ACS Publications. Most Trusted. Most Cited. Most Read
How Solvent Dynamics Controls the Schlenk Equilibrium of Grignard Reagents: A Computational Study of CH3MgCl in Tetrahydrofuran
My Activity

Figure 1Loading Img
  • Open Access
Article

How Solvent Dynamics Controls the Schlenk Equilibrium of Grignard Reagents: A Computational Study of CH3MgCl in Tetrahydrofuran
Click to copy article linkArticle link copied!

View Author Information
Department of Chemistry and Centre for Theoretical and Computational Chemistry (CTCC), University of Oslo, Postbox 1033 Blindern, 0315 Oslo, Norway
Institut Charles Gerhardt, UMR 5253 CNRS-Université de Montpellier, Université de Montpellier, cc 1501, Place E. Bataillon, 34095 Montpellier, France
*E-mail: [email protected] (A.N.).
*E-mail: [email protected] (M.C.).
Open PDFSupporting Information (1)

The Journal of Physical Chemistry B

Cite this: J. Phys. Chem. B 2017, 121, 16, 4226–4237
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.jpcb.7b02716
Published March 30, 2017

Copyright © 2017 American Chemical Society. This publication is licensed under these Terms of Use.

Abstract

Click to copy section linkSection link copied!

The Schlenk equilibrium is a complex reaction governing the presence of multiple chemical species in solution of Grignard reagents. The full characterization at the molecular level of the transformation of CH3MgCl into MgCl2 and Mg(CH3)2 in tetrahydrofuran (THF) by means of ab initio molecular dynamics simulations with enhanced-sampling metadynamics is presented. The reaction occurs via formation of dinuclear species bridged by chlorine atoms. At room temperature, the different chemical species involved in the reaction accept multiple solvation structures, with two to four THF molecules that can coordinate the Mg atoms. The energy difference between all dinuclear solvated structures is lower than 5 kcal mol–1. The solvent is shown to be a direct key player driving the Schlenk mechanism. In particular, this study illustrates how the most stable symmetrically solvated dinuclear species, (THF)CH3Mg(μ-Cl)2MgCH3(THF) and (THF)CH3Mg(μ-Cl)(μ-CH3)MgCl(THF), need to evolve to less stable asymmetrically solvated species, (THF)CH3Mg(μ-Cl)2MgCH3(THF)2 and (THF)CH3Mg(μ-Cl)(μ-CH3)MgCl(THF)2, in order to yield ligand exchange or product dissociation. In addition, the transferred ligands are always departing from an axial position of a pentacoordinated Mg atom. Thus, solvent dynamics is key to successive Mg–Cl and Mg–CH3 bond cleavages because bond breaking occurs at the most solvated Mg atom and the formation of bonds takes place at the least solvated one. The dynamics of the solvent also contributes to keep relatively flat the free energy profile of the Schlenk equilibrium. These results shed light on one of the most used organometallic reagents whose structure in solvent remains experimentally unresolved. These results may also help to develop a more efficient catalyst for reactions involving these species.

Copyright © 2017 American Chemical Society

Introduction

Click to copy section linkSection link copied!

In 1900, the presentation by Victor Grignard of a new compound, completely soluble in ether and formed upon reaction of magnesium with alkyl halides, (1) paved the way to establish one of the most widely used organometallic reagents in organic synthesis, worth the Noble prize in 1912. (2) The so-called Grignard reagent, with nominal formula RMgX, is used in nucleophilic addition or substitution reactions, taking advantage of the enriched electron density in the R group coordinated to the Mg atom.
Initial limitations to the direct use of this reagent were found due to its relatively modest tolerance to different functional groups compared to that of corresponding boron and zinc metals. However, the use of Ni catalysts in coupling reactions using the Grignard reagent, independently developed by Corriu and Massé and Kumada and Tamao, increased the relevance of this reagent. (3, 4) Nowadays, interest in the functionalization of phenol-based derivatives involving earth-abundant metal catalysts and the development of efficient synthetic protocols allowing work at very mild conditions and with large functional tolerance make organomagnesium reagents very attractive. (5-12)
Despite the use of Grignard reagents in organic synthesis, their molecular structure in solution has remained elusive so far, even if considerable research effort has been spent on this topic. In an illuminating essay, Seyferth described the difficulties in characterizing the structures of Grignard reagents, which can be affected by both the solvent, and the halide and organic groups bonded to Mg. (13) Solid-state structures determined by X-ray diffraction studies revealed a diversity of coordination modes for Mg, and of nuclearity for the molecular system. Tetrahedral coordination is commonly observed in crystals structures obtained from ether solutions as EtMgBr(OEt2)2, (14) where two molecules of solvent are bound to Mg. Instead, both trigonal-bipyramidal and square-pyramidal structures were found for CH3MgBrTHF3 when working in tetrahydrofuran (THF) solvent. (15) X-ray data of aggregates of two EtMg2Cl3 moieties showed a greater than four coordination for the Mg centers. (16, 17) Due to lattice packing restrictions, the structures of Grignard reagents observed in the crystalline solid state may not be necessarily representative of the stable species present in solution. Investigations using molecular weight, (18) calorimetric measurements, (19) and NMR and IR spectroscopy (20) support in fact the existence of structures other than those detected by X-ray crystallography.
Determination of the structures of Grignard reagents in solution is particularly complex because of the presence of multiple chemical species at thermodynamic equilibrium. This complication has been evident since 1929, when Schlenk father and son (21) proposed that redistribution of ligands yielding MgR2 and MgX2 from RMgX could take place (original Schlenk equilibrium in Scheme 1). At present, it is commonly recognized that the halide groups in RMgX and MgX2 tend to form bridges between magnesium atoms to yield dimers and oligomeric structures, which are in equilibrium with monomeric species (generalized Schlenk equilibrium in Scheme 1). This causes association between the various moieties, eventually favoring ligand exchange. The degree of association among the Mg centers is dependent on the solvent, as shown by Ashby and Walker, who found a degree of association of around ∼1 in THF and between 1 and 4, depending on the nature of R and X, in diethyl ether. (22) Further studies suggested that some complex equilibrium between RMgCl(THF)n, MgR2(μ-Cl)2Mg(THF)5/4 and RMg2(μ-Cl)3(THF)5 better describes the traditional Schlenk equilibrium. (23, 24)

Scheme 1

Scheme 1. Schlenk Equilibria
Given the complexity of the problem, computational modeling offers an excellent tool to determine in detail the chemical nature of the species involved in the Schlenk equilibrium. In the past, examination of the Cambridge Crystallographic Database and density functional theory (DFT) (25-27) calculations for isolated molecules led to establishing the overwhelming prevalence of structures with two over three bridging ligands and a preference for halide over alkyl ligands in that position. Moreover, they observed that the maximum stabilizing effect of the coordinating solvent is obtained for dimeric moieties. (28) Including the solvent as a continuum model in a systematic way and combining generalized valence bond, second-order Møller–Plesset perturbation theory (MP2), and DFT calculations indicated that radical and charged species may be present at equilibrium, although in small concentrations. (29) Explicit solvent was limited to the first coordination sphere of the Mg atom, and in many cases, the total number of ligands at Mg including the solvent was fixed to four. However, both crystallographic (30-32) and computational studies (33) showed that Mg could expand its coordination sphere over more than four ligands, up to hexacoordination. This limitation is probably even more inappropriate when trying to understand the dynamic processes involved in the Schlenk equilibrium (Scheme 1).
In consideration of the longstanding and recently renewed importance of Grignard in organic synthesis, (34-36) it is important to reach a better understanding of its structure and dynamic behavior in solvent. In this respect, dynamic modeling approaches may constitute the optimal choice for an accurate description of the Grignard reagent. In the past decades, ab initio molecular dynamics simulations (AIMD) have been shown to be especially appropriate to study processes in the liquid phase. (37, 38) For example, this technique was used to study structural properties of various metal ions in solution, including Na, (39) Mg, (40, 41) or Co and Ag. (42) In addition, AIMD with enhanced sampling methods were effectively used to provide an accurate description of solvent effects in various organic (43, 44) and organometallic reactions. (45, 46) Larger metallo-organic systems have also been been treated by coupling AIMD to molecular mechanics schemes. (47) This approach was used to describe, for example, electronic solute–solvent coupling in photoexcitation dynamics of ruthenium metallo-organic compounds, (48) the reaction mechanisms of hydrolysis in zinc β-lactamases, (49, 50) the redox properties of copper in azurin, (51) the functional role of Mg2+ ions in ribonuclease H, (52) and the design of organoruthenium anticancer complexes. (53)
In this work, this methodology was used to investigate the equilibria that take place when CH3MgCl, a model of the Grignard reagent, is dissolved in THF. Our study was also complemented by DFT static calculations on selected structures identified as minima by ab initio molecular dynamics. The data presented in this work show that multiple structures with different solvation patterns are in equilibrium and evidence how the dynamic exchange of solvent molecules in the Mg coordination sphere is key to evolution from the reactant to product in the Schlenk equilibrium.

Computational Methods

Click to copy section linkSection link copied!

System Setup

The Schlenk equilibrium was investigated through a set of simulations describing monomeric reactants, monomeric products, and different intermediate dimeric states.

Monomeric Species

We built three different systems, each containing 1 molecule of either CH3MgCl, MgCl2, or Mg(CH3)2 surrounded by 25 molecules of THF in a periodic box of dimensions 15.0 × 15.0 × 15.0 Å3. The initial coordinates of THF were obtained from a previous 20 ps long equilibration run in a similar box containing the pure solvent. The number of solvent molecules was set to be consistent with the experimental density of THF at room temperature. (54) All systems were first relaxed for 15 ps in the microcanonical ensemble at energies matching an average temperature of 300 K. Relaxation at the target temperature was performed using a canonical sampling/velocity rescaling (CSVR) thermostat with a time constant of 10 fs until the temperature of the system oscillated around the target value. Production runs were simulated in the NVT ensemble at 300 K. The Nosé–Hoover chain thermostat with a chain length of 3 and time constant of 1 ps was used for data production. (55-57)

Dimeric Species

The (MgCH3Cl)2 species were simulated in an orthorhombic periodic box of dimensions 25.2 × 15.0 × 15.0 Å3, containing 42 THF molecules (Figure 1). A larger box of dimensions 25.2 × 20.0 × 20.0 Å3 with 74 THF molecules was also simulated to verify that there is no significant bias due to finite-size effects (see the Supporting Information). The systems were thermalized following the same protocol as that described for the monomers.

Figure 1

Figure 1. Simulation box used in this study. The atoms of the Grignard reagent are represented by spheres and the THF solvent molecules by sticks in licorice and red. Hydrogen atoms of THF are not shown for clarity.

Ab Initio Molecular Dynamics Simulations

The electronic problem was solved by DFT (25, 26) using the Perdew–Burke–Ernzerhof approximation to the exchange–correlation functional. (58) Kohn–Sham orbitals were expanded over mixed Gaussian and plane-wave basis functions. (59) The DZVP basis set for first and second row elements and Mg and a molecularly optimized basis set for the chlorine atoms were employed. (60) The auxiliary plane-wave basis set was expanded to a 200 Ry cutoff. The core electrons were integrated using pseudopotentials of Goedecker–Teter–Hutter type. (61) Dispersion forces were accounted for using the D3 Grimme approximation. (62) AIMD simulations were run over the Born–Oppenheimer surface, with a time step of 0.25 fs, optimizing the energy gradient to a threshold of 10–5 au.

Free-Energy Calculations

Exploration of the conformational and reactive landscape and determination of the corresponding free-energy surface (FES) were performed by coupling AIMD to metadynamics simulations. (37, 63) Solvation of different chemical species observed during the AIMD runs, as well as the transmetalation reaction, was investigated by independent metadynamics runs. All collective variables (CVs) employed in this study were defined as the coordination number of specific ligand species to individual Mg atoms. The coordination number of a species X around Mg at a given time t (CN[X](t)) was evaluated according to the formula shown in eq 1, as defined in previous works (38)(1)where NX is the number of species X present in the system, di is the distance of the ith atom X from Mg, and d0, p, and q are free parameters (see the Supporting Information).
The time-dependent bias potential was formed by sets of Gaussians of 0.25 kcal mol–1 height and 0.04 width for coordination variables and added every 50 steps of AIMD for the dichloride bridged system. Additional details of the metadynamics parameters used as well as the associated errors are given in Table S1 of the Supporting Information.
The metadynamics simulations convergence was confirmed by checking that the calculations reached the diffusion limit within their wall constraints; statistical errors, computed according to refs 64 and 65, are within 1–3kT (see the Supporting Information).
AIMD runs were computed using the QUICKSTEP (66, 67) module of the CP2K 2.5.1 package. Trajectory analysis was performed using the tools available in the VMD 1.9.2 package. (68)

Static Calculations and Electronic Structure Analysis

Chemically relevant geometries sampled by AIMD were fully optimized at the DFT(PBE+GD3) level by using the Gaussian09 software package. (69) Mg, C, H, and O were described with the all-electron double-ζ 6-31+G** basis set. (70-72) Vibrational frequencies were computed analytically to verify that the stationary points found were energy minima. In addition to the solvent molecules bound to Mg, implicit solvation was modeled by using the SMD solvation model. (73) In selected cases, Gibbs energies were obtained for T = 298.15 K and p = 1 atm. In the bimolecular steps, these energies were corrected for the 1 M standard state (T = 298.15 K and p = 24.465 atm). Donor–acceptor interactions were explored by means of natural bond orbital (NBO) calculations (NBO6 version). (74) The nature of these interactions was determined by computing the associated natural localized molecular orbitals (NLMOs). (75)

Results

Click to copy section linkSection link copied!

We restricted our investigation to monomers and dinuclear species, excluding higher aggregation states, because experimental studies showed that in THF the aggregation is very low, mainly consisting of monomeric species. (22)

Monomeric Structures

The solvation of CH3MgCl, Mg(CH3)2, and MgCl2 species was investigated over 40 ps of AIMD simulations at 300 K. CH3MgCl and Mg(CH3)2 were found with only two THF molecules in the solvation sphere of Mg and tetrahedral coordination at this atom (Figure 2). On the contrary, both tetrameric and pentameric coordination were found for MgCl2. The nature of the solvation of MgCl2 was then further investigated by metadynamics simulations, using the coordination number of oxygen atoms of THF to Mg as the CV. These simulations showed that the most likely solvation state is MgCl2(THF)3 arranged in trigonal-bipyramidal coordination at Mg. MgCl2(THF)3 is ∼1.6 and ∼2.8 kcal mol–1 more stable than the tetrahedral or octahedral complexes with two and four coordinating THF molecules, respectively. Thus, at room temperature, the different solvation states are in equilibrium in solution in a 0.07:1.0:0.01 ratio for the di-, tri-, and tetra-solvated complexes, yielding an average MgCl2(THF)x solvation with value x = 2.9.

Figure 2

Figure 2. Most likely solvation structures for MgCl2 (left), Mg(CH3)2 (middle), and (CH3)MgCl (right). The dotted black lines represent coordination of the ligands to the Mg center.

DFT optimization of the structures located as minima on the FES showed that they are also minima on the potential energy surface (PES) with similar structural features (these structures are shown in Figure 2). NBO analysis on these optimized structures showed higher electron donation of CH3 to Mg compared to Cl. This is shown by the Mg contribution on the lone pairs’ NLMOs of CH3 and Cl (χMg), which are 10.2 and 5.3%, respectively, in CH3MgCl (see Figure S1 and Table S1 for further details). This may explain the need for additional coordination of THF in MgCl2 to improve the screening of the Mg charge.

Dinuclear Structures

The formation of dinuclear species was first studied by performing an AIMD simulation at 300 K starting from two molecules of CH3MgCl placed with their respective Mg atoms at a distance of 9.5 Å. In order to accelerate the dimerization, a bias restraining potential to induce the approach of the two molecules was used (see the Computational Methods section). A dimeric structure with two bridging chlorides was formed as soon as the two Mg atoms were at a distance of 5.5 Å.

Cl/CH3 Exchange Process

Chloride/methyl exchange between the two Mg centers was investigated combining AIMD and metadynamics simulations, using as CVs the THF solvation number on one arbitrary Mg atom (CV1) and the difference in methyl coordination between the two Mg atoms (CV2). The CVs where chosen to monitor the free energy change upon methyl transfer between the Mg atoms as a function of the local solvation of Mg.
The resulting FES (Figure 3) shows five separated wells (A, B, C, D, and E) that can be classified according to the nature of the Mg bridging groups. The most representative structures captured in each well are shown in Figure 3. Wells A and B correspond to dichloride bridged dinuclear species, C and D correspond to the mixed Me and Cl bridged species, and E encloses monochloride bridged structures of formula ClMg(μ-Cl)MgMe2. While wells A and B represent the structures of the reactants, well E represents the preproduct state prior to final dissociation into MgCl2 and Mg(CH3)2.

Figure 3

Figure 3. FES of the Schlenk equilibrium. The CVs for this representation are the difference in Mg–CH3 coordination number between Mg2 and Mg1 (CV1) and the THF coordination number to Mg1 (CV2). The chemical structures drawn in the figure depict the most representative species obtained for wells AE.

We introduce a notation for describing the nature of the dimer D. The notation DijXY stands for dimer (D) where i and j are the numbers of THF molecules on Mg1 and Mg2, respectively, and X and Y describe the nature of the bridging ligands. DXY describes the entire family of X and Y bridging species regardless of the number of THF molecules on either Mg. The FES plot in Figure 3 shows that the ensemble of DClCl structures is more stable by ∼3.5 kcal mol–1 than the one for DClMe. In addition, the FES shows a clear path connecting the AE wells, thus giving information on the ligand exchange reaction. The transformation from wells A and B to D goes preferentially via well C, while exchange directly connecting B to D does not constitute a similarly viable path. The energy barriers for the AC and CD transition are ∼8 and ∼6 kcal mol–1, respectively, while the free energy barrier between B and D is 13 kcal mol–1 higher than that of well A. The preference for this path was analyzed by studying more closely the solvation structures of DClCl and DClMe.

Solvation of Dichloride Bridged Structures

The accessible solvation states of DClCl in the A and B wells (Figure 3) were investigated by metadynamics simulations using the coordination number of the oxygen atoms of THF to each of the Mg atoms as CVs. The simulations revealed the existence of three solvation states, corresponding to distinct free-energy minima reported in Figure 4a–c. The solvation state with the lowest free energy corresponds to a symmetric dimeric structure in which each Mg atom is solvated by a single THF (D11ClCl). The second most stable solvation state, only 1.3 kcal mol–1 higher in energy, corresponds to another symmetric adduct with two THF molecules coordinating each Mg (D22ClCl). In the later, each Mg is five-coordinated. The interconversion between these two structures occurs through a metastable asymmetric solvation state D12ClCl with one THF solvent on one Mg and two on the other. This asymmetric solvation state D12ClCl, which is 2.7 kcal mol–1 higher in energy than D11ClCl, is expected to be statistically present in 1% of room-temperature samples, also taking into account degenerate states. Higher solvation states with three THF molecules in the first solvation shell of a single Mg center are more than 5 kcal mol–1 higher in energy than D11ClCl and therefore are expected to be present only in minor amounts in room-temperature samples. Both the relative stability and activation energies for the interchange between D11ClCl, D12ClCl, and D22ClCl are low (less than 5 kcal mol–1), implying fast interconversion between the different structures at room temperature.

Figure 4

Figure 4. Solvated DClCl structures found by metadynamics simulations (left) and the corresponding FES (right). CV1 and CV2 are defined as the coordination numbers of THF at Mg1 and Mg2, respectively, following eq 1. The two minima b correspond to the chemically equivalent D12ClCl and D21ClCl structures. Only D12ClCl is shown for simplicity.

Analysis of the structural changes from D11ClCl to D12ClCl showed that the addition of THF to Mg2 (Figure 4) occurs always anti to one of the chlorides, producing local trigonal-bipyramidal coordination with the bridging Cl’s in axial and equatorial positions. AIMD simulations indicated significant lengthening of the Mg–Clax bond relative to the Mg–Cleq one. Reversible cleavage of the axial Cl–Mg bond and the consequential formation of transient single chloride bridged structures were observed during the analysis of AIMD trajectories. In order to analyze the influence of the solvent on Mg–Cl cleavage, the statistical distribution of the Mg–Cl bond lengths in D11ClCl, D12ClCl, and D22ClCl was computed (Figure 5). As expected, an increase in solvation is associated with statistical elongation of the Mg–Cl bonds.

Figure 5

Figure 5. (Top) Mg–Cl distance distributions in DClCl dimeric structures. (Bottom) Mg1–Cl (blue) and Mg2–Cl (red) distance distributions in D12ClCl. Mg–Cl bond cleavage is observed when the Mg–Cl distance is larger than 3.7 Å.

Strikingly, Mg–Cl bond cleavage was observed only in the asymmetric D12ClCl structure with statistically significant probability of finding Mg–Cl distances with longer values than 3.7 Å, as shown in the inset of Figure 5 top. On the contrary, both the Mg–Cl bond length distributions of the least and the most solvated D11ClCl and D22ClCl species vanish above 3.7 Å, indicating statistical irrelevance of Mg–Cl bond-cleaved structures. Analysis of the AIMD trajectory evidenced that in all cases the cleavage of the Mg–Cl bond occurs between the pentacoordinated Mg2 and the Clax bridged atom in axial position in the trigonal bipypramid, as shown in Figure 5 bottom. The presence in thermal samples of the monochloride bridged, unsymmetrically solvated structures appears to be an essential ingredient in ligand exchange at each Mg.
Activation of the Mg–Cl bond in the axial position is also consistent with NBO analysis of DClCl, which shows that the pentacoordinated Mg atoms in both D12ClCl and D22ClCl receive the smallest electron donation from Cl (χMg ≈ 6%). In particular, the bridging chloride, axially coordinated to Mg2, Clax, has the lowest contribution to this magnesium (χMg ≈ 2%). Such a small donation implies an overpolarization of the Clax–Mg2 bond, which in turn facilitates opening of the dichloride bridged moiety of D12ClCl, as observed during AIMD. The asymmetry in the donation between the axial and the equatorial Cl explains also the selective opening of the ring at the Clax–Mg2 bond.
The trigonal-bipyramidal geometry on Mg2 in D12ClCl was not found in D22ClCl. Instead, the coordination geometry of Mg atoms took a distorted square-pyramidal geometry, with an apical THF, and equatorial Cl atoms, a methyl group, and the remaining THF. In all D11ClCl, D12ClCl, and D22ClCl, the equilibrium values for the Cl–Mg–Cl bond angle are close to 90° (Table 1), even though the coordination geometries of the Mg atoms in the structures are different from each other. As an additional structural feature, it is observed that during AIMD simulations the four atoms of the Mg(μ-Cl)2Mg moiety do not lie in the same plane. The average angles between the two MgCl2 planes are reported in Table 1.
Table 1. Average Angles (αdyn, β1dyn, and β2dyn, in degrees) with Associated Standard Deviation (in parentheses) Obtained from Cluster Analysis of the Metadynamics Trajectory and Angles from DFT Optimization with Implicit Solvent (αst, β1st, and β2st, in degrees)
 αdynαstβ1dynβ1stβ2dynβ2st
D11ClCl167(8)169.789(5)91.389(5)91.6
D12ClCl166(9)165.491(7)92.182(6)83.9
D22ClCl165(11)160.694(6)84.194(6)84.2
DFT optimization of D11ClCl, D12ClCl, and D22ClCl confirmed that the minima on the FES correspond to minima on the PES. However, structural discrepancies are obtained for the species with the largest number of THF molecules in the coordination sphere of the Mg centers. Thus, the Mg centers in D22ClCl are trigonal-bipyramidal on the PES, while they are distorted square-pyramidal on the FES. This indicates that the solvent cage has an increasing influence on the structure and dynamics of the more flexible, solvated species. AIMD simulations also revealed the existence of interchange of THF molecules in the Mg solvation shell of D22ClCl taking place through an associative mechanism. In this case, a short-lived octahedral structure was formed. The statistical abundance of these structures is however marginal.

Solvation of Methyl Chloride Bridged Structures

The accessible solvation states of the DClMe species in C and D wells (Figure 3) were analyzed by metadynamics simulations using the coordination number of the oxygen atoms of THF to each of the Mg atoms as CVs (Figure 6). Three low-energy solvation structures were obtained. The lowest in energy, D11ClMe, has two tetrahedral-coordinated Mg centers bridged by a Cl atom and a methyl group (Figure 6a). This structure is favored over the other ones (Figure 6b–d) by about 3 kcal mol–1. The two other major structures are dinuclear species with three coordinated solvent molecules, D21ClMe and D11ClMeTHF. In the former species, the additional THF is bound to the Mg–Cl moiety, while in the latter, THF bridges the two Mg centers (Figure 6). The free energies of triply bridged species D11ClMeTHF and the asymmetrically solvated D21ClMe differ by only 0.5 kcal mol–1, separated by a barrier on the order of 2kT. A well corresponding to the asymmetric structure with two THF molecules coordinated to the Mg–CH3 moiety (D12ClMe) was also obtained about 5 kcal mol–1 above D11.ClMe Overall, fast interchange between D21ClMe, D11ClMeTHF, and D12ClMe is observed with low activation energy barriers of around 4 kcal mol–1.

Figure 6

Figure 6. FES for the methyl bridged dimer DClMe equilibria using the THF coordination number to Mg1 (CV1) and the THF coordination number to Mg2 (CV2) as variables, together with the most representative species obtained for wells ad.

All DClMe structures show tetrahedral and trigonal-bipyramidal geometries for the Mg centers solvated by one and two THF molecules, respectively. The bridging methyl (μ-CH3) interacts differently with the two Mg atoms in the ClMg(μ-CH3)(μ-Cl)MgCH3 dimer. This is due to the asymmetric distribution of the ligands at the two Mg atoms because Mg1 has a terminal chloride while Mg2 has a terminal methyl group. Moreover, the two Mg atoms can have different solvation states at room temperature, as shown by the FES in Figure 6. In order to quantify the way the bridging CH3 group interacts with the two Mg centers in the different structures, the direction of the bridging pz orbital axis relative to the two C–Mg directions was monitored (Figure 7).

Figure 7

Figure 7. Orientation of the methyl group in DClMe as a function of the solvation state, represented by φ1 and φ2. A larger φ angle is indicative of a stronger Mg–CH3 interaction.

In both D11ClMe and D12ClMe, the pz orbital is oriented toward the Mg1–Cl moiety, with average angles φ1 of approximately 141 and 149°, respectively. In these two structures, φ2 has average values of 127 and 128°, indicating a poorer interaction of the bridging methyl with the Mg2 pz orbital. In D11ClMeTHF, the μ-CH3 bridging group is more equally shared between the two Mg centers, as indicated by the values of φ1 and φ2 that oscillate around 145 and 136°, respectively. A larger solvation for the Mg bound to the terminal Cl than the other Mg atom (D21ClMe) results in practically equal sharing of the bridging methyl between the two Mg centers. In this case, both angles φ1 and φ2 oscillate around an average value of ∼137°.
The orientation of the bridging methyl group is thus highly sensitive to the chemical groups and the solvation at each Mg center. For equal solvation, the methyl interacts more with the more electron deficient Mg1 center, that is, the one with the terminal chloride. Increasing solvation of Mg1 by either bridging or terminal THF results in an increased interaction of the bridging methyl group with Mg2 bearing the terminal methyl group. In this way, the solvent helps the bridging methyl group weaken its interaction with Mg1 and increase that to Mg2, assisting the transformation of CH3MgCl into MgCl2 and Mg(CH3)2.
Geometry optimizations on the PES with DFT methods of the minima D11ClMe, D12ClMe, D11ClMeTHF, and D21ClMe on the FES gave minima with similar structures, as indicated in Table 2. In all cases, trigonal-bipyramid geometries were found on the pentacoordinated Mg atoms with the bridging CH3 and a terminal THF in the axial positions. NBO analysis on the optimized structures showed slightly higher electron donation of the bridged CH3 lone pair to the Mg of the Mg–Cl moiety (χMg1 = 5.7, 5.2, and 4.6%) compared to that for the Mg–CH3 moiety (χMg2 = 3.7, 3.8, and 3.6%) for D11ClMe, D12ClMe, and D11ClMeTHF, respectively. With increasing solvation of Mg1, as in D21ClMe, donation of the bridging methyl group to the two Mg centers becomes equivalent (χMg1 = 4.3%; χMg2 = 4.2%), consistent with the distribution of the φ1 and φ2 angles during AIMD simulations (Figure 8). Although small, the increased electron donation of the Me group to Mg2, which already bears the other methyl group, is in agreement with D21ClMe being the most prone of the methyl chloride bridged dimers to yield the final products MgCl2 and Mg(CH3)2.
Table 2. Average angles (αdyn, β1dyn, and β2dyn, in degrees) with the Standard Deviations (in parentheses) Obtained from Cluster Analysis of the Metadynamics Trajectory and Optimized Angles from DFT Calculations with Implicit Solvent (αst, β1st, and β2st, in degrees)
 αdynαstβ1dynβ1stβ2dynβ2st
D11ClMe167(11)175.6101(10)103.794(7)100.0
D12ClMe166(8)179.4102(7)110.689(7)93.2
D11ClMeTHF165(6)154.4103(7)102.096(7)99.6
D21ClMe161(7)172.394(7)100.4101(7)102.5

Figure 8

Figure 8. Intermediates involved in the Schlenk equilibrium according to dynamic simulations. Arrows indicate the chemical transformations along the main pathway leading from monomeric reactants to products (inside of solid squares). The most stable dichloride and methyl chloride bridged dinuclear species are inside of dashed squares.

Reaction Pathway of the Schlenk Equilibrium

The computational results presented above allow reconstruction of the full reaction pathway of the Schlenk equilibrium in a THF solution. Figure 8 presents the set of structures, localized as minima on the FES, that are involved in the transformation of the reactants CH3MgCl(THF)2 to the products Mg(CH3)2(THF)2 and MgCl2(THF)n (n = 2–4). The scheme also highlights the key role played by the solvent in assisting the Cl/CH3 ligand exchange.
The Schlenk equilibrium starts by dimerization of (CH3)MgCl(THF)2 via the two chlorides to form the dichloride bridged species D22ClCl. Static calculations show that this reaction is slightly endoergic by 4.9 kcal mol–1, consistent with experimental results. (22, 76)
The coordination geometry of the pentacoordinated Mg atoms in D22ClCl was found to be distorted square-pyramidal when exploring the FES but trigonal-bipyramidal when performing ab initio geometry optimizations on the PES. Such discrepancy evidences a strong influence of the surrounding solvent molecules on the structural geometry of D22ClCl. THF exchange on the Mg atom may occur through addition of one THF to form an octahedral species. (77-79) These last transient structures are very short lived and thus have not been included in Figure 8.
Desolvation of D22ClCl by one THF molecule yields an asymmetrical dinuclear complex (D12ClCl) with tetrahedral and pyramidal-trigonal geometries for Mg1 and Mg2, respectively. This trisolvated species can lose one THF molecule to produce the disolvated (D11ClCl), which was found to be the most stable species in solution. Tetrahedral coordination of Mg is also the one preferred in the solid state, as shown by reported crystallographic structures of related Mg2X2R2 species. (28, 80) The difference in energy between D11ClCl and the least stable intermediate D12ClCl is 2.7 kcal/mol–1 with an activation energy barrier of less than 5 kcal mol–1 (Figure 4). Therefore, all D11ClCl, D12ClCl, and D22ClCl species are expected to coexist and undergo interconversion at room temperature.
Interestingly, not all species favor the Mg–Cl bond cleavage required for the Cl/Me ligand exchange. Analysis of the Mg–Cl bond distances in D11ClCl, D22ClCl, and D12ClCl (Figure 5 top) shows that the formation of monochloride bridged species takes place preferentially from D12ClCl. NBO analysis indicates that asymmetric solvation of D12ClCl favors Mg2–Clax bond cleavage (step (ii) in Figure 8). In this complex, the higher solvation on Mg2 weakens the Mg2–Cl bond, while the lower solvation on Mg1 makes it prone to accept an extra anionic ligand during the Cl/CH3 exchange process. Analysis on several trajectories, such as the one represented in Figure 9, reaffirms that the chloride involved in the cleavage is the one located in the axial position of Mg2 (Figure 5 bottom). The monobridged species generated by Mg–Cl bond cleavage (snapshot 2 in Figure 9) is short-lived and rapidly evolves into methyl chloride bridged structures (DClMe). The methyl group occupying the bridging position comes from Mg1, as shown in snapshot 3; this is consistent with the higher electron density on the MgCl(THF)(μ-Cl) fragment compared to that on Mg(THF)2(μ-Cl) calculated for the structure shown in snapshot 2.

Figure 9

Figure 9. Snapshots for the methyl transfer reaction in D12ClCl (Mg1 on the left-hand side and Mg2 on the right for all snapshots): (1) initial D12ClCl structure, (2) transition state of the transmetalation reaction, (3) D12ClMe, (4) solvent loss to form D11ClMe, and (5) solvent addition to form D21ClMe and (6) D11ClMeTHF. The atoms for the Grignard reagent and the coordinating THF molecules are depicted as balls and/or sticks and colored according to standard color codes. Selected neighboring solvent molecules are drawn with thin lines.

The first DClMe species generated after Mg–Cl bond cleavage is the D12ClMe complex (snapshot 3 in Figure 9). This species is however the least stable of the four solvation states observed by AIMD, with a free energy difference of +5 kcal mol–1 (Figure 4) over the most stable solvation structure. As in the case of DClCl species, the most stable DClMe complex is the disolvated D11ClMe, in which both Mg atoms have tetrahedral coordination geometry. However, DClMe incorporates other trisolvated species, D11ClMeTHF and D21ClMe, which are both 3 kcal mol–1 above D11ClMe. Exchange between these species has low activation energy barriers (smaller than 4 kcal mol–1) and involves the addition of an external THF either at the bridging position (D11ClMeD11ClMeTHF) or at the terminal position of Mg1(D11ClMeD21ClMe). Internal rearrangement of THF from bridging to terminal is also observed (D11ClMeTHFD21ClMe) with an even lower energy barrier (<0.5 kcal mol–1). The higher stability of D21ClMe compared to that of D12ClMe is consistent with the higher solvation of the most electrophilic Mg atom (Mg1), that is, the one with a terminal chloride.
In order to form the final products, both Mg2–Cl and Mg1–CH3 bonds need to be cleaved. This however does not take place simultaneously but consecutively, and solvation plays an important role. Because Cl is a much better ligand than CH3, (28) the Mg–CH3 bond is the first to be broken. Analysis of the orientation of the formal lone pair in the bridging methyl ligand shows that at equal solvation of Mg atoms in DClMe, μ-CH3 is more strongly bonded to Mg1 than to Mg2 because Mg1 is more electrophilic (Figure 7). This preference is however modified by increasing the solvation of Mg1 relative to Mg2. Thus, in D21ClMe, the methyl is equally bonded to Mg1 and Mg2. In addition, in this species, the axial position occupied by the bridged methyl group in the coordination sphere of Mg1 also favors the Mg1–CaxH3 bond cleavage. An increase in solvation of Mg1 thus yields the preproduct of the Schlenk equilibrium (P21 and minimum E in Figure 3), consisting of a Mg1Cl2 species bridged by way of a single chloride to Mg2(CH3)2. An increase in solvation of both Mg centers is expected to favor the final release of the mononuclear MgCl2 and Mg(CH3)2 products. The dissociation energy from P21 to CH3MgCl(THF)2 and the most stable MgCl2(THF)3 has been estimated to be −8.7 kcal mol–1 in this study (see Figure S4 for further details).
Despite their relatively low abundance in solution, as also experimentally observed, (22) formation of dimeric adducts is key to the evolution of the Schlenk equilibrium reaction. The reaction pathway identified in this study highlights the crucial role of the solvent in assisting Cl/Me exchange in the Schlenk equilibrium. Most importantly, it shows that the tetracoordinated Mg species proposed in most computational studies (28, 81, 82) are indeed the most stable structures but are not on the reactive pathway. Instead, asymmetric solvation on Mg atoms is needed to promote the Mg–Cl and Mg–CH3 bond cleavage to go from MgCH3Cl to Mg(CH3)2 and MgCl2. These tetra/pentacoordinated Mg dimers (D12ClCl and D21ClMe) are transient intermediates that interchange with the most stable solvation structures at room temperature. In addition, the bridging ligand involved in the bond-breaking process is located always in the axial position of the pentacoordinated Mg atom. This detailed information on the Schlenk equilibrium mechanism will be useful for a better understanding of the reactivity of the Grignard reagents. In addition, this study may help to better understand other transmetalation processes, such as those involved in the Kumada and Negishi cross-coupling reactions, (4, 83-86) which are also assisted by solvent.

Conclusions

Click to copy section linkSection link copied!

The here presented AIMD study of the Schlenk equilibrium reaction, in which two molecules of CH3MgCl exchange methyl and chloride groups to yield Mg(CH3)2 and MgCl2, showed how the ether solvent (THF here) has a crucial role in assisting the reaction. Although coordination of the solvent is needed for stabilizing the various mono- and dinuclear Mg complexes, the stabilizing effect is not the only factor at work.
The reaction goes via the formation of a dinuclear (CH3)Mg(μ-Cl)2Mg(CH3) intermediate whose most stable solvated form (with a single THF at each Mg) is unreactive. The chloride/methyl exchange is promoted by making the two Mg atoms electronically different. The calculations show that these differences are created by different solvations of the two Mg centers. The cleavage of the Mg–Cl bond and associated shift of the methyl group from the terminal to bridging position is assisted by increasing the solvation at the Mg involved in the bond cleavage while keeping the other Mg less solvated.
A similar process occurs at the dinuclear (CH3)Mg(μ-CH3)(μ-Cl)MgCl; the cleavage of the bond between Mg and the bridging CH3 group to form solvated Mg(CH3)2 and MgCl2 requires the chloride-rich Mg atom to be more solvated than the methyl-rich Mg one. Increasing solvation at one Mg favors bond cleavage, while decreasing solvation on the other one favors the formation of new terminal bonds.
The species selected by the dynamics to be on the reaction pathway are not necessarily minima on the PES, which emphasizes the need for the former method. Our findings highlight the need of including explicit solvent dynamics in the modeling of such reactions because this is crucial in allowing the Cl/CH3 group exchange to occur with a low energy barrier.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpcb.7b02716.

  • Metadynamics parameters, natural bond orbital analysis, DFT optimized geometries, and solvation properties for different sizes of the simulation box (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

Click to copy section linkSection link copied!

  • Corresponding Authors
    • Ainara Nova - Department of Chemistry and Centre for Theoretical and Computational Chemistry (CTCC), University of Oslo, Postbox 1033 Blindern, 0315 Oslo, Norway Email: [email protected]
    • Michele Cascella - Department of Chemistry and Centre for Theoretical and Computational Chemistry (CTCC), University of Oslo, Postbox 1033 Blindern, 0315 Oslo, Norway Email: [email protected]
  • Authors
    • Raphael M. Peltzer - Department of Chemistry and Centre for Theoretical and Computational Chemistry (CTCC), University of Oslo, Postbox 1033 Blindern, 0315 Oslo, Norway
    • Odile Eisenstein - Department of Chemistry and Centre for Theoretical and Computational Chemistry (CTCC), University of Oslo, Postbox 1033 Blindern, 0315 Oslo, NorwayInstitut Charles Gerhardt, UMR 5253 CNRS-Université de Montpellier, Université de Montpellier, cc 1501, Place E. Bataillon, 34095 Montpellier, FranceOrcidhttp://orcid.org/0000-0001-5056-0311
  • Notes
    The authors declare no competing financial interest.

Acknowledgment

Click to copy section linkSection link copied!

This work was supported by the Research Council of Norway (RCN) through the CoE Centre for Theoretical and Computational Chemistry (CTCC) Grant No. 179568/V30 and 171185/V30 and by the Norwegian Supercomputing Program (NOTUR) (Grant No. NN4654K). A.N. thanks the RCN for Grants 221801/F20 and 250044/F20. The authors thank Elisa Rebolini for enlightening discussion.

References

Click to copy section linkSection link copied!

This article references 86 other publications.

  1. 1
    Grignard, V. Sur Quelques Nouvelles Combinaisons Organométalliques du Magnésium et Leur Application à des Synthèses d’Alcools et d’Hydrocabures C. R. Acad. Sci. 1900, 130, 1322 1324
  2. 2
    Grignard, V. The Use of Organomagnesium Compounds in Preparative Organic Chemistry–Nobel Lecture 1912 Nobel Lectures Chemistry 1921, 1966, 234 246
  3. 3
    Corriu, R. J. P.; Massé, J. P. Activation of Grignard Reagents by Transition-Metal Complexes. A New and Simple Synthesis of Trans-Stilbenes and Polyphenyls J. Chem. Soc., Chem. Commun. 1972, 144a 144a DOI: 10.1039/c3972000144a
  4. 4
    Tamao, K.; Sumitani, K.; Kumada, M. Selective Carbon-Carbon Bond Formation by Cross-Coupling of Grignard Reagents with Organic Halides. Catalysis by Nickel-Phosphine Complexes J. Am. Chem. Soc. 1972, 94, 4374 4376 DOI: 10.1021/ja00767a075
  5. 5
    Fürstner, A.; Leitner, A.; Méndez, M.; Krause, H. Iron-Catalyzed Cross-Coupling Reactions J. Am. Chem. Soc. 2002, 124, 13856 13863 DOI: 10.1021/ja027190t
  6. 6
    Frisch, A. C.; Beller, M. Catalysts for Cross-Coupling Reactions with Non-activated Alkyl Halides Angew. Chem., Int. Ed. 2005, 44, 674 688 DOI: 10.1002/anie.200461432
  7. 7
    Terao, J.; Kato, Y.; Kambe, N. Titanocene-Catalyzed Regioselective Alkylation of Styrenes with Grignard Reagents Using β-Bromoethyl Ethers, Thioethers, or Amines Chem. - Asian J. 2008, 3, 1472 1478 DOI: 10.1002/asia.200800134
  8. 8
    Vechorkin, O.; Barmaz, D.; Proust, V.; Hu, X. Ni-Catalyzed Sonogashira Coupling of Nonactivated Alkyl Halides: Orthogonal Functionalization of Alkyl Iodides, Bromides, and Chlorides J. Am. Chem. Soc. 2009, 131, 12078 12079 DOI: 10.1021/ja906040t
  9. 9
    Adrio, J.; Carretero, J. C. Functionalized Grignard Reagents in Kumada Cross-Coupling Reactions ChemCatChem 2010, 2, 1384 1386 DOI: 10.1002/cctc.201000237
  10. 10
    Jana, R.; Pathak, T. P.; Sigman, M. S. Advances in Transition Metal (Pd,Ni,Fe)-Catalyzed Cross-Coupling Reactions Using Alkyl-organometallics as Reaction Partners Chem. Rev. 2011, 111, 1417 1492 DOI: 10.1021/cr100327p
  11. 11
    Cong, X.; Tang, H.; Zeng, X. Regio- and Chemoselective Kumada–Tamao–Corriu Reaction of Aryl Alkyl Ethers Catalyzed by Chromium Under Mild Conditions J. Am. Chem. Soc. 2015, 137, 14367 14372 DOI: 10.1021/jacs.5b08621
  12. 12
    Neufeld, R.; Teuteberg, T. L.; Herbst-Irmer, R.; Mata, R. A.; Stalke, D. Solution Structures of Hauser Base iPr2NMgCl and Turbo-Hauser Base iPr2NMgCl·LiCl in THF and the Influence of LiCl on the Schlenk-Equilibrium J. Am. Chem. Soc. 2016, 138, 4796 4806 DOI: 10.1021/jacs.6b00345
  13. 13
    Seyferth, D. The grignard reagents Organometallics 2009, 28, 1598 1605 DOI: 10.1021/om900088z
  14. 14
    Guggenberger, L. J.; Rundle, R. E. Crystal Structure of the Ethyl Grignard Reagent, Ethylmagnesium Bromide Dietherate J. Am. Chem. Soc. 1968, 90, 5375 5378 DOI: 10.1021/ja01022a007
  15. 15
    Vallino, M. Structure Cristalline de CH3MgBr·3 C4H8O J. Organomet. Chem. 1969, 20, 1 10 DOI: 10.1016/S0022-328X(00)80080-7
  16. 16
    Toney, J.; Stucky, G. D. The Stereochemistry of Polynuclear Compounds of the Main Group Elements [C2H5Mg2Cl3(C4H8O)3]2, a Tetrameric Grignard Reagent J. Organomet. Chem. 1971, 28, 5 20 DOI: 10.1016/S0022-328X(00)81569-7
  17. 17
    Blasberg, F.; Bolte, M.; Wagner, M.; Lerner, H.-W. An Approach to Pin Down the Solid-State Structure of the “Turbo Grignard Organometallics 2012, 31, 1001 1005 DOI: 10.1021/om201080t
  18. 18
    Smith, M. B.; Becker, W. E. The constitution of the grignard reagent—III:The reaction between R2Mg and MgX2 in tetrahydrofuran Tetrahedron 1967, 23, 4215 4227 DOI: 10.1016/S0040-4020(01)88819-0
  19. 19
    Smith, M. B.; Becker, W. E. The constitution of the Grignard Reagent - I. The reaction between diethyl magnesium and magnesium bromide in diethyl ether Tetrahedron Lett. 1965, 6, 3843 3847 DOI: 10.1016/S0040-4039(01)89135-8
  20. 20
    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, 3162 3171 DOI: 10.1021/ja00844a040
  21. 21
    Schlenk, W.; Schlenk, W. Über die Konstitution der Grignardschen Magnesiumverbindungen Ber. Dtsch. Chem. Ges. B 1929, 62, 920 924 DOI: 10.1002/cber.19290620422
  22. 22
    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, 3845 3850 DOI: 10.1021/ja01042a027
  23. 23
    Sobota, P.; Duda, B. Influence of MgCl2 on Grignard Reagent Composition in Tetrahydrofuran. III J. Organomet. Chem. 1987, 332, 239 245 DOI: 10.1016/0022-328X(87)85090-8
  24. 24
    Sakamoto, S.; Imamoto, T.; Yamaguchi, K. Constitution of Grignard Reagent RMgCl in Tetrahydrofuran Org. Lett. 2001, 3, 1793 1795 DOI: 10.1021/ol010048x
  25. 25
    Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas Phys. Rev. B 1964, 136, 864 871 DOI: 10.1103/PhysRev.136.B864
  26. 26
    Kohn, W.; Sham, L. J. Self-Consistent Equations Including Exchange and Correlation Effects Phys. Rev. A 1965, 140, 1133 1138 DOI: 10.1103/PhysRev.140.A1133
  27. 27
    Jiménez-Halla, J. O. C.; Bickelhaupt, F. M.; Solà, M. Organomagnesium clusters: Structure, stability, and bonding in archetypal models J. Organomet. Chem. 2011, 696, 4104 4111 DOI: 10.1016/j.jorganchem.2011.06.014
  28. 28
    Lioe, H.; White, J. M.; O’Hair, R. A. J. Preference for bridging versus terminal ligands in magnesium dimers J. Mol. Model. 2011, 17, 1325 1334 DOI: 10.1007/s00894-010-0834-1
  29. 29
    Henriques, A. M.; Barbosa, A. G. H. Chemical Bonding and the Equilibrium Composition of Grignard Reagents in Ethereal Solutions J. Phys. Chem. A 2011, 115, 12259 12270 DOI: 10.1021/jp202762p
  30. 30
    Ramirez, F.; Sarma, R.; Chaw, F.; McCaffrey, T. M. Magnesium bromide-tetrahydrofuran complexes: bis(tetrahydrofuran)magnesium bromide, tris(tetrahydrofuran)magnesium bromide, tetrakis(tetrahydrofuran)magnesium bromide, and diaquotetrakis(tetrahydrofuran)magnesium bromide. A reagent for the preparation of anhydrous magnesium phosphodiester salts J. Am. Chem. Soc. 1977, 99, 5285 5289 DOI: 10.1021/ja00458a010
  31. 31
    Pirinen, S.; Koshevoy, I. O.; Denifl, P.; Pakkanen, T. T. A Single-Crystal Model for MgCl2 – Electron Donor Support Materials: [Mg3Cl5(THF)4Bu]2 (Bu = n-Butyl) Organometallics 2013, 32, 4208 4213 DOI: 10.1021/om400407p
  32. 32
    Ashby, E. C.; Becker, W. E. Concerning the Structure of the Grignard Reagent J. Am. Chem. Soc. 1963, 85, 118 119 DOI: 10.1021/ja00884a032
  33. 33
    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, 133 139 DOI: 10.1021/jp035653r
  34. 34
    Tobisu, M.; Chatani, N. Cross-Couplings Using Aryl Ethers via C–O Bond Activation Enabled by Nickel Catalysts Acc. Chem. Res. 2015, 48, 1717 1726 DOI: 10.1021/acs.accounts.5b00051
  35. 35
    Cahiez, G.; Moyeux, A.; Cossy, J. Grignard Reagents and Non-Precious Metals: Cheap and Eco-Friendly Reagents for Developing Industrial Cross-Couplings. A Personal Account Adv. Synth. Catal. 2015, 357, 1983 1989 DOI: 10.1002/adsc.201400654
  36. 36
    Tasker, S. Z.; Standley, E. A.; Jamison, T. F. Recent Advances in Homogeneous Nickel Catalysis Nature 2014, 509, 299 309 DOI: 10.1038/nature13274
  37. 37
    Laio, A.; Parrinello, M. Escaping Free-Energy Minima Proc. Natl. Acad. Sci. U. S. A. 2002, 99, 12562 12566 DOI: 10.1073/pnas.202427399
  38. 38
    Iannuzzi, M.; Laio, A.; Parrinello, M. Efficient Exploration of Reactive Potential Energy Surfaces Using Car-Parrinello Molecular Dynamics Phys. Rev. Lett. 2003, 90, 23 26 DOI: 10.1103/PhysRevLett.90.238302
  39. 39
    Vuilleumier, R.; Sprik, M. Electronic Properties of Hard and Soft Ions in Solution: Aqueous Na+ and Ag+ Compared J. Chem. Phys. 2001, 115, 3454 3468 DOI: 10.1063/1.1388901
  40. 40
    Lightstone, F. C.; Schwegler, E.; Hood, R. Q.; Gygi, F.; Galli, G. A First Principles Molecular Dynamics Simulation of the Hydrated Magnesium Ion Chem. Phys. Lett. 2001, 343, 549 555 DOI: 10.1016/S0009-2614(01)00735-7
  41. 41
    Bernasconi, L.; Baerends, E. J.; Sprik, M. Long-Range Solvent Effects on the Orbital Interaction Mechanism of Water Acidity Enhancement in Metal Ion Solutions: A Comparative Study of the Electronic Structure of Aqueous Mg and Zn Dications J. Phys. Chem. B 2006, 110, 11444 11453 DOI: 10.1021/jp0609941
  42. 42
    Blumberger, J.; Bernasconi, L.; Tavernelli, I.; Vuilleumier, R.; Sprik, M. Electronic Structure and Solvation of Copper and Silver Ions: A Theoretical Picture of a Model Aqueous Redox Reaction J. Am. Chem. Soc. 2004, 126, 3928 3938 DOI: 10.1021/ja0390754
  43. 43
    Guido, C. A.; Pietrucci, F.; Gallet, G. A.; Andreoni, W. The Fate of a Zwitterion in Water from Ab Initio Molecular Dynamics: Monoethanolamine (MEA)-CO2 J. Chem. Theory Comput. 2013, 9, 28 32 DOI: 10.1021/ct301071b
  44. 44
    Boero, M.; Ikeshoji, T.; Liew, C. C.; Terakura, K.; Parrinello, M.; Boero, M.; Ikeshoji, T.; Liew, C. C.; Terakura, K. Hydrogen Bond Driven Chemical Reactions: Beckmann Rearrangement of Cyclohexanone Oxime into ε-Caprolactam in Supercritical Water Hydrogen Bond Driven Chemical Reactions: Beckmann Rearrangement of Cyclohexanone Oxime into E-Caprolactam in Supercritical J. Am. Chem. Soc. 2004, 126, 6280 6286 DOI: 10.1021/ja049363f
  45. 45
    Vidossich, P.; Lledós, A.; Ujaque, G. Realistic Simulation of Organometallic Reactivities in Solution by Means of First-Principles Molecular Dynamics. In Computational Studies in Organometallic Chemistry; Macgregor, S. A.; Eisenstein, O., Eds.; Structure and Bonding; Springer International Publishing: Berlin, Germany, 2016; Vol. 167, pp 81 106.
  46. 46
    Vidossich, P.; Lledós, A.; Ujaque, G. First-Principles Molecular Dynamics Studies of Organometallic Complexes and Homogeneous Catalytic Processes Acc. Chem. Res. 2016, 49, 1271 1278 DOI: 10.1021/acs.accounts.6b00054
  47. 47
    Laio, A.; VandeVondele, J.; Rothlisberger, U. A Hamiltonian Electrostatic Coupling Scheme for Hybrid Car–Parrinello Molecular Dynamics Simulations J. Chem. Phys. 2002, 116, 6941 6947 DOI: 10.1063/1.1462041
  48. 48
    Moret, M.-E.; Tavernelli, I.; Chergui, M.; Rothlisberger, U. Electron Localization Dynamics in the Triplet Excited State of [Ru(bpy)3]2+ in Aqueous Solution Chem. - Eur. J. 2010, 16, 5889 5894 DOI: 10.1002/chem.201000184
  49. 49
    Dal Peraro, M.; Llarrull, L. I.; Rothlisberger, U.; Vila, A. J.; Carloni, P. Water-Assisted Reaction Mechanism of Monozinc β-Lactamases J. Am. Chem. Soc. 2004, 126, 12661 12668 DOI: 10.1021/ja048071b
  50. 50
    Dal Peraro, M.; Vila, A. J.; Carloni, P.; Klein, M. L. Role of Zinc Content on the Catalytic Efficiency of B1Metallo β-Lactamases J. Am. Chem. Soc. 2007, 129, 2808 2816 DOI: 10.1021/ja0657556
  51. 51
    Cascella, M.; Magistrato, A.; Tavernelli, I.; Carloni, P.; Rothlisberger, U. Role of Protein Frame and Solvent for the Redox Properties of Azurin from Pseudomonas Aeruginosa Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 19641 19646 DOI: 10.1073/pnas.0607890103
  52. 52
    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, 10955 10962 DOI: 10.1021/ja8005786
  53. 53
    Gossens, C.; Tavernelli, I.; Rothlisberger, U. Rational Design of Organo-Ruthenium Anticancer Compounds Chimia 2005, 59, 81 84 DOI: 10.2533/000942905777676795
  54. 54
    Metz, D. J.; Glines, A. Density, Viscosity, and Dielectric Constant of Tetrahydrofuran between −78 and 30° J. Phys. Chem. 1967, 71, 1158 1158 DOI: 10.1021/j100863a067
  55. 55
    Hoover, W. G. Canonical Dynamics: Equilibrium Phase-Space Distributions Phys. Rev. A: At., Mol., Opt. Phys. 1985, 31, 1695 1697 DOI: 10.1103/PhysRevA.31.1695
  56. 56
    Martyna, G. J.; Klein, M. L.; Tuckerman, M. Nose–Hoover chains: The Canonical Ensemble via Continuous Dynamics J. Chem. Phys. 1992, 97, 2635 2643 DOI: 10.1063/1.463940
  57. 57
    Nosé, S. A Unified Formulation of the Constant Temperature Molecular Dynamics Methods J. Chem. Phys. 1984, 81, 511 519 DOI: 10.1063/1.447334
  58. 58
    Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple Phys. Rev. Lett. 1996, 77, 3865 3868 DOI: 10.1103/PhysRevLett.77.3865
  59. 59
    Lippert, G.; Hutter, J.; Parrinello, M. The Gaussian and Augmented-Plane-Wave Density Functional Method for Ab Initio Molecular Dynamics Simulations Theor. Chem. Acc. 1999, 103, 124 140 DOI: 10.1007/s002140050523
  60. 60
    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
  61. 61
    Goedecker, S.; Teter, M.; Hutter, J. Separable Dual-Space Gaussian Pseudopotentials Phys. Rev. B: Condens. Matter Mater. Phys. 1996, 54, 1703 1710 DOI: 10.1103/PhysRevB.54.1703
  62. 62
    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
  63. 63
    Hutter, J.; Iannuzzi, M.; Schiffmann, F.; VandeVondele, J. Atomistic Simulations of Condensed Matter Systems WIREs 2014, 4, 15 25 DOI: 10.1002/wcms.1159
  64. 64
    Laio, A.; Gervasio, F. L. Metadynamics: a Method to Simulate Rare Events and Reconstruct the Free Energy in Biophysics, Chemistry and Material Science Rep. Prog. Phys. 2008, 71, 126601 DOI: 10.1088/0034-4885/71/12/126601
  65. 65
    Laio, A.; Rodriguez-Fortea, A.; Gervasio, F. L.; Ceccarelli, M.; Parrinello, M. Assessing the Accuracy of Metadynamics J. Phys. Chem. B 2005, 109, 6714 6721 DOI: 10.1021/jp045424k
  66. 66
    Vandevondele, J.; Krack, M.; Mohamed, F.; Parrinello, M.; Chassaing, T.; Hutter, J. r. Quickstep: Fast and Accurate Density Functional Calculations Using a Mixed Gaussian and Plane Waves Approach Comput. Phys. Commun. 2005, 167, 103 128 DOI: 10.1016/j.cpc.2004.12.014
  67. 67
    Krack, M.; Parrinello, M. In QUICKSTEP: Make the Atoms Dance; NIC Series; Forschungszentrum Jülich, 2004; p 29.
  68. 68
    Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual Molecular Dynamics J. Mol. Graphics 1996, 14, 33 38 DOI: 10.1016/0263-7855(96)00018-5
  69. 69
    Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Gaussian 09; Gaussian, Inc.: Wallingford, CT, 2009.
  70. 70
    Hehre, W. J.; Ditchfield, R.; Pople, J. A. Self—Consistent Molecular Orbital Methods. XII. Further Extensions of Gaussian—Type Basis Sets for Use in Molecular Orbital Studies of Organic Molecules J. Chem. Phys. 1972, 56, 2257 2261 DOI: 10.1063/1.1677527
  71. 71
    Clark, T.; Chandrasekhar, J.; Spitznagel, G. W.; Schleyer, P. V. R. Efficient Diffuse Function-Augmented Basis Sets for Anion Calculations. III. The 3-21+G Basis Set for First-Row Elements, Li–F J. Comput. Chem. 1983, 4, 294 301 DOI: 10.1002/jcc.540040303
  72. 72
    Frisch, M. J.; Pople, J. A.; Binkley, J. S. Self-Consistent Molecular Orbital Methods 25. Supplementary Functions for Gaussian Basis Sets J. Chem. Phys. 1984, 80, 3265 DOI: 10.1063/1.447079
  73. 73
    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, 6378 6396 DOI: 10.1021/jp810292n
  74. 74
    (a) Glendening, E. D.; Landis, C. R.; Weinhold, F. NBO 6.0: Natural Bond Orbital Analysis Program J. Comput. Chem. 2013, 34, 1429 1437 DOI: 10.1002/jcc.23266
    (b) Glendening, E. D., Jr.; Badenhoop, K.; Reed, A. E.; Carpenter, J. E.; Bohmann, J. A.; Morales, C. M.; Landis, C. R.; Weinhold, F.NBO 6.0; Theoretical Chemistry Institute, University of Wisconsin, Madison, WI, 2013.
  75. 75
    Weinhold, F.; Landis, C. R. Discovering Chemistry with Natural Bond Orbitals; Wiley, 2012.
  76. 76
    Silverman, G. S.; Rakita, P. E. Handbook of Grignard Reagents; CRC Press: New York, 1996.
  77. 77
    Vestergren, M.; Eriksson, J.; Håkansson, M. Absolute Asymmetric Synthesis of “Chiral-at-Metal” Grignard Reagents and Transfer of the Chirality to Carbon Chem. - Eur. J. 2003, 9, 4678 4686 DOI: 10.1002/chem.200305003
  78. 78
    Vestergren, M.; Eriksson, J.; Håkansson, M. Chiral cis-Octahedral Grignard reagents J. Organomet. Chem. 2003, 681, 215 224 DOI: 10.1016/S0022-328X(03)00616-8
  79. 79
    Vestergren, M.; Gustafsson, B.; Davidsson, Ö.; Håkansson, M. Octahedral Grignard Reagents Can Be Chiral at Magnesium Angew. Chem., Int. Ed. 2000, 39, 3435 3437 DOI: 10.1002/1521-3773(20001002)39:19<3435::AID-ANIE3435>3.0.CO;2-A
  80. 80
    Harrison-Marchand, A.; Mongin, F. Mixed AggregAte (MAA): A Single Concept for All Dipolar Organometallic Aggregates. 1. Structural Data Chem. Rev. 2013, 113, 7470 7562 DOI: 10.1021/cr300295w
  81. 81
    Yamazaki, S.; Yamabe, S. A Computational Study on Addition of Grignard Reagents to Carbonyl Compounds J. Org. Chem. 2002, 67, 9346 9353 DOI: 10.1021/jo026017c
  82. 82
    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, 6158 6165 DOI: 10.1021/jp9009788
  83. 83
    Hölzer, B.; Hoffmann, R. W. Kumada-Corriu Coupling of Grignard reagents, Probed with a Chiral Grignard Reagent Chem. Commun. 2003, 2, 732 733 DOI: 10.1039/b300033h
  84. 84
    King, A. O.; Okukado, N.; Negishi, E.-i. Highly General Stereo-, Regio-, and Chemo-Selective Synthesis of Terminal and Internal Conjugated Enynes by the Pd-Catalysed Reaction of Alkynylzinc Reagents with Alkenyl Halides J. Chem. Soc., Chem. Commun. 1977, 683 684 DOI: 10.1039/c39770000683
  85. 85
    Negishi, E.-i. Palladium- or Nickel-Catalyzed Cross Coupling. A New Selective Method for Carbon-Carbon Bond Formation Acc. Chem. Res. 1982, 15, 340 348 DOI: 10.1021/ar00083a001
  86. 86
    García-Melchor, M.; Fuentes, B.; Lledós, A.; Casares, J. A.; Ujaque, G.; Espinet, P. Cationic Intermediates in the Pd-Catalyzed Negishi Coupling. Kinetic and Density Functional Theory Study of Alternative Transmetalation Pathways in the Me–Me Coupling of ZnMe2 and trans-[PdMeCl(PMePh2)2] J. Am. Chem. Soc. 2011, 133, 13519 13526 DOI: 10.1021/ja204256x

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 72 publications.

  1. Evan A. Patrick, Jeremy D. Erickson, R. Morris Bullock, Ba L. Tran. Synthesis and Structural Investigation of Rigid Naphthyridine-Bis(carbene) for Trigonal Planar Coordination of Coinage Metals. Organometallics 2025, 44 (2) , 373-384. https://doi.org/10.1021/acs.organomet.4c00383
  2. Kaidi Li, Baptiste Leforestier, Amalia I. Poblador-Bahamonde, Céline Besnard, Laure Guénée, Svetlana Kucher, Clément Mazet. Ni-Catalyzed Enantioconvergent Kumada–Corriu Cross-Coupling between β-Bromostyrenes and Secondary Grignard Reagents: Reaction Development, Scope and Mechanistic Investigations. ACS Catalysis 2025, 15 (1) , 392-402. https://doi.org/10.1021/acscatal.4c06360
  3. Lucrezia Margherita Comparini, Giulio Gallorini, Lucilla Favero, Francesca Sardelli, Valeria Di Bussolo, Sebastiano Di Pietro, Mauro Pineschi. Ring-Opening Reactions of Imidazolidines and Hexahydropyrimidines with Grignard Reagents. The Journal of Organic Chemistry 2024, 89 (21) , 15652-15664. https://doi.org/10.1021/acs.joc.4c01769
  4. Russell F. Algera, Sergei Tcyrulnikov, Giselle P. Reyes. Mechanism-Based Regiocontrol in SNAr: A Case Study of Ortho-Selective Etherification with Chloromagnesium Alkoxides. Journal of the American Chemical Society 2024, 146 (41) , 28095-28109. https://doi.org/10.1021/jacs.4c07427
  5. David A. Vaccaro, Ran Yan, Gerard Parkin. Reactivity of [TismPriBenz]MgH and [TismPriBenz]MgMe towards Carbonyl Compounds: Access to Terminal Alkoxide and Enolate Complexes. Organometallics 2024, 43 (7) , 746-763. https://doi.org/10.1021/acs.organomet.3c00541
  6. Joseph Q. Nguyen, Olaf Nachtigall, Brynn N. Turpin, Joseph W. Ziller, William J. Evans. Phenylsilyl-Substituted Tetramethylcyclopentadienide Uranium Chemistry: Magnesocenes vs Grignard Reagents. Organometallics 2023, 42 (24) , 3474-3482. https://doi.org/10.1021/acs.organomet.3c00420
  7. 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
  8. 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
  9. 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
  10. 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
  11. Weidong Tong, George Zhou, Jacob H. Waldman. Real-Time and In Situ Monitoring of Transmetalation of Grignard with Manganese(II) Chloride by Raman Spectroscopy. Organic Process Research & Development 2022, 26 (4) , 1184-1190. https://doi.org/10.1021/acs.oprd.1c00446
  12. Kerry E. Jones, Bohyun Park, Nicolle A. Doering, Mu-Hyun Baik, Richmond Sarpong. Rearrangements of the Chrysanthenol Core: Application to a Formal Synthesis of Xishacorene B. Journal of the American Chemical Society 2021, 143 (48) , 20482-20490. https://doi.org/10.1021/jacs.1c10804
  13. 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
  14. 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
  15. 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
  16. 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
  17. 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
  18. Elçin Içten, Andrew J. Maloney, Matthew G. Beaver, Xiaoxiang Zhu, Dongying E. Shen, Jo Anna Robinson, Andrew T. Parsons, Ayman Allian, Seth Huggins, Roger Hart, Pablo Rolandi, Shawn D. Walker, Richard D. Braatz. A Virtual Plant for Integrated Continuous Manufacturing of a Carfilzomib Drug Substance Intermediate, Part 2: Enone Synthesis via a Barbier-Type Grignard Process. Organic Process Research & Development 2020, 24 (10) , 1876-1890. https://doi.org/10.1021/acs.oprd.0c00188
  19. Ethan R. Curtis, Matthew D. Hannigan, Andrew K. Vitek, Paul M. Zimmerman. Quantum Chemical Investigation of Dimerization in the Schlenk Equilibrium of Thiophene Grignard Reagents. The Journal of Physical Chemistry A 2020, 124 (8) , 1480-1488. https://doi.org/10.1021/acs.jpca.9b09985
  20. Raphael Mathias Peltzer, Jürgen Gauss, Odile Eisenstein, Michele Cascella. The Grignard Reaction – Unraveling a Chemical Puzzle. Journal of the American Chemical Society 2020, 142 (6) , 2984-2994. https://doi.org/10.1021/jacs.9b11829
  21. Adam A. Pollit, Shuyang Ye, Dwight S. Seferos. Elucidating the Role of Catalyst Steric and Electronic Effects in Controlling the Synthesis of π-Conjugated Polymers. Macromolecules 2020, 53 (1) , 138-148. https://doi.org/10.1021/acs.macromol.9b02098
  22. Lavrenty G. Gutsev, Gennady L. Gutsev, Katharine Moore Tibbetts, Puru Jena. Homocoupling and Heterocoupling of Grignard Perfluorobenzene Reagents via Aryne Intermediates: A DFT Study. The Journal of Physical Chemistry A 2019, 123 (45) , 9693-9700. https://doi.org/10.1021/acs.jpca.9b05623
  23. Philippe Bertus. From Dialkyltitanium Species to Titanacyclopropanes: An Ab Initio Study. Organometallics 2019, 38 (21) , 4171-4182. https://doi.org/10.1021/acs.organomet.9b00509
  24. Jeremy N. Harvey, Fahmi Himo, Feliu Maseras, Lionel Perrin. Scope and Challenge of Computational Methods for Studying Mechanism and Reactivity in Homogeneous Catalysis. ACS Catalysis 2019, 9 (8) , 6803-6813. https://doi.org/10.1021/acscatal.9b01537
  25. Joseane A. Mendes, Pedro Merino, Tatiana Soler, Eduardo J. Salustiano, Paulo R. R. Costa, Miguel Yus, Francisco Foubelo, Camilla D. Buarque. Enantioselective Synthesis, DFT Calculations, and Preliminary Antineoplastic Activity of Dibenzo 1-Azaspiro[4.5]decanes on Drug-Resistant Leukemias. The Journal of Organic Chemistry 2019, 84 (4) , 2219-2233. https://doi.org/10.1021/acs.joc.8b03203
  26. Patrick Kielty, Dennis A. Smith, Peter Cannon, Michael P. Carty, Michael Kennedy, Patrick McArdle, Richard J. Singer, Fawaz Aldabbagh. Selective Methylmagnesium Chloride Mediated Acetylations of Isosorbide: A Route to Powerful Nitric Oxide Donor Furoxans. Organic Letters 2018, 20 (10) , 3025-3029. https://doi.org/10.1021/acs.orglett.8b01060
  27. Juan del Pozo, María Pérez-Iglesias, Rosana Álvarez, Agustí Lledós, Juan A. Casares, Pablo Espinet. Speciation of ZnMe2, ZnMeCl, and ZnCl2 in Tetrahydrofuran (THF), and Its Influence on Mechanism Calculations of Catalytic Processes. ACS Catalysis 2017, 7 (5) , 3575-3583. https://doi.org/10.1021/acscatal.6b03636
  28. Marcos A. Loroño-González, Daniel J. Loroño-González. Magnesium and potassium scorpionate complexes based on dihydrobis(pyrazolyl)borate. Acta Crystallographica Section C Structural Chemistry 2025, 81 (3) , 131-139. https://doi.org/10.1107/S2053229625000750
  29. Annabel Rae, Alan R. Kennedy, Stuart D. Robertson. Magnesium 4, 5, and 6 coordinate complexes with ligands bound via sp or sp2 hybridized atoms. Polyhedron 2025, 266 , 117257. https://doi.org/10.1016/j.poly.2024.117257
  30. Lorenzo Restaino, Riccardo Mincigrucci, Markus Kowalewski. Distinguishing Organomagnesium Species in the Grignard Addition to Ketones with X‐Ray Spectroscopy. Chemistry – A European Journal 2024, 30 (70) https://doi.org/10.1002/chem.202402099
  31. Marinella de Giovanetti, Sondre Hilmar Hopen Eliasson, Sigbjørn Løland Bore, Odile Eisenstein, Michele Cascella. Morphology of lithium halides in tetrahydrofuran from molecular dynamics with machine learning potentials. Chemical Science 2024, 15 (48) , 20355-20364. https://doi.org/10.1039/D4SC04957H
  32. Erika Mooney, Matthias Tacke, Helge Müller-Bunz, Julia Bruno-Colmenárez, Gordon Cooke, Emma Caraher, Fintan Kelleher, Bernadette S. Creaven. Hybrid silver(I) coumarin-carbene and coumarin-triphenylphosphine complexes: Towards more effective antimicrobial therapies. Inorganica Chimica Acta 2024, 572 , 122222. https://doi.org/10.1016/j.ica.2024.122222
  33. Manabu Hatano, Kisara Kuwano, Riho Asukai, Ayako Nagayoshi, Haruka Hoshihara, Tsubasa Hirata, Miho Umezawa, Sahori Tsubaki, Takeshi Yoshikawa, Ken Sakata. Zinc chloride-catalyzed Grignard addition reaction of aromatic nitriles. Chemical Science 2024, 15 (22) , 8569-8577. https://doi.org/10.1039/D4SC01659A
  34. Aurélien Alix. Diastereoselective Transformation Using Group 2 and 13 Metal Salts. 2024, 357-431. https://doi.org/10.1016/B978-0-32-390644-9.00134-7
  35. Odile Eisenstein. Nucleophilic addition to carbonyl groups from qualitative to quantitative computational studies. A historical perspective. Comptes Rendus. Chimie 2024, 27 (S2) , 5-19. https://doi.org/10.5802/crchim.298
  36. 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
  37. 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
  38. Etienne V. Brouillet, Scott A. Brown, Alan R. Kennedy, Annabel Rae, Heather P. Walton, Stuart D. Robertson. Atom-economic access to cationic magnesium complexes. Dalton Transactions 2023, 52 (37) , 13332-13338. https://doi.org/10.1039/D3DT02669H
  39. 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
  40. Meng‐Yang Chang, Kuan‐Ting Chen, Hsing‐Yin Chen. Grignard Reagent‐Mediated Regioselective 1,8‐Addition of α‐(2‐Thienylidene)‐β‐ketosulfones. Advanced Synthesis & Catalysis 2023, 365 (13) , 2264-2270. https://doi.org/10.1002/adsc.202300227
  41. 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
  42. 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
  43. Lucas Loir-Mongazon, Carmen Antuña-Hörlein, Christophe Deraedt, Yann Cornaton, Jean-Pierre Djukic. Activation Barriers for Cobalt(IV)-Centered Reductive Elimination Correlate with Quantified Interatomic Noncovalent Interactions. Synlett 2023, 34 (10) , 1169-1173. https://doi.org/10.1055/a-1937-9296
  44. 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
  45. 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
  46. 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
  47. Aude Salamé, Jordan Rio, Ilaria Ciofini, Lionel Perrin, Laurence Grimaud, Pierre-Adrien Payard. Copper-Catalyzed Homocoupling of Boronic Acids: A Focus on B-to-Cu and Cu-to-Cu Transmetalations. Molecules 2022, 27 (21) , 7517. https://doi.org/10.3390/molecules27217517
  48. 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
  49. Lucas A. Freeman, Jacob E. Walley, Robert J. Gilliard. Synthesis and reactivity of low-oxidation-state alkaline earth metal complexes. Nature Synthesis 2022, 1 (6) , 439-448. https://doi.org/10.1038/s44160-022-00077-6
  50. 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
  51. 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
  52. 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
  53. Michael S. Hill, Anne‐Frédérique Pécharman, Andrew S. S. Wilson. Turbo Charging Group 2 Reagents for Metathesis, Metalation, and Catalysis. 2022, 97-131. https://doi.org/10.1002/9781119448877.ch3
  54. 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
  55. 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
  56. 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
  57. 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
  58. S. Chantal E. Stieber. Computational Methods in Organometallic Chemistry. 2022, 176-210. https://doi.org/10.1016/B978-0-12-820206-7.00099-8
  59. Rengui Weng, Xuebin Lu, Na Ji, Atsushi Fukuoka, Abhijit Shrotri, Xiaoyun Li, Rui Zhang, Ming Zhang, Jian Xiong, Zhihao Yu. Taming the butterfly effect: modulating catalyst nanostructures for better selectivity control of the catalytic hydrogenation of biomass-derived furan platform chemicals. Catalysis Science & Technology 2021, 11 (24) , 7785-7806. https://doi.org/10.1039/D1CY01708J
  60. Kieren J. Evans, Paul A. Morton, Calum Sangster, Stephen M. Mansell. One-step synthesis of heteroleptic rare-earth amide complexes featuring fluorenyl-tethered N-heterocyclic carbene ligands. Polyhedron 2021, 197 , 115021. https://doi.org/10.1016/j.poly.2021.115021
  61. 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
  62. . Miscellaneous Reactions. 2021, 161-291. https://doi.org/10.1002/9783527828166.ch4
  63. Sonia Bajo, Macarena G. Alférez, María M. Alcaide, Joaquín López‐Serrano, Jesús Campos. Metal‐only Lewis Pairs of Rhodium with s , p and d ‐Block Metals. Chemistry – A European Journal 2020, 26 (70) , 16833-16845. https://doi.org/10.1002/chem.202003167
  64. 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
  65. Lingyi Shen, Yanxia Zhao, Dihua Dai, Ying-Wei Yang, Biao Wu, Xiao-Juan Yang. Stabilization of Grignard reagents by a pillar[5]arene host – Schlenk equilibria and Grignard reactions. Chemical Communications 2020, 56 (9) , 1381-1384. https://doi.org/10.1039/C9CC08728A
  66. Sjoerd Harder. Introduction to Early Main Group Organometallic Chemistry and Catalysis. 2020, 1-29. https://doi.org/10.1002/9783527818020.ch1
  67. Odile Eisenstein, Gregori Ujaque, Agustí Lledós. What Makes a Good (Computed) Energy Profile?. 2020, 1-38. https://doi.org/10.1007/3418_2020_57
  68. Odile Eisenstein. Concluding remarks for “Mechanistic Processes in Organometallic Chemistry”: the importance of a multidisciplinary approach. Faraday Discussions 2019, 220 , 489-495. https://doi.org/10.1039/C9FD00101H
  69. Clara Aupic, Amel Abdou Mohamed, Carlotta Figliola, Paola Nava, Béatrice Tuccio, Gaëlle Chouraqui, Jean-Luc Parrain, Olivier Chuzel. Highly diastereoselective preparation of chiral NHC-boranes stereogenic at the boron atom. Chemical Science 2019, 10 (26) , 6524-6530. https://doi.org/10.1039/C9SC01454C
  70. Rachel D. Davidson, Yenny Cubides, Justin L. Andrews, Chelsea M. McLain, Homero Castaneda, Sarbajit Banerjee. Magnesium Nanocomposite Coatings for Protection of a Lightweight Al Alloy: Modes of Corrosion Protection, Mechanisms of Failure. physica status solidi (a) 2019, 216 (13) https://doi.org/10.1002/pssa.201800817
  71. Satish K. Nune, David B. Lao, Mark E. Bowden, Herbert T. Schaef, Rama S. Vemuri, Radha Kishan Motkuri, B. Peter McGrail. Investigation of reactive intermediates during the synthesis of di-n-butylmagnesium. Inorganica Chimica Acta 2019, 489 , 150-154. https://doi.org/10.1016/j.ica.2019.01.035
  72. Amalia I. Poblador Bahamonde, Stéphanie Halbert. Computational Study of the Cu‐Free Allylic Alkylation Mechanism with Grignard Reagents: Role of the NHC Ligand. European Journal of Organic Chemistry 2017, 2017 (39) , 5935-5941. https://doi.org/10.1002/ejoc.201701010

The Journal of Physical Chemistry B

Cite this: J. Phys. Chem. B 2017, 121, 16, 4226–4237
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.jpcb.7b02716
Published March 30, 2017

Copyright © 2017 American Chemical Society. This publication is licensed under these Terms of Use.

Article Views

19k

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Scheme 1

    Scheme 1. Schlenk Equilibria

    Figure 1

    Figure 1. Simulation box used in this study. The atoms of the Grignard reagent are represented by spheres and the THF solvent molecules by sticks in licorice and red. Hydrogen atoms of THF are not shown for clarity.

    Figure 2

    Figure 2. Most likely solvation structures for MgCl2 (left), Mg(CH3)2 (middle), and (CH3)MgCl (right). The dotted black lines represent coordination of the ligands to the Mg center.

    Figure 3

    Figure 3. FES of the Schlenk equilibrium. The CVs for this representation are the difference in Mg–CH3 coordination number between Mg2 and Mg1 (CV1) and the THF coordination number to Mg1 (CV2). The chemical structures drawn in the figure depict the most representative species obtained for wells AE.

    Figure 4

    Figure 4. Solvated DClCl structures found by metadynamics simulations (left) and the corresponding FES (right). CV1 and CV2 are defined as the coordination numbers of THF at Mg1 and Mg2, respectively, following eq 1. The two minima b correspond to the chemically equivalent D12ClCl and D21ClCl structures. Only D12ClCl is shown for simplicity.

    Figure 5

    Figure 5. (Top) Mg–Cl distance distributions in DClCl dimeric structures. (Bottom) Mg1–Cl (blue) and Mg2–Cl (red) distance distributions in D12ClCl. Mg–Cl bond cleavage is observed when the Mg–Cl distance is larger than 3.7 Å.

    Figure 6

    Figure 6. FES for the methyl bridged dimer DClMe equilibria using the THF coordination number to Mg1 (CV1) and the THF coordination number to Mg2 (CV2) as variables, together with the most representative species obtained for wells ad.

    Figure 7

    Figure 7. Orientation of the methyl group in DClMe as a function of the solvation state, represented by φ1 and φ2. A larger φ angle is indicative of a stronger Mg–CH3 interaction.

    Figure 8

    Figure 8. Intermediates involved in the Schlenk equilibrium according to dynamic simulations. Arrows indicate the chemical transformations along the main pathway leading from monomeric reactants to products (inside of solid squares). The most stable dichloride and methyl chloride bridged dinuclear species are inside of dashed squares.

    Figure 9

    Figure 9. Snapshots for the methyl transfer reaction in D12ClCl (Mg1 on the left-hand side and Mg2 on the right for all snapshots): (1) initial D12ClCl structure, (2) transition state of the transmetalation reaction, (3) D12ClMe, (4) solvent loss to form D11ClMe, and (5) solvent addition to form D21ClMe and (6) D11ClMeTHF. The atoms for the Grignard reagent and the coordinating THF molecules are depicted as balls and/or sticks and colored according to standard color codes. Selected neighboring solvent molecules are drawn with thin lines.

  • References


    This article references 86 other publications.

    1. 1
      Grignard, V. Sur Quelques Nouvelles Combinaisons Organométalliques du Magnésium et Leur Application à des Synthèses d’Alcools et d’Hydrocabures C. R. Acad. Sci. 1900, 130, 1322 1324
    2. 2
      Grignard, V. The Use of Organomagnesium Compounds in Preparative Organic Chemistry–Nobel Lecture 1912 Nobel Lectures Chemistry 1921, 1966, 234 246
    3. 3
      Corriu, R. J. P.; Massé, J. P. Activation of Grignard Reagents by Transition-Metal Complexes. A New and Simple Synthesis of Trans-Stilbenes and Polyphenyls J. Chem. Soc., Chem. Commun. 1972, 144a 144a DOI: 10.1039/c3972000144a
    4. 4
      Tamao, K.; Sumitani, K.; Kumada, M. Selective Carbon-Carbon Bond Formation by Cross-Coupling of Grignard Reagents with Organic Halides. Catalysis by Nickel-Phosphine Complexes J. Am. Chem. Soc. 1972, 94, 4374 4376 DOI: 10.1021/ja00767a075
    5. 5
      Fürstner, A.; Leitner, A.; Méndez, M.; Krause, H. Iron-Catalyzed Cross-Coupling Reactions J. Am. Chem. Soc. 2002, 124, 13856 13863 DOI: 10.1021/ja027190t
    6. 6
      Frisch, A. C.; Beller, M. Catalysts for Cross-Coupling Reactions with Non-activated Alkyl Halides Angew. Chem., Int. Ed. 2005, 44, 674 688 DOI: 10.1002/anie.200461432
    7. 7
      Terao, J.; Kato, Y.; Kambe, N. Titanocene-Catalyzed Regioselective Alkylation of Styrenes with Grignard Reagents Using β-Bromoethyl Ethers, Thioethers, or Amines Chem. - Asian J. 2008, 3, 1472 1478 DOI: 10.1002/asia.200800134
    8. 8
      Vechorkin, O.; Barmaz, D.; Proust, V.; Hu, X. Ni-Catalyzed Sonogashira Coupling of Nonactivated Alkyl Halides: Orthogonal Functionalization of Alkyl Iodides, Bromides, and Chlorides J. Am. Chem. Soc. 2009, 131, 12078 12079 DOI: 10.1021/ja906040t
    9. 9
      Adrio, J.; Carretero, J. C. Functionalized Grignard Reagents in Kumada Cross-Coupling Reactions ChemCatChem 2010, 2, 1384 1386 DOI: 10.1002/cctc.201000237
    10. 10
      Jana, R.; Pathak, T. P.; Sigman, M. S. Advances in Transition Metal (Pd,Ni,Fe)-Catalyzed Cross-Coupling Reactions Using Alkyl-organometallics as Reaction Partners Chem. Rev. 2011, 111, 1417 1492 DOI: 10.1021/cr100327p
    11. 11
      Cong, X.; Tang, H.; Zeng, X. Regio- and Chemoselective Kumada–Tamao–Corriu Reaction of Aryl Alkyl Ethers Catalyzed by Chromium Under Mild Conditions J. Am. Chem. Soc. 2015, 137, 14367 14372 DOI: 10.1021/jacs.5b08621
    12. 12
      Neufeld, R.; Teuteberg, T. L.; Herbst-Irmer, R.; Mata, R. A.; Stalke, D. Solution Structures of Hauser Base iPr2NMgCl and Turbo-Hauser Base iPr2NMgCl·LiCl in THF and the Influence of LiCl on the Schlenk-Equilibrium J. Am. Chem. Soc. 2016, 138, 4796 4806 DOI: 10.1021/jacs.6b00345
    13. 13
      Seyferth, D. The grignard reagents Organometallics 2009, 28, 1598 1605 DOI: 10.1021/om900088z
    14. 14
      Guggenberger, L. J.; Rundle, R. E. Crystal Structure of the Ethyl Grignard Reagent, Ethylmagnesium Bromide Dietherate J. Am. Chem. Soc. 1968, 90, 5375 5378 DOI: 10.1021/ja01022a007
    15. 15
      Vallino, M. Structure Cristalline de CH3MgBr·3 C4H8O J. Organomet. Chem. 1969, 20, 1 10 DOI: 10.1016/S0022-328X(00)80080-7
    16. 16
      Toney, J.; Stucky, G. D. The Stereochemistry of Polynuclear Compounds of the Main Group Elements [C2H5Mg2Cl3(C4H8O)3]2, a Tetrameric Grignard Reagent J. Organomet. Chem. 1971, 28, 5 20 DOI: 10.1016/S0022-328X(00)81569-7
    17. 17
      Blasberg, F.; Bolte, M.; Wagner, M.; Lerner, H.-W. An Approach to Pin Down the Solid-State Structure of the “Turbo Grignard Organometallics 2012, 31, 1001 1005 DOI: 10.1021/om201080t
    18. 18
      Smith, M. B.; Becker, W. E. The constitution of the grignard reagent—III:The reaction between R2Mg and MgX2 in tetrahydrofuran Tetrahedron 1967, 23, 4215 4227 DOI: 10.1016/S0040-4020(01)88819-0
    19. 19
      Smith, M. B.; Becker, W. E. The constitution of the Grignard Reagent - I. The reaction between diethyl magnesium and magnesium bromide in diethyl ether Tetrahedron Lett. 1965, 6, 3843 3847 DOI: 10.1016/S0040-4039(01)89135-8
    20. 20
      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, 3162 3171 DOI: 10.1021/ja00844a040
    21. 21
      Schlenk, W.; Schlenk, W. Über die Konstitution der Grignardschen Magnesiumverbindungen Ber. Dtsch. Chem. Ges. B 1929, 62, 920 924 DOI: 10.1002/cber.19290620422
    22. 22
      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, 3845 3850 DOI: 10.1021/ja01042a027
    23. 23
      Sobota, P.; Duda, B. Influence of MgCl2 on Grignard Reagent Composition in Tetrahydrofuran. III J. Organomet. Chem. 1987, 332, 239 245 DOI: 10.1016/0022-328X(87)85090-8
    24. 24
      Sakamoto, S.; Imamoto, T.; Yamaguchi, K. Constitution of Grignard Reagent RMgCl in Tetrahydrofuran Org. Lett. 2001, 3, 1793 1795 DOI: 10.1021/ol010048x
    25. 25
      Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas Phys. Rev. B 1964, 136, 864 871 DOI: 10.1103/PhysRev.136.B864
    26. 26
      Kohn, W.; Sham, L. J. Self-Consistent Equations Including Exchange and Correlation Effects Phys. Rev. A 1965, 140, 1133 1138 DOI: 10.1103/PhysRev.140.A1133
    27. 27
      Jiménez-Halla, J. O. C.; Bickelhaupt, F. M.; Solà, M. Organomagnesium clusters: Structure, stability, and bonding in archetypal models J. Organomet. Chem. 2011, 696, 4104 4111 DOI: 10.1016/j.jorganchem.2011.06.014
    28. 28
      Lioe, H.; White, J. M.; O’Hair, R. A. J. Preference for bridging versus terminal ligands in magnesium dimers J. Mol. Model. 2011, 17, 1325 1334 DOI: 10.1007/s00894-010-0834-1
    29. 29
      Henriques, A. M.; Barbosa, A. G. H. Chemical Bonding and the Equilibrium Composition of Grignard Reagents in Ethereal Solutions J. Phys. Chem. A 2011, 115, 12259 12270 DOI: 10.1021/jp202762p
    30. 30
      Ramirez, F.; Sarma, R.; Chaw, F.; McCaffrey, T. M. Magnesium bromide-tetrahydrofuran complexes: bis(tetrahydrofuran)magnesium bromide, tris(tetrahydrofuran)magnesium bromide, tetrakis(tetrahydrofuran)magnesium bromide, and diaquotetrakis(tetrahydrofuran)magnesium bromide. A reagent for the preparation of anhydrous magnesium phosphodiester salts J. Am. Chem. Soc. 1977, 99, 5285 5289 DOI: 10.1021/ja00458a010
    31. 31
      Pirinen, S.; Koshevoy, I. O.; Denifl, P.; Pakkanen, T. T. A Single-Crystal Model for MgCl2 – Electron Donor Support Materials: [Mg3Cl5(THF)4Bu]2 (Bu = n-Butyl) Organometallics 2013, 32, 4208 4213 DOI: 10.1021/om400407p
    32. 32
      Ashby, E. C.; Becker, W. E. Concerning the Structure of the Grignard Reagent J. Am. Chem. Soc. 1963, 85, 118 119 DOI: 10.1021/ja00884a032
    33. 33
      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, 133 139 DOI: 10.1021/jp035653r
    34. 34
      Tobisu, M.; Chatani, N. Cross-Couplings Using Aryl Ethers via C–O Bond Activation Enabled by Nickel Catalysts Acc. Chem. Res. 2015, 48, 1717 1726 DOI: 10.1021/acs.accounts.5b00051
    35. 35
      Cahiez, G.; Moyeux, A.; Cossy, J. Grignard Reagents and Non-Precious Metals: Cheap and Eco-Friendly Reagents for Developing Industrial Cross-Couplings. A Personal Account Adv. Synth. Catal. 2015, 357, 1983 1989 DOI: 10.1002/adsc.201400654
    36. 36
      Tasker, S. Z.; Standley, E. A.; Jamison, T. F. Recent Advances in Homogeneous Nickel Catalysis Nature 2014, 509, 299 309 DOI: 10.1038/nature13274
    37. 37
      Laio, A.; Parrinello, M. Escaping Free-Energy Minima Proc. Natl. Acad. Sci. U. S. A. 2002, 99, 12562 12566 DOI: 10.1073/pnas.202427399
    38. 38
      Iannuzzi, M.; Laio, A.; Parrinello, M. Efficient Exploration of Reactive Potential Energy Surfaces Using Car-Parrinello Molecular Dynamics Phys. Rev. Lett. 2003, 90, 23 26 DOI: 10.1103/PhysRevLett.90.238302
    39. 39
      Vuilleumier, R.; Sprik, M. Electronic Properties of Hard and Soft Ions in Solution: Aqueous Na+ and Ag+ Compared J. Chem. Phys. 2001, 115, 3454 3468 DOI: 10.1063/1.1388901
    40. 40
      Lightstone, F. C.; Schwegler, E.; Hood, R. Q.; Gygi, F.; Galli, G. A First Principles Molecular Dynamics Simulation of the Hydrated Magnesium Ion Chem. Phys. Lett. 2001, 343, 549 555 DOI: 10.1016/S0009-2614(01)00735-7
    41. 41
      Bernasconi, L.; Baerends, E. J.; Sprik, M. Long-Range Solvent Effects on the Orbital Interaction Mechanism of Water Acidity Enhancement in Metal Ion Solutions: A Comparative Study of the Electronic Structure of Aqueous Mg and Zn Dications J. Phys. Chem. B 2006, 110, 11444 11453 DOI: 10.1021/jp0609941
    42. 42
      Blumberger, J.; Bernasconi, L.; Tavernelli, I.; Vuilleumier, R.; Sprik, M. Electronic Structure and Solvation of Copper and Silver Ions: A Theoretical Picture of a Model Aqueous Redox Reaction J. Am. Chem. Soc. 2004, 126, 3928 3938 DOI: 10.1021/ja0390754
    43. 43
      Guido, C. A.; Pietrucci, F.; Gallet, G. A.; Andreoni, W. The Fate of a Zwitterion in Water from Ab Initio Molecular Dynamics: Monoethanolamine (MEA)-CO2 J. Chem. Theory Comput. 2013, 9, 28 32 DOI: 10.1021/ct301071b
    44. 44
      Boero, M.; Ikeshoji, T.; Liew, C. C.; Terakura, K.; Parrinello, M.; Boero, M.; Ikeshoji, T.; Liew, C. C.; Terakura, K. Hydrogen Bond Driven Chemical Reactions: Beckmann Rearrangement of Cyclohexanone Oxime into ε-Caprolactam in Supercritical Water Hydrogen Bond Driven Chemical Reactions: Beckmann Rearrangement of Cyclohexanone Oxime into E-Caprolactam in Supercritical J. Am. Chem. Soc. 2004, 126, 6280 6286 DOI: 10.1021/ja049363f
    45. 45
      Vidossich, P.; Lledós, A.; Ujaque, G. Realistic Simulation of Organometallic Reactivities in Solution by Means of First-Principles Molecular Dynamics. In Computational Studies in Organometallic Chemistry; Macgregor, S. A.; Eisenstein, O., Eds.; Structure and Bonding; Springer International Publishing: Berlin, Germany, 2016; Vol. 167, pp 81 106.
    46. 46
      Vidossich, P.; Lledós, A.; Ujaque, G. First-Principles Molecular Dynamics Studies of Organometallic Complexes and Homogeneous Catalytic Processes Acc. Chem. Res. 2016, 49, 1271 1278 DOI: 10.1021/acs.accounts.6b00054
    47. 47
      Laio, A.; VandeVondele, J.; Rothlisberger, U. A Hamiltonian Electrostatic Coupling Scheme for Hybrid Car–Parrinello Molecular Dynamics Simulations J. Chem. Phys. 2002, 116, 6941 6947 DOI: 10.1063/1.1462041
    48. 48
      Moret, M.-E.; Tavernelli, I.; Chergui, M.; Rothlisberger, U. Electron Localization Dynamics in the Triplet Excited State of [Ru(bpy)3]2+ in Aqueous Solution Chem. - Eur. J. 2010, 16, 5889 5894 DOI: 10.1002/chem.201000184
    49. 49
      Dal Peraro, M.; Llarrull, L. I.; Rothlisberger, U.; Vila, A. J.; Carloni, P. Water-Assisted Reaction Mechanism of Monozinc β-Lactamases J. Am. Chem. Soc. 2004, 126, 12661 12668 DOI: 10.1021/ja048071b
    50. 50
      Dal Peraro, M.; Vila, A. J.; Carloni, P.; Klein, M. L. Role of Zinc Content on the Catalytic Efficiency of B1Metallo β-Lactamases J. Am. Chem. Soc. 2007, 129, 2808 2816 DOI: 10.1021/ja0657556
    51. 51
      Cascella, M.; Magistrato, A.; Tavernelli, I.; Carloni, P.; Rothlisberger, U. Role of Protein Frame and Solvent for the Redox Properties of Azurin from Pseudomonas Aeruginosa Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 19641 19646 DOI: 10.1073/pnas.0607890103
    52. 52
      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, 10955 10962 DOI: 10.1021/ja8005786
    53. 53
      Gossens, C.; Tavernelli, I.; Rothlisberger, U. Rational Design of Organo-Ruthenium Anticancer Compounds Chimia 2005, 59, 81 84 DOI: 10.2533/000942905777676795
    54. 54
      Metz, D. J.; Glines, A. Density, Viscosity, and Dielectric Constant of Tetrahydrofuran between −78 and 30° J. Phys. Chem. 1967, 71, 1158 1158 DOI: 10.1021/j100863a067
    55. 55
      Hoover, W. G. Canonical Dynamics: Equilibrium Phase-Space Distributions Phys. Rev. A: At., Mol., Opt. Phys. 1985, 31, 1695 1697 DOI: 10.1103/PhysRevA.31.1695
    56. 56
      Martyna, G. J.; Klein, M. L.; Tuckerman, M. Nose–Hoover chains: The Canonical Ensemble via Continuous Dynamics J. Chem. Phys. 1992, 97, 2635 2643 DOI: 10.1063/1.463940
    57. 57
      Nosé, S. A Unified Formulation of the Constant Temperature Molecular Dynamics Methods J. Chem. Phys. 1984, 81, 511 519 DOI: 10.1063/1.447334
    58. 58
      Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple Phys. Rev. Lett. 1996, 77, 3865 3868 DOI: 10.1103/PhysRevLett.77.3865
    59. 59
      Lippert, G.; Hutter, J.; Parrinello, M. The Gaussian and Augmented-Plane-Wave Density Functional Method for Ab Initio Molecular Dynamics Simulations Theor. Chem. Acc. 1999, 103, 124 140 DOI: 10.1007/s002140050523
    60. 60
      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
    61. 61
      Goedecker, S.; Teter, M.; Hutter, J. Separable Dual-Space Gaussian Pseudopotentials Phys. Rev. B: Condens. Matter Mater. Phys. 1996, 54, 1703 1710 DOI: 10.1103/PhysRevB.54.1703
    62. 62
      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
    63. 63
      Hutter, J.; Iannuzzi, M.; Schiffmann, F.; VandeVondele, J. Atomistic Simulations of Condensed Matter Systems WIREs 2014, 4, 15 25 DOI: 10.1002/wcms.1159
    64. 64
      Laio, A.; Gervasio, F. L. Metadynamics: a Method to Simulate Rare Events and Reconstruct the Free Energy in Biophysics, Chemistry and Material Science Rep. Prog. Phys. 2008, 71, 126601 DOI: 10.1088/0034-4885/71/12/126601
    65. 65
      Laio, A.; Rodriguez-Fortea, A.; Gervasio, F. L.; Ceccarelli, M.; Parrinello, M. Assessing the Accuracy of Metadynamics J. Phys. Chem. B 2005, 109, 6714 6721 DOI: 10.1021/jp045424k
    66. 66
      Vandevondele, J.; Krack, M.; Mohamed, F.; Parrinello, M.; Chassaing, T.; Hutter, J. r. Quickstep: Fast and Accurate Density Functional Calculations Using a Mixed Gaussian and Plane Waves Approach Comput. Phys. Commun. 2005, 167, 103 128 DOI: 10.1016/j.cpc.2004.12.014
    67. 67
      Krack, M.; Parrinello, M. In QUICKSTEP: Make the Atoms Dance; NIC Series; Forschungszentrum Jülich, 2004; p 29.
    68. 68
      Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual Molecular Dynamics J. Mol. Graphics 1996, 14, 33 38 DOI: 10.1016/0263-7855(96)00018-5
    69. 69
      Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Gaussian 09; Gaussian, Inc.: Wallingford, CT, 2009.
    70. 70
      Hehre, W. J.; Ditchfield, R.; Pople, J. A. Self—Consistent Molecular Orbital Methods. XII. Further Extensions of Gaussian—Type Basis Sets for Use in Molecular Orbital Studies of Organic Molecules J. Chem. Phys. 1972, 56, 2257 2261 DOI: 10.1063/1.1677527
    71. 71
      Clark, T.; Chandrasekhar, J.; Spitznagel, G. W.; Schleyer, P. V. R. Efficient Diffuse Function-Augmented Basis Sets for Anion Calculations. III. The 3-21+G Basis Set for First-Row Elements, Li–F J. Comput. Chem. 1983, 4, 294 301 DOI: 10.1002/jcc.540040303
    72. 72
      Frisch, M. J.; Pople, J. A.; Binkley, J. S. Self-Consistent Molecular Orbital Methods 25. Supplementary Functions for Gaussian Basis Sets J. Chem. Phys. 1984, 80, 3265 DOI: 10.1063/1.447079
    73. 73
      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, 6378 6396 DOI: 10.1021/jp810292n
    74. 74
      (a) Glendening, E. D.; Landis, C. R.; Weinhold, F. NBO 6.0: Natural Bond Orbital Analysis Program J. Comput. Chem. 2013, 34, 1429 1437 DOI: 10.1002/jcc.23266
      (b) Glendening, E. D., Jr.; Badenhoop, K.; Reed, A. E.; Carpenter, J. E.; Bohmann, J. A.; Morales, C. M.; Landis, C. R.; Weinhold, F.NBO 6.0; Theoretical Chemistry Institute, University of Wisconsin, Madison, WI, 2013.
    75. 75
      Weinhold, F.; Landis, C. R. Discovering Chemistry with Natural Bond Orbitals; Wiley, 2012.
    76. 76
      Silverman, G. S.; Rakita, P. E. Handbook of Grignard Reagents; CRC Press: New York, 1996.
    77. 77
      Vestergren, M.; Eriksson, J.; Håkansson, M. Absolute Asymmetric Synthesis of “Chiral-at-Metal” Grignard Reagents and Transfer of the Chirality to Carbon Chem. - Eur. J. 2003, 9, 4678 4686 DOI: 10.1002/chem.200305003
    78. 78
      Vestergren, M.; Eriksson, J.; Håkansson, M. Chiral cis-Octahedral Grignard reagents J. Organomet. Chem. 2003, 681, 215 224 DOI: 10.1016/S0022-328X(03)00616-8
    79. 79
      Vestergren, M.; Gustafsson, B.; Davidsson, Ö.; Håkansson, M. Octahedral Grignard Reagents Can Be Chiral at Magnesium Angew. Chem., Int. Ed. 2000, 39, 3435 3437 DOI: 10.1002/1521-3773(20001002)39:19<3435::AID-ANIE3435>3.0.CO;2-A
    80. 80
      Harrison-Marchand, A.; Mongin, F. Mixed AggregAte (MAA): A Single Concept for All Dipolar Organometallic Aggregates. 1. Structural Data Chem. Rev. 2013, 113, 7470 7562 DOI: 10.1021/cr300295w
    81. 81
      Yamazaki, S.; Yamabe, S. A Computational Study on Addition of Grignard Reagents to Carbonyl Compounds J. Org. Chem. 2002, 67, 9346 9353 DOI: 10.1021/jo026017c
    82. 82
      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, 6158 6165 DOI: 10.1021/jp9009788
    83. 83
      Hölzer, B.; Hoffmann, R. W. Kumada-Corriu Coupling of Grignard reagents, Probed with a Chiral Grignard Reagent Chem. Commun. 2003, 2, 732 733 DOI: 10.1039/b300033h
    84. 84
      King, A. O.; Okukado, N.; Negishi, E.-i. Highly General Stereo-, Regio-, and Chemo-Selective Synthesis of Terminal and Internal Conjugated Enynes by the Pd-Catalysed Reaction of Alkynylzinc Reagents with Alkenyl Halides J. Chem. Soc., Chem. Commun. 1977, 683 684 DOI: 10.1039/c39770000683
    85. 85
      Negishi, E.-i. Palladium- or Nickel-Catalyzed Cross Coupling. A New Selective Method for Carbon-Carbon Bond Formation Acc. Chem. Res. 1982, 15, 340 348 DOI: 10.1021/ar00083a001
    86. 86
      García-Melchor, M.; Fuentes, B.; Lledós, A.; Casares, J. A.; Ujaque, G.; Espinet, P. Cationic Intermediates in the Pd-Catalyzed Negishi Coupling. Kinetic and Density Functional Theory Study of Alternative Transmetalation Pathways in the Me–Me Coupling of ZnMe2 and trans-[PdMeCl(PMePh2)2] J. Am. Chem. Soc. 2011, 133, 13519 13526 DOI: 10.1021/ja204256x
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpcb.7b02716.

    • Metadynamics parameters, natural bond orbital analysis, DFT optimized geometries, and solvation properties for different sizes of the simulation box (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.