 3850, DOI: 10.1021/ja01042a027Google Scholar54https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaF1MXktlerur0%253D&md5=98ef0f28cd3bdbeaec2ff3c655aee557Composition of Grignard compounds. VI. Nature of association in tetrahydrofuran and diethyl ether solutionsWalker, Frank W.; Ashby, E. C.Journal of the American Chemical Society (1969), 91 (14), 3845-50CODEN: JACSAT; ISSN:0002-7863.Ebullioscopic data are presented for tetrahydrofuran (I) and Et2O solns. of several Grignard and related Mg compounds over a wide concn. range. Anal. of the data is accomplished by observing the change in assocn. with concn. and by consideration of the constancy of the equil. consts. calcd. for several possible descriptions of the assocd. system. The expected nonideality of the solns. studied was considered in the interpretation of the data. While all the compds. studied were monomeric in I, the alkyl- and arylmagnesium bromides and iodides were monomeric in Et2O only at low concn. (<0.1 m), exhibiting in general an increase in assocn. with concn. These compds. are assocd. in a polymeric fashion. In contrast, the alkylmagnesium chlorides assoc. in Et2O to form stable dimers with the assocn. insensitive to concn. changes. Comparison of the data for Mg halides and dialkylmagnesium compds. in Et2O indicates that, except for the Me compd., assocn. is considerably stronger for the Mg halides than for the dialkylmagnesium compds. Thus, except for methylmagnesium halides, Grignard compds. assoc. with bridging mainly through the halogen atom. The methylmagnesium halides are exceptional since Me bridging is strong enough in Et2O to permit assocn. by bridging through either the Me group or the halogen atom. Although the steric and electronic nature of the alkyl group has some effect on the assocn. of Grignard compds., the effect is generally small compared to to the effect of halogen or solvent.
- 55Holm, T. Thermochemical bond dissociation energies of carbon-magnesium bonds. J. Chem. Soc., Perkin Trans. 2 1981, 2, 464– 467, DOI: 10.1039/P29810000464Google ScholarThere is no corresponding record for this reference.
- 56We consider here thermodynamically equilibrated solutions of the Grignard reagent. In fact, the formation of the Grignard reagent may occur via a radical mechanism; see, for instance, the recent study:Henriques, A. M.; Barbosa, A. G. H. Chemical bonding and the equilibrium composition of Grignard reagents in ethereal solution. J. Phys. Chem. A 2011, 115, 12259– 12270, DOI: 10.1021/jp202762pGoogle Scholar56https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhtlWjt7bJ&md5=dd04ac72cb7f1d42b5ed2e58265d14e7Chemical bonding and the equilibrium composition of Grignard reagents in ethereal solutionsHenriques, Andre M.; Barbosa, Andre G. H.Journal of Physical Chemistry A (2011), 115 (44), 12259-12270CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)A thorough anal. of the electronic structure and thermodn. aspects of Grignard reagents and its assocd. equil. compn. in ethereal solns. is performed. Considering methylmagnesium halides contg. fluorine, chlorine, and bromine, we studied the neutral, charged, and radical species assocd. with their chem. equil. in soln. The ethereal solvents considered, THF and di-Et ether, were modeled using the polarizable continuum model (PCM) and also by explicit coordination to the Mg atoms in a cluster. The chem. bonding of the species that constitute the Grignard reagent is analyzed in detail with generalized valence bond (GVB) wave functions. Equil. consts. were calcd. with the DFT/M06 functional and GVB wave functions, yielding similar results. According to our calcns. and existing kinetic and electrochem. evidence, the species R·, R-, ·MgX, and RMgX2- must be present in low concn. in the equil. We conclude that depending on the halogen, a different route must be followed to produce the relevant equil. species in each case. Chloride and bromide must preferably follow a "radical-based" pathway, and fluoride must follow a "carbanionic-based" pathway. These different mechanisms are contrasted against the available exptl. results and are proven to be consistent with the existing thermodn. data on the Grignard reagent equil.
- 57Ziegler, D. S.; Wei, B.; Knochel, P. Improving the halogen–magnesium exchange by using new turbo-Grignard reagents. Chem. - Eur. J. 2019, 25, 2695– 2703, DOI: 10.1002/chem.201803904Google Scholar57https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXisFektbnE&md5=7f13713cfdbc0ca7df641474933c6bb6Improving the Halogen-Magnesium Exchange by using New Turbo-Grignard ReagentsZiegler, Dorothee S.; Wei, Baosheng; Knochel, PaulChemistry - A European Journal (2019), 25 (11), 2695-2703CODEN: CEUJED; ISSN:0947-6539. (Wiley-VCH Verlag GmbH & Co. KGaA)A review. This Minireview describes the scope of the halogen-magnesium exchange. It shows that the use of the turbo-Grignard reagent (iPrMgCl·LiCl) greatly enhances the rate of the Br- and I-Mg exchange. Furthermore, this magnesium reagent allows the performance of a fast sulfoxide-magnesium exchange. Also, the use of s-BuMgOR·LiOR (R=2-ethylhexyl) enables a Br-Mg exchange in toluene leading to various Grignard reagents in toluene. Addnl., the new exchange reagent s-Bu2Mg·2LiOR (R = 2-ethylhexyl) further increases the rate of the halogen-magnesium exchange allowing one to perform a chlorine-magnesium exchange for arom. chlorides bearing an ortho-methoxy substituent in toluene.
- 58Harutyunyan, S. R.; Lopez, F.; Browne, W. R.; Correa, A.; Pena, D.; Badorrey, R.; Meetsma, A.; Minaard, A. J.; Feringa, B. L. On the mechanism of the copper-catalyzed enantioselective 1,4-addition of Grignard reagents to α, β-unsaturated carbonyl compounds. J. Am. Chem. Soc. 2006, 128, 9103– 9118, DOI: 10.1021/ja0585634Google Scholar58https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28Xmt12msbw%253D&md5=5c3b262d1e881545199120f63476f146On the Mechanism of the Copper-Catalyzed Enantioselective 1,4-Addition of Grignard Reagents to α,β-Unsaturated Carbonyl CompoundsHarutyunyan, Syuzanna R.; Lopez, Fernando; Browne, Wesley R.; Correa, Arkaitz; Pena, Diego; Badorrey, Ramon; Meetsma, Auke; Minnaard, Adriaan J.; Feringa, Ben L.Journal of the American Chemical Society (2006), 128 (28), 9103-9118CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The mechanism of the enantioselective 1,4-addn. of Grignard reagents to α,β-unsatd. carbonyl compds. promoted by copper complexes of chiral ferrocenyl diphosphines is explored through kinetic, spectroscopic, and electrochem. anal. On the basis of these studies, a structure of the active catalyst is proposed. The roles of the solvent, copper halide, and the Grignard reagent have been examd. Kinetic studies support a reductive elimination as the rate-limiting step in which the chiral catalyst, the substrate, and the Grignard reagent are involved. The thermodn. activation parameters were detd. from the temp. dependence of the reaction rate. The putative active species and the catalytic cycle of the reaction are discussed.
- 59Lopez, F.; Minaard, A. J.; Feringa, B. L. Catalytic enantioselective conjugate addition with Grignard reagents. Acc. Chem. Res. 2007, 40, 179– 188, DOI: 10.1021/ar0501976Google Scholar59https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XhtlaksrrL&md5=418c3907b4d4aed60135f06abc29c389Catalytic enantioselective conjugate addition with Grignard reagentsLopez, Fernando; Minnaard, Adriaan J.; Feringa, Ben L.Accounts of Chemical Research (2007), 40 (3), 179-188CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)A review on recent advances in Cu-catalyzed asym. conjugate addn. of Grignard reagents, asym. SN2' substitution reactions of allylic bromides with Grignard reagents, and application to synthesis of natural products.
- 60Seitz, T. A.; Seitz, J. A. A general two-metal-ion mechanism for catalytic RNA. Proc. Natl. Acad. Sci. U. S. A. 1993, 90, 6498– 6502, DOI: 10.1073/pnas.90.14.6498Google ScholarThere is no corresponding record for this reference.
- 61De Vivo, M.; Dal Peraro, M.; Klein, M. L. Phosphodiester cleavage in ribonuclease H occurs via an associative two-metal-aided catalytic mechanism. J. Am. Chem. Soc. 2008, 130, 10955– 10962, DOI: 10.1021/ja8005786Google Scholar61https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXptVGisr8%253D&md5=db9ad0480dccbdb588f0f86ecc2d710dPhosphodiester Cleavage in Ribonuclease H Occurs via an Associative Two-Metal-Aided Catalytic MechanismDe Vivo, Marco; Dal Peraro, Matteo; Klein, Michael L.Journal of the American Chemical Society (2008), 130 (33), 10955-10962CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)RNase H belongs to the nucleotidyl-transferase (NT) superfamily and hydrolyzes the phosphodiester linkages that form the backbone of the RNA strand in RNA•DNA hybrids. This enzyme is implicated in replication initiation and DNA topol. restoration and represents a very promising target for anti-HIV drug design. Structural information has been provided by high-resoln. crystal structures of the complex RNase H/RNA•DNA from Bacillus halodurans (Bh), which reveals that two metal ions are required for formation of a catalytic active complex. Here, we use classical force field-based and quantum mechanics/mol. mechanics calcns. for modeling the nucleotidyl transfer reaction in RNase H, clarifying the role of the metal ions and the nature of the nucleophile (water vs. hydroxide ion). During the catalysis, the two metal ions act cooperatively, facilitating nucleophile formation and stabilizing both transition state and leaving group. Importantly, the two Mg2+ metals also support the formation of a meta-stable phosphorane intermediate along the reaction, which resembles the phosphorane intermediate structure obtained only in the debated β-phosphoglucomutase crystal (Lahiri, S. D.; et al. Science 2003, 299 (5615), 2067-2071). The nucleophile formation (i.e., water deprotonation) can be achieved in situ, after migration of one proton from the water to the scissile phosphate in the transition state. This proton transfer is actually mediated by solvation water mols. Due to the highly conserved nature of the enzymic bimetal motif, these results might also be relevant for structurally similar enzymes belonging to the NT superfamily.
- 62Rosta, E.; Nowotny, M.; Yang, W.; Hummer, G. Catalytic mechanism of RNA backbone cleavage by ribonuclease H from quantum mechanics/molecular mechanics simulations. J. Am. Chem. Soc. 2011, 133, 8934– 8941, DOI: 10.1021/ja200173aGoogle Scholar62https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXmsFymt7Y%253D&md5=0819abe0e0cedd6f352f09dee5c1c6f2Catalytic Mechanism of RNA Backbone Cleavage by Ribonuclease H from Quantum Mechanics/Molecular Mechanics SimulationsRosta, Edina; Nowotny, Marcin; Yang, Wei; Hummer, GerhardJournal of the American Chemical Society (2011), 133 (23), 8934-8941CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)We use quantum mechanics/mol. mechanics simulations to study the cleavage of the RNA (RNA) backbone catalyzed by RNase H. This protein is a prototypical member of a large family of enzymes that use two-metal catalysis to process nucleic acids. By combining Hamiltonian replica exchange with a finite-temp. string method, we calc. the free energy surface underlying the RNA-cleavage reaction and characterize its mechanism. We find that the reaction proceeds in two steps. In a first step, catalyzed primarily by magnesium ion A and its ligands, a water mol. attacks the scissile phosphate. Consistent with thiol-substitution expts., a water proton is transferred to the downstream phosphate group. The transient phosphorane formed as a result of this nucleophilic attack decays by breaking the bond between the phosphate and the ribose oxygen. In the resulting intermediate, the dissocd. but unprotonated leaving group forms an alkoxide coordinated to magnesium ion B. In a second step, the reaction is completed by protonation of the leaving group, with a neutral Asp132 as a likely proton donor. The overall reaction barrier of ∼15 kcal mol-1, encountered in the first step, together with the cost of protonating Asp132, is consistent with the slow measured rate of ∼1-100/min. The two-step mechanism is also consistent with the bell-shaped pH dependence of the reaction rate. The nonmonotonic relative motion of the magnesium ions along the reaction pathway agrees with X-ray crystal structures. Proton-transfer reactions and changes in the metal ion coordination emerge as central factors in the RNA-cleavage reaction.
- 63Casalino, L.; Palermo, G.; Rothlisberger, U.; Magistrato, A. Who activates the nucleophile in ribozyme catalysis? An answer from the splicing mechanism of group II introns. J. Am. Chem. Soc. 2016, 138, 10374– 10377, DOI: 10.1021/jacs.6b01363Google Scholar63https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XpvFylsb0%253D&md5=e01f4d5f916c9ea688b90d8baeeb0a69Who Activates the Nucleophile in Ribozyme Catalysis? An Answer from the Splicing Mechanism of Group II IntronsCasalino, Lorenzo; Palermo, Giulia; Rothlisberger, Ursula; Magistrato, AlessandraJournal of the American Chemical Society (2016), 138 (33), 10374-10377CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Group II introns are Mg2+-dependent ribozymes that are considered to be the evolutionary ancestors of the eukaryotic spliceosome, thus representing an ideal model system to understand the mechanism of conversion of premature mRNA (mRNA) into mature mRNA. Neither in splicing nor for self-cleaving ribozymes has the role of the two Mg2+ ions been established, and even the way the nucleophile is activated is still controversial. Here we employed hybrid quantum-classical QM(Car-Parrinello)/MM mol. dynamics simulations in combination with thermodn. integration to characterize the mol. mechanism of the first and rate-detg. step of the splicing process (i.e., the cleavage of the 5'-exon) catalyzed by group II intron ribozymes. Remarkably, our results show a new RNA-specific dissociative mechanism in which the bulk water accepts the nucleophile's proton during its attack on the scissile phosphate. The process occurs in a single step with no Mg2+ ion activating the nucleophile, at odds with nucleases enzymes. We suggest that the novel reaction path elucidated here might be an evolutionary ancestor of the more efficient two-metal-ion mechanism found in enzymes.
- 64Genna, V.; Vidossich, P.; Ippoliti, E.; Carloni, P.; De Vivo, M. A self-activated mechanism for nucleic acid polymerization catalyzed by DNA/RNA polymerases. J. Am. Chem. Soc. 2016, 138, 14592– 14598, DOI: 10.1021/jacs.6b05475Google Scholar64https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XhtlCrsLzE&md5=f8ee07a32f8b12a2978ab9c974a701b9A Self-Activated Mechanism for Nucleic Acid Polymerization Catalyzed by DNA/RNA PolymerasesGenna, Vito; Vidossich, Pietro; Ippoliti, Emiliano; Carloni, Paolo; Vivo, Marco DeJournal of the American Chemical Society (2016), 138 (44), 14592-14598CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)The enzymic polymn. of DNA and RNA is at the basis of genetic inheritance for all living organisms. It is catalyzed by the DNA/RNA polymerase (Pol) superfamily. Here, bioinformatics anal. revealed that the incoming nucleotide substrate always forms an H-bond between its 3'-OH and β-phosphate moieties upon formation of the Michaelis complex. This previously unrecognized H-bond implies a novel self-activated mechanism (SAM), which synergistically connects the in situ nucleophile formation with subsequent nucleotide addn. and, importantly, nucleic acid translocation. Thus, SAM allows an elegant and efficient closed-loop sequence of chem. and phys. steps for Pol catalysis. This is markedly different from previous mechanistic hypotheses. This proposed mechanism was corroborated via ab initio QM/MM simulations on a specific Pol, human DNA polymerase-η, an enzyme involved in repairing damaged DNA. The structural conservation of DNA and RNA Pols supports the possible extension of SAM to Pol enzymes from the 3 domains of life.
- 65Genna, V.; Donati, E.; De Vivo, M. The catalytic mechanism of DNA and RNA polymerases. ACS Catal. 2018, 8, 11103– 11118, DOI: 10.1021/acscatal.8b03363Google Scholar65https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXhvFCks7nI&md5=79eb79ea021ae74916df15ef0bbc0c17The Catalytic Mechanism of DNA and RNA PolymerasesGenna, Vito; Donati, Elisa; De Vivo, MarcoACS Catalysis (2018), 8 (12), 11103-11118CODEN: ACCACS; ISSN:2155-5435. (American Chemical Society)A review. DNA and RNA polymerases (Pols) catalyze nucleic acid biosynthesis in all domains of life, with implications for human diseases and health. Pols carry out nucleic acid extension through the addn. of one incoming nucleotide trisphosphate at the 3'-OH terminus of the growing primer strand, at every catalytic cycle. Thus, Pol catalysis involves chem. reactions for nucleophile 3'-OH deprotonation and nucleotide addn., as well as major protein conformational motions and structural rearrangements for nucleotide selection, binding, and nucleic acid translocation to complete the overall catalytic cycle. In this respect, quantum and mol. mechanics simulations, integrated with exptl. data, have advanced our mechanistic understanding of how Pols operate at the at. level. This Perspective outlines how modern mol. simulations can further deepen our understanding of Pol catalytic reactions and fidelity, which may help in devising strategies for designing drugs and artificial Pols for biotechnol. and clin. purposes.
Supporting Information
Supporting Information
ARTICLE SECTIONSThe Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.9b11829.
Reference coupled-cluster calculations and benchmark BDE calculations, and structures of Mg(CH3)2, Mg(CH3)2(CH3CHO), and CH3Mg(OCH(CH3)2) in THF (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.