ACS Publications. Most Trusted. Most Cited. Most Read
My Activity
CONTENT TYPES

Figure 1Loading Img

The Grignard Reaction – Unraveling a Chemical Puzzle

  • Raphael Mathias Peltzer
    Raphael Mathias Peltzer
    Department of Chemistry and Hylleraas Centre for Quantum Molecular Sciences, University of Oslo, P.O. Box 1033 Blindern, Oslo 0315, Norway
  • Jürgen Gauss
    Jürgen Gauss
    Department Chemie, Johannes Gutenberg-Universität Mainz, Duesbergweg 10-14, Mainz 55128, Germany
  • Odile Eisenstein*
    Odile Eisenstein
    Department of Chemistry and Hylleraas Centre for Quantum Molecular Sciences, University of Oslo, P.O. Box 1033 Blindern, Oslo 0315, Norway
    ICGM, Université de Montpellier, CNRS, ENSCM, Montpellier 34095 Cedex 5, France
    *[email protected]
  • , and 
  • Michele Cascella*
    Michele Cascella
    Department of Chemistry and Hylleraas Centre for Quantum Molecular Sciences, University of Oslo, P.O. Box 1033 Blindern, Oslo 0315, Norway
    *[email protected]
Cite this: J. Am. Chem. Soc. 2020, 142, 6, 2984–2994
Publication Date (Web):January 17, 2020
https://doi.org/10.1021/jacs.9b11829

Copyright © 2020 American Chemical Society. This publication is licensed under CC-BY-NC-ND.

  • Open Access

Article Views

80844

Altmetric

-

Citations

LEARN ABOUT THESE METRICS
PDF (2 MB)
Supporting Info (1)»

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

ARTICLE SECTIONS
Jump To

The Grignard reaction is a prominent textbook process to form carbon–carbon bonds. (1) In this reaction, the so-called Grignard reagent, an organomagnesium species RMgX where R is an organic residue and X is a halogen (usually Cl or Br), promotes the addition of its organic residue to an electrophilic substrate. Discovered by Victor Grignard (2) already in 1900, Grignard reagents rapidly became important compounds for organic synthesis in research laboratories and in industry. (3) Even though the reaction today is applied to a large variety of electrophiles, the prototypic substrates are carbonyl moieties R1R2(C═O) yielding a magnesium-alcoholate (eq 1). The R1 and R2 residues usually include hydrogen as well as alkyl-, vinyl-, or aryl- groups. The Grignard reaction usually takes place in ethereal solvents, the final alcohol product being obtained by hydrolysis.
(1)
Despite more than 100 years of extensive studies, the mechanism of this reaction has remained elusive, with little quantitative information and missing consensus. Difficulties in elucidating this mechanism are prominently related to the fact that the ethereal solutions of Grignard reagents contain a variety of chemical species. (4,5) In fact, the nominal reactant RMgX is just a condensed representation of numerous mono- and polymetallic molecules that coexist at equilibrium (eq 2) and whose relative abundance depends on the nature of R and X as well as the experimental conditions. (4,6)
(2)
Furthermore, while the polarity of the C(δ−)–Mg(δ+) bond suggests that the Grignard reagent acts as a nucleophile, it has been argued that an electron-transfer mechanism may occur in the presence of appropriate substrates. (5)
This multifaceted aspect of the Grignard chemistry is per se fascinating; in fact, the lack of understanding of its molecular mechanisms has slowed progress in the rational design of more efficient variants of the reaction. Indeed, while the efficiency of Grignard reagents has been heuristically improved, for example by the synergistic effect of additives like alkali ions, the reasons for these improvements remain unclear. (7) Elucidation of the mechanism of the basic Grignard reaction is the required step to better understand a wide variety of related reactions.
Although the nucleophilic mechanism (also known as the polar mechanism) is currently recognized in the literature as a route to the alcohol product, no experimental study was able to assign the relative reactivity of the various forms of the organomagnesium species. In parallel, experimental evidence consistently supported the presence of an alternative mechanism (called radical or single electron-transfer mechanism) involving the formation of radical intermediates (Figure 1). In particular, products were detected, which could not originate from a nucleophilic addition but could be only obtained through a combination of organic radicals. (8−15) Furthermore, electron spin resonance spectroscopy identified the presence of ketyl radicals in benzophenone/Grignard reagent mixtures. (16) Factors favoring the radical mechanism were searched. Thus, using the distribution of products as an indicator for the mechanism, Ashby and co-workers inferred that an electron-transfer mechanism depends on the reduction potential of the substrate, the oxidation potential of the Grignard reagent, and the polarity of the solvent. Ketones with low reduction potential favored the radical mechanism. Indeed, a radical-type reaction was established for aromatic and conjugated ketones (such as benzophenone and fluorenenone) in the case of tertiary as well as primary Grignard reagents. (14)

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.

Chiral Grignard reagents were also used to probe the reaction pathway. It was in this way shown that chiral secondary Grignard reagents react largely, but not exclusively, by a radical mechanism with benzophenone, but essentially by nucleophilic addition with benzaldehyde. (17,18) The secondary H/D kinetic isotope effect indicated that the competition between the nucleophilic and the single electron-transfer mechanisms depends on the reagent (i.e., magnesium or lithium organometallic species) and the substrate. (19) Recently, it was shown that the Grignard reaction with aliphatic aldehydes does not involve single electron transfer. (20) These last results point at a clear preference for nucleophilic addition for aliphatic aldehyde and a possible competition between the radical and nucleophilic pathways for substrates with lower reduction potential, that is, aromatic or conjugated aldehydes or ketones.
While various aspects of the chemistry of Grignard species, including their formation, (21−23) their structure, and their thermodynamics in solution, (24−27) were addressed by computational approaches, the Grignard reaction itself has been the subject of only a few studies. (28) In particular, in 2002 Yamazaki and Yamabe used density functional theory (DFT) calculations together with implicit solvation models to suggest that dinuclear Grignard moieties are more reactive in a nucleophilic addition than are monomeric species. (29) Similar results were obtained with second-order Møller–Plesset perturbation theory together with implicit solvation models. (30) Yamazaki and Yamabe suggested that a mechanism going through Mg–C homolytic cleavage could be preferred in the case of bulky organic residues, for which the nucleophilic addition is sterically hindered. (29)
Recently, some of us carried out an ab initio molecular dynamics (AIMD) study of CH3MgCl in tetrahydrofuran (THF) to get insight on the role of the solvent in the Schlenk equilibrium. (31) Among other findings, this study showed that the magnesium centers of the mononuclear and dinuclear species can accommodate a variable number of solvent molecules in their first coordination spheres and that the interchange of the methyl and chloride groups between the two magnesium centers is associated with a change in the solvation number. (31) This study thus revealed that it was essential to allow a synergy between the solvent contribution and the chemical event at the Mg centers and to represent the associated solvent reorganization, an effect that cannot be properly described by a static approach.
The purpose of the present study is to use computational approaches to determine the factors that drive the Grignard reaction toward preferred pathways (in particular nucleophilic vs single electron transfer) in the presence of prototypical carbonyl substrates, as well as to elucidate the mechanistic details for each specific transformation. The employment of ab initio molecular dynamics and enhanced-sampling methods in explicit solvent facilitates an unbiased exploration of the conformational space. In particular, it allows a thorough sampling of the solvation states of all organomagnesium species and a reliable description of solvent dynamics. Finally, state-of-the-art quantum-chemical calculations provide solid data to understand the factors that promote a single electron-transfer pathway.

Computational Methods

ARTICLE SECTIONS
Jump To

Ab Initio Molecular Dynamics Simulations

Simulation Protocol

The thermodynamically equilibrated structures of the Grignard reagent in THF in different protomeric and solvated forms were determined in a previous study. (31) In this work, the initial geometry of each reactive complex was built starting from that of the respective Grignard reagent by replacing one THF coordinated to Mg with a substrate molecule. All systems were simulated in an orthorhombic periodic box of dimensions 25.2 × 15.0 × 15.0 Å3, containing 41 THF molecules. The quantum-chemical problem was solved at the DFT (32,33) level using the Perdew–Burke–Ernzerhof (PBE) exchange correlation (xc) functional. (34) The Kohn–Sham orbitals were expanded over mixed Gaussian and plane-wave basis functions. The DZVP basis set for first and second row elements and a molecular optimized basis set for chlorine were employed. (35) The auxiliary plane-wave basis set was expanded up to a 200 Ry cutoff. The core electrons were integrated out using pseudopotentials of the Goedecker–Teter–Hutter type (GTH). (36) Dispersion forces were included using the D3 Grimme approximation. (37) AIMD simulations were run over the ground-state potential energy surface, adopting a threshold for the energy gradient of 10–5 au as numerical convergence criterion for the electronic energy. The equations of motion were integrated employing the velocity Verlet algorithm (38) with a time-step of 0.25 fs. Relaxation to the target temperature was first performed using a canonical sampling/velocity rescaling (CSVR) thermostat (39) with a time constant of 10 fs until the temperature of the system oscillated around the target value. A Nosé–Hoover chain thermostat with a chain length of 3 and a time constant of 1 ps then was used for data production. (40−42) Trajectory analysis was performed using the tools available in the VMD 1.9.2 package. (43)

Polar Mechanism – Activation Energies

The activation barriers for the nucleophilic addition were estimated by constraint dynamics in the blue-moon ensemble (44) using the distance between the methyl group and the carbonyl carbon as reaction coordinate λ. The constraint force f(λ) was collected over independent AIMD runs at nine fixed values of λ spaced by 0.2–0.1 Å. The average value of f at each λ point (⟨fλ) converged with steady mean values of ∼0.5 kcal mol–1 Å–1 within ∼2 ps of simulations. The activation free energy ΔA was computed by trapezoidal integration of ⟨fλ between λ values corresponding to the reactant and the TS. This last point was identified as the value of λ at which ⟨fλ = 0. The work reports Helmholtz free energies as AIMD are performed at constant volume conditions.

Radical Formation Energy

The dissociation energy associated with the homolytic cleavage of the Mg–methyl bond for different Grignard reagent/substrate complexes was estimated as the difference between the total energy of the reactant and that of the individual fragments. THF molecules directly bound to Mg were explicitly considered in the model, while the rest of the solvent was treated implicitly using a continuum dielectric model. (45) The electronic problem was solved at the DFT level using the M06-2X xc functional (46) and the 6-31+G(d,p) basis set. (47) Calculations were performed using the Gaussian 09 package. (48) The choice of the M06-2X functional was done after assessing the quality of different xc functionals in reproducing coupled-cluster (CC) data for the radical formation energy in a simplified model (Table S1). The reference energies were obtained using the CC singles and doubles (CCSD) approximation augmented by a perturbative treatment of triple excitations (CCSD(T)) (49) with an energy convergence threshold of 10–9 hartree, using a cc-pVTZ basis set. (50) The CCSD(T) calculations were carried out using the CFOUR package. (51)

Results and Discussion

ARTICLE SECTIONS
Jump To

Polar Mechanism

Alkyl carbonyl substrates have been shown to prefer a nucleophilic pathway in their reaction with Grignard reagents. (20) For this reason, we modeled the reaction of acetaldehyde (ACA hereafter) with CH3MgCl in THF. All forms of CH3MgCl identified in our previous study as stable species coexisting at thermal equilibrium in THF (31) were considered as potential reactive systems after replacing one coordinated THF by ACA. (31)
Figure 2 shows the labels used throughout this work to identify the organometallic mononuclear and binuclear species and the associated possible pathways, which are characterized by the position of the reactive groups on the Mg centers. The pathway is called geminal if the two groups are on the same magnesium atom and is called vicinal if they are on two different magnesiums. Finally, a pathway where the methyl group bridges the two magnesium atoms was also considered.

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

Both mononuclear and dinuclear Mg species can be associated with geminal pathways. Of the two mononuclear systems, Bgem gives the energetically lowest transition state (TS), with an ΔA of 6.5 ± 0.7 kcal mol–1 (Figure 2). The pathway starting from CH3MgCl (Agem) has a higher lying TS with ΔA = 8.7 ± 0.8 kcal mol–1. Egem is the most reactive dinuclear complex via a geminal pathway. This species is a variant of Bgem in which the coordinated THF is replaced by a Cl-coordinated MgCl2(THF)2 moiety. It is thus not surprising that the TSs associated with Bgem and Egem have similar activation energies. However, our earlier study on the Schlenk equilibrium (31) indicated that Egem is less abundant than Bgem. This suggests that Egem may contribute only marginally to the formation of the product. Species Cgem and Dgem are instead characterized by a significantly higher activation energy (both around 13 kcal mol–1, Figure 2). Finally, the bridging methyl group is unlikely to act as a nucleophile, as indicated by the high activation energy of ΔA = 22.6 kcal mol–1 for Dbrd. As could be expected, a nucleophile is less reactive at a bridging than at a terminal position.

Structural Features of the Geminal Reaction

The reaction in Bgem is used here as a representative case to describe the structural details of the geminal pathway (Figure 3). In THF, Mg(CH3)2 forms Mg(CH3)2(THF)2. This complex has a distorted tetrahedral structure with an angle of 135 ± 9° between the two Mg–CH3 bonds, larger than that formed by the two Mg–O bonds (92 ± 8°, Figure S1). Replacement of one THF molecule by ACA yields compound Bgem, which also has a distorted tetrahedral structure with obtuse CH3–Mg–CH3 and O(ACA)–Mg–O(THF) angles (153 ± 9° and 137 ± 9°, respectively) and an O(ACA)–Mg–CH3 angle of 92 ± 11° (Figure S2). This arrangement leads to a close proximity between the nucleophilic and the electrophilic groups.

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.

In the reactant, the methyl group is tightly bound to Mg with an average Mg–C bond distance of 2.15 ± 0.08 Å. In contrast, ACA is more loosely bound, with a Mg–O distance of 2.4 ± 0.2 Å, which is longer than the Mg–THF distance of 2.2 ± 0.1 Å. The Mg atom lies in the plane of ACA with a C–O–Mg angle of 118 ± 9°, indicating that the magnesium ion interacts with the two lone pairs of the carbonyl oxygen.
At the TS, the addition of one THF molecule from the bulk solvent leads to a complex with a distorted trigonal-bipyramidal structure where ACA and the approaching THF reside at the apical sites with an average O(ACA)–Mg–O(THF) angle of 164 ± 10° (Figure 4). Reaching the TS also requires a rotation of ACA around the C═O axis to establish the interaction of the nucleophile with the carbonyl π system (Figure 4). The formation of the bond between the two carbon atoms is also promoted by a slight bending of ACA toward the methyl group, indicated by a decrease in the C–O–Mg angle to a value of 97 ± 3°. On the contrary, the CH3–Mg–O(ACA) angle of (93 ± 3°) remains similar to that in Bgem.

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.

The geminal reaction occurs via a concerted mechanism, with the TS showing partial formation and cleavage of the C–C and Mg–CH3 bonds, respectively. An informative parameter is the direction of the C3 axis of the methyl group, which is significantly displaced away from the CH3–Mg direction, but not yet aligned to the new C–C direction (ϕ1, ϕ2 angles in Figure 4). The four atoms involved in the bond rearrangement are coplanar (Figure 4), with a CH3···C–O angle of attack of 107 ± 4°, consistent with expected values for a nucleophilic addition to a carbonyl group. (52,53) The four-centered TS has a reactant-like geometry, with a relatively long C···C bond distance of 2.55 Å and the ACA carbon with an out-of-plane angle (defined by the improper CH3–C–O–H dihedral angle) of 161 ± 7°, closer to the ideal value of 180° for an sp2 atom than 120° for an sp3 one. Both ACA and the methyl groups are slightly farther away from the Mg center than in the reactant geometry, as indicated by the Mg–CH3 and Mg–O distances of 2.23 ± 0.07 and 2.5 ± 0.3 Å, respectively.
In the relaxed product CH3Mg(OCH(CH3)2)(THF)2, the Mg atom is again tetracoordinated, with the CH3–Mg–O(iso-propanol) and O(THF)–Mg–O(THF) angles similar to those in the reactant geometry (134 ± 8° and 93 ± 7°, respectively, Figure S3).
The coordination of the solvent assists the heterolytic cleavage of the Mg–CH3 bond by donating electron density to the Mg atom. It appears that the reactivity of the Mg complex increases when the coordinated ligands are better electron donors and/or more ligands coordinate to the magnesium center. These electron-donating ligands screen the charge on the Mg center, making the nucleophilic CH3 a better leaving group. In particular, Mg(CH3)2 is found to be slightly more reactive than CH3MgCl, which is consistent with the known greater electron-donating ability of a methyl group relative to chloride in these magnesium complexes as it also appeared in our previous study. (31) Moreover, the slightly higher reactivity of the mononuclear over dinuclear species correlates to the fact that terminal ligands are better electron donors than are the bridging ones.

Reaction with Vicinal Groups

The vicinal reaction was studied for species Cvic and Fvic (Figure 2). The most solvated dinuclear complex Fvic, with two pentacoordinated Mg atoms, has an activation energy for the C–C bond formation of ΔA ∼ 4.8 kcal mol–1. This is the lowest activation energy that has been calculated for all mononuclear and dinuclear species considered in this study. According to our previous study, Fvic is also one of the most abundant species in the Schlenk equilibrium, with a free energy of only 1.3 kcal mol–1 higher than that of the most stable dichloro-bridged compound (Cvic). (31)Fvic is therefore a key contributor to the formation of product by way of the nucleophilic addition. The most reactive species has ACA and the nucleophilic methyl group initially coordinated to two distinct magnesium atoms (Mg1 and Mg2, respectively) and thus rather far from each other. Considerable structural reorganization is occurring on the way to the transition state. At the TS, characterized by a C···C distance of 2.6 Å, ACA enters in the coordination sphere of Mg2 while still interacting with Mg1, as shown by the similar Mg1–O and Mg2–O distances of 2.12 ± 0.07 and 2.27 ± 0.19 Å, respectively (Figure 4). Consequently, Mg1 retains a pentacoordinated structure at the TS. In parallel, Mg2 acquires an octahedral coordination (Figure 5), with an O(ACA)−Mg2−CH3 angle between the two reacting cis-ligands of ∼90°.

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.

The four-center TS shares strong similarities with that found for the geminal reaction, with an elongated Mg2–CH3 bond (∼2.3 Å), a methyl lone-pair axis twisted ∼30° away from the Mg2–CH3 direction, and a favorable CH3···C═O angle (52,53) of 114 ± 10° (Figure 4). Also, in this case, the ACA carbon essentially retains an sp2 hybridization at the transition, similar to that found earlier for Bgem.
Relaxation from the transition state yields a product where the alcoholate bridges the two magnesium atoms (Figure 5). The chlorine atom that is anti to the nucleophilic CH3 in the reactant leaves its bridging position to occupy a terminal site at Mg2. Reorganization of the bridging ligands yields the final product where Mg1 is tetracoordinated and Mg2 is pentacoordinated (Figure 5).
The significantly higher activation energy (ΔA ∼ 13.0 kcal mol–1) required for the most stable dinuclear species Cvic (Figure 2) to undergo the nucleophilic reaction highlights the role of the coordinated solvent. Both Fvic and Cvic have two bridged chloride ligands and a terminal methyl group on each magnesium atom. Thus, the difference between them resides in the number of THF solvent molecules coordinated to each magnesium, the more reactive species being more solvated. The least solvated Cvic is in fact a rigid species in which ACA cannot approach the nucleophilic methyl group coordinated to the other magnesium. Interestingly, the calculations indicate that an additional THF binds Cvic on the ACA-bonded magnesium prior to reaching the transition state. This event occurs when the carbon–carbon distance is ∼4 Å, still more than 1 Å longer than at the TS. The formation of a pentacoordinated species on the ACA-coordinated Mg facilitates the migration of ACA toward the methyl bearing Mg, enabling the reaction. The solvent thus appears to be a key factor in the reaction by increasing the flexibility of the Mg complex, thus lowering the energy needed for getting the electrophile and nucleophile close to each other (Figure 5).
A previous work proposed that the Grignard reaction would occur starting from a dinuclear structure characterized by a single bridging chloride and two methyl groups each bound to a different Mg atom. (30) Nonetheless, the calculations showed that this complex was not stable in solution and it rapidly evolved into a dichloro-bridged structure. This is consistent with what has been already observed in our previous work: the monochloride bridged structures were not associated with free-energy minima, but they corresponded to the transition states during the transmetalation step in the Schlenk equilibrium. (31) Because of their instability, and the already high reactivity of more stable compounds, it is unlikely that such moieties play a significant role in the Grignard reaction.

Role of the Schlenk Equilibrium in the Nucleophilic Grignard Reaction

Figure 6 compiles the activation energies of the reaction pathways considered in this work, together with the relative stability of the various organomagnesium species involved in the Schlenk equilibrium as previously derived. (31)

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.

The Schlenk equilibrium is driven by the dynamics of the coordinating solvent in the Grignard reagent dimers, (31) and its mechanism shares structural similarities with the vicinal pathway. Specifically, in the Schlenk equilibrium, the increased coordination of the solvent assists the displacement of the ionic ligands from the terminal to the bridging positions by increasing the flexibility of the organomagnesium complex. In the Grignard reaction, the most reactive structure Fvic has two pentacoordinated Mg centers, Mg1 with one THF and ACA, and Mg2 with two THF molecules. This high coordination of the two Mg centers facilitates the shift of ACA from a terminal to a bridging position while minimizing the motion of the other ligands (Figure 5). At the bridging position, the electrophilicity of ACA is significantly increased. In the same time, the coordination of Mg2 augments from five to six, which in turn helps in the departure of the nucleophilic methyl group. Thus, through different coordination to the Mg centers, the solvent has contributed to make the substrate ACA more electrophilic and the methyl group more nucleophilic. The entropic factor also favors the vicinal reaction in this dinuclear complex relative to the most favorable geminal reaction. Indeed, in the latter, the coordination of an additional THF is necessary at the transition state, while in the former, there is no need for a THF from the bulk of the solvent to enter the coordination sphere of any of the two magnesium centers. The dinuclear complex needs only to modify the position of the various ligands and coordinated solvent to reach the transition state. Thus, cleaving the terminal Mg2–CH3 bond is accompanied by the displacement of a bridging chloride to a terminal position to ensure appropriate anionic coordination at Mg2. Two coordinating THF solvent molecules are sufficient to stabilize the pentacordinated Mg2 linked also to a terminal Cl and to the bridging alcoholate and chloride ligands.
In practice, the Grignard reaction can be viewed as a variant of the Schlenk equilibrium where the nonreactive solvent is replaced by an unsaturated substrate that becomes a strong electrophile when entering the coordination sphere of Mg. The energetically easy exchange between terminal and bridging ligands under the assistance of the solvent is at the heart of this chemistry.
While the vicinal pathway through the dinuclear Mg complex (Fvic) looks intrinsically the most reactive, the geminal pathway should nevertheless be considered as a competitive alternative. In particular, mononuclear species have activation energies only slightly higher than that of the most reactive dinuclear complex. Thus, the preferred pathway for the Grignard reaction would be determined by the relative abundance of the mononuclear and dinuclear compounds. Likely, several forms may be active simultaneously in the complex mixture formed by the Grignard reagent in solution. It should be kept in mind that the Grignard reagent forms clusters whose nature and relative abundance depend on the solvent and the halide ligand in a nontrivial manner. (6,54)
Our findings stress the importance of an unbiased exploration of the conformational space and, in particular, of the explicit treatment of the solvent. In fact, the solvent is a direct player in all reactive pathways identified in the present work, in particular by structurally and dynamically affecting the coordination sphere of Mg. This is particularly relevant for the determination of the free energy barrier in the geminal reaction where the C–C bond formation, associated with a cleavage of the Mg–CH3 bond, occurs with the assistance of an incoming solvent entering the coordination sphere of Mg. This ensures that the magnesium is coordinated to at least four ligands as it was found to be required. (31) The direct role of the solvent in modulating the reaction is the main reason for the discrepancies found with the preceding studies by Yamazaki and Yamabe (29) and by Mori and Kato. (30) In particular, our study contrasts with their hypothesis that mononuclear species may be significantly less able to promote the Grignard reaction. In fact, such divergences arise because of, in these previous investigations, a fixed number of ligands at the magnesium centers, in particular disregarding the dynamic role of the solvent and its capability of entering in the coordination sphere during the reaction.

Radical Mechanism

Quantum-chemical calculations indicate that the homolytic cleavage of the Mg–CH3 bond in the solvated Grignard reagent CH3MgCl(THF)n requires a high energy of 66.6 kcal mol–1 for n = 2, which is not lowered by increasing solvation (66.4 kcal mol–1 for n = 3). The calculated values compare well with those determined experimentally for CH3MgBr in diethyl ether (55) (∼61 kcal mol–1), suggesting a marginal influence of the halide and solvent on the Mg–CH3 bond dissociation energy (BDE).
The high value of the Mg–CH3 BDE indicates that alkyl radicals cannot be formed in pure solutions of the Grignard reagent. (56) Indeed, the experiment reports the presence of organic radicals only when electrophilic substrates of low reduction potential are present. (14) This suggests that the substrate may promote the homolytic cleavage.
The M–CH3 BDEs were thus calculated for complexes with coordinated substrates. Figure 7 reports the Mg–CH3 BDE for substrates of decreasing reduction potential (acetaldehyde > formaldehyde > carbonyl difluoride) with one or two coordinated THF molecules. Indeed, the nature of the coordinated substrate influences significantly the Mg–CH3 BDE, as shown by its decrease from acetaldehyde (37.9 kcal mol–1) to carbonyl difluoride (24.8 kcal mol–1) (Figure 7, values for two bound THFs). The strong influence of the substrate on the BDE can be rationalized by considering the localization of the unpaired electron on the organomagnesium radical product. As shown in Figure 7, the electron spin density mostly resides on the substrate carbonyl group and not on the Mg atom regardless of the nature of the substrate and the number of bound THFs. Consequently, the BDE correlates to the energy of the singly occupied π*(CO) Kohn–Sham orbital in the radical product, as indicated by the corresponding orbital energies of −0.18057, −0.18846, and −0.25703 au for acetaldehyde, formaldehyde, and carbonyl difluoride, respectively.

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.

The role of a low-lying empty π* orbital at the substrate in the stabilization of the radical species rationalizes the high BDE values in the absence of any coordinated carbonyl substrate. In fact, in the case of MgCl(THF)2 radical, the unpaired electron remains located on Mg, with about 0.8 e of the spin density. The previous analysis also explains why the degree of solvation has a stronger influence for the BDE in the substrate-bound species than in the substrate-free Grignard reagent. In the presence of the substrate, the delocalization of the unpaired electron produces an effective increase of the electrostatic charge on the Mg atom, which can be stabilized by additional electron-donating ligands. Thus, increased solvation decreases the Mg–CH3 BDE. The effect of the solvent depends only slightly on the nature of the coordinated substrate because the stabilization is from 13 to 10 kcal mol–1 for acetaldehyde to carbonyl difluoride, respectively. For a similar reason, Mg(CH3)2 promotes the homolytic dissociation of the methyl radical more than does CH3MgCl, with a decrease in the BDE by roughly 10%.

Nucleophilic Attack versus Homolytic Cleavage: The Case of Fluorenone

The quantum-chemical calculations of the Mg–CH3 BDE presented here show that the formation of radical species by cleavage of the Mg–CH3 bond is unlikely in the absence of a coordinated substrate with a low-lying empty orbital. This observation accounts for previous conclusions by Otte and Woerpel (20) that the Grignard reaction does not proceed via a radical mechanism with alkyl aldehydes and ketones. The same finding is also consistent with the experimental evidence that a radical mechanism is often preferred for aromatic ketones and, in particular, for benzophenone and fluorenone. (8−15)
To evaluate the competition between the radical and nucleophilic addition pathways in aromatic substrates, we considered the case of the fluorenone. The calculations were carried out with monomeric Mg(CH3)2. This form of the Grignard reagent was selected because it is only slightly less reactive than the most reactive dinuclear complex, while it allows faster ab initio MD simulations.
The Mg–CH3 BDE in CH3MgCl(fluorenone)(THF)2 is 16.5 kcal mol–1, a value more than 20 kcal mol–1 smaller than that for ACA. Such a drastic lowering of the BDE is associated with a significant delocalization of the unpaired electron over the extended π system of fluorenone (Figure 8). In particular, the spin density surrounding the carbonyl carbon integrates to only 0.5 e, as compared to values larger than 0.8 e for the other substrates (Figure 7), with the rest of the spin density being distributed on the conjugated aromatic ring (Figure 8). Considering an entropic contribution to the free energy (−TΔS) for a bond dissociation on the order of 10 kcal mol–1, the formation of a methyl radical would reach an overall free energy cost of only ∼6 kcal mol–1, which is in the same range of the most favorable free energy barriers for the nucleophilic addition mechanism with ACA.

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.

Interestingly, ab initio MD simulations reveal that fluorenone is significantly less reactive than ACA via the nucleophilic pathway. For Mg(CH3)2(fluorenone)(THF), we estimated an activation energy of 14.5 ± 1.2 kcal mol–1 (Figure 8), a value around 8 kcal mol–1 higher than that for ACA.
The TS for the nucleophilic addition to fluorenone has structural features similar to those for ACA, as indicated by a C···C bond distance of 2.5 Å, a C out-of-plane angle of 164 ± 4°, and an approach angle of 106 ± 4°. However, the geometries of the reactant states are significantly different in the two species. For ACA, the Mg–CH3 and C═O bonds in the reactant state are essentially eclipsed, as shown by a CH3–Mg–O–C dihedral angle of 14 ± 15° (Figure 9). Therefore, a rotation of the ACA around the C═O axis by ∼90° is the main structural change required to reach the TS (Figure 9). For fluorenone, the bulky aryl group precludes an eclipsed conformation for the Mg–CH3 and C═O bonds, as indicated by an average CH3–Mg–O–C dihedral of 114 ± 30° (Figure 9). Consequently, in this case, an additional rotation around the Mg–O bond is required to reach the TS. This suggests that the higher barrier calculated in the case of fluorenone originates in good part from a larger structural reorganization going from reactant to TS.

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.

The present calculations help in understanding why the radical mechanism was observed with substrates characterized by a low reduction potential. Interestingly, such substrates are often bulky due to the presence of aryl groups and other conjugated substituents. These large groups may thus require larger structural rearrangement on the way to the TS for the nucleophilic addition, which can further disfavor such a mechanism against the radical one, as shown in the representative case of fluorenone.

Conclusions

ARTICLE SECTIONS
Jump To

The activation of both the nucleophilic and the radical mechanisms depends in a complex way on the nature of the substrate, on its binding to the Mg center, and on solvent dynamics. The combination of these subtle effects is the cause of the difficulties in determining and predicting the preferred pathways for the Grignard reaction.
The nucleophilic reaction occurs between the substrate and the nucleophile, which are both in the coordination sphere of Mg centers. Several organomagnesium species that coexist at thermal equilibrium in solution can promote the nucleophilic reaction with similar activation energies. This implies that several pathways involving different forms of the Grignard reagent may occur in parallel. The number of Mg atoms in the Mg species does not appear to influence significantly its reactivity because the two most reactive species with very similar activation energies are a dinuclear and a mononuclear compound. Likewise, similar activation energies are found when the substrate and the nucleophile are initially coordinated on the same or on nearby Mg atoms.
The solvent plays an essential role in the reaction. More solvated Mg species are more reactive. This seems to be due to the higher flexibility of the entire Mg species, which favors the structural reorganization from the reactant to the transition state. In less solvated species, additional coordination of solvent molecules coming from the bulk is still needed to promote the reaction, but the associated entropic cost makes these species less reactive. Thus, in all pathways, the solvent should be considered as one of the reactants. Similar effects probably apply in numerous organometallic reactions notably involving alkali, alkali-earth metals, and related lanthanide complexes, which are all known to be highly sensitive to solvent effects.
Like for the nucleophilic addition, the radical mechanism is strongly dependent on the coordination of the Mg center. The pure Grignard reagent is unlikely to release an organic radical even if heavily solvated. However, when one of the coordinated ligands is a substrate with a low-lying empty π orbital and when the solvation is high, the Mg–carbon bond breaks much more easily. Interestingly, the substituents able to lower the reduction potential of the substrate are often rather bulky, which in turn disfavors the nucleophilic pathway. The combination of electronic effects stabilizing the radical species and steric hindrance hampering the nucleophilic addition explains why a radical pathway has been observed with substrates like fluorenone or benzophenone but not with benzaldehyde.
The coexistence of multiple species in rapid exchange due to the Schlenk equilibrium and the evidence that the activation energy range of all possible reactions is relatively modest indicate that the Grignard reaction should not be described by an individual process. Instead, it should rather be thought of as an ensemble of transformations that can occur simultaneously in solution. It is likely that improvements of the Grignard reaction, for example, by alkali (7,57) or copper salt additives, (58,59) occur by interference with one or more of the possible pathways. In fact, the determination of all possible reaction pathways for the Grignard reaction and the key factors regulating them establishes the foundation for rational improvements of this process.
In this respect, it is noticeable that most reactive dinuclear species found in this study share strong similarities with the two-metal ion mechanism universally found in nature in the general class of catalytic RNA or metalloendonuclease enzymes. (60) In these cases, hydrolysis of a phosphate ester by a nucleophilic water involves binding to two Mg’s with nonequal 5-, 6-fold coordination, and a vicinal mechanism where the substrate moves into a bridging position with a trigonal geometry, to form a four-center TS with the nucleophile hydroxyl group attacking from its Mg-coordinating position. (61−65) The structural analogy between the two reactions suggests that the evolutionary pressure has selected among the many chemically possible pathways the most advantageous organization of the Mg ions and their coordinating groups to catalyze a nucleophilic attack having both the substrate and the nucleophile as initial coordinating ligands to the metal centers. The Grignard reaction puzzle is finally assembling together with potential impacts on related chemistry.

Supporting Information

ARTICLE SECTIONS
Jump To

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.

Author Information

ARTICLE SECTIONS
Jump To

  • Corresponding Authors
  • Authors
    • Raphael Mathias Peltzer - Department of Chemistry and Hylleraas Centre for Quantum Molecular Sciences, University of Oslo, P.O. Box 1033 Blindern, Oslo 0315, Norway
    • Jürgen Gauss - Department Chemie, Johannes Gutenberg-Universität Mainz, Duesbergweg 10-14, Mainz 55128, Germany
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

ARTICLE SECTIONS
Jump To

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

ARTICLE SECTIONS
Jump To

This article references 65 other publications.

  1. 1
    Vollhardt, K.; Schore, N. Organic Chemistry: Structure and Function; W. H. Freeman: New York, 2014.
  2. 2
    Grignard, 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, 132224
  3. 3
    Rappoport, Z., Marek, I., Eds. The Chemistry of Organomagnesium Compounds; Wiley-VCH: Weinheim, Germany, 2008.
  4. 4
    Schlenk, W.; Schlenk, W. Über die Konstitution der Grignardschen Magnesiumverbindungen. Ber. Dtsch. Chem. Ges. B 1929, 62, 920924,  DOI: 10.1002/cber.19290620422
  5. 5
    Seyferth, D. The Grignard Reagents. Organometallics 2009, 28, 15981605,  DOI: 10.1021/om900088z
  6. 6
    Silverman, G. S.; Rakita, P. E. Handbook of Grignard Reagents; CRC Press: New York, 1996.
  7. 7
    Robertson, S. D.; Uzelac, M.; Mulvey, R. E. Alkali-metal-mediated synergistic effects in polar main group organometallic chemistry. Chem. Rev. 2019, 119, 83328405,  DOI: 10.1021/acs.chemrev.9b00047
  8. 8
    Fauvarque, J.; Rouget, E. Compt. Rend. Hebd. Séances Acad. Sci., Ser. C 1968, 257, 1355
  9. 9
    Blomberg, C.; Mosher, H. A radical process in a reaction of a Grignard compound. J. Organomet. Chem. 1968, 13, 519522,  DOI: 10.1016/S0022-328X(00)82781-3
  10. 10
    Blomberg, C.; Grootveld, H.; Gerner, T.; Bickelhaupt, F. Radical formation during reactions of Grignard reagents with quinones. J. Organomet. Chem. 1970, 24, 549553,  DOI: 10.1016/S0022-328X(00)84483-6
  11. 11
    Ashby, 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, 31623171,  DOI: 10.1021/ja00844a040
  12. 12
    Ashby, 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, 19641966,  DOI: 10.1021/ja00840a066
  13. 13
    Ashby, E. C. A detailed description of the mechanism of reaction of Grignard reagents with ketones. Pure Appl. Chem. 1980, 52, 545569,  DOI: 10.1351/pac198052030545
  14. 14
    Ashby, 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, 22422250,  DOI: 10.1021/ja00399a018
  15. 15
    Lund, T.; Pedersen, M. L.; Frandsen, L. A. Does the reaction between fluorenone and grignard reagents involve free fluorenone anion radicals?. Tetrahedron Lett. 1994, 35, 92259226,  DOI: 10.1016/0040-4039(94)88472-2
  16. 16
    Blomberg, 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, 23852388,  DOI: 10.1021/jo01260a028
  17. 17
    Hoffmann, R. W.; Hölzer, B. Concerted and stepwise Grignard additions, probed with a chiral Grignard reagent. Chem. Commun. 2001, 491492,  DOI: 10.1039/b009678o
  18. 18
    Hoffmann, R. W. The quest for chiral Grignard reagents. Chem. Soc. Rev. 2003, 32, 225230,  DOI: 10.1039/b300840c
  19. 19
    Gajewski, 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, 326334,  DOI: 10.1021/ja982504r
  20. 20
    Otte, 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, 39063909,  DOI: 10.1021/acs.orglett.5b01893
  21. 21
    Garst, J. F.; Soriaga, M. P. Grignard reagent formation. Coord. Chem. Rev. 2004, 248, 623652,  DOI: 10.1016/j.ccr.2004.02.018
  22. 22
    Chen, Z.-N.; Fu, G.; Xu, X. Theoretical studies on Grignard reagent formation: radical mechanism versus non-radical mechanism. Org. Biomol. Chem. 2012, 10, 94919500,  DOI: 10.1039/c2ob26658j
  23. 23
    Shao, Y.; Liu, Z.; Huang, P.; Liu, B. A unified model of Grignard reagent formation. Phys. Chem. Chem. Phys. 2018, 20, 1110011108,  DOI: 10.1039/C8CP01031E
  24. 24
    Axten, 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, 99108,  DOI: 10.1007/BF02265351
  25. 25
    Ehlers, 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, 186194,  DOI: 10.1007/s0089400060186
  26. 26
    Tammiku, J.; Burk, P.; Tuulmets, A. 1,10-phenanthroline and its complexes with magnesium compounds. Disproportionation equilibria. J. Phys. Chem. A 2001, 105, 85548561,  DOI: 10.1021/jp011476n
  27. 27
    Tammiku-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, 133139,  DOI: 10.1021/jp035653r
  28. 28
    Yamabe, S.; Yamazaki, S. In The Chemistry of Organomagnesium Compounds; Rappoport, Z., Marek, I., Eds.; Wiley-VCH: Weinheim, Germany, 2008; pp 369402.
  29. 29
    Yamazaki, S.; Yamabe, S. A computational study on addition of grignard reagents to carbonyl compounds. J. Org. Chem. 2002, 67, 93469353,  DOI: 10.1021/jo026017c
  30. 30
    Mori, 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, 61586165,  DOI: 10.1021/jp9009788
  31. 31
    Peltzer, 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, 42264237,  DOI: 10.1021/acs.jpcb.7b02716
  32. 32
    Hohenberg, P.; Kohn, W. Inhomogeneous electron gas. Phys. Rev. 1964, 136, B864B871,  DOI: 10.1103/PhysRev.136.B864
  33. 33
    Kohn, W.; Sham, L. J. Self-consistent equations including exchange and correlation effects. Phys. Rev. 1965, 140, 1133,  DOI: 10.1103/PhysRev.140.A1133
  34. 34
    Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 38653868,  DOI: 10.1103/PhysRevLett.77.3865
  35. 35
    VandeVondele, J.; Hutter, J. Gaussian basis sets for accurate calculations on molecular systems in gas and condensed phases. J. Chem. Phys. 2007, 127, 114105,  DOI: 10.1063/1.2770708
  36. 36
    Goedecker, S.; Teter, M.; Hutter, J. Separable dual-space Gaussian pseudopotentials. Phys. Rev. B 1996, 54, 17031710,  DOI: 10.1103/PhysRevB.54.1703
  37. 37
    Grimme, 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: 10.1063/1.3382344
  38. 38
    Swope, 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: 10.1063/1.442716
  39. 39
    Bussi, G.; Donadio, D.; Parrinello, M. Canonical sampling through velocity rescaling. J. Chem. Phys. 2007, 126, 014101,  DOI: 10.1063/1.2408420
  40. 40
    Nosé, S. A unified formulation of the constant temperature molecular dynamics methods. J. Chem. Phys. 1984, 81, 511519,  DOI: 10.1063/1.447334
  41. 41
    Hoover, W. G. Canonical dynamics: Equilibrium phase-space distributions. Phys. Rev. A 1985, 31, 16951697,  DOI: 10.1103/PhysRevA.31.1695
  42. 42
    Martyna, G. J.; Klein, M. L.; Tuckerman, M. Nosé—Hoover chains: The canonical ensemble via continuous dynamics. J. Chem. Phys. 1992, 97, 26352643,  DOI: 10.1063/1.463940
  43. 43
    Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual molecular dynamics. J. Mol. Graphics 1996, 14, 3338,  DOI: 10.1016/0263-7855(96)00018-5
  44. 44
    Carter, E. A.; Ciccotti, G.; Heynes, J. T.; Kapral, R. Constrained reaction coordinate dynamics for the simulation of rare events. Chem. Phys. Lett. 1989, 156, 472477,  DOI: 10.1016/S0009-2614(89)87314-2
  45. 45
    Marenich, 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, 63786396,  DOI: 10.1021/jp810292n
  46. 46
    Zhao, 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, 215241,  DOI: 10.1007/s00214-007-0310-x
  47. 47
    Ditchfield, 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, 724728,  DOI: 10.1063/1.1674902
  48. 48
    Frisch, 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.
  49. 49
    Raghavachari, K.; Trucks, G. W.; Pople, J. A.; Head-Gordon, M. A fifth-order perturbation comparison of electron correlation theories. Chem. Phys. Lett. 1969, 157, 479483,  DOI: 10.1016/S0009-2614(89)87395-6
  50. 50
    Dunning, T. H. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 10071023,  DOI: 10.1063/1.456153
  51. 51
    CFOUR, 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.
  52. 52
    Buergi, H. B.; Dunitz, J. D.; Shefter, E. Geometrical reaction coordinates. II. Nucleophilic addition to a carbonyl group. J. Am. Chem. Soc. 1973, 95, 50655067,  DOI: 10.1021/ja00796a058
  53. 53
    Buergi, H. B.; Lehn, J. M.; Wipff, G. Ab initio study of nucleophilic addition to a carbonyl group. J. Am. Chem. Soc. 1974, 96, 19561957,  DOI: 10.1021/ja00813a062
  54. 54
    Walker, 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, 38453850,  DOI: 10.1021/ja01042a027
  55. 55
    Holm, T. Thermochemical bond dissociation energies of carbon-magnesium bonds. J. Chem. Soc., Perkin Trans. 2 1981, 2, 464467,  DOI: 10.1039/P29810000464
  56. 56
    We 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, 1225912270,  DOI: 10.1021/jp202762p
  57. 57
    Ziegler, D. S.; Wei, B.; Knochel, P. Improving the halogen–magnesium exchange by using new turbo-Grignard reagents. Chem. - Eur. J. 2019, 25, 26952703,  DOI: 10.1002/chem.201803904
  58. 58
    Harutyunyan, 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, 91039118,  DOI: 10.1021/ja0585634
  59. 59
    Lopez, F.; Minaard, A. J.; Feringa, B. L. Catalytic enantioselective conjugate addition with Grignard reagents. Acc. Chem. Res. 2007, 40, 179188,  DOI: 10.1021/ar0501976
  60. 60
    Seitz, T. A.; Seitz, J. A. A general two-metal-ion mechanism for catalytic RNA. Proc. Natl. Acad. Sci. U. S. A. 1993, 90, 64986502,  DOI: 10.1073/pnas.90.14.6498
  61. 61
    De 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, 1095510962,  DOI: 10.1021/ja8005786
  62. 62
    Rosta, 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, 89348941,  DOI: 10.1021/ja200173a
  63. 63
    Casalino, 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, 1037410377,  DOI: 10.1021/jacs.6b01363
  64. 64
    Genna, 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, 1459214598,  DOI: 10.1021/jacs.6b05475
  65. 65
    Genna, V.; Donati, E.; De Vivo, M. The catalytic mechanism of DNA and RNA polymerases. ACS Catal. 2018, 8, 1110311118,  DOI: 10.1021/acscatal.8b03363

Cited By

ARTICLE SECTIONS
Jump To

This article is cited by 78 publications.

  1. 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
  2. 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
  3. 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
  4. 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
  5. 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
  6. 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
  7. 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
  8. 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
  9. 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
  10. 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
  11. 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
  12. 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
  13. 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
  14. 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
  15. 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
  16. 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
  17. 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
  18. 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
  19. 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
  20. 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
  21. 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
  22. 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
  23. 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
  24. 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
  25. 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
  26. 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
  27. 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
  28. 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
  29. 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
  30. 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
  31. 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
  32. 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
  33. 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
  34. 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
  35. 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
  36. 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
  37. 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
  38. 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
  39. 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
  40. 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
  41. 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
  42. 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
  43. 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
  44. 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
  45. 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
  46. Alexander Düfert. Carbonylchemie. 2023, 39-186. https://doi.org/10.1007/978-3-662-65244-2_2
  47. 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
  48. 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
  49. 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
  50. 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
  51. 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
  52. 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
  53. 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
  54. 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
  55. 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
  56. 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
  57. 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
  58. 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
  59. 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
  60. 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
  61. 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
  62. 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
  63. 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
  64. S. Chantal E. Stieber. Computational Methods in Organometallic Chemistry. 2022, 176-210. https://doi.org/10.1016/B978-0-12-820206-7.00099-8
  65. 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
  66. 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
  67. 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
  68. 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
  69. Eamonn F. Healy. Organic chemistry as representation. Foundations of Chemistry 2021, 23 (1) , 59-68. https://doi.org/10.1007/s10698-020-09379-z
  70. 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
  71. 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
  72. 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
  73. 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
  74. 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
  75. 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
  76. 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
  77. 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
  78. 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 SECTIONS
    Jump To

    This article references 65 other publications.

    1. 1
      Vollhardt, K.; Schore, N. Organic Chemistry: Structure and Function; W. H. Freeman: New York, 2014.
    2. 2
      Grignard, 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, 132224
    3. 3
      Rappoport, Z., Marek, I., Eds. The Chemistry of Organomagnesium Compounds; Wiley-VCH: Weinheim, Germany, 2008.
    4. 4
      Schlenk, W.; Schlenk, W. Über die Konstitution der Grignardschen Magnesiumverbindungen. Ber. Dtsch. Chem. Ges. B 1929, 62, 920924,  DOI: 10.1002/cber.19290620422
    5. 5
      Seyferth, D. The Grignard Reagents. Organometallics 2009, 28, 15981605,  DOI: 10.1021/om900088z
    6. 6
      Silverman, G. S.; Rakita, P. E. Handbook of Grignard Reagents; CRC Press: New York, 1996.
    7. 7
      Robertson, S. D.; Uzelac, M.; Mulvey, R. E. Alkali-metal-mediated synergistic effects in polar main group organometallic chemistry. Chem. Rev. 2019, 119, 83328405,  DOI: 10.1021/acs.chemrev.9b00047
    8. 8
      Fauvarque, J.; Rouget, E. Compt. Rend. Hebd. Séances Acad. Sci., Ser. C 1968, 257, 1355
    9. 9
      Blomberg, C.; Mosher, H. A radical process in a reaction of a Grignard compound. J. Organomet. Chem. 1968, 13, 519522,  DOI: 10.1016/S0022-328X(00)82781-3
    10. 10
      Blomberg, C.; Grootveld, H.; Gerner, T.; Bickelhaupt, F. Radical formation during reactions of Grignard reagents with quinones. J. Organomet. Chem. 1970, 24, 549553,  DOI: 10.1016/S0022-328X(00)84483-6
    11. 11
      Ashby, 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, 31623171,  DOI: 10.1021/ja00844a040
    12. 12
      Ashby, 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, 19641966,  DOI: 10.1021/ja00840a066
    13. 13
      Ashby, E. C. A detailed description of the mechanism of reaction of Grignard reagents with ketones. Pure Appl. Chem. 1980, 52, 545569,  DOI: 10.1351/pac198052030545
    14. 14
      Ashby, 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, 22422250,  DOI: 10.1021/ja00399a018
    15. 15
      Lund, T.; Pedersen, M. L.; Frandsen, L. A. Does the reaction between fluorenone and grignard reagents involve free fluorenone anion radicals?. Tetrahedron Lett. 1994, 35, 92259226,  DOI: 10.1016/0040-4039(94)88472-2
    16. 16
      Blomberg, 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, 23852388,  DOI: 10.1021/jo01260a028
    17. 17
      Hoffmann, R. W.; Hölzer, B. Concerted and stepwise Grignard additions, probed with a chiral Grignard reagent. Chem. Commun. 2001, 491492,  DOI: 10.1039/b009678o
    18. 18
      Hoffmann, R. W. The quest for chiral Grignard reagents. Chem. Soc. Rev. 2003, 32, 225230,  DOI: 10.1039/b300840c
    19. 19
      Gajewski, 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, 326334,  DOI: 10.1021/ja982504r
    20. 20
      Otte, 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, 39063909,  DOI: 10.1021/acs.orglett.5b01893
    21. 21
      Garst, J. F.; Soriaga, M. P. Grignard reagent formation. Coord. Chem. Rev. 2004, 248, 623652,  DOI: 10.1016/j.ccr.2004.02.018
    22. 22
      Chen, Z.-N.; Fu, G.; Xu, X. Theoretical studies on Grignard reagent formation: radical mechanism versus non-radical mechanism. Org. Biomol. Chem. 2012, 10, 94919500,  DOI: 10.1039/c2ob26658j
    23. 23
      Shao, Y.; Liu, Z.; Huang, P.; Liu, B. A unified model of Grignard reagent formation. Phys. Chem. Chem. Phys. 2018, 20, 1110011108,  DOI: 10.1039/C8CP01031E
    24. 24
      Axten, 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, 99108,  DOI: 10.1007/BF02265351
    25. 25
      Ehlers, 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, 186194,  DOI: 10.1007/s0089400060186
    26. 26
      Tammiku, J.; Burk, P.; Tuulmets, A. 1,10-phenanthroline and its complexes with magnesium compounds. Disproportionation equilibria. J. Phys. Chem. A 2001, 105, 85548561,  DOI: 10.1021/jp011476n
    27. 27
      Tammiku-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, 133139,  DOI: 10.1021/jp035653r
    28. 28
      Yamabe, S.; Yamazaki, S. In The Chemistry of Organomagnesium Compounds; Rappoport, Z., Marek, I., Eds.; Wiley-VCH: Weinheim, Germany, 2008; pp 369402.
    29. 29
      Yamazaki, S.; Yamabe, S. A computational study on addition of grignard reagents to carbonyl compounds. J. Org. Chem. 2002, 67, 93469353,  DOI: 10.1021/jo026017c
    30. 30
      Mori, 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, 61586165,  DOI: 10.1021/jp9009788
    31. 31
      Peltzer, 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, 42264237,  DOI: 10.1021/acs.jpcb.7b02716
    32. 32
      Hohenberg, P.; Kohn, W. Inhomogeneous electron gas. Phys. Rev. 1964, 136, B864B871,  DOI: 10.1103/PhysRev.136.B864
    33. 33
      Kohn, W.; Sham, L. J. Self-consistent equations including exchange and correlation effects. Phys. Rev. 1965, 140, 1133,  DOI: 10.1103/PhysRev.140.A1133
    34. 34
      Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 38653868,  DOI: 10.1103/PhysRevLett.77.3865
    35. 35
      VandeVondele, J.; Hutter, J. Gaussian basis sets for accurate calculations on molecular systems in gas and condensed phases. J. Chem. Phys. 2007, 127, 114105,  DOI: 10.1063/1.2770708
    36. 36
      Goedecker, S.; Teter, M.; Hutter, J. Separable dual-space Gaussian pseudopotentials. Phys. Rev. B 1996, 54, 17031710,  DOI: 10.1103/PhysRevB.54.1703
    37. 37
      Grimme, 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: 10.1063/1.3382344
    38. 38
      Swope, 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: 10.1063/1.442716
    39. 39
      Bussi, G.; Donadio, D.; Parrinello, M. Canonical sampling through velocity rescaling. J. Chem. Phys. 2007, 126, 014101,  DOI: 10.1063/1.2408420
    40. 40
      Nosé, S. A unified formulation of the constant temperature molecular dynamics methods. J. Chem. Phys. 1984, 81, 511519,  DOI: 10.1063/1.447334
    41. 41
      Hoover, W. G. Canonical dynamics: Equilibrium phase-space distributions. Phys. Rev. A 1985, 31, 16951697,  DOI: 10.1103/PhysRevA.31.1695
    42. 42
      Martyna, G. J.; Klein, M. L.; Tuckerman, M. Nosé—Hoover chains: The canonical ensemble via continuous dynamics. J. Chem. Phys. 1992, 97, 26352643,  DOI: 10.1063/1.463940
    43. 43
      Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual molecular dynamics. J. Mol. Graphics 1996, 14, 3338,  DOI: 10.1016/0263-7855(96)00018-5
    44. 44
      Carter, E. A.; Ciccotti, G.; Heynes, J. T.; Kapral, R. Constrained reaction coordinate dynamics for the simulation of rare events. Chem. Phys. Lett. 1989, 156, 472477,  DOI: 10.1016/S0009-2614(89)87314-2
    45. 45
      Marenich, 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, 63786396,  DOI: 10.1021/jp810292n
    46. 46
      Zhao, 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, 215241,  DOI: 10.1007/s00214-007-0310-x
    47. 47
      Ditchfield, 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, 724728,  DOI: 10.1063/1.1674902
    48. 48
      Frisch, 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.
    49. 49
      Raghavachari, K.; Trucks, G. W.; Pople, J. A.; Head-Gordon, M. A fifth-order perturbation comparison of electron correlation theories. Chem. Phys. Lett. 1969, 157, 479483,  DOI: 10.1016/S0009-2614(89)87395-6
    50. 50
      Dunning, T. H. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 10071023,  DOI: 10.1063/1.456153
    51. 51
      CFOUR, 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.
    52. 52
      Buergi, H. B.; Dunitz, J. D.; Shefter, E. Geometrical reaction coordinates. II. Nucleophilic addition to a carbonyl group. J. Am. Chem. Soc. 1973, 95, 50655067,  DOI: 10.1021/ja00796a058
    53. 53
      Buergi, H. B.; Lehn, J. M.; Wipff, G. Ab initio study of nucleophilic addition to a carbonyl group. J. Am. Chem. Soc. 1974, 96, 19561957,  DOI: 10.1021/ja00813a062
    54. 54
      Walker, 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, 38453850,  DOI: 10.1021/ja01042a027
    55. 55
      Holm, T. Thermochemical bond dissociation energies of carbon-magnesium bonds. J. Chem. Soc., Perkin Trans. 2 1981, 2, 464467,  DOI: 10.1039/P29810000464
    56. 56
      We 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, 1225912270,  DOI: 10.1021/jp202762p
    57. 57
      Ziegler, D. S.; Wei, B.; Knochel, P. Improving the halogen–magnesium exchange by using new turbo-Grignard reagents. Chem. - Eur. J. 2019, 25, 26952703,  DOI: 10.1002/chem.201803904
    58. 58
      Harutyunyan, 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, 91039118,  DOI: 10.1021/ja0585634
    59. 59
      Lopez, F.; Minaard, A. J.; Feringa, B. L. Catalytic enantioselective conjugate addition with Grignard reagents. Acc. Chem. Res. 2007, 40, 179188,  DOI: 10.1021/ar0501976
    60. 60
      Seitz, T. A.; Seitz, J. A. A general two-metal-ion mechanism for catalytic RNA. Proc. Natl. Acad. Sci. U. S. A. 1993, 90, 64986502,  DOI: 10.1073/pnas.90.14.6498
    61. 61
      De 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, 1095510962,  DOI: 10.1021/ja8005786
    62. 62
      Rosta, 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, 89348941,  DOI: 10.1021/ja200173a
    63. 63
      Casalino, 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, 1037410377,  DOI: 10.1021/jacs.6b01363
    64. 64
      Genna, 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, 1459214598,  DOI: 10.1021/jacs.6b05475
    65. 65
      Genna, V.; Donati, E.; De Vivo, M. The catalytic mechanism of DNA and RNA polymerases. ACS Catal. 2018, 8, 1110311118,  DOI: 10.1021/acscatal.8b03363
  • Supporting Information

    Supporting Information

    ARTICLE SECTIONS
    Jump To

    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.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect